Você está na página 1de 544

Mechanical Effects on Extruded Dielectric

Cables and Joints Installed in Underground


Transmission Systems in North America
Technical Report
L
I
C
E
N
S
E D
M
A T
E
R
I
A
L
WARNING:
Please read the License Agreement
on the back cover before removing
the Wrapping Material.
EPRI Project Manager
W. Zenger
EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com
Mechanical Effects on Extruded
Dielectric Cables and Joints
Installed in Underground
Transmission Systems in North
America

1001849
Final Report, March 2004




DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN
ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH
INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE
ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)
WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER
(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.
ORGANIZATION(S) THAT PREPARED THIS DOCUMENT
Cable Consulting International Ltd.










ORDERING INFORMATION
Requests for copies of this report should be directed to EPRI Orders and Conferences, 1355 Willow
Way, Suite 278, Concord, CA 94520, (800) 313-3774, press 2 or internally x5379, (925) 609-9169,
(925) 609-1310 (fax).
Electric Power Research Institute and EPRI are registered service marks of the Electric Power
Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power
Research Institute, Inc.
Copyright 2004 Electric Power Research Institute, Inc. All rights reserved.


iii
CITATIONS
This report was prepared by
Cable Consulting International Ltd.
P.O. Box 1
Sevenoaks, Kent, TN14 7EN
United Kingdom
Principal Investigators
B. Gregory
S. Galloway
This report describes research sponsored by EPRI.
The report is a corporate document that should be cited in the literature in the following manner:
Mechanical Effects on Extruded Dielectric Cables and Joints Installed in Underground
Transmission Systems in North America, EPRI, Palo Alto, CA: 2004. 1001849.





v
PRODUCT DESCRIPTION

In other parts of the world cable is directly buried or installed in tunnels; but underground cable
systems in North America have traditionally been installed in duct/manhole systems or in pipes.
In these settings, cables expand and contract during load cycling; and the resulting mechanical
stresses can have an impact on the reliable performance of the cable system. With the increased
use of extruded dielectric transmission cable systems, EPRI member utilities are faced with the
challenge of evaluating system performance under these stresses, a task made more difficult by
limited service experience and the unavailability of calculations for these complex mechanical
systems. This report provides information on the mechanical effects on extruded dielectric cables
in underground transmission systems.
Results & Findings
The thermal expansion of loaded cable generates high axial force and initiates longitudinal and
transverse movement of the cable within the pipe and manholes. In particular, the cable forms
wave patterns that absorb thermal strain and reduce axial force. Unless measures are taken in the
design of the circuit, forces and movements can be excessive, damaging the cable and joints and
thus reducing reliability and service life. However, duct manhole and pipe system designs can
now be optimized to achieve reliable and predictable service performance.
This report describes the thermo-mechanical parameters that limit the life of cross-linked
polyethylene (XLPE) cable systems and gives the design values of the limiting parameters. It
includes a sensitivity study that identifies and ranks the key design measures necessary to limit
axial force. The report provides criteria for selecting the type of straight jointstandard straight-
through or anchor typeand gives the design withstand values for the cable differential force
and the minimum bending radius for a one-piece premolded 138kV joint. Designs of manhole
constraining systems are given, together with the measured withstand value of a prototype clamp
cleat fitted on samples of XLPE cable.
Traffic-induced vibration in heated cable can result in the avalanche effect, a previously
unknown mechanism. Cables and manhole constraints must be designed to withstand this force
and movement. Designers also need to take into account the dynamic movement that occurs
during and after short circuits.
The report provides an application guide for the design of duct-manhole and pipe systems,
including design objectives, calculation methods, design parameters, selection criteria for cable
and joint types, and information on the positioning and design of manholes.

vi
Challenges & Objectives
This report is relevant to anyone involved in the planning, selection, detailed design, installation
and maintenance of pipe and duct-manhole cable circuits containing XLPE cable for
transmission voltages of 69kV and above. The information it contains can benefit the bottom line
by providing cost savings in the selection of cables, joints, and manhole constraints necessary to
withstand thermomechanical forces and by reducing the unplanned outages and rectification
work that follow thermomechanical damage or circuit failure.
Applications, Values & Use
EPRI is planning a project to demonstrate the optimized solutions described in the report. In
addition, a user-friendly software tool to aid in using the normalized span method is under
development along with a stand-alone users guide. EPRI will also provide finite element analysis
calculations as a service for its members.
EPRI Perspective
High voltage and extra high voltage XLPE cable applications are in increasing use but have
limited service experience. Understanding how to minimize the risk of thermomechanical
damage will lead to greater confidence in the planning and approval of major projects using
these cables. This EPRI sponsored research is a major advance in the understanding and design
of reliable extruded dielectric cable systems installed in pipe and duct-manhole systems. The
report provides a textbook on a new branch of cable engineering knowledge, describing and
explaining the subject from first principles to applications.
Because of their greater axial and bending stiffness, cables and joints in XLPE systems behave
differently than those in high-pressure fluid-filled (HPFF) systems. A separate EPRI project is
looking at thermomechanical stresses in HPFF systems.
Approach
The project team developed two calculation methods, a finite element (FEA) modeling method to
be used for the detailed design of a new cable route and a normalized span calculation method
that is faster and requires less expertise for planning purposes. These new methods were made
possible by the development of a pipe and duct modeling technique and the adaptation of a
special type of FEA software to solve the dynamic interactions that occur within pipe and duct
routes. The cable characteristics used in the models were derived from tests performed on full-
size samples of 138kV and 230kV XLPE cables.
Keywords
Transmission cables
Cross-linked polyethylene
Cable accessories
Pipe
Duct
Manhole
Thermomechanical design

EPRI Licensed Material
vii
ABSTRACT
This work is a landmark in the design of duct and pipe systems. For the first time a detailed
understanding of thermomechanical behavior has been gained. Design methods are produced to
give a predictable and reliable performance for transmission class circuits containing XLPE
insulated cable. The advances were made possible by the explicit finite element analysis
(FEA) technique, to model the complex route geometry and the dynamic interactions between
the cables and the pipe/duct. An application guide lists the limiting design parameters and
describes the calculation steps to be followed for a) detailed design of the system, using the FEA
modelling method and b) circuit planning using the normalized span technique (NS).
When the cable carries load, it first expands to occupy the free space in the preformed bends.
Once full, the cable is constrained from longitudinal expansion and behaves as a rigid-bar,
generating maximum axial force as the temperature rises. At a comparatively low temperature,
the force and cable curvature in the bend produce a bending moment that forms a pattern of
cylindrical sinusoidal waves. Helical waves can also form in duct systems. A second transition
occurs at a critical turn-over temperature at which the amplitude and storage capacity of the
waves is suddenly increased. Locked-in thermal strain is released from the adjacent cable into
the patterns, permitting the force to be limited and to fall to low level. The amount of strain that
is initially released into the pattern is limited by the cable-pipe/duct frictional gradient.
Wave patterns can form wherever there is curvature, i.e. at bends, at manhole entrances and at
cable kinks. Pipe spans contain a high number of horizontal and vertical bends, such that at cable
operating temperature the patterns occupy over 90% of the length.
The thermomechanical mechanism is controlled by twelve parameters (geometric and cable
properties). The key design parameter is the clearance between one cable and the pipe wall; an
increase a) reduces sidewall force in bends, thereby limiting the conductor movement through
the XLPE insulation and b) reduces force in manholes, thereby limiting disturbance to the joint
insulation and deflection of the cable. An optimum clearance exists, beyond which the bending
strain exceeds the cyclic fatigue limit of the metallic sheath/foil moisture barrier. Pipe systems
inherently have a large clearance and benefit from low axial force. Duct systems do not have
this benefit. In both cases the optimum clearance needs to be calculated.
The key design parameter for the straight joint is the differential force at which interface slip
occurs. The joint is protected by a) designing the manhole cleating system to minimize
differential load and b) locating the manholes in the route at positions of thermomechanical
symmetry, free of cable movement.
EPRI Licensed Material

viii
The application of either a short circuit, or of traffic induced vibration, has a dramatic effect in
reducing frictional constraint and in suddenly releasing locked-in thermal strain from the span
length into the wave patterns. This mechanism is new discovery and is named the avalanche
effect.



EPRI Licensed Material
ix
ACKNOWLEDGEMENTS
A number of organizations and individuals helped to make this project a success.
In particular, thanks are due to Reza Ghafurian of Consolidated Edison, Vince Curci of LADWP
and Takashi Kojima of BC Hydro, who had an active involvement with the project throughout its
duration. Their help, advice, guidance and provision of cable samples is greatly appreciated.
Prior to publication the report was reviewed by Reza Ghafurian of Consolidated Edison and
Vince Curci of LADWP. Their constructive comments and ideas were invaluable.
Acknowledgment is made of the significant contribution of the engineering software house
ABAQUS UK Ltd, who provided the software, pioneered the modelling techniques and
performed the computer runs. In particular thanks are due to Stephen King and Jonathan Roy for
their constructive work throughout the duration of the project.
Acknowledgment is made of the work of the EPRIsolutions Laboratory at Haslet in measuring
the mechanical properties of XLPE cable samples and the performance of cable cleats. In
particular Tip Goodwin and Jeff Young who supervised the measurements.
EPRI Licensed Material
xi
CONTENTS
1 INTRODUCTION....................................................................................................... 1-1
General .................................................................................................................... 1-1
Pipe Systems........................................................................................................... 1-2
Duct-Manhole Systems............................................................................................ 1-4
Scope of the Project................................................................................................. 1-6
Layout of the Report ................................................................................................ 1-7
Background of the Investigators............................................................................... 1-9
Nomenclature and System of Units.......................................................................... 1-9
2 THERMOMECHANICAL AND GRAVITATIONAL EFFECTS: BASIC THEORY..... 2-1
General .................................................................................................................... 2-1
Explanation of the Thermomechanical Effects......................................................... 2-1
Start of Cable Heating ......................................................................................... 2-4
Straight Route ................................................................................................. 2-4
Route with Bends ............................................................................................ 2-5
In the Bend, Position B Figure 2-1................................................................... 2-5
First Transition State: Inception and Longitudinal Growth of Waves ............... 2-7
Second Transition State: Turn-over Force and Turn-over temperature........ 2-9
Wave Shapes................................................................................................ 2-11
The Anchored End of the Route, End A in Figure 2-6 ................................... 2-12
Turn-over Force Due to Frictional Slip at the Span End A........................... 2-13
Route with Slopes ......................................................................................... 2-14
Effects of Short Circuits, Traffic Vibration and Load Cycling ......................... 2-14
Avalanche Effect ....................................................................................... 2-14
Short Circuits............................................................................................. 2-15
Traffic Vibration......................................................................................... 2-15
EPRI Licensed Material

xii
Load Cycling ............................................................................................. 2-15
Individual Basic Theory Cases............................................................................... 2-16
Sliding Under Gravity and Thermomechanical Force ........................................ 2-16
Sliding with no Thermomechanical Force...................................................... 2-16
Sliding Downhill with Downward Thermomechanical Force .......................... 2-17
Sliding Downhill with Upward Thermomechanical Force............................... 2-17
Sliding Uphill with Upward Thermomechanical Force ................................... 2-18
Horizontal Sliding under Thermomechanical Force....................................... 2-18
Length of Cable Storage within the Pipe Route................................................. 2-19
Length of Cable Storage within a Bend......................................................... 2-19
Length of Cable Storage within a Straight Pipe............................................. 2-20
Length of Cable Stored within Lateral Deformation Patterns......................... 2-20
Length of Cable Stored in a Cylindrical Sinusoid Wave ................................ 2-21
Length of Cable Stored in a Helix.................................................................. 2-22
Side Wall Force and Axial Thrust....................................................................... 2-22
Sidewall Force and Axial Thrust in a Bend with No Friction .......................... 2-22
Axial Thrust in a Bend with Friction ............................................................... 2-22
Sidewall Force in a Bend with Friction........................................................... 2-24
3 RESULTS OF TESTS ON THE MECHANICAL PERFORMANCE OF CABLES
AND OF CONSTRAINTS IN JOINT CHAMBERS....................................................... 3-1
Selection of Cable Samples for Test........................................................................ 3-1
Tests on Cable Samples to Derive Mechanical Parameters. ................................... 3-3
Measurement of Transverse Bending Stiffness ................................................... 3-4
Finite Element Modelling of the Bending Stiffness Test Rig ................................ 3-8
Measurement of Torsional Stiffness .................................................................... 3-9
Measurement of Axial Stiffness ......................................................................... 3-13
Measurement of Axial stiffness Under Tension ............................................. 3-13
Measurement of Axial Stiffness under Compression..................................... 3-16
Thermal Tests.................................................................................................... 3-17
Measurement of Axial Stiffness..................................................................... 3-17
Measurement of Coefficient of Thermal Expansion....................................... 3-20
Measurement of Thermal Relaxation ............................................................ 3-21
EPRI Licensed Material

xiii
List of Parameters Derived from the Mechanical Tests on the 138kV
1500kcmil Cable................................................................................................ 3-23
List of Parameters Derived from the Mechanical Tests on the 230kV
2500kcmil Cable................................................................................................ 3-24
Tests on Types of Cable Constraints..................................................................... 3-24
Measurement of Internal Friction between Core and Outer Layers. .................. 3-24
Measurement of Pull-through Strength of Cable Cleats .................................. 3-26
Cleat Ultimate Strength Test ......................................................................... 3-26
Cleat Withstand Test ..................................................................................... 3-28
Cleat Withstand Test Result .......................................................................... 3-28
Multiple Cleat Test......................................................................................... 3-28
List of Parameters Derived from Tests on Constraining Methods for 138kV
XLPE Cable....................................................................................................... 3-30
List of Parameters Derived from Tests on Constraining Methods for 230 KV
XLPE Cable....................................................................................................... 3-31
Discussion.............................................................................................................. 3-31
Conclusions ........................................................................................................... 3-33
4 PARAMETERS LIMITING SYSTEM PERFORMANCE............................................ 4-1
Introduction.............................................................................................................. 4-1
Limiting Values of Cable Bending Radii ................................................................... 4-1
Limiting Value of Tensile Force................................................................................ 4-4
Limiting Value of Compressive Force ...................................................................... 4-6
Limiting Values of Sheath Strain.............................................................................. 4-7
Calculation of Sheath Strain from Bending Radius.............................................. 4-9
Limiting Value of Sheath Fatigue Strain................................................................. 4-10
Limiting Value of Sidewall Force............................................................................ 4-13
Limiting Value of Lateral Pressure on the Insulation and Jacket............................ 4-15
Limiting Value of Axial Thermomechanical Force .................................................. 4-18
Limiting Value of Pipe to Cable Clearance............................................................. 4-18
Pipe Jam Ratio .................................................................................................. 4-19
Bibliography ........................................................................................................... 4-19
5 CABLE AND ROUTE MODELLING: METHODS AND ANALYSIS ......................... 5-1
EPRI Licensed Material

xiv
Summary.................................................................................................................. 5-1
Contents................................................................................................................... 5-3
Modeling: Objectives ........................................................................................... 5-3
Modeling the Cable.................................................................................................. 5-4
Calculation of the Effective Axial and Bending Modulii of XLPE Insulation at
Different Temperatures........................................................................................ 5-6
Modeling the Pipe or Duct ................................................................................. 5-12
Modeling the Joint and Casing .......................................................................... 5-12
Modeling Friction ............................................................................................... 5-14
Loading Conditions: Heating and Load Cycling ..................................................... 5-14
First Computational Step: Dropping under Gravity ............................................ 5-14
Second Computational Step: Heating................................................................ 5-15
Third Computational Step: Constant Temperature ............................................ 5-15
Fourth Computational Time Step: Cooling and Completion of Load Cycle........ 5-15
Loading Conditions: Vibration and Short Circuit..................................................... 5-15
Computational Step: Vibration........................................................................... 5-16
Computational Step: Short Circuit...................................................................... 5-16
Process of Modeling a Cable Route....................................................................... 5-16
Solving the Thermomechanical Behavior of the Model Route................................ 5-20
Inputting Information into the Solver .................................................................. 5-20
Checking the Success of the Run...................................................................... 5-21
Engineering Analysis of the Model Output ............................................................. 5-22
Charts and Videos ................................................................................................. 5-23
Computing Requirements ...................................................................................... 5-29
Description of the Finite Element Analysis Modeling Methods:.............................. 5-29
Modeling: Implicit Analysis Method ...................................................................... 5-30
Modeling: Explicit Analysis Method ...................................................................... 5-31
Validation of the Explicit Finite Element Analysis Modelling Technique ............... 5-32
Validation Results.............................................................................................. 5-33
Description of thermomechanical phenomena .............................................. 5-33
Comparison of explicit and implicit solutions.............................................. 5-35
Comparison with In-Service Results ...................................................................... 5-37
EPRI Licensed Material

xv
6 THERMOMECHANICAL PERFORMANCE OF CABLES IN A STRAIGHT
ROUTE......................................................................................................................... 6-1
Introduction.............................................................................................................. 6-1
Theoretically Perfect Straight Cable......................................................................... 6-2
Straight Route Containing a Single Kink.................................................................. 6-3
Single Cable Model .................................................................................................. 6-3
Single Kink- Deformation Patterns....................................................................... 6-4
Single Kink Force Distribution Along the Route................................................ 6-9
Phase A, 15
o
C-42.5
o
C.................................................................................. 6-10
Phase B, 42.5
o
C-65
o
C.................................................................................. 6-11
Phase C, 65
o
C -90
o
C................................................................................... 6-12
Phase D, 90
o
C-125
o
C.................................................................................. 6-12
Single Kink - Cable Length Absorption With Temperature................................. 6-13
Single Kink - Force Increase With Temperature................................................ 6-14
Multiple Kinks - Deformation Patterns................................................................ 6-16
Multiple Kinks Force Distribution Along the Route.......................................... 6-18
Multiple Kinks - Force Increase With Temperature............................................ 6-18
Three Cable Model ................................................................................................ 6-20
Three Cable Case: Single Kink-Deformation Patterns....................................... 6-20
Three Cable Case: Single and Multiple Kinks-Force Increase with
Temperature...................................................................................................... 6-22
Conclusions ........................................................................................................... 6-24
7 THERMOMECHANICAL PERFORMANCE OF CABLE IN A ROUTE
CONTAINING BENDS AND SLOPES: SENSITIVITY STUDY.................................... 7-1
Introduction.............................................................................................................. 7-1
Reference Model...................................................................................................... 7-2
Normalization of the Sensitivity Study...................................................................... 7-3
Verification of Normalization on Pipe Clearance...................................................... 7-4
Sensitivity Study: Variations in Geometry of Model Route....................................... 7-8
Clearance Between Cable and Pipe.................................................................... 7-8
Pipe Length ......................................................................................................... 7-8
Bend Radius and Bend Angle.............................................................................. 7-9
Angle of Inclination ............................................................................................ 7-10
EPRI Licensed Material

xvi
Coefficient of Friction......................................................................................... 7-10
Cable Designs ................................................................................................... 7-11
Calculation of Cable Component Stiffnesses..................................................... 7-12
EI: Bending Stiffness..................................................................................... 7-13
Jacket/Sheath EI
s
...................................................................................... 7-13
Insulation EI
x
............................................................................................. 7-13
Conductor EI
c
............................................................................................ 7-14
Total Cable EI ........................................................................................... 7-14
EA Axial Stiffness.......................................................................................... 7-15
GJ Torsional Stiffness ................................................................................... 7-15
Cable Parameters Used in the Sensitivity Study.................................................... 7-16
Weights for the Range of Cables....................................................................... 7-16
Coefficient of Thermal Expansion for the Range of Cables ............................... 7-17
Stiffness Parameters for the Range of Cables................................................... 7-17
Load Case for the Sensitivity Study ....................................................................... 7-21
Formation of Thermomechanical Patterns in the 5 Bend Model ........................... 7-21
Formation of Thermomechanical Patterns in the Reference Duct Model ............... 7-24
Comparison of 5
o
and 90
o
Bend Models ................................................................ 7-27
Formation of Thermomechanical Patterns in the Reference Pipe Model ............... 7-28
Method of Analysis of the Sensitivity Study Results............................................... 7-29
Sensitivity Study Results........................................................................................ 7-31
Duct System Sensitivity Results ........................................................................ 7-31
Back-up Information: ..................................................................................... 7-31
Pipe System Sensitivity Results......................................................................... 7-31
Back-up Information: ..................................................................................... 7-31
Sensitivity Study Results for Cyclic Loading...................................................... 7-31
Ranking of Sensitivity Study Results ................................................................. 7-32
Common Rankings in Pipe and Duct Systems .................................................. 7-33
Duct System: Ranking of Parameters at Position A........................................... 7-36
Duct System: Ranking of Parameters at Position B........................................... 7-38
Pipe System: Ranking of Parameters at Position A........................................... 7-40
Pipe System: Ranking of Parameters at Position B........................................... 7-42
Duct and Pipe System: Comparison of Thermomechanical Wave Patterns ...... 7-44
EPRI Licensed Material

xvii
Pipe System: Thermomechanical Wave Patterns.............................................. 7-46
Parameters that Limit the Life of Duct and Pipe Systems...................................... 7-48
Axial Force......................................................................................................... 7-48
Sidewall Force in Bend...................................................................................... 7-49
Minimum Bending Radius in the Straight Section.............................................. 7-51
Sheath Strain in the Straight Section................................................................. 7-52
Discussion.............................................................................................................. 7-54
Conclusions ........................................................................................................... 7-58
Sensitivity Study Bar Charts and Graphs............................................................... 7-64
8 PERFORMANCE OF SELECTED DESIGNS OF XLPE CABLES IN THE
SYSTEM VOLTAGE RANGE 69-345KV..................................................................... 8-1
Introduction.............................................................................................................. 8-1
Parameters .............................................................................................................. 8-2
Cable Axial Stiffness EA...................................................................................... 8-2
Coefficient of Thermal Expansion ..................................................................... 8-3
Coefficient of Friction ........................................................................................ 8-3
Axial Force Coefficient EA................................................................................. 8-3
Duct System Cable Models...................................................................................... 8-3
Duct Systems Results.............................................................................................. 8-6
230kV 2500kcmil Lead Sheath Cable, (EA,,)
4

, Clearance 37.7mm................ 8-7
Position B, Bend.............................................................................................. 8-7
Position A, Span End ...................................................................................... 8-8
Summary......................................................................................................... 8-8
69kV 1500kcmil CSA, (EA,,)
4
Clearance 40.5mm........................................... 8-9
Position B, Bend Entrance............................................................................... 8-9
Position A, Span End .................................................................................... 8-10
Summary....................................................................................................... 8-10
345kV 3000kcmil CSA Sheath Cable, (EA,,)
4
Clearance 79.8mm................. 8-11
Position B, Bend Entrance............................................................................. 8-11
Position A, Span End .................................................................................... 8-11
138kV 1500kcmil Foil Laminate, (EA,,)
4
Clearance 150mm.......................... 8-13
Position B, Bend............................................................................................ 8-13
EPRI Licensed Material

xviii
Position A, Span End .................................................................................... 8-14
Summary....................................................................................................... 8-14
138kV 1500kcmil Foil Laminate: Effect of Varying EA, and ......................... 8-15
Duct System Summary...................................................................................... 8-16
Pipe System Cable Models.................................................................................... 8-17
Pipe System Results.............................................................................................. 8-18
69kV 1000kcmil Aluminum Foil Sheath Cable, (EA,,)
4
, clearance 132mm.... 8-19
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,,)
4
, Clearance
123mm............................................................................................................... 8-21
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,,)
4
, Clearance
150mm............................................................................................................... 8-23
Position B, Bend............................................................................................ 8-23
Position A, Span End .................................................................................... 8-23
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,,)
ref
, Clearance
123mm............................................................................................................... 8-24
Pipe System Summary ...................................................................................... 8-25
Normalized Span Calculation Method.................................................................... 8-26
NS Calculation Method...................................................................................... 8-26
Example NS Calculation for a 69kV Duct System 1500kcmil CSA Sheathed
Cable ................................................................................................................. 8-28
Calculation of Axial Force at Position A in the 200m Hockey Stick Route..... 8-28
Comparison of NS Answer with Model Analysis............................................ 8-29
Position B in 200m Hockey Stick Route ........................................................ 8-30
Force Factors for the Duct Models..................................................................... 8-32
Force Factors for the Pipe Models..................................................................... 8-32
Comparison of NS Calculation Method Results with FEA Model Analysis......... 8-33
Accuracy of NS Calculation for the Pipe Systems......................................... 8-33
Accuracy of Calculation for the Duct System................................................. 8-35
Conclusions ........................................................................................................... 8-37
9 EFFECTS OF VIBRATIONS..................................................................................... 9-1
Introduction.............................................................................................................. 9-1
Published Work on Traffic Vibration......................................................................... 9-1
Previous Investigational Work on Cable Problems .............................................. 9-1
EPRI Licensed Material

xix
Previous Work to Quantify the Magnitude of Traffic Induced Vibrations.............. 9-2
Applying Vibration to Duct and Pipe System Models ............................................... 9-5
Model Geometry .................................................................................................. 9-5
Modelling Traffic Vibration ................................................................................... 9-6
Load Cases ......................................................................................................... 9-8
Avalanche Cable Movement .................................................................................... 9-9
Avalanche Mechanism......................................................................................... 9-9
Response of Pipe and Duct Systems to Vibration ............................................... 9-9
Equilibrium Before the Application of Vibration.............................................. 9-10
Equilibrium After the Application of Vibration................................................. 9-13
Angle of Slope............................................................................................... 9-13
Uphill Slopes ................................................................................................. 9-14
Downhill Slopes............................................................................................. 9-14
Results from Model Spans Subject to Vibration..................................................... 9-16
Duct and Pipe Systems: Summary of Effect of Vibrations on Different Slope
Angles................................................................................................................ 9-16
Duct System: Horizontal Span........................................................................... 9-17
Duct System: Uphill Spans ................................................................................ 9-19
Uphill Span +7.5
o
........................................................................................... 9-19
Before the Truck Applies Traffic Vibration................................................. 9-19
When the Truck Applies Traffic Vibration.................................................. 9-21
Uphill Span +3
o
.............................................................................................. 9-22
Duct System: Downhill Span ............................................................................. 9-23
Downhill Span -7.5
o
....................................................................................... 9-23
Before the Truck Applies Traffic Vibration................................................. 9-23
When the Truck Applies Traffic Vibration.................................................. 9-25
Downhill Span -3
o
.......................................................................................... 9-25
Duct System: Repeat Truck Journeys ............................................................... 9-26
Pipe System: Horizontal Span........................................................................... 9-27
Before the Truck Applies Traffic Vibration..................................................... 9-27
When the Truck Applies Traffic Vibration ...................................................... 9-30
Pipe System: +7.5
o
Uphill Span......................................................................... 9-31
Pipe System: -7.5
o
Downhill Span ..................................................................... 9-32
EPRI Licensed Material

xx
Conclusions ........................................................................................................... 9-33
10 EFFECT OF SHORT CIRCUITS........................................................................... 10-1
Introduction............................................................................................................ 10-1
Pipe and Duct Spans Models................................................................................. 10-1
Short Circuit Force................................................................................................. 10-3
Basic Theory of Short Circuit Forces ................................................................. 10-3
Application of Short Circuit Force to the Model Cables...................................... 10-4
Temporal Displacement of Current................................................................ 10-5
Spatial Displacement of the Conductor Positions.......................................... 10-5
Verification of Modeled Short Circuit Forces ................................................. 10-5
Load Cases............................................................................................................ 10-6
Description of Thermomechanical and Electrodynamic Effects ............................. 10-7
Ducts in Flat Formation: 50kA Flat Route .............................................................. 10-8
Ducts in Trefoil Formation-50 kA Flat Route........................................................ 10-18
Three Cables in Pipe - 50kA Flat Route............................................................... 10-23
Effect of Slope Angles 50kA Short Circuit ......................................................... 10-28
Flat Formation - Reduced Short Circuit Level of 20kA......................................... 10-31
Ducts in Trefoil Formation-Reduced Short Circuit Level of 20kA......................... 10-35
Three Cables in the Pipe System - Reduced Short Circuit Level of 20kA............ 10-37
Conclusion........................................................................................................... 10-38
11 MECHANICAL PERFORMANCE OF A ONE-PIECE STRAIGHT JOINT ............ 11-1
Introduction............................................................................................................ 11-1
Joint Performance in a Full Sized Route................................................................ 11-2
138kV One-Piece Joint Design .............................................................................. 11-8
Description of Joint ............................................................................................ 11-8
Description of Assembly Sequence................................................................. 11-11
Limiting Parameters......................................................................................... 11-12
Limiting Parameter: Interfacial Pressure...................................................... 11-12
Limiting Parameter: Longitudinal Movement ............................................... 11-14
Detailed Model of One-Piece Joint....................................................................... 11-14
Bending Model................................................................................................. 11-15
EPRI Licensed Material

xxi
Cable Pull-Through Model ............................................................................... 11-16
Bending Behavior of One-Piece Joint .................................................................. 11-16
Results: Bending Radii .................................................................................... 11-16
Results: Interfacial Pressure............................................................................ 11-18
Straight Cable and Joint .............................................................................. 11-18
Cable and Joint Bending ............................................................................. 11-20
Bending Sensitivity Study............................................................................ 11-20
Results: Cable Pull-Through Withstand Strength ............................................ 11-24
Interface Slip.................................................................................................... 11-25
Void Formation at the Interface ....................................................................... 11-26
Outline Test Schedule.......................................................................................... 11-29
Conclusions ......................................................................................................... 11-30
12 DESIGN OF CONSTRAINTS IN MANHOLES...................................................... 12-1
Introduction............................................................................................................ 12-1
Pipe System Constraints........................................................................................ 12-1
Pipe System: In-Line Joint Constraints .................................................................. 12-2
In-Line Joint Model ............................................................................................ 12-2
In-Line Joint: Improved Constraint Design......................................................... 12-3
Pipe Systems: Offset Joints ................................................................................... 12-7
Offset Joint Model.............................................................................................. 12-7
Offset Joint: Sensitivity Study on Offset Geometry and Cleat Strength ........... 12-11
Offset Joint: Design of Offset for 138kV 1500kcmil Pipe System.................... 12-16
Duct System Constraints ..................................................................................... 12-18
Duct System: Expansion Cleat System........................................................... 12-19
Model of Thermal Expansion Cleat Constraining System................................ 12-20
Duct System: Clamp Cleat Constraint.................................................................. 12-24
Internal Constraint Within the Cable .................................................................... 12-26
Conclusions ......................................................................................................... 12-27
13 APPLICATION GUIDE FOR THE DESIGN OF DUCT-MANHOLE SYSTEMS
AND PIPE SYSTEMS CONTAINING UNDERGROUND HV XLPE CABLES........... 13-1
Route Geometry and Thermomechanical Effects .................................................. 13-2
Route Geometry Modelling by FEA (Finite Element Analysis) Technique ............. 13-7
EPRI Licensed Material

xxii
Route Analysis by the Normalized Span Technique .............................................. 13-8
Gravitational Sliding Effects During Installation.................................................... 13-11
Thermal Expansion and Movement ..................................................................... 13-12
Thermomechanical Axial Force: Rigid Bar Behavior............................................ 13-12
Key Design Parameter: Clearance Between Cable and Pipe/Duct ...................... 13-14
Absorption of Thermal Strain and Locking Within Bends..................................... 13-15
Absorption of Strain in Bends .......................................................................... 13-15
Locking Effect of Bends................................................................................... 13-15
Thermomechanical Patterns: Absorption of Strain, Turn-over Temperature
and Turn-Over Force.......................................................................................... 13-16
Wave Shapes .................................................................................................. 13-17
Cylindrical Sinusoids ................................................................................... 13-20
Inception and Longitudinal Growth of Waves .................................................. 13-21
Limiting Effect of Pipe Friction on Strain Absorption........................................ 13-24
Transition to Large Amplitude Waves.............................................................. 13-25
Turn-Over Force Due to Frictional Slip at the Span End................................ 13-28
Thermomechanical Patterns in the Full sized Span............................................. 13-29
Positioning of Manholes along the Route............................................................. 13-33
Calculation of Thermomechanical Effects............................................................ 13-35
Calculation: FEA Method................................................................................. 13-35
Calculation: NS Method................................................................................... 13-36
Preparation of Input Data................................................................................. 13-36
Cable and Pipe Data ................................................................................... 13-36
Route Data .................................................................................................. 13-37
Ranking of Key Design Parameters..................................................................... 13-37
Duct System Ranking of Parameters at End of Span, A.................................. 13-37
Duct System: Ranking of Parameters at Entrance to Bend, B......................... 13-38
Pipe System: Ranking of Parameters at Span End, A..................................... 13-39
Pipe System: Ranking of Parameters at Bend Entrance, B............................. 13-40
Design Parameters for Cables in Pipe and Duct Systems ................................... 13-41
Choice of Cable Type ...................................................................................... 13-42
Bending Radius of Preformed Bends............................................................... 13-44
Bending Radius in Manholes........................................................................... 13-44
EPRI Licensed Material

xxiii
Sidewall Pressure on XLPE Insulation and PE Jacket..................................... 13-44
Cable Axial Force ............................................................................................ 13-45
Cable Radius of Curvature .............................................................................. 13-46
Sheath Fatigue Strain...................................................................................... 13-46
Design Parameters for Joint Manholes in Pipe and Duct Systems ...................... 13-47
Choice of Joint ................................................................................................. 13-48
Prefabricated Premolded Joint .................................................................... 13-49
Prefabricated Composite Anchor Joint ........................................................ 13-50
Field Molded Joint ....................................................................................... 13-51
Cleat Types ..................................................................................................... 13-51
Clamp Cleat................................................................................................. 13-51
Guide Cleat ................................................................................................. 13-52
Expansion Cleat .......................................................................................... 13-53
Short Circuit Cleat ....................................................................................... 13-53
Traffic Vibration and Short Circuit Forces........................................................ 13-54
In-line Joint Constraint in a Pipe System......................................................... 13-54
Offset Joint Constraint in a Pipe System......................................................... 13-55
Clamped Offset Constraint in a Duct System.................................................. 13-57
Expansion Cleat and S-Bend Constraint in a Duct System............................. 13-58
14 CONCLUSIONS.................................................................................................... 14-1
15 FURTHER WORK................................................................................................. 15-1



EPRI Licensed Material
xxv
LIST OF FIGURES
Figure 1-1 138kV 1500kcmil and 230kV 2500kcmil cables from North American
installations ..................................................................................................................... 1-2
Figure 1-2 138kV 1500kcmil XLPE cables in a pipe with an example clearance of
117mm (4.6)................................................................................................................... 1-3
Figure 1-3 230kV 2500kcmil lead sheathed XLPE cable in a 142mm duct with an
example clearance of 25mm............................................................................................ 1-5
Figure 2-1 Diagram of a 200m duct route with one bend........................................................ 2-2
Figure 2-2 Coil spring analogy of thermomechanical effects in a route with a straight
cable section and one bend............................................................................................. 2-3
Figure 2-3 Axial force-temperature characteristic at the entrance to the bend, position B ...... 2-6
Figure 2-4 Bending moment formation at the entrance to the bend........................................ 2-8
Figure 2-5 Graph of cable absorbed in a sine wave for different large ratios of amplitude
to wavelength.................................................................................................................2-10
Figure 2-6 Axial force-temperature characteristic at the fixed cable end, position A..............2-13
Figure 2-7 Cable in an inclined pipe......................................................................................2-16
Figure 2-8 Sidewall force in a bend.......................................................................................2-23
Figure 3-1 Diagram of the beam method of measuring bending stiffness............................... 3-4
Figure 3-2 Photograph of the bending stiffness test rig .......................................................... 3-5
Figure 3-3 Displacement-force output for 138kV cable at ambient temperature ..................... 3-6
Figure 3-4 Bending stiffness results for 138kV cable.............................................................. 3-7
Figure 3-5 Bending stiffness for the 230kV cable ................................................................... 3-7
Figure 3-6 Output from the finite element model of the bending stiffness rig on 138kV
cable ............................................................................................................................... 3-9
Figure 3-7 Diagram of the torsional stiffness test ..................................................................3-10
Figure 3-8 Photograph of the torsional stiffness test rig.........................................................3-11
Figure 3-9 Output of angular rotation with torque for the 138kV cable...................................3-12
Figure 3-10 Results of the torsional stiffness tests for the 138kV cable.................................3-12
Figure 3-11 Diagram of measurement of axial stiffness under tension ..................................3-13
Figure 3-12 Output of axial stiffness under tension measurements for the 138kV cable........3-14
Figure 3-13 Derived values of axial stiffness under tension for the 138kV cable ...................3-15
Figure 3-14 Photograph of axial stiffness under compression rig ..........................................3-15
Figure 3-15 Output from axial stiffness test on 138kV cable..................................................3-16
EPRI Licensed Material

xxvi
Figure 3-16 Derived values of axial stiffness under compression for 138kV and 230kV
cables.............................................................................................................................3-17
Figure 3-17 Measurement of thrust and axial stiffness in fully constrained heated cable.......3-18
Figure 3-18 Output of force with temperature from constrained thermal test .........................3-19
Figure 3-19 Photograph of the thermal expansion test on 138kV cable.................................3-20
Figure 3-20 Output of thermal extension with temperature for the 138kV cable.....................3-21
Figure 3-21 Output of thermal relaxation test on 138kV cable...............................................3-22
Figure 3-22 Photograph of the internal friction test rig, 230kV cable......................................3-25
Figure 3-23 Photograph of the cleat pull-through test rig on the 230kV cable........................3-27
Figure 3-24 Output of the cleat pull-through test on the 138kV cable ....................................3-28
Figure 3-25 Output of the cleat pull through test on 138 and 230kV cables movement
at hold points..................................................................................................................3-29
Figure 3-26 Output of cleat withstand test for 230kV cable ...................................................3-29
Figure 4-1 Flexing life characteristic for corrugated and smooth sided aluminum sheaths ....4-11
Figure 5-1 Elastic modulus of XLPE in tension (linear/linear scale) ........................................ 5-7
Figure 5-2 Elastic modulus of XLPE in compression versus temperature (linear/linear
scale) .............................................................................................................................. 5-8
Figure 5-3 Effective axial and bending stiffnesses for XLPE insulation in a 138kV
1500kcmilcable................................................................................................................ 5-9
Figure 5-4 Effective axial and bending stiffnesses for XLPE insulation in a 230kV
2500kcmil cable..............................................................................................................5-11
Figure 5-5 View of modeled joint casing in an in-line 138kV pipe system..............................5-13
Figure 5-6 Example profile of a joint casing in a 138kV pipe system.....................................5-20
Figure 5-7 Three dimensional stilt diagram of two 138kV pipe spans totaling 763m..............5-21
Figure 5-8 Plan view of the 200m normalized reference span at 90
o
C with wave patterns
adjacent to the bend at B................................................................................................5-23
Figure 5-9 Rotated perspective view of 200m normalized pipe span at 90
o
C ........................5-24
Figure 5-10 Thermomechanical pattern distributions in a double span 138kV 1500kcmil
pipe system at 15
o
C, 30
o
C, 45
o
C, 60
o
C, 75
o
C and 90
o
C...................................................5-24
Figure 5-11 Deformation line plot with 200m duct model rotated and lateral scale
increase (before and during journey of vibrating truck over duct)....................................5-25
Figure 5-12 Close-up view of the end 25m of the 200m pipe span next to bend B at
90
o
C, to observe individual wave patterns ......................................................................5-25
Figure 5-13 Rotated close-up view of the end 25m of the 200m pipe span at 125
o
C to
show change in wave shape at bend entrance B............................................................5-26
Figure 5-14 End view along ducts in trefoil at 90
o
C during 50kA short circuit.........................5-27
Figure 5-15 History plot of axial force with temperature in the three cables in the 200m
normalized route at the end of the straight span at A and adjacent to the bend at B.......5-28
Figure 5-16 Path plot of axial conductor force along a double span of a pipe system at
90
o
C ...............................................................................................................................5-28
EPRI Licensed Material

xxvii
Figure 5-17 Cable thermomechanical patterns in the 30m duct validation model ..................5-33
Figure 5-18 Reaction forces at each end of the 30m length in the validation model ..............5-34
Figure 5-19 Length of cable inside the 30m duct with temperature for the validation
model .............................................................................................................................5-35
Figure 5-20 Comparison of 'explicit' and 'implict' cable deformation patterns for the
validation model .............................................................................................................5-36
Figure 5-21 Comparison of cable bending radii in the explicit and explicit FEA models.........5-36
Figure 5-22 Comparison of axial force distribution in 'explicit' and implicit FEA
validation model .............................................................................................................5-37
Figure 5-23 Deflection of 138kV joints, top is at 90
o
C, bottom is at 15
o
C at end of 3
rd

cycle...............................................................................................................................5-38
Figure 5-24 Deflection of 138kV joints, top is at 90
o
C, bottom is at 15
o
C at end of 3
rd

cycle...............................................................................................................................5-39
Figure 6-1 Thermomechanical pattern at 50
o
C initiated by one kink. ...................................... 6-4
Figure 6-2 Initial deformation at 15 C and at 35 C-single kink case ..................................... 6-5
Figure 6-3 Thermomechanical deformation at 42.5
o
C and 45
o
C-single kink case ................. 6-5
Figure 6-4 Cable thermomechanical patterns at 45
o
C, 50
o
C and 55
o
C-single kink ................ 6-7
Figure 6-5 Growth of thermomechanical patterns between 15
o
C and 125
o
C-single kink......... 6-8
Figure 6-6 Lengthwise view inside the pipe from the 100m central position ........................... 6-9
Figure 6-7 Force distribution along route, Phase A, 20
o
C-42.5
o
C..........................................6-10
Figure 6-8 Force distribution along route, Phase B, 42.5 65
o
C...........................................6-11
Figure 6-9 Force distribution along route, Phase C, 65-90C ................................................6-12
Figure 6-10 Force distribution along route, Phase D, 90-125C ............................................6-13
Figure 6-11 Variation of cable absorption length (extension) with temperature......................6-14
Figure 6-12 Increase of cable axial force at the ends of the route with temperature..............6-15
Figure 6-13 Increase of cable axial force at the central kink with temperature.......................6-16
Figure 6-14 Growth of deformation patterns-29 kinks............................................................6-17
Figure 6-15 Deformation patterns in three cables at 15
o
C and 35
o
C-single kink ....................6-20
Figure 6-16 Thermomechanical pattern in three cables at 50
o
C-single kink...........................6-21
Figure 6-17 Deformation patterns in the three cables at 90
o
C-single kink..............................6-22
Figure 6-18 Increase of axial force at the ends of three cables-1 and 29 kinks .....................6-23
Figure 7-1 Dimensions of reference model............................................................................. 7-3
Figure 7-2 Print-out of reference model (200m straight section and 2.9m radius bend).......... 7-3
Figure 7-3 Axial force/temperature characteristics for the 69kV cable (left) and 345kV
cable (right) ..................................................................................................................... 7-6
Figure 7-4 Identical thermomechanical pattern formed at the turn-over temperature/peak
force position for the 69kV and 345kV duct system models............................................. 7-7
Figure 7-5 Duct system with 5
o
bend, showing a pattern forming at 37.5
o
C...........................7-22
EPRI Licensed Material

xxviii
Figure 7-6 Duct system with 5
o
bend, showing the pattern at the turn-over temperature
of 50
o
C ...........................................................................................................................7-22
Figure 7-7 Reference duct system: force-temperature characteristic at position B................7-25
Figure 7-8 Reference duct system: incipient wave at 65
o
C....................................................7-25
Figure 7-9 Reference duct system: pattern at turn-over temperature of 72.5
o
C.....................7-26
Figure 7-10 Reference duct system: pattern at operating temperature of 90
o
C......................7-26
Figure 7-11 Reference duct system: pattern at 90
o
C showing helical loops and sine
waves.............................................................................................................................7-27
Figure 7-12 Reference pipe system: patterns at 90
o
C ...........................................................7-28
Figure 7-13 Reference pipe system: view of patterns adjacent to bend at 125
o
C ..................7-29
Figure 7-14 Diagrammatic force-temperature plot used to characterize sensitivity study
output .............................................................................................................................7-30
Figure 7-15 Duct system: Force - % clearance......................................................................7-56
Figure 7-16 Pipe system: Force - % clearance......................................................................7-56
Figure 7-17 Duct and pipe systems: Force - % clearance .....................................................7-57
Figure 7-18 Duct and pipe systems: Force - % clearance, all cables plotted.........................7-57
Figure 7-19 Duct position A: Bar chart ranking of turn-over force for % change in
parameters.....................................................................................................................7-64
Figure 7-20 Duct position A: Bar chart ranking of turn-over temperature for % change in
parameters.....................................................................................................................7-64
Figure 7-21 Duct position A: Angle of route inclination ..........................................................7-65
Figure 7-22 Duct position A: Duct to cable clearance............................................................7-65
Figure 7-23 Duct position A: Length of straight span.............................................................7-66
Figure 7-24 Duct position A: Weight......................................................................................7-66
Figure 7-25 Duct position A: Angle of duct bend ...................................................................7-67
Figure 7-26 Duct position A: Coefficient of friction, cable to duct ...........................................7-67
Figure 7-27 Duct position A: Coefficient thermal expansion ..................................................7-68
Figure 7-28 Duct position A: Bending stiffness, EI.................................................................7-68
Figure 7-29 Duct position A: Axial stiffness, EA.....................................................................7-69
Figure 7-30 Duct position A: Radius of duct bend .................................................................7-69
Figure 7-31 Duct position A: Torsional stiffness, GJ..............................................................7-70
Figure 7-32 Duct position B: Bar chart ranking of turn-over force for % change in
parameters.....................................................................................................................7-70
Figure 7-33 Duct position B: Bar chart ranking of turn-over temperature for % change in
parameters.....................................................................................................................7-71
Figure 7-34 Duct position B: Pipe to cable clearance............................................................7-71
Figure 7-35 Duct position B: Coefficient of thermal expansion ..............................................7-72
Figure 7-36 Duct position B: Bending stiffness, EI.................................................................7-72
Figure 7-37 Duct position B: Length of straight span.............................................................7-73
EPRI Licensed Material

xxix
Figure 7-38 Duct position B: Weight......................................................................................7-73
Figure 7-39 Duct position B: Axial stiffness, EA.....................................................................7-74
Figure 7-40 Duct position B: Angle of duct bend ...................................................................7-74
Figure 7-41 Duct position B: Coefficient of friction, cable to duct ...........................................7-75
Figure 7-42 Duct position B: Angle of route inclination ..........................................................7-75
Figure 7-43 Duct position B: Radius of duct bend .................................................................7-76
Figure 7-44 Duct position B: Torsional stiffness, GJ..............................................................7-76
Figure 7-45 Duct positions A,B,C, Bar Chart part 1: Axial force variation with parameters ....7-77
Figure 7-46 Duct positions A,B,C, Bar Chart part 2: Axial force variation with parameters ....7-77
Figure 7-47 Duct positions A,B,C, Bar Chart part 1: Turn-over temperature..........................7-78
Figure 7-48 Duct positions A,B,C, Bar Chart part 2: Turn-over temperature..........................7-78
Figure 7-49 Duct, Bar Chart part 1: Numbers of total and helical waves along straight
route...............................................................................................................................7-79
Figure 7-50 Duct, Bar Chart part 2: Numbers of total and helical waves along straight
route...............................................................................................................................7-79
Figure 7-51 Duct, Bar Chart part 1: Minimum cable bending radius in straight section..........7-80
Figure 7-52 Duct, Bar Chart part 2: Minimum cable bending radius in straight section..........7-80
Figure 7-53 Duct, Bar Chart part 1: Maximum sheath strain in straight section .....................7-81
Figure 7-54 Duct, Bar Chart part 2: Maximum sheath strain in straight section .....................7-81
Figure 7-55 Duct, Bar Chart part 1: Maximum sidewall force in bend....................................7-82
Figure 7-56 Duct, Bar Chart part 2: Maximum sidewall force in bend....................................7-82
Figure 7-57 Pipe position A cable 1: Bar chart ranking of turn-over force for % change in
parameters.....................................................................................................................7-83
Figure 7-58 Pipe position A cable 2 or 3: Bar chart ranking of turn-over force for %
change in parameters.....................................................................................................7-83
Figure 7-59 Pipe position A: Clearance pipe to one cable.....................................................7-84
Figure 7-60 Pipe position A: Angle of route inclination ..........................................................7-84
Figure 7-61 Pipe position A: Coefficient of thermal expansion...............................................7-85
Figure 7-62 Pipe position A: Weight ......................................................................................7-85
Figure 7-63 Pipe position A: Length of straight span.............................................................7-86
Figure 7-64 Pipe position A: Axial stiffness, EA.....................................................................7-86
Figure 7-65 Pipe position A: Radius of pipe bend..................................................................7-87
Figure 7-66 Pipe position A: Coefficient of friction, pipe to cable...........................................7-87
Figure 7-67 Pipe position A: Angle of pipe bend....................................................................7-88
Figure 7-68 Pipe position A: Bending stiffness, EI.................................................................7-88
Figure 7-69 Pipe position A: Coefficient of friction, cable to cable .........................................7-89
Figure 7-70 Pipe position A: Torsional stiffness, GJ ..............................................................7-89
Figure 7-71 Pipe position B cable 3: Bar chart ranking of turn-over force for % change in
parameters.....................................................................................................................7-90
EPRI Licensed Material

xxx
Figure 7-72 Pipe position B cable 3: Bar chart ranking of turn-over temperature for %
change in parameters.....................................................................................................7-90
Figure 7-73 Pipe position B: Pipe clearance to one cable .....................................................7-91
Figure 7-74 Pipe position B: Coefficient of thermal expansion...............................................7-91
Figure 7-75 Pipe position B: Axial stiffness, EA.....................................................................7-92
Figure 7-76 Pipe position B: Bending stiffness, EI.................................................................7-92
Figure 7-77 Pipe position B: Weight ......................................................................................7-93
Figure 7-78 Pipe position B: Coefficient of friction, cable to cable .........................................7-93
Figure 7-79 Pipe position B: Angle of route inclination ..........................................................7-94
Figure 7-80 Pipe position B: Coefficient of friction, pipe to cable...........................................7-94
Figure 7-81 Pipe position B: Length of straight span.............................................................7-95
Figure 7-82 Pipe position B: Radius of pipe bend..................................................................7-95
Figure 7-83 Pipe position B: Angle of pipe bend....................................................................7-96
Figure 7-84 Pipe position B: Torsional stiffness, GJ ..............................................................7-96
Figure 7-85 Pipe, Bar Chart position A, cables 1,2,3, part 1: Axial force variation with
parameters.....................................................................................................................7-97
Figure 7-86 Pipe, Bar Chart position A, cables 1,2,3, part 2: Axial force variation with
parameters.....................................................................................................................7-97
Figure 7-87 Pipe, Bar Chart position A, cables 1,2,3, part 1: Turn-over temperature
variation with parameters ...............................................................................................7-98
Figure 7-88 Pipe, Bar Chart position, cables 1,2,3, A part 2: Turn-over temperature
variation with parameters ...............................................................................................7-98
Figure 7-89 Pipe, Bar Chart position B, cables 1,2,3, part 1: Axial force variation with
parameters.....................................................................................................................7-99
Figure 7-90 Pipe, Bar Chart position B, cables 1,2,3, part 2: Axial force variation with
parameters.....................................................................................................................7-99
Figure 7-91 Pipe, Bar Chart position B, cables 1,2,3, part 1: Turn-over temperature
variation with parameters .............................................................................................7-100
Figure 7-92 Pipe, Bar Chart position B, cables 1,2,3, part 2: Turn-over temperature
variation with parameters .............................................................................................7-100
Figure 7-93 Pipe, Bar Chart, cables 1,2,3, part 1, Cable minimum bending radius in
straight section.............................................................................................................7-101
Figure 7-94 Pipe, Bar Chart, cables 1,2,3, part 2, Cable minimum bending radius in
straight span.................................................................................................................7-101
Figure 7-95 Pipe, Bar Chart, cables 1,2,3, part 1, Maximum sheath strain..........................7-102
Figure 7-96 Pipe, Bar Chart, cables 1,2,3, part 2, Maximum sheath strain..........................7-102
Figure 7-97 Pipe, Bar Chart, cables 1,2,3, part 1, Maximum sidewall force in pipe band.....7-103
Figure 7-98 Pipe, Bar Chart, cables 1,2,3, part 2, Maximum sidewall force in pipe band.....7-103
Figure 7-99 Duct system: Variation in axial force at position B during load cycling 15-
90
o
C .............................................................................................................................7-104
EPRI Licensed Material

xxxi
Figure 7-100 Duct system: Comparison of the 1st and 30
th
patterns at 90
o
C
Foreshortened view, scale 1:1......................................................................................7-104
Figure 8-1 Photographic comparison of 138kV 1500kcm and 230kV 2500kcm cables........... 8-1
Figure 8-2 138kV 1500kcmil cable EA
4
axial stiffness values compared to the previous
EA
ref
reference set ............................................................................................................ 8-2
Figure 8-3 Force-temperature characteristics of 69kV-345kV cables in the rigid-bar
condition.......................................................................................................................... 8-4
Figure 8-4 230kV 2500kcmil lead sheathed cable in a 152mm duct, with 25mm
clearance......................................................................................................................... 8-6
Figure 8-5 230kV 2500kcmil lead cable, (EA, , )
4
: Force-temperature graphs at A and
B ..................................................................................................................................... 8-8
Figure 8-6 230kV 2500kcmil lead cable, (EA, , )
4
: Patterns at 125

C................................. 8-9
Figure 8-7 69kV 1500kcmil CSA cable: Force-temperature characteristics at A and B..........8-10
Figure 8-8 69kV 1500kcmil CSA cable: Thermomechanical wave pattern at 125
o
C ..............8-11
Figure 8-9 345kV 3000kcmil CSA cable,(EA,,)
4
: Force-temperature graphs at A and
B, ...................................................................................................................................8-12
Figure 8-10 345kV 3000kcmil CSA cable,(EA,,)
4
: Thermomechanical pattern at 90
o
C,......8-13
Figure 8-11 138kV 1500kcmil foil cable, EA
4
: Force-temperature graphs at A and B.............8-14
Figure 8-12 138kV 1500kcmil foil cable, EA
4
: Pattern at 90
o
C ...............................................8-15
Figure 8-13 138kV 1500kcmil XLPE cables in a 200mm diameter pipe with a clearance
of 117m..........................................................................................................................8-18
Figure 8-14 Pipe, 69kV 1000kcmil foil, clearance 132mm, (EA,,)
4
: Force temperature
graph at A and B ............................................................................................................8-20
Figure 8-15 Pipe, 69kV 1000kcmil foil, clearance 132mm, (EA,,)
4
: Thermomechanical
pattern at 90

C................................................................................................................8-21
Figure 8-16 Pipe, 138kV 1500kcmil foil, clearance 123mm, (EA, , )
4
: Force
temperature graph at A and B ........................................................................................8-22
Figure 8-17 Pipe, 138kV 1500kcmil foil, clearance 123mm, (EA, , )
4
:
Thermomechanical pattern at 90C................................................................................8-22
Figure 8-18 Pipe, 138kV 1500kcmil foil, 150mm clearance, (EA,,)
4
: Force
temperature graph at A and B ........................................................................................8-24
Figure 8-19 Pipe, 138kV 1500kcmil foil, 150mm clearance, (EA,,)
4
: Pattern at 90C.........8-24
Figure 8-20 Duct: Force- % clearance characteristic at A......................................................8-31
Figure 8-21 Duct: Force- % bending stiffness EI characteristic at A......................................8-31
Figure 9-1 Road surface test profiles ..................................................................................... 9-3
Figure 9-2 Waveform of ground vertical velocity representing traffic vibration........................ 9-5
Figure 9-3 Pipe system: model of traffic vibration caused by a moving truck.......................... 9-6
Figure 9-4 Spring analogy of duct system at 90
o
C: before application of truck vibration ........9-10
Figure 9-5 Spring analogy of duct system: transition diagram at initial application of
vibrations........................................................................................................................9-12
EPRI Licensed Material

xxxii
Figure 9-6 Spring analogy of duct system: after disturbance by vibrations ............................9-13
Figure 9-7 Duct system (horizontal): thermomechanical patterns before and after 20m
truck movement..............................................................................................................9-17
Figure 9-8 Duct system (horizontal): axial force before and after 200m truck movement.......9-18
Figure 9-9 Duct system (horizontal): release of cable free length during a) heating to
90
o
C and b) 200m truck movement.................................................................................9-19
Figure 9-10 Duct system (+7.5
o
uphill route): growth of thermomechanical patterns as
truck drives along span...................................................................................................9-20
Figure 9-11 Duct system (+7.5
o
uphill route): details of thermomechanical patterns
before and after truck moves 40m, 55m and 75m...........................................................9-20
Figure 9-12 Duct system (+7.5
o
uphill route): axial force before and after 200m truck
movement ......................................................................................................................9-21
Figure 9-13 Duct system (+3.0
o
uphill route): axial force before and after 200m truck
movement ......................................................................................................................9-23
Figure 9-14 Duct system (-7.5
o
downhill route): axial force before and after 200m truck
movement ......................................................................................................................9-24
Figure 9-15 Duct system (-7.5
o
downhill route): details of thermomechanical patterns
before and after truck has moved 20m...........................................................................9-25
Figure 9-16 Duct system (-3.0
o
downhill route): axial force before and after 200m truck
movement ......................................................................................................................9-26
Figure 9-17 Duct system (-1.0
o
downhill route): thermomechanical patterns before and
after 1, 2 and 3 truck journeys along span......................................................................9-27
Figure 9-18 Pipe system (horizontal): thermomechanical patterns at 90
o
C before truck
movement ......................................................................................................................9-28
Figure 9-19 Pipe system (horizontal): axial force in cable 1 (top) before and after 200m
truck movement..............................................................................................................9-29
Figure 9-20 Pipe system (horizontal): axial force in cable 2 before and after 200m truck
movement ......................................................................................................................9-29
Figure 9-21 Pipe system (horizontal): axial force in cable 3 before and after 200m truck
movement ......................................................................................................................9-30
Figure 9-22 Pipe system (horizontal): thermomechanical patterns at 90
o
C after truck
movement along 200m span ..........................................................................................9-31
Figure 9-23 Pipe system (+7.5
o
uphill route): details of thermomechanical patterns
before and after truck has moved 100m.........................................................................9-32
Figure 9-24 Pipe system (-7.5
o
downhill route): details of thermomechanical patterns
before and after truck has moved 10m...........................................................................9-33
Figure 10-1 Diagram of model route......................................................................................10-2
Figure 10-2 Short circuit force diagram.................................................................................10-3
Figure 10-3 Ducts in flat formation. Thermomechanical patterns at 90
o
C in 45m length
next to bend ...................................................................................................................10-9
Figure 10-4 Ducts in flat formation before the 50kA short circuit, position B........................10-10
Figure 10-5 Ducts in flat formation during the 50kA short circuit, position B........................10-11
EPRI Licensed Material

xxxiii
Figure 10-6 Ducts in flat formation during the 50kA short circuit, position B........................10-12
Figure 10-7 Ducts in flat formation after the 50kA short circuit, position B...........................10-13
Figure 10-8 Ducts flat formation. Distribution of axial force in the conductor of cable 3
span before, during and after the 50kA short circuit......................................................10-14
Figure 10-9 Ducts in flat formation. Axial force in cable 3 at position A before, during
and after 50kA short circuit ...........................................................................................10-15
Figure 10-10 Ducts in flat formation. Axial force in cable 3 at position B before, during
and after 50kA short circuit ...........................................................................................10-15
Figure 10-11 Ducts in trefoil formation. Thermomechanical patterns at 90
o
C before the
short circuit in 45m next to bend...................................................................................10-19
Figure 10-12 Ducts in trefoil formation during the 50kA short circuit....................................10-19
Figure 10-13 Ducts in trefoil formation during the 50kA short circuit....................................10-20
Figure 10-14 Ducts in trefoil formation after the 50kA short circuit.......................................10-21
Figure 10-15 Ducts in trefoil formation. Distribution of axial force along span in cable 3
before, during and after 50kA short circuit ....................................................................10-21
Figure 10-16 Ducts in trefoil formation. Axial force in cable 3 at position A before, during
and after 50kA short circuit ...........................................................................................10-22
Figure 10-17 Ducts in trefoil formation. Axial force in cable 3 at position B before, during
and after 50kA short circuit ...........................................................................................10-23
Figure 10-18 Three cables in a pipe system. Thermomechanical patterns at 90
o
C in the
45m next to bend..........................................................................................................10-23
Figure 10-19 Pipe system during 50kA short circuit at position B........................................10-24
Figure 10-20 Pipe system during 50kA short circuit at position B........................................10-24
Figure 10-21 Pipe system after 50kA short circuit at position B...........................................10-25
Figure 10-22 Three cables in pipe. Distribution of axial force in conductor along the
cable 3 span before, during and after 50kA short circuit ...............................................10-26
Figure 10-23 Ducts in trefoil formation. Axial force in cable 3 at position A before, during
and after the 50kA short circuit .....................................................................................10-27
Figure 10-24 Ducts in trefoil formation. Axial force in cable 3 at position B before, during
and after 50kA short circuit ...........................................................................................10-27
Figure 10-25 Ducts in flat formation, -3
o
slope, during 20kA short circuit at position B.........10-31
Figure 10-26 Ducts in flat formation, -3
o
slope, thermomechanical patterns during 20kA
short circuit at position B ..............................................................................................10-32
Figure 10-27 Ducts in flat formation, -3
o
slope. Distribution of force along the cable 3
span, before, during and after 20kA short circuit...........................................................10-33
Figure 10-28 Ducts in flat formation, -3
o
slope. Axial force in cable 3 at position A during,
before and after 20kA short circuit ................................................................................10-34
Figure 10-29 Ducts in flat formation, -3
o
slope. axial force cable 3 at position B during,
before and after 20kA short circuit. ...............................................................................10-34
Figure 10-30 Ducts in trefoil; during 20kA short circuit at position B....................................10-35
EPRI Licensed Material

xxxiv
Figure 10-31 Ducts in trefoil formation, -3
o
slope, thermomechanical pattern during 20
kA short circuit at position B .........................................................................................10-36
Figure 10-32 Three cables in pipe, -3
o
slope during 20kA short circuit ................................10-37
Figure 10-33 Three cables in pipe, -3
o
slope. Thermomechanical patterns at 90
o
C in
150m next to bend........................................................................................................10-38
Figure 11-1 138kV pipe system route profile showing joint position.......................................11-2
Figure 11-2 Thermomechanical deformation inside 138kV joint casing at 90
o
C: elevation
view................................................................................................................................11-3
Figure 11-3 Thermomechanical deformation inside 138kV joint casing at 90
o
C: plan view....11-3
Figure 11-4 Axial force from cable 3 on A-X side of joint .......................................................11-5
Figure 11-5 Axial force from cable 3 on X-B side of joint .......................................................11-5
Figure 11-6 Differential force across joint 3, (negative movement is from joint to bend) ........11-6
Figure 11-7 Differential force across all three joints, (negative movement is from joint to
bend)..............................................................................................................................11-7
Figure 11-8 Maximum differential force, (F
A
-F
B
), where F
A
is the force at the end of the
straight section and F
B
is the force at the entrance to the bend in the reference route....11-7
Figure 11-9 Outline dimensions of 138kV one-piece joint showing key positions ..................11-9
Figure 11-10 Reference positions in 138kV one-piece joint.................................................11-10
Figure 11-11 FEA model of 138kV joint stretched onto the cable for the bending study ......11-15
Figure 11-12 FEA model of 138kV joint for the cable pull-through study ..............................11-16
Figure 11-13 FEA model of 138kV joint bent to 2m radius ..................................................11-17
Figure 11-14 Distribution of radii within 138kV joint for different applied bending radii..........11-17
Figure 11-15 Interfacial pressure contours within 138kV joint: reference case straight........11-19
Figure 11-16 Interfacial pressure distribution within 138kV joint: reference case straight ....11-19
Figure 11-17 Interfacial pressure contours within 138kV joint: bent to 2m radius ................11-20
Figure 11-18 Interfacial pressure distribution within 138kV joint: bent to 2m radius.............11-21
Figure 11-19 Variation of interfacial pressure at cable LV shield termination with
different bending radii under different conditions ..........................................................11-22
Figure 11-20 Variation of Interfacial pressure at HV electrode to EPR insulation
interface with different bending radii under different conditions.....................................11-22
Figure 11-21 Interfacial pressure contours in 138kV joint bent to 2m radius: minimum
stretch sensitivity case..................................................................................................11-23
Figure 11-22 Variation of pull-through force with cable movement and interfacial slip; the
data points are distributed uniformly at 3.44mm intervals. RH and LH refer to right
and left hand ends of molding.......................................................................................11-25
Figure 11-23 Distortion of one-piece molding during cable pull-through with cable
coefficient of friction of 0.5............................................................................................11-27
Figure 11-24 Interfacial pressure in one-piece molding with coefficient of friction of 0.5,
prior to applying pull-through force ...............................................................................11-27
Figure 11-25 Interfacial pressure in one-piece molding with coefficient of friction of 0.5;
formation of void during cable pull-through...................................................................11-28
EPRI Licensed Material

xxxv
Figure 12-1 Cable deflection with cleats at ends of joint moldings.........................................12-2
Figure 12-2 Cable deflection with trefoil cleats at ends of casing ..........................................12-3
Figure 12-3 General assembly showing positions of cleats and spiders................................12-3
Figure 12-4 Trefoil cleat in smallest sleeve ...........................................................................12-4
Figure 12-5 Centralizing spider in medium sleeve.................................................................12-5
Figure 12-6 Joint constraint in large sleeve, style A ..............................................................12-6
Figure 12-7 Perspective view of model of offset joints in pipe system...................................12-8
Figure 12-8 100mm long clamp cleat, as tested at EPRI solutions, with springs added.........12-8
Figure 12-9 Modeled withstand characteristic of clamp cleat from test results ......................12-9
Figure 12-10 Deflection of cables between clamp cleats, elevation view.............................12-10
Figure 12-11 Deflection of cables between clamp cleats, plan view....................................12-10
Figure 12-12 Model of generic clamped offset used in sensitivity study ..............................12-11
Figure 12-13 Force distribution along offset, 19 cleats, 90
0
angle, 15D bend radius ............12-13
Figure 12-14 Effect of a) increasing the number of cleats and b) addition of stronger
clamp cleat(s), 60
0
angle and 15D bend radius .............................................................12-14
Figure 12-15 Summary of sensitivity study: Percentage of cleats holding without slip.........12-15
Figure 12-16 Summary of sensitivity study: Number of cleats holding without slip ..............12-15
Figure 12-17 Summary of sensitivity study: Variation of output force with offset
geometry and cleat characteristic .................................................................................12-16
Figure 12-18 Pipe System: Improved design of offset joints in vertical alignment on walls
of manhole, two pipe circuits are shown.......................................................................12-17
Figure 12-19 Input/output force characteristic across ten cleats, 30
0
15D model, 138kV
cable, 6500/10833N cleat design .................................................................................12-18
Figure 12-20 230kV duct manhole with expansion cleat, cable stocking and anchor joint
constraints....................................................................................................................12-20
Figure 12-21 Diagram showing operation of expansion cleat ..............................................12-21
Figure 12-22 Plan view of deflection of cable at expansion cleat offset, on left-hand end
of manhole ...................................................................................................................12-21
Figure 12-23 Plan view of deflection of cable at stocking offset, on right-hand end of
manhole .......................................................................................................................12-22
Figure 12-24 Expansion cleat equilibrium diagram..............................................................12-23
Figure 12-25 Chart of force at entry to the manholes at 90
o
C in the first and fifth load
cycles...........................................................................................................................12-23
Figure 12-26 Manhole layout showing clamp cleats, 60
0
, 12D radius offset and joint
guide cleat....................................................................................................................12-25
Figure 12-27 Diagram showing input, output and differential forces across clamped
manhole .......................................................................................................................12-25
Figure 12-28 Input/output force characteristic across 14 cleats in the 60
0
, 15D clamped
offset model, with cleat withstand strengths of 6500N ..................................................12-26
Figure 13-1 Two spans in a 138kV pipe system showing horizontal and vertical bends ........13-3
EPRI Licensed Material

xxxvi
Figure 13-2 Distribution of wave patterns in two spans of a 138kV pipe system at 90
o
C .......13-3
Figure 13-3 Thermomechanical clearance of 25mm in an example 230kV XLPE
2500kcmil duct system...................................................................................................13-5
Figure 13-4 Example of thermomechanical clearance of 117mm in an example 138kV
1500kcmil pipe system...................................................................................................13-5
Figure 13-5 Effect of varying duct to cable clearance for the 138kV reference cable.............13-6
Figure 13-6 Effect of varying pipe to cable clearance for the 138kV reference cable.............13-7
Figure 13-7 Reference span..................................................................................................13-8
Figure 13-8 Force-temperature characteristic for XLPE cables constrained as rigid bars ..13-14
Figure 13-9 Helical and sine wave patterns at 90
o
C in a 138kV duct with 150mm
clearance......................................................................................................................13-18
Figure 13-10 Waves at 90
o
C in a 138kV 1500kcmil pipe system with 150mm clearance ....13-18
Figure 13-11 Waves at 125
o
C in a 138kV 1500kcmil pipe system with 150mm clearance ..13-19
Figure 13-12 Incipient half wave formed in a 138kV cable by a 5
o
bend at 37.5
0
C...............13-21
Figure 13-13 Application of bending moments and growth of waves from pipe bend...........13-22
Figure 13-14 Frictional constraint force and length, after formation of incipient wave..........13-25
Figure 13-15 Nonlinear absorption of strain by increase in sine wave amplitude...............13-26
Figure 13-16 Large amplitude sinusoidal wave patterns at a turn-over temperature of
50
0
C .............................................................................................................................13-27
Figure 13-17 Force-temperature characteristic, 138kV 1500kcmil cables with pipe
clearance of 123mm. Span end A on left. Pipe entry B on right ....................................13-27
Figure 13-18 Force-temperature characteristic 230kV 2500kcmil cable with duct
clearance of 38mm. Span end A on left. Pipe entry B on right. .....................................13-28
Figure 13-19 Thermomechanical pattern distributions in the double span at 15
o
C, 30
o
C,
45
o
C, 60
o
C, 75
o
C and 90
o
C..........................................................................................13-30
Figure 13-20 A close-up view of patterns in left hand side of route at 45
0
C and 90
o
C...........13-31
Figure 13-21 A close-up view of patterns in right hand side of route at 45
0
C and at 90
o
C....13-31
Figure 13-22 Deformation patterns inside 138kV joint casing have spread to form wave
patterns in adjacent pipes. Perspective view at 90
o
C....................................................13-32
Figure 13-23 Preferred locations of manholes.....................................................................13-34
Figure 13-24 A differential force and slip interface in a 138kV one-piece straight joint .........13-49
Figure 13-25 400kV anchor straight joint of the prefabricated composite type, showing
differential force taken by anchor..................................................................................13-50
Figure 13-26 Generic design of clamp cleat for XLPE cable................................................13-51
Figure 13-27 Pipe System: In-line joint casing.....................................................................13-55
Figure 13-28 Pipe System: Joints offset on manhole wall: plan, elevation and end view.....13-56
Figure 13-29 Pipe System: Plan view of manhole showing cable offset angle and cleat
positions.......................................................................................................................13-57
Figure 13-30 Duct System: Offset clamped cleat constraining system.................................13-58
Figure 13-31 Duct System: Expansion cleat and S bend constraining system......................13-59
EPRI Licensed Material
xxxvii
LIST OF TABLES
Table 2-1 Critical angles for cable sliding..............................................................................2-17
Table 3-1 List of available cables ........................................................................................... 3-2
Table 3-2 Parameters derived from tests on a 138kV 1500kcmil XLPE cable .......................3-23
Table 3-3 Parameters derived from tests on a 230kV 2500kcmil XLPE cable .......................3-24
Table 3-4 Parameters derived from tests on types of constraint on 138kV 1500kcmil
XLPE cable ....................................................................................................................3-30
Table 3-5 Parameters derived from types of constraint on 230kV 2500kcmil XLPE cable.....3-31
Table 4-1 Cable minimum bending radii ................................................................................. 4-2
Table 4-2 Minimum bend radii for cable inside preformed pipe and duct bends ..................... 4-3
Table 4-3 Bend radii for design of joint manholes................................................................... 4-4
Table 4-4 Limiting value of bend radii for the main run of 138kV and 230kV XLPE cables ..... 4-4
Table 4-5 Limiting bend radii in joint chambers for 138kV and 230kV XLPE cables ............... 4-4
Table 4-6 Limiting value of pulling-in force for 138kV and 230kV XLPE cables ...................... 4-5
Table 4-7 Properties of sheath materials................................................................................ 4-8
Table 4-8 Limiting values of strain for lead sheathing materials ............................................4-10
Table 4-9 Sheath fatigue limits..............................................................................................4-13
Table 4-10 Insulation deformation and pressure measurements on XLPE cables.................4-16
Table 4-11 Lateral force limits for 138kV and 230kV XLPE cables........................................4-18
Table 5-1 Modulus of XLPE in tension at different temperatures............................................ 5-7
Table 5-2 Elastic modulus of XLPE in compression ............................................................... 5-8
Table 5-3 Effective axial and bending stiffnesses for XLPE insulation in a 138kV
1500kcmil cable..............................................................................................................5-10
Table 5-4 Effective axial and bending stiffnesses for XLPE insulation in a 230kV
2500kcmil cable..............................................................................................................5-11
Table 5-5 Example route geometry sheet for 138kV pipe span .............................................5-17
Table 5-6 Example input sheet for the joint casing in a 138kV pipe system...........................5-20
Table 6-1 Example calculation of maximum theoretical cable thrust....................................... 6-3
Table 6-2 Dimensions of the incipient thermomechanical pattern........................................... 6-6
Table 6-3 Length of route experiencing reduced force-single kink.........................................6-11
Table 6-4 Wavelengths at 90
o
C.............................................................................................6-17
Table 6-5 Temperatures at which thermomechanical patterns reach the ends of the
span ...............................................................................................................................6-18
EPRI Licensed Material

xxxviii
Table 6-6 Axial force at the ends of the cable span-multiple kinks.........................................6-19
Table 6-7 Axial Force at 90
o
C at the ends of the cable span-multiple kinks. ..........................6-19
Table 6-8 Axial forces at the span ends, three cables with multiple kinks..............................6-23
Table 7-1 Test parameters to validate normalization.............................................................. 7-5
Table 7-2 Calibration between time and temperature............................................................. 7-6
Table 7-3 Test case: comparison of axial forces .................................................................... 7-7
Table 7-4 Cable to Pipe Clearances....................................................................................... 7-8
Table 7-5 Lengths of straight pipe modeled ........................................................................... 7-9
Table 7-6 Bend Radii ............................................................................................................. 7-9
Table 7-7 Bend Angles..........................................................................................................7-10
Table 7-8 Slope Angle...........................................................................................................7-10
Table 7-9 Coefficient of friction..............................................................................................7-11
Table 7-10 Total Cable Weights............................................................................................7-17
Table 7-11 Values of coefficients of thermal expansion.........................................................7-17
Table 7-12 Bending stiffness EI for the cable designs...........................................................7-18
Table 7-13 Axial Stiffness EA for cable designs ....................................................................7-19
Table 7-14 Torsional stiffness GJ for cable designs ..............................................................7-20
Table 7-15 Axial stiffness EA for the complete 138kV and 230kV Cables .............................7-21
Table 7-16 Duct and Pipe Systems: Comparison of peak force and turn-over
temperatures for pattern initiation by either a cable kink, a 5
o
bend or a 90
o
bend...........7-23
Table 7-17 Duct at position A: Ranking list of parameters by peak force at turn-over
temperature....................................................................................................................7-37
Table 7-18 Duct System at A: Ranking by type of effect........................................................7-38
Table 7-20 Duct at position B: Ranking list of parameters by peak force at turn-over
temperature....................................................................................................................7-39
Table 7-21 Duct System at B: Ranking by type of effect........................................................7-40
Table 7-22 Pipe cable 3 at position A: Ranking list of parameters by peak force at
turnover temperature......................................................................................................7-41
Table 7-23 Pipe System at A: Ranking by type of effect........................................................7-42
Table 7-24 Pipe cable 3 at position B: Ranking list of parameters by peak force...................7-43
Table 7-24 Pipe System at B: Ranking by type of effect.........................................................7-44
Table 7-26 Duct System: Examples of thermomechanical wave patterns..............................7-46
Table 7-27 Pipe System: Examples of thermomechanical wave patterns in cable 2
(bottom)..........................................................................................................................7-47
Table 7-28 Duct system: Examples of peak axial force levels ...............................................7-49
Table 7-29 Pipe System: Examples of peak axial force levels...............................................7-49
Table 7-30 Duct system: Examples of maximum sidewall force in the bend...........................7-50
Table 7-31 Pipe system; Examples of maximum sidewall forces in the bend ........................7-51
Table 7-32 Duct System: Examples of minimum cable bending radii in the straight duct.......7-52
EPRI Licensed Material

xxxix
Table 7-33 Pipe System: Examples of minimum cable bending radii in the straight pipe.......7-52
Table 7-34 Duct System: Examples of maximum sheath strain in straight section ................7-53
Table 7-35 Pipe system: Examples of maximum sheath strain in straight section .................7-54
Table 8-1 Details of cables for the duct system studies.......................................................... 8-4
Table 8-2 Duct systems: axial forces at positions A and B ..................................................... 8-7
Table 8-3 138kV duct system with 150mm clearance, effect of varying axial stiffness EA.....8-16
Table 8-4 Axial forces in duct systems listed in order of duct clearance................................8-16
Table 8-5 Details of cables for the pipe system studies.........................................................8-17
Table 8-6 Pipe systems: peak axial forces at positions A and B............................................8-19
Table 8-7 Axial forces in pipe systems listed in order of increasing clearance.......................8-25
Table 8-8 Geometric parameter ratios for both duct and pipe systems..................................8-27
Table 8-9 Cable parameter ratios and graphs for duct and pipe systems..............................8-27
Table 8-10 Peak axial force factors for the 69kv 1500kcmil duct system at position A...........8-30
Table 8-11 Peak axial force factors for the 69kV 1500kcmil duct system at position B..........8-30
Table 8-12 Force factors for duct system calculations...........................................................8-32
Table 8-13 Force factors for pipe system calculations...........................................................8-33
Table 8-14 Pipe systems: difference in peak axial force between calculated and model
methods .........................................................................................................................8-34
Table 8-15 Duct system: difference in peak axial force between calculation and model
methods .........................................................................................................................8-36
Table 9-1 Vertical displacement of model cable resulting from vibration of the duct ............... 9-8
Table 9-2 Effect of traffic vibrations on duct routes with different slope angles......................9-16
Table 9-3 Effect of traffic vibrations on pipe routes with different slope angles......................9-17
Table 10-1 Thermomechanical pattern lengths-Ducts in flat formation 50kA case.................10-8
Table 10-2 Length of thermomechanical patterns at 90
o
C before, during and after the
short circuit...................................................................................................................10-11
Table 10-3 Peak axial forces in cable 3 at end of span, position A......................................10-16
Table 10-4 Peak axial force in cable 3 at B, entrance to bend.............................................10-16
Table 10-5 Sidewall force cable 1 (inside cable)-duct flat ....................................................10-17
Table 10-6 Sidewall force cable 2 (centre cable)-duct flat ...................................................10-18
Table 10-7 Sidewall force cable 3 (outside cable)-duct flat..................................................10-18
Table 10-8 Pattern lengths at +7.5 degree slope for a 50kA short circuit.............................10-28
Table 10-9 Pattern lengths at +3 degrees slope for a 50kA short circuit..............................10-29
Table 10-10 Pattern lengths at 0 degree slope for a 50kA short circuit................................10-29
Table 10-11 Pattern lengths -3 degree slope for a 50kA short circuit ..................................10-29
Table 10-12 Pattern lengths at -7.5 degree slope for a 50kA short circuit............................10-30
Table 11-1 Cable strain stored in the casing and movement through casing.........................11-4
Table 11-2 Ranking of bending parameters on interfacial pressure at taper base ...............11-24
EPRI Licensed Material

xl
Table 11-3 Cable movement and force for slip at different coefficients of friction ................11-26
Table 11-4 Void formation under LV stress cone in molding during cable pull-through........11-28
Table 12-1 Sizes of clamp cleat used in the modeling work ..................................................12-9
Table 12-2 Geometry of offsets modeled in the sensitivity study .........................................12-11
Table 13-1 Geometric parameter ratios to calculate axial force for both duct and pipe
systems........................................................................................................................13-10
Table 13-2 Cable parameter ratios to calculate axial force for duct and pipe systems.........13-10
Table 13-3 Reference axial forces.......................................................................................13-11
Table 13-4 Locking effect of bend angle in withstanding differences in axial force ..............13-16
Table 13-5 Values of bending stiffness EI ...........................................................................13-24
Table 13-6 Details of thermomechanical patterns in Figure 13-19.......................................13-30
Table 13-7 Duct at position A: Ranking list of parameters by peak force at turn-over
temperature..................................................................................................................13-38
Table 13-8 Duct at position B: Ranking list of parameters by peak force at turn-over
temperature..................................................................................................................13-39
Table 13-9 Pipe cable 3 at position A: Ranking list of parameters by peak force at
turnover temperature....................................................................................................13-40
Table 13-10 Pipe cable 3 at position B: Ranking list of parameters by peak force...............13-41
Table 13-11 Minimum bend radii for preformed pipe and duct bends..................................13-44
Table 13-12 Bend radii in joint manholes ............................................................................13-44
Table 13-13 Pressure Limits on XLPE insulation and PE Jacket .........................................13-45
Table 13-14 Minimum radii of curvature for free cable in ducts and pipes ...........................13-46
Table 13-15 Sheath fatigue limits........................................................................................13-47


EPRI Licensed Material
1-1
1
INTRODUCTION
General
This report describes the thermomechanical and gravitational performance of extruded XLPE
insulated, transmission class cables installed in both pipe and duct-manhole systems. Advanced
engineering modelling and computation techniques have been applied to analyse the complex
movements and forces that occur within ducts and pipes in long length routes of transmission
class, XLPE cable. Thermomechanical and gravitational phenomena are identified and the
reasons for their creation are described. The engineering means of reducing and controlling their
effects on the service performance of cables and joints are given. The contributory effects of load
cycling, road traffic vibration and short circuit forces are quantified. Based on this work a guide
for the design of pipe and duct-manhole systems has been compiled. This work is a major step
forward in the understanding and design of pipe and duct-manhole systems.
The merits of pipe and duct-manhole systems are a) ease and speed of cable installation with
minimum disturbance to the public and b) good protection from third party damage. Once
installed, ducts and pipes are reusable assets, as the cables can be pulled out and replaced without
the need to disturb roads, sidewalks, and countryside. These advantages have been long
recognized in North America. This led to the evolution and extensive use over the past 40 years
of paper and paper laminate insulated cables in high pressure fluid filled (HPFF) and high
pressure gas systems (HPGF).
Five years ago service experience commenced with pipe and duct systems containing dry designs
of solid dielectric, XLPE cables at 138kV and 230kV system voltages. In these cables the
insulation and shields are formed from extruded cross-linked polyethylene (XLPE)
[i]
. Examples
of two constructions are shown in Figure 1-1. The 138kV cable has a 1500kcmil circular
stranded copper conductor, triple extruded XLPE insulation and semiconducting shields, a
double copper tape shield conductor (now superseded by copper wires) and an aluminum foil co-
polymer water barrier bonded to an extruded polyethylene jacket; with an overall diameter of
83mm. The 230kV cable has a 2500kcmil Milliken conductor construction comprised of five
segments of stranded copper, triple extruded XLPE insulation and shields, two temperature
sensing optical fibers housed in stainless steel micro-sheaths, an extruded lead alloy sheath and
an extruded polyethylene jacket; with an overall diameter of 125mm. The constructions of XLPE
cables and their joints are significantly different from those of the service proven paper cables,
their material properties and temperature dependences. The differences are such that XLPE
cables in pipes and ducts cannot be expected to exhibit the same mechanical performance
characteristics and service experience as paper cables.
EPRI Licensed Material

Introduction
1-2

Figure 1-1
138kV 1500kcmil and 230kV 2500kcmil cables from North American installations
After pulling-in the XLPE cable(s) into the pipe or duct, the free space is left empty. The
consequences are that the cables are free to experience forces and movements due to a)
gravitational sliding down slopes and b) thermal expansion in the longitudinal and lateral
directions. As described in this report the movements and forces should be allowed for in the
engineering design of the cable system to ensure that reliable operation and the required service
life are achieved. The behavior of cables in pipes and ducts in long route lengths of complex
geometry, containing multiple bends and slopes, is a difficult engineering problem, which has
previously defied investigation by experimental measurement and by analytical mathematics.
The reasons will become apparent in the report.
Pipe Systems
A pipe system comprises one large diameter pipe into which are pulled three single core cables,
Figure 1-2. The pipe span is typically 1500ft long, this length being determined by a) the
maximum length of cable that can be transported on the shipping reel, b) the maximum permitted
pulling load on the nose of the cable and c) the maximum sidewall load exerted at the bends. The
pipe size is selected to provide a typical clearance of 1.5 to 2 inches in excess of the diameter of
the circumscribed circle encompassing the three cables in trefoil formation. This clearance is
selected to permit the three cables to be pulled without damage through the pipe around the
horizontal and vertical bends present in the particular route. Larger clearances are avoided to
prevent the cables departing from their trefoil formation and crossing in line abreast to jam
within the pipe
[ii,iii]
. The overall consequence in a pipe system is that the eccentric clearance
between the outer diameter of each individual cable and the pipe inner diameter is large (4 to 6
inches).
EPRI Licensed Material

Introduction
1-3
The pipe span is formed from steel tubes which are welded together in up to 40ft lengths. The
ends of the pipe span are terminated within a jointing chamber, termed a manhole. In the
manhole the cable lengths are connected together by hand assembled straight joints. Upon
completion, the joints are bound together in trefoil formation. They are then sealed within a
casing formed of telescoped steel tubes welded to each other and to the smaller diameter pipes
containing the cable spans. Thus the joints are a mechanical continuation of the cable system and
so experience the gravitational sliding and thermomechanical forces exerted by the adjacent
cable spans. The joints are not usually mechanically anchored to the casing and so are
comparatively free to move longitudinally and radially within the casing.

Figure 1-2
138kV 1500kcmil XLPE cables in a pipe with an example clearance of 117mm (4.6)
The design of the pipe systems containing XLPE cables evolved from the HPFF system which
contains a cable core lapped with paper tape insulation, metal shielding tapes, sealing tapes and
skid wires. Because of the role of the steel pipe it was unnecessary to provide the cable with an
extruded metal sheath, a ground return conductor, or an anti-corrosion sheath, thus the cores and
the steel pipe were comparatively small in diameter. In the HPFF system a key part of the
electrical insulation is the hydrocarbon fluid which is maintained at a 200psi pressure to prevent
the formation of gaseous voids between the paper tapes. The steel pipe therefore has to be
pressure tight both to contain the insulating fluid and to prevent contamination of the insulation
by the entry of water and air from the surrounding ground. The steel pipe also provides a return
path for the fault current that flows following an electrical short circuit.
XLPE cables are provided with a thicker layer of insulation than paper cables and this results in
an overall increase in both cable and pipe diameter. This is one of the reasons for XLPE pipe
installations having been limited to date to system voltages of 138kV, whereas HPFF cables have
been installed and operated at 345kV. The wish to install 138kV XLPE cables in smaller pipes
EPRI Licensed Material

Introduction
1-4
and ducts, particularly for retrofitting, is one of the main drivers for the development of thin
wall, higher electrically stressed cables
[iv]
.
In a pipe system containing XLPE cable it is unnecessary to fill the pipe with insulating fluid at
high pressure. This introduces the possibility that water may enter the pipe undetected and
diffuse into the XLPE insulation to impair its electrical strength. Each cable is therefore provided
with a watertight barrier, usually a longitudinally applied, co-polymer coated, metallic foil tape
which is encapsulated within and bonded to an extruded polymeric jacket. The use of the foil a)
reduces cable cost compared to an extruded metallic sheath, b) reduces the cable and pipe
diameter, permitting the shipping reel length to be increased and c) reduces cable weight. The
jacket protects the thin foil from corrosion and provides it with the mechanical strength
necessary to resist buckling during bending and puncture under sidewall loads. The jacket
material is usually selected to be a medium to high density grade of polyethylene as this has a
good combination of water resistance, abrasion resistance, mechanical strength and elevated
temperature performance. It is unnecessary to apply a skid wire to protect the jacket from
abrasion during installation.
The insulating property of the jacket prevents the transfer to the steel pipe of a) capacitive
currents flowing radially through the XLPE insulation and b) fault currents attempting to return
along the cable. It is thus necessary to provide the XLPE cable with a ground-return conductor
formed from an annular layer of copper wires or tapes. In manufacture, the extruded cable core is
first wrapped with semiconducting cushioning tapes onto which the ground-return conductor
(also termed a shield conductor) is applied. A final layer comprising semiconducting cushioning
tapes is applied, prior to the application of the copolymer foil and jacket.
For simplicity, XLPE cables in pipe systems are usually solidly bonded, that is, at each joint
position the ground-return conductors of the three phase cables are connected together and to
ground via connections to the steel pipe.
Duct-Manhole Systems
A duct-manhole system comprises three separate smaller diameter ducts which act as conduits
for the cable, Figure 1-3. One single core XLPE cable of typically 1500ft length is pulled into
each duct. The duct route is formed from short lengths, typically 20ft, of non-ferrous tube, these
usually being fiber reinforced epoxy (FRE), or long lengths of extruded medium/high density
polyethylene (MDPE/HDPE). The ducts may be connected together by hand assembled, push-in
fittings, or in the case of MDPE/HDPE, by thermal welds. The three ducts are spaced apart either
in a flat, trefoil or rectangular formation. The three ducts are usually formed into a duct bank
by being cast in-situ within a block of concrete, or a low thermal resistivity fill, such as a
fluidized fill or a compacted cement bound sand (CBS). HDPE ducting may alternatively be
supplied to site on a reel in a continuous length of up to 300ft, this usually being the case when a
directional drilling technique is used to provide a route underneath an obstruction such as a road,
railroad or river. After the drilling operation, the hole is reamed and the ducting pulled in.
The duct simply provides a conduit for the cable. The XLPE cable thus has to be of the self-
contained design, complete with robust layers of anti-corrosion extruded PE jacket, radial
EPRI Licensed Material

Introduction
1-5
metallic water barrier and ground-return conductor. The design of ground-return conductor can
be selected, in order of decreasing robustness, from a) extruded corrugated aluminum sheath, b)
welded stainless steel sheath and copper wires, c) welded copper sheath, d) welded aluminum
sheath and copper wires as necessary, e) lead sheath, or f) a concentric layer of copper wires
within a copolymer aluminum or copper foil water barrier.
As in pipe systems, cable reel lengths are connected together by hand assembled straight joints
housed within a manhole joint chamber. Unlike the pipe system it is usual to locate the joints,
against one wall of the chamber, offset from the line of the duct. Various methods of
constraining the cables and joints have been employed, these are described in detail in later
chapters. The manhole is not completely watertight and so the straight joint is provided with
joint protection, comprising heat shrink sleeves, self-fusing (self-amalgamating) elastomeric
tapes, or compound filled FRE boxes. In specially bonded systems the joint protection also
electrically insulates the joint shield from ground potential.

Figure 1-3
230kV 2500kcmil lead sheathed XLPE cable in a 142mm duct with an example clearance of
25mm
The manhole also provides access for maintenance purposes to cabinets containing the link
connections required a) in a solidly bonded system to connect the metallic sheath/ground
conductors between the phase cables and to ground and b) in a specially bonded system to
EPRI Licensed Material

Introduction
1-6
connect the sheath/ground conductors to sheath voltage limiters (SVLs). In a cross bonded
system links are provided to transpose the sheath/ground conductors between each phase cable. It
is also usual to transpose the cables between the ducts at the entry and exit to the manhole, this
balances electromagnetic induction between the phase conductors. Special bonding is employed
to maintain the current rating by preventing the circulation of sheath/ground conductor currents.
These currents are higher in duct systems because of the wider phase spacings necessary.
Duct systems are suited to transmission class XLPE cable circuits in which a) larger diameter,
higher system voltage cables, such as 230kV and 345kV, are to be installed and b) higher heat
dissipation and current ratings are required to be achieved by spacing the ducts wider apart.
Scope of the Project
The project objectives were to mathematically model, replicate and study the thermomechanical
and gravitational behavior of pipe and duct-manhole systems and, based on the knowledge
gained, to produce a design guide. The following systems were studied:
Single core XLPE cables in the system voltage range of 69kV to 345kV. Cable designs
would be selected from those installed in North America. A wider range of cable
designs/mechanical parameters would be selected for the sensitivity study model to provide
data for the design guide.
Pipe and duct spans with geometries, bends and slopes based on those already employed in
North American installations. Values outside this range would be applied to a sensitivity
study model to provide design information. A range of spans from an in service 138kV pipe
route and a 230kV duct-manhole route would be selected and modeled. Complete routes
comprising many spans would not be modeled, although the modelling technique was
expected to be applicable to these. Vertical pipe and duct-manhole routes are very unusual
and would not be studied.
The thermomechanical performance of straight joint designs would be studied comprising a)
the one-piece factory pre-molded type and b) the thermomechanical anchor type. Cable
terminations and riser pipes would not be modeled, although the findings of the study would
also be applicable to these.
Methods of constraint of the cable and joints within the manhole would be studied. Generic
designs with improved performance would be produced.
The study would model both continuous circuit loadings and load cycling. The steady state
thermal resistances of the cable components would be modeled. Transient loading conditions
with rapid changes temperature would not be considered (e.g. for a 138kV 1500kmil XLPE
cable, rates of temperature change of 0.1
o
C.min
-1
, this being 20% of the cable time
constant).
The tensions and sidewall loadings that occurred during the cable pulling-in operation would
not be studied as methods of calculation were already available.
The performance of the cable span models would be studied from an ambient temperature of
15
o
C, through a rise of 75
o
C, up to the design operating temperature of 90
o
C. This study
EPRI Licensed Material

Introduction
1-7
would be based on material properties derived from measurements from nominally 15
o
C to
90
o
C on a) full size cable samples and b) plaque samples of XLPE insulation
[v]
. Temperature
rises up to 110
o
C would be applied to the model in the sensitivity study to provide data on the
thermomechanical effects of a) starting at a lower ambient temperature of -20
o
C to 90
o
C and
b) of heating from 15
o
C to 125
o
C. The performance under emergency temperature operation
would not be modeled in detail as this was being studied in a separate EPRI project.
Test methods would be written to permit the mechanical parameters of selected XLPE cable
samples to be measured. The measurements would be analysed and the parameters derived
within the thermomechanical study. (EPRI subsequently went to competitive tender and
subcontracted the test work to the EPRIsolutions mechanical laboratory at Haslet TX.)
The effects of traffic vibration and short circuit forces on the thermomechanical and
gravitational performance would be studied. The effect of earthquake seismic vibration
would not be included.
Layout of the Report
The layout is given in the form of a text book and not as a historical account of the work. This
approach was selected to give a clear and logical explanation for the complex distributions of
thermomechanical and gravitational forces and thermomechanical patterns that form inside long
length pipe and duct routes. A historical account of the start of the project is described in the
Interim Report
[vi]
in which the modelling technique and the first modelling results of 138kV pipe
and 230kV duct routes are presented. The distributions of thermomechanical patterns and forces
were recorded for the first time. The phenomenon was observed that cable deformation patterns
were localized to the bends in the route, in particular to the start of the bends and not within the
bends themselves. An understanding of these and other phenomena was gained by designing a
simplified pipe and duct-manhole route with normalized geometry. A detailed sensitivity study
was then performed with variations in cable and pipe parameters. The phenomena were analysed
and explanations were sought based upon the building blocks of basic engineering theory.
The report is written in the form of a text book a) describing the new branch of cable knowledge
and b) recording the results of the cable measurements and thermomechanical studies. The layout
and contents of the report are as follows:
In Chapter 2 the basic theory of individual gravitational and thermomechanical phenomena is
given.
Chapter 3 describes a) two programs of tests on samples of 138kV and 230kV XLPE cables
and b) how the results were analysed to derive the cable parameters for the modelling. The
tests were formulated to derive the mechanical properties of cables and cable cleats. These
tests were undertaken over a nine month period at the EPRIsolutions laboratory at Haslet TX.
Chapter 4 records the limiting parameters that may inhibit service reliability and reduce the
specified service life of cable circuits. This information was obtained from published
literature and from the knowledge of the Investigators.
Chapter 5 describes the modelling procedure commencing with the preparation of geometric
route data and ending with the post processing of the output data necessary to a) visually
EPRI Licensed Material

Introduction
1-8
display the cable deformation patterns and b) graphically display the parameters (of force and
movements etc). The finite element software and methodology is described.
Chapter 6 analyses the thermomechanical behavior of a model comprising a) a one 138kV
cable in a straight duct and b) three 138kV cables in a straight pipe. Different numbers of
perturbations (kinks) are introduced into the model to represent the presence of minor
asymmetries likely to be present in a real cable and pipe span. The growth of
thermomechanical deformation patterns initiated by the kinks and the consequent reduction
of force are studied.
Chapter 7 is the core of the report. It analyses the thermomechanical behavior of a reference
route with normalized geometry containing a straight section and a bend. The reduction is
studied of force resulting from a) the presence of the bend and b) the growth of deformation
patterns from the bend. A comprehensive sensitivity study is undertaken to examine the
thermomechanical performance variations resulting from the ranges of cable properties, pipe
geometries and slope angles likely to be encountered in transmission class XLPE cable
routes. The parameters are then ranked in order of those that produced the highest force at
selected positions along the model span. The sensitivity study data is used as the basis of the
Normalized Span calculation technique.
Chapter 8 analyses the thermomechanical performance of five specific constructions of 69kV
to 345kV system voltage cables, with large and small conductor sizes and with different
combinations of sheaths and jackets. The cables were modeled in the reference route inside
practical pipe and duct diameters. The thermomechanical behavior is compared with that
extrapolated from the sensitivity study in Chapter 7. The suitabilities of a) the explicit finite
element analysis (FEA) modelling technique and b) the normalized span (NS) interpolation
method are evaluated as calculation methods for use in the Chapter 13 design guide.
Chapter 9 considers the dynamic effects of road traffic vibration when applied to the pipe and
ducts in the normalized model, at different angles of slope. The cables were pre-heated to
exhibit thermomechanical cable patterns before the application of the vibrations.
Chapter 10 considers the dynamic effects of short circuit fault current applied to the
thermomechanical patterns in the normalized pipe and duct models at different angles of
slope. The cables were pre-heated to exhibit thermomechanical cable patterns before the
application of the short circuits.
Chapter 11 analyses the mechanical performance of a 138kV pre-molded EPR joint using a
different type of three dimensional model for solution by the implicit FEA technique. The
joint is modeled to possess different levels of radial stretch when assembled onto the cable. It
is then subjected to the range of bending conditions calculated in the previous
thermomechanical studies. The pull-through strength of the molding, when subjected to
thermomechanical cable force is calculated. The objectives were to a) study the degree of
change in the interfacial pressure at the critical electrically stressed interface between the
joint and cable insulations and b) assess whether the electrical performance would be
influenced by the thermomechanical deformation modeled within the joint casing.
Chapter 12 analyzes the methods used to constrain cables and joints within manholes in pipe
and duct systems. The behavior is analyzed using the explicit FEA modeling technique of
existing designs of constraints in use in service installations. Improved designs of constraint
EPRI Licensed Material

Introduction
1-9
are then developed based on a) the compressive and differential cable force levels derived
from the thermomechanical studies in Chapter 7 and 8 and b) the measured withstand
performance of the prototype cleat recorded in Chapter 3.
Chapter 13 is an application guide to the design of duct-manhole and pipe systems, which
includes a) selection of cables and joints, b) methods of calculation, c) selection of the
optimum pipe and duct dimensions, d) positioning of manholes and e) design of manhole
constraints.
Chapter 14 summarizes the overall conclusions of the study; detailed conclusions are given at
the end of each chapter.
Chapter 15 gives recommendations for further work.
Background of the Investigators
The two Investigators have 20-35 years of detailed experience as practicing engineers in the
design of major transmission class cable systems. The relevant parts of their experience to this
project are a) thermomechanical design of XLPE circuits, b) design and manufacture of XLPE
joints to withstand thermomechanical forces and c) the development and application of
engineering modelling techniques to complex mechanical, electrical, thermal and manufacturing
processes. In 1996 they were involved in the development, manufacture and prequalification
testing of 400kV 1600mm
2
XLPE cable and anchor joints in a long term test program for
BEWAG at the CESI laboratories in Italy
[iv].
In the one year test the cables were installed in route
sections containing a duct, a tunnel, close cleating in air and direct burial in the ground. In the
mid to late 1990s they were involved in a) the design of the first installation of a major 230kV
XLPE duct-manhole system in the USA, b) re-cabling a pipe routes in the UK and c) a number
of 132kV XLPE duct installations in the UK.
The subcontracting software house participating in this project provided two engineers
experienced in modelling cables using advanced engineering software. They had previously
worked with the two Investigators over a period of 10 years as part of the same design and
development team.
Nomenclature and System of Units
To avoid the repetition of the phrase pipe and duct-manhole, the generic word pipe will be
used in the description of the thermomechanical phenomena.
The mathematics of the engineering finite element analysis modelling employs the unified SI
(System International dUnites) form of metric units. This is consistent with EPRI practice. The
exception is that the cables will be referred to by the North American unit of conductor cross
sectional area, kcmil. cable design will refer to the conductor cross sectional area in kcmil.
The main SI units are:
Mass: kilogram [kg]
EPRI Licensed Material

Introduction
1-10
Length: meter [m]
Time: second [s]
Temperature: degrees Celsius [
o
C]
Force: Newton [N]



i
Cable System Technology Review of XLPE EHV Cables, 220kV to 500kV, EPRI, Palo Alto, CA:2002.1001846
ii
Underground Transmission Systems Reference Book, 1992 Edition, EPRI, Palo Alto, CA: 1992.7909-01
iii
Rifenburg R.C, Pipe-Line Design for Pipe-Type Feeders, 1275-1287,AIEE Transaction. Vol.9,December 1953
iv
Attwood J.R, Gregory B, Dickinson M, Hampton R.N, Svoma R, Development of High Stress HV and EHV
XLPE Cable Systems, Session Paper No 21-108, CIGRE, 1998.
v
Research to determine the acceptable emergency operating temperatures for extruded dielectric cables, EPRI Palo
Alto CA, Report EL-938, Project 933-1,November 1978.
vi
Mechanical effects on extruded dielectric cables and joints installed in underground transmission systems in North
America, Technical Update Report, August 2002, EPRI, Palo Alto, CA:2002: 1001848.
EPRI Licensed Material
2-1
2
THERMOMECHANICAL AND GRAVITATIONAL
EFFECTS: BASIC THEORY
General
The basic theory given in this chapter was derived to explain the phenomena seen during the
studies of the output from the finite element models.
The chapter first explains the overall thermomechanical and gravitational phenomena and how
they affect the rate of rise of axial thrust and the growth of cable patterns at different positions
along a pipe route.
The basic phenomena are then described as individual cases. An understanding of these cases
will assist the reader in the interpretation of the output from the finite element models in later
chapters. The finite element analysis method does not use these cases at all, but solves Newtons
fundamental laws of motion for each small element of the cable as it moves under the actions of
a) gravity down a slope, b) thermal expansion and c) differential forces generated by adjacent
cables. In the model output the phenomena are found to overlap and interact with each other to
varying degrees along the route. However one of the cases will be found to predominate at each
position in the span for a particular a) temperature rise, b) set of cable parameters and c) pipe
geometry.
It is normal practice for engineers to construct a simple model to solve the forces and movements
in a solid structure of defined geometry. A cable within in a long length of pipe is a more
complex problem. The essence of a cable is that it is designed to be flexible. It is wound onto a
reel for shipping and then is pulled around bends during installation. In consequence it can easily
change its shape into different patterns in service within the pipe. The thermomechanical
problem requires to be solved by a dynamic modelling technique. It is not possible to apply
simple static models of the structure to the problem. The shape of the structure formed by the
pipe route poses a particular challenge to the dynamic modelling technique as the aspect ratio of
span length to cable diameter is very large; for example 30,000:1.
Explanation of the Thermomechanical Effects
Figure 2-1 shows an example span of a duct route containing one straight section, A to B, and
one bend, B to C. The cable ends at A and C are fixed perpendicular to the duct ends, such that
the cables are straight and do not possess curvature. The geometry of the span is explained in
Chapter 7, where it is adopted as the reference normalized span in the sensitivity study.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-2

Figure 2-1
Diagram of a 200m duct route with one bend
For explanatory purposes the pipe system can be considered to be an absorber of thermal
expansion as shown in Figure 2-2, comprised of:
Coil springs connected in series with different elastic compressibilities, each representing a
region along the route comprising a) straight cable, b) cable wave patterns and c) bends.
Cable having characteristics similar to a steel coil spring, in that it can be elastically
compressed to absorb thermal expansion. The degree to which the cable can absorb
expansion depends upon whether it remains straight, or whether lateral deformation patterns
form (wave patterns). Wave patterns are elastically more compressible than the straight cable
and absorb more thermal expansion strain. The cable in a wave region can be considered to
be a coil spring possessing high elastic compressibility.
Geometric positions where cable expansion length can be absorbed, for example within a)
preformed pipe bends, b) the bends in manholes and c) bends in joint casings.
The spring analogy in Figure 2-2 shows a number of different sized coil springs joined together
in series. This is a diagram depicting the compressibility of the wave patterns, it is not a drawing
of the shape of the wave patterns. The diagram shows the regions of different compressive
stiffness. It shows three stages of temperature rise,
1
,
2
and
3
, in which if free, the springs
would expand by L
1,
L
2
and L
3.
In each case the series of springs are compressed by being
pushed back to their original length as it was at ambient temperature:
First stage temperature rise
1
. The small diameter spring represents the low axial
compressibility of the straight cable. This is called the rigid-bar state. Elastic energy is
stored in the axial compression of the cable. The large diameter spring represents the high
apparent compressibility of the cable within the pipe bend. This spring represents the cable
being moved radially outwards and also upwards against gravity, thus storing gravitational
potential energy.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-3
Second stage temperature rise
2
. The bend is now full and the cable can no longer be pushed
into the bend. The cable is held from moving by a combination of high side wall forces and
friction, so it now behaves as being incompressible. The cable in the bend is depicted as a
straight line strut with no compressibility. Small lateral wave patterns now form in the
straight section adjacent to the bend. The wave patterns absorb thermal expansion strain from
the cable system. The waves are represented by the increased compressibility of the medium
diameter coil spring.
Third stage temperature rise
3
. A sudden transition to large amplitude wave patterns has
occurred. The waves have a high absorption of thermal strain. They have the highest
compressibility and are represented by the large diameter spring. As the temperature and
force is increased the waves spread along the cable from the right to the left hand end.

Figure 2-2
Coil spring analogy of thermomechanical effects in a route with a straight cable section
and one bend

EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-4
Start of Cable Heating
Straight Route
As the temperature of the cable increases from the background ambient temperature it attempts
to expand longitudinally and to increase its length by L.
L L = Equation 2-1
Where,
L: length of cable [m]
: coefficient of thermal expansion of the cable in the axial direction [
o
C
-1
]
: temperature rise [
o
C]
However the ends of the cable are fixed (for example at the cable terminations) and this prevents
the cable from expanding freely. In consequence a thermomechanical axial force F in
compression is developed inside the cable. The force is calculated by imagining that the cable
had been allowed to expand freely by L and that a force F had been applied to compress the
cable back to its original length L. The elastic compressibility of the cable is given by Hookes
Law:
L
L EA
F

= Equation 2-2
Where:
F: thermomechanical compressive force [N]
L: cable deformation in compression [m]
L: cable length [m]
E: modulus of elasticity of the cable in compression [N m
-2
]
A: effective cross sectional area of the cable [m
2
]
EA: axial stiffness of the cable in compression [N]
The axial stiffness of the cable EA is the effective value, which is best obtained by measurement.
Alternatively it may be obtained by summing the calculated values for each layer in the cables
construction. The axial stiffness EA has temperature dependence arising from a) the thermal non-
linearity of the XLPE insulation, b) the compressive grip by the XLPE insulation of the
conductor and c) the polymeric jacketing materials.
If the cable is completely prevented from expanding, the maximum rigid-bar force is obtained
by combining Equation 2-1 and Equation 2-2 :
EA F = Equation 2-3
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-5
It is seen that the maximum force F in Equation 2-3 is independent of length. (Although the
locked in strain is dependent on length, as given in Equation 2-1). F is the maximum force that
can be generated in a horizontal route, without slopes.
This force can only occur in a perfectly straight route with no bends in either the pipe or the
cable, so that bending moments do not exist. In pipe and duct systems this is an idealized and
impractical condition that can not occur. Even in a straight route there will be a) slight angular
misalignments at the welded pipe connections, b) kinks in the cable (curvature from the reels
used in manufacture and shipping) and c) curvature in the cable at the cable termination and joint
manhole positions. The closest theoretical case to a perfectly straight cable occurs in a duct-
manhole system in which the cable is a very tight fit in the duct and thus cannot deflect laterally.
In practice a clearance has to be provided to accommodate variations in a) cable and pipe ovality,
b) diameter of the duct joint and c) diameter of the cable pulling-eye.
Route with Bends
Routes with bends are the normal condition in both pipe and duct-manhole systems. Figure 2-1
shows a simple route with a straight section and one bend (this route is used as the basis for the
normalized sensitivity studies described in Chapter 7).
The following thermomechanical events occur in step with the cable temperature rise. The way
they occur results from the cable a) laying horizontally, b) being constrained and controlled
within the cylindrical shape of the pipe and c) being forced to rise up the pipe walls against the
force of gravity when the cable is deflected laterally:
The movement into the pre-formed pipe bend is smooth and continuous.
The first transition from straight cable to incipient waves is smooth and continuous.
The second transition is a sudden dynamic event in which large amplitude waves form. The
initiation of the event is the same as the elastic buckling of a strut in compression
[i,ii,iii,iv]
,
however structural collapse cannot occur because the cable is constrained both by the pipe
walls and by gravity
In the Bend, Position B Figure 2-1
The thermomechanical force pushes the cable into the bend. As it enters, it will initially move
with little resistance across the floor of the pipe outwards to a larger bend radius (the angle of
inclination of the pipe curvature to the horizontal is initially very small). The movement absorbs
the initial thermal expansion strain and limits the increase of the thermomechanical force to a
small value, as shown between 0 and a in the force/temperature graph in Figure 2-3. Note that
in the graph the force is shown in the downward direction (negative), this is the convention used
in the finite element modelling analysis to denote a compressive force.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-6

Figure 2-3
Axial force-temperature characteristic at the entrance to the bend, position B
As the cable moves further outwards, the angle of inclination of the pipe curvature to the
horizontal increases rapidly. In consequence a progressively greater thermomechanical force is
required to lift the weight of the cable against gravity. The furthest outwards the cable can move
occurs when it has been lifted to the 3
o
/c - 9
o
/c axis of the pipe. The maximum distance the cable
can move outwards and upwards is the clearance (Rp-Rc) between the pipe inner radius Rp and
the cable outer radius Rc. The maximum length of cable that can be absorbed in the bend is:
L=(Rp-Rc) Equation 2-4
Where:
L: length of cable that can be absorbed by the bend [m]
Rp: inner radius of the pipe [m]
Rc: outer radius of the cable [m]
: angle of the pipe bend [radians]

It will be seen that the length is dependent on both the clearance (Rp-Rc) and the angle of the
pipe bend. The length is independent of the pipe bend radius Rb, which does not appear in
Equation 2-4. Thus the choice of a pipe with a large cable to pipe clearance (Rp-Rc) and with
either one bend with a large angle , or many bends with a cumulative angle , will
significantly increase the capacity of the route to absorb thermal expansion. For example, one
pipe bend with a radial clearance of 75mm and angle of 90
o
(1.57 radians) will be able to absorb
the thermal expansion of a 85m length of 138kV XLPE cable, heated through a 75
o
C temperature
rise, with a coefficient of expansion of 18.5 x 10
-6

o
C
-1
.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-7
After the cable has moved into the bend, the force will rise in a near linear way with temperature,
as shown in Figure 2-3 from a to b. The slope of the graph
( )
( ) a b
Fa Fb

, will be close to the


maximum force given in Equation 2-3. From an external viewpoint the cable still behaves as
though it is a rigid-bar with low compressibility. However internally the cable has formed
small incipient waves, which are beginning to absorb the thermal strain.
During the temperature rise, the force F rises to the first of two critical transition states:
First Transition State: Inception and Longitudinal Growth of Waves
At the first transition state, incipient waves grow from the bend entrance. They become
perceptible at a very low rise in temperature, ie less than 15
o
C. These waves are in the form of
one to three wave half cycles of very small amplitude, ie less than 15mm. These incipient
waves can only absorb a small amount of thermal strain and so the slope
( )
( ) a b
Fa Fb

becomes
slightly less than the rigid bar value, ie ~97%. The incipient waves cause the cable to appear to
be slightly more elastic in compression and so the rate of rise of force is also slightly reduced.
The transition in the slope of the force/temperature characteristic is smooth and continuous and is
not seen graphically. Examples of the slopes and waveforms computed from the models are
given in Chapter 6 for waves formed from cable kinks and in Chapter 7 for waves formed from
pipe bends. In the Figure 2-3 diagram the first transition can be considered to have occurred at
point a, ie as soon as the cable has finished moving into the pipe bend.
Figure 2-4 shows diagrammatically that, in the section of cable immediately adjacent to the bend,
the cable has become curved in shape. This offset together with the thermomechanical thrust F
1

forms a bending moment. The curve grows in magnitude and forms a half wave:
During thermal expansion the cable enters the bend and is forced to curve upwards and
outwards (following a similar racing curve to that taken on an inclined bank in a motor
racing track). This bending radius R
1
extends outside the pipe bend into the straight section
and so applies a bending moment M
1
to the straight cable. It can be seen that the lever arm
of the moment is the outward movement of the cable in the bend, ie. (Rp-Rc). The bending
moment M
1
is approximately equal to F
1
(Rp-Rc).
The rise of thermomechanical thrust with temperature to F
2
and F
3
further increases the
bending moments to M
2
and M
3
and produces a longitudinal progression of the half waves
along the pipe with radii R
2
and R
3
.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-8

Figure 2-4
Bending moment formation at the entrance to the bend
The cable bending radii are related to the magnitude of the bending moments by the basic
relationship for the elastic bending of a beam
[i,ii,iii,iv]
:
M
EI
R = Equation 2-5
Where,
R: radius of curvature of the cable [m]
M: bending moment causing the lateral deflection [Nm]
E: modulus of elasticity of the cable in bending [Nm
-2
]
I: second moment of area of the cable cross section [m
4
]
EI: cable bending stiffness [Nm
2
]
Cables with a lower bending stiffness EI (i.e. a greater flexibility) will deflect more readily and
form a smaller bending radius. It has been found experimentally (Chapter 3) that the bending
stiffness of an XLPE cable increases with conductor size. It is also increased by the grip of the
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-9
XLPE insulation, which is high at ambient temperature and reduces as the temperature rises.
Overall, cables with smaller conductors will bend more readily.
The progressive longitudinal growth of the half waves is explained by Equation 2-5. When the
bending radius R
1
in the first loop at the bend entrance deepens, its curvature applies a bending
moment M
2
to the adjacent straight cable. A second loop with bending radius R
2
grows to form a
third loop with M
3
and R
3
and so on. A pattern of small amplitude sine waves has been formed.
In the first transition state the thermal strain is stored in the cable in two geometrically different
forms. The first is the locked-in strain in the straight cable, this being equivalent to the stiff
spring analogy shown in Figure 2-2 (top). The second is the strain released into the longitudinal
growth of small amplitude waves represented by the medium stiffness spring in Figure 2-2
(middle). The axial forces in the two springs must always be in equilibrium and, if there is no
friction with the floor of the pipe, the axial forces will be equal. (As will be seen in Chapters 6
and 7 the frictional force is present and is an important parameter in cable thermomechanical
behavior). In the condition of equilibrium the thermal strain is shared between the two springs
with proportionally more strain being stored in the incipient wave pattern. The magnitude of the
force is reduced, but is still at a significant level.
Second Transition State: Turn-over Force and Turn-over temperature
The second transition state is a sudden dynamic event which causes the cable axial force to fall at
point b in Figure 2-3. The temperature b is called the turn-over temperature and the force Fb
is called the turn-over force.
The second transition state occurs when the energy of the system is minimized by the sudden
release of locked-in thermal strain from the straight cable into the sudden growth of large
amplitude waves. The large amplitude waves are represented in Figure 2-2 by the sudden
physical change of the medium stiffness spring (middle) into the large diameter spring of low
stiffness (bottom). The large spring absorbs the strain from the cable (small spring) and thus, by
Hookes law, the axial force falls in proportion. The physical mechanism of how and why this
occurs is explained below:
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-10

Figure 2-5
Graph of cable absorbed in a sine wave for different large ratios of amplitude to
wavelength
For example, taking a small-medium wave amplitude of 25mm* and a wavelength of 5m (a ratio
of 0.5%) permits the absorption of 0.025% of the wavelength. This is an absorption of 1.3mm of
the 5m length. An absorption of 1.3mm is equivalent to the thermal expansion of a 6.8m length
of cable heated through 10
o
C. Although this appears to be small, it is a beneficial capability as it
will release and absorb all of the strain that was previously locked within its own 5m length. The
locked-in strain from the adjacent cable is still of sufficiently high magnitude to flow into the
sine wave and to maintain a high axial force. If cable waves were able to form along the
complete length of the pipe span, all of the locked-in strain would be released for the 10
o
C
temperature rise, (the waves are initially restricted from doing so by friction). * The wave is
resting within the circumference of the pipe. The axial center of the cable can move around a
circle of radius equal to the clearance (Cp-Cc). If the radial clearance is 75mm, the wave
amplitude of 25mm is equivalent to an angular change in amplitude of 19
o
.
The rising temperature increases the axial force to progressively higher magnitudes. The
amplitude of the wave is increased by the cable being pushed up and around the circumference of
the pipe. Taking the extreme example of a large wave of 100mm* amplitude and the same
wavelength of 5m (a ratio of 2%), shows the absorption will increase x 16 to 0.4% of the
wavelength. This is an absorption of 20mm in the 5m length. An absorption of 20mm is
equivalent to the thermal expansion of a 108m length of cable heated through 10
o
C. In
consequence a high proportion of the locked-in strain is released from the rigid cable and the
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-11
axial force falls to a low level. * For a radial clearance (Cp-Cc) of 75mm, the 100mm
amplitude is equivalent to an angular change of 76
o
.
Figure 2-3 shows that the formation of the large amplitude waves has caused the force to
suddenly fall from point b to point c. At point c equilibrium is again reached between the
axial forces within the rigid cable stiff spring and wave flexible spring shown
diagrammatically in Figure 2-2 (bottom).
The formation of the large amplitude waves has prevented the force from rising above Fb. In the
cases studied of pipes and ducts with larger radial clearances of ~75mm, the force did not regain
or rise above this value within the 15
o
C- 90
o
C operating range of the cable. In this circumstance
the axial thrust Fb is the maximum force that can be experienced by a) the cable in the bend and
b) an adjacent manhole. (For the purposes of design, the maximum axial force occurs at position
A at the opposite end of the span, as explained in a following section).
The ratio of the turn-over force Fb to the equilibrium level at Fc is an approximate measure of
the apparent reduction in the axial stiffness EA of the sine wave pattern compared to the straight
cable and is typically 1.5-3.0.
As the temperature is increased further, the family of sine waves grow in number and progress
along the straight section in Figure 2-1 from the bend at B to the end of the cable at A,
continuing to absorb thermal strain as they do so. During the growth of the waves the cable
within the pipe acts as a constant force spring. This is shown diagrammatically in Figure 2-3 by
the force remaining almost horizontal between c and d, despite the large increases in
temperature from b to d. At point d the patterns at the bend B are unable to absorb more
strain and so the force again rise at close to the rigid-bar rate with temperature, such that d-e is
a similar gradient to a-b.
Wave Shapes
The predominant wave shape is a cylindrical sinusoid. Only sine waves can form within pipe
systems, within the range of radial clearances studied of 50-125mm. In duct systems, which
contain a single cable, other types of wave shapes can form:
Large Amplitude Cylindrical Sinusoidal Waves. The sine wave rises on one side of the duct,
passes over the top 12
o
/c position and descends to reach its peak amplitude on the opposite
side, or at the bottom 6
o
/c position. The wave then alternates in direction and returns back
over the top of the pipe. In this way the amplitude of the sine wave is not limited by the
clearance (Rp-Rc) and can alternate through an amplitude of 2(Rp-Rc). From a
downwards plan view of the duct, the cable appears to have formed a sine wave of half the
wavelength. From a longitudinal perspective view the cable appears to have formed a helical
loop. Neither of these appearances are true, the pattern has remained that of a true sine wave
of unchanged wavelength and can be treated in calculations as such.
Helical Waves. The cable does not slide around the duct wall as occurs with a sine wave. A
convex beam of cable rises vertically off the floor of the duct. Looking longitudinally
through the duct, the beam initially takes the shape of a collapsed circular loop (see Chapter
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-12
6). As the temperature and axial thrust increase, the loop becomes a circular helix with an
expanding diameter that opens out until it firmly contacts the pipe wall. Either one single
helical loop may form amongst the sine waves, or many helical waves may form to fill the
duct end to end, depending upon the cable parameters (Chapter 7). Similarly the helical loop
may start to form immediately at the second transition turn-over temperature b, or it may
form amongst patterns of sine waves at higher temperature. Even though the torsional
stiffness of the cable is small (see the Sensitivity Study in Chapter 7), the helical patterns
tend to be neutrally torque balanced. Thus the presence of one clockwise rotating loop is
usually accompanied by a counter clockwise loop, not necessarily immediately adjacent to
each other. (Sine waves are naturally torque balanced as they rotate clockwise and
counterclockwise on each alternating half cycle). The helical wave can absorb more strain
than a sine wave of the same length, the formula is given in Equation 2-17. Sufficient force
and energy is required to lift the cable vertically upwards rather than to slide it around the
duct wall. The competition between the formation of large amplitude sine waves and helical
waves is dictated by the minimum energy condition of the cable system (ie the minimization
of elastic and gravitational potential energy). The ratio of the number of helical to sinusoidal
waves increases significantly when reductions are made in a) the cable to pipe clearance and
b) the cable weight (see Chapter 7).
In pipe systems, three cables are present. The three cables also expand together into the bend
and, as with a single cable, form sinusoidal patterns growing from the entrance of the bend into
the straight section. The three cables each contrive to use the maximum radial clearance (Rp-Rc)
between their individual diameter and the internal diameter of the pipe. This is achieved by the
sine waves inter-twining, such that the gap left by the base of one sine wave is occupied by the
peak amplitude of another cable. An important difference between pipe and duct systems is that
the intertwining sinusoidal patterns prevents the formation of helical loops. Thus the
thermomechanical deformation patterns are more regular in pipe systems than duct-manhole
systems.
The Anchored End of the Route, End A in Figure 2-6
End A is partially insolated from the cable movement into the bend at B and the low thrust that is
present there by the constraint of friction between the cable length and the pipe wall. If the
straight section is long enough, end A will be completely insolated and will behave as though it
was rigidly cleated, with no room to expand. The cable would then exhibit the maximum
increase of thermomechanical force with temperature rise F/ as given in Equation 2-3.
The graph of the force-temperature characteristic at A is given in Figure 2-6. A comparison of
the F/ characteristic at A in Figure 2-6 with that at B in Figure 2-3 shows the difference to be
that the force rises immediately at A. This is because the cable at A is in the rigid-bar condition
with no bend to absorb strain. At B the absorption of the bend delays the force rise until the
temperature has risen by a between 0 and a. Up to the temperature rise b at point b in
Figure 2-6 , the force is always higher at end A of the span compared to end B.
The longitudinal difference in cable axial force between A and B is supported by the distributed
friction between the cable and the pipe wall. The span length and cable properties limit the force
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-13
difference that can be withstood. For example a length of 200m of light cable of 10kg/m weight,
with a low coefficient of friction of 0.1, will support a difference of up to 1,960 N ie a maximum
force gradient of 9.8N/m. When the difference in force between A and B attempts to exceeds this
value, friction can no longer restrain movement. A portion of the locked-in thermal expansion
strain is released from the complete cable length into B, where it will move into the bend. In
consequence the force will be reduced at A.

Figure 2-6
Axial force-temperature characteristic at the fixed cable end, position A
Once the bend is filled, the force starts to rise and incipient waves form. These reduce the force
slightly at B. In some of the cases studied this suddenly increases the cable force differential
above the limit of the frictional constraint, causing the strain to be released by a sudden slipping
movement, accompanied by a momentary reduction in the force-temperature characteristic at
end A.
Turn-over Force Due to Frictional Slip at the Span End A
When the large amplitude waves form at B, the local axial force is reduced, causing the force
differential between A and B to suddenly and significantly increase in magnitude. The frictional
constraint is exceeded and a further length of thermal strain is fed from the straight cable into the
waves at B, causing a rapid growth, both in amplitude and length. This causes a turn-over in
force to occur at end A, as shown in Figure 2-6 at point b. It should be noted that the turn-
over temperature and force at end A will be higher than those at end B ( e.g. ~2.5
o
C and by
~frictional constraint force given in Equation 2-7). The turn-over force at end A is usually found
to be the maximum force that can occur in the span up to 90
o
C (particularly for large clearance
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-14
pipes and ducts). In this circumstance the force is taken as the limiting value for the design of the
cable system.
As the temperature continues to rise, the wave patterns grow towards end A and in most cases
will reach A, having occupied more than 90% of the span length. They drain the strain out of end
A as they approach and cause the force to fall to a lower level, creating a similar force-
temperature characteristic to that at end B.
The waves at end B will become saturated with strain causing the local force to rise, whilst the
force at end A is still falling. A temperature is reached at which the two force-temperature
characteristics cross and the force at end B rises above that at A. At this temperature the
frictional gradients restraining cable movement reverse.
Route with Slopes
The effects of slopes and friction are described in the section below dealing with individual cases
of basic theory. If the route has a slope which is less than the critical sliding angle, for example
5.7
o
for a coefficient of friction of 0.1, then the cable will not slide after installation.
If the route in Figure 2-1 is inclined with end A uppermost, then the rising thermomechanical
thrust will reach a sufficient magnitude to overcome friction and the cable will slide into the
bend. Thus the compressive force at B will increase causing deformation patterns to form at a
lower temperature rise.
If the route is inclined with end A downwards, then the thermomechanical force will have to rise
to a higher level before frictional forces can be overcome to release thermal strain uphill into the
bend. The thermomechanical force at the entrance to the bend will be reduced.
The effect of slopes is modeled in detail in Chapters 7, 9 and 10.
Effects of Short Circuits, Traffic Vibration and Load Cycling
Avalanche Effect
The third temperature stage
3
in Figure 2-2 depicts the cable as being comprised of two
compressed coil springs. This is the usual thermomechanical condition at the cable operating
temperature. The diagram represents the low compressibility of the remaining length of straight
cable and the high compressibility of the region with sinusoidal patterns. However they are not in
balance with each other. The cable shown depicted by the coil spring on the left hand side is
under higher compression and is prevented from moving to balance the loads by the distributed
frictional constraint along the pipe. This is shown diagrammatically in Chapter 9 in figures 9-4,
9-5 and 9-6, in which each turn of the springs is held on a grooved surface. The application of
vertical movement to the pipe or cable, momentarily lifts the cable off the pipe floor and reduces
the frictional constraint. This suddenly releasing a portion of the thermal strain along the cable
length into the wave patterns. This has been named the avalanche effect because of the close
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-15
similarity of the initiation mechanism in which a small vibration overcomes interlayer
friction/shear force to release a weight of snow in an avalanche movement.
Short Circuits
The application of high short circuit forces had a dramatic effect on the sudden release of locked-
in thermal strain and in the growth of cable thermomechanical deformation patterns. Each cable
was thrown from its position on the pipe floor to be compressed outwards on the wall of the pipe.
The compression was particularly pronounced in the wave patterns adjacent to the pipe entrance.
At the end of the short circuit the electromagnetic forces holding the cable in compression were
released. The cable sprang back to its original position on the opposite side of the pipe. In so
doing, the cable was momentarily freed from the pipe surface along the complete length of the
span. This completely eliminated the longitudinal friction constraint and immediately released
the locked-in thermal strain to move from the straight cable into the deformation patterns. The
patterns grew immediately along the 200m length towards end A. Short circuit effects on
horizontal and inclined spans is studied in detail in Chapter 10.
Traffic Vibration
Traffic vibration had a similar effect. The vertical vibration applied to the pipe was small in
magnitude. During the vibration the cable was lifted vertically off the floor of the pipe, thus
momentarily overcoming friction. The effect was to release some of the locked-in thermal strain
from the straight cable at end A into the waves at B, permitting the cable to move longitudinally
over long distances e.g. 150m. This movement helps to explain the phenomenon sometimes
called cable walking or cable surfing observed in both pipe systems and self-contained cable
systems (with loose fitting corrugated aluminum sheaths). It was previously thought that cable
movement only occurred in the direction of the traffic movement. The work in this report shows
that the movement is from a region of high axial force to low axial force. In the majority of cases
the region of low axial force will be adjacent to a bend, manhole or termination. The effect of
traffic induced vibration on horizontal and inclined spans is studied in detail in Chapter 9.
Load Cycling
The application was made of 30 extreme temperature cycles of 15
o
C to 90
o
C to the normalized
span containing one bend, shown in Figure 2-1. A small cyclic variation in maximum force
occurred, as shown in Chapter 7 in figure 7-99. The number of helical waves present amongst the
sine waves reduced from two to one. These small changes were attributed to the presence of a
long straight section of 200m and one bend. In consequence the absorption capacity of the bend
was quickly filled and movement within the bend became locked by sidewall force, thus
providing no scope for cumulative movement during temperature cycling. Spans from service
applications containing many bends were studied (the 138kV pipe span in figure 13-1 in Chapter
13 contains 22 bends and unconstrained joints capable of sliding). The variations in movement
and force occurring at a straight joint during three temperature cycles is given in Chapter 11 in
figures 11-4 to 11-7.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-16
Individual Basic Theory Cases
Sliding Under Gravity and Thermomechanical Force
Figure 2-7 represents one cable element of unit length (black outline) within a length (dotted
outline) resting in an inclined pipe.

Figure 2-7
Cable in an inclined pipe
( ) + = cos sin Wg F F
tm T
Equation 2-6
Where:
F
T
: resultant total force acting on the cable down the slope [N]
F
tm
: cable thermomechanical force acting on the cable down the slope [N]
W: unit weight of cable [kg m
-1
]
g: acceleration due to gravity, 9.81 [m s
-2
]
: angle of slope [degrees]
: coefficient of friction [per unit]
R: reaction force experienced by cable [N]
Sliding with no Thermomechanical Force
This is the case when the cable has just been installed and before it has been jointed and heated,
thus the thermomechanical force F
tm
is zero.
The reaction force,
R=Wg cos
Sliding will occur if,
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-17
Wg sin R Wg cos Equation 2-7
The critical angle above which sliding occurs is,
tan
-1
Equation 2-8
It can be seen that the angle is only dependent on the coefficient of friction and is independent of
the weight of the cable. Generally pipes follow the incline of public roads, which are unlikely to
have slopes of steeper than 25% (1in 4). Table 2-1 shows that if the cable has a coefficient of
friction of 0.1 it will slide at an angle of 5.7
o
. A cable with a PE jacket coated with an installation
lubricant inside a smooth pipe will exhibit a value of 0.1.
Table 2-1
Critical angles for cable sliding
Critical Angle for Sliding =
tan
-1
[degrees]
Slope = sin [%]
0.1 5.7 10
0.2 11.3 20
0.3 16.7 30
0.4 21.8 40
0.5 26.6 50
1.0 45 100
Sliding Downhill with Downward Thermomechanical Force
From Figure 2-7 sliding will occur when
Wg sin + Ftm Wg cos Equation 2-9
Or

Wg
Ftm
cos - sin
Taking the previous case of = 0.1, and if the thermomechanical force was 10% of the weight of
the cable, then the critical angle for slip would be reduced from 5.7
o
to 0
o
.
Sliding Downhill with Upward Thermomechanical Force
From Figure 2-7 sliding will occur when
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-18
Wg sin -Ftm Wg cos Equation 2-10
Or

Wg
Ftm
sin - cos
Taking the previous case of = 0.1, and if the thermomechanical force of 10% of the weight of
the cable was applied, then the critical angle for slip would be doubled from 5.7
o
to 11.5
o
.
Sliding Uphill with Upward Thermomechanical Force
Sliding upwards will occur when
Ftm Wg cos +Wg sin Equation 2-11
Or

Wg
Ftm
cos +sin
Taking the previous case of = 0.1 and the critical angle of downwards slip under gravity being
5.7
o
, the thermomechanical force would have to be greater than 20% of the weight of the cable to
move it up the slope.
Horizontal Sliding under Thermomechanical Force
This is the simple case where the angle of the slope is zero. Movement occurs when the
thermomechanical force exceeds the frictional reaction force,
Ftm Wg Equation 2-12
Or

Wg
Ftm

Thus a thermomechanical force of greater than 10% of the cable weight would be required to
overcome a coefficient of friction of 0.1.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-19
Length of Cable Storage within the Pipe Route
Length of Cable Storage within a Bend
When the cable is first installed it will lie along the neutral axis of the pipe both in the straight
sections and in the pipe bends. This assumes that the pipe wall is smooth, the cable is well
lubricated and the pulling-in tension is low, thus residual axial forces and strains will be absent.
If the slope of the route is above the critical sliding angle then the cable will move downhill
under the constraint of its axial elastic stiffness. It will be under compressive axial force at the
bottom of the slope. Should a bend be present, the cable within the bend will be pushed outwards
to a larger bend diameter. The cable at the top of the slope will be under tensile force and should
a bend be present this will pull it inwards to a smaller bend diameter. These change in bending
diameters permit extra cable to be stored in the bend at the bottom of the slope and will permit
cable to be released from the bend at the top of the slope.
If the angle of inclination of the route is less than the critical sliding angle, then the cable will not
move into the bend until it is heated. The thermal expansion strain (increase in length) will push
the cable outwards into each bend. The cable will be moved into the bend until it reaches the
furthest outwards distance at the horizontal axis of the pipe at the 3
o
/c - 9
o
/c position.
The length L of cable along the axis of the bend is,
L=Rb
The maximum length L
1
of cable that can be fitted into the bend is,
L
1
=(Rb+Rp-Rc)
Thus the additional length that can be stored is,
L= L
1
-L=(Rp-Rc) Equation 2-13
Where,
Rb: radius of pipe bend [m]
Rp: internal radius of the pipe [m]
Rc: outer radius of the cable [m]
: angle of bend [radians]
L: length of cable in bend [m]
L: additional cable length stored in bend [m]
The storage length is only dependent on the radial clearance and on the angle of the bend. It is
independent of the bending radius Rb.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-20
In a pipe with a large radial clearance and a large angle of bend, the cable storage capability will
be large and the bend will absorb a significant proportion of the thermal expansion. For example,
taking a large pipe radial clearance of 75 mm and a 90
o
bend angle (1.57 radians) an expansion
length of 118mm can be accommodated. This is equivalent to the unconstrained thermal
expansion of a 85m length of typical 138kV XLPE cable with a 1500kcmil copper conductor
heated from 15 to 90
o
C, with a coefficient of thermal expansion of 18.5x10
-6

o
C
-1
.
A high cable thermomechanical thrust is required for the bend to absorb the maximum possible
expansion length. The axial compressive force F
tm
has to be of sufficient magnitude to be able to
lift the mass of the cable through the height (Rp-Rc) and to overcome the frictional forces at the
pipe wall. Thus work is expended in a) frictional heating and b) potential energy, this being
recoverable.
Length of Cable Storage within a Straight Pipe
Cable can be stored within its own straight length in the form of elastic compressibility.
From Hookes law of elastic extension/compression,
EA
FL
L = Equation 2-14
Where,
L: cable storage (deformation) in compression [N]
F: compressive axial force [N]
L: cable length [m]
E: modulus of elasticity of the cable in compression [N m
-2
]
A: cross sectional area of the particular cable component [m
2
]
EA: axial stiffness of the cable in compression [N]
It should be noted that EA is the effective axial stiffness of the total cable and is comprised of the
addition of the EA values for each of the annular layers in the particular construction.
Length of Cable Stored within Lateral Deformation Patterns
The length of cable that can be stored by lateral elastic deformation in the shape of an arc of a
perfect circle is given by the basic relationship:
EI
M
R
1
= Equation 2-15
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-21
Where,
R: radius of curvature of the cable [m]
M: bending moment causing the lateral deflection [Nm]
E: modulus of elastic deformation of the cable in bending [Nm
-2
]
I: second moment of area of the cable component [m
4
]
EI: bending stiffness of the cable component [Nm
2
]
The cable within a pipe does not deflect in the shape of a perfect arc. The deformation patterns
that form first are lateral waves in the shape of cylindrical sinusoids that oscillate from one side
of the pipe wall to the other. The patterns that form at higher temperatures, or with certain
combinations of cable parameters and pipe geometries, are helices that rotate around the walls of
the pipe. The helical pattern has a constant radius of curvature over its shape, but not at each of
its ends where it rejoins the sinusoidal patterns. The sine wave in comparison has a radius of
curvature that varies along its complete shape.
Length of Cable Stored in a Cylindrical Sinusoid Wave
A formula to calculate the length of the curve ( y = sin x) does no exist. Lengths can be
calculated by a numerical approximation technique and the results given in tabular or graphical
form, Figure 2-5. It should be noted that the amplitude is the arc length that the cable describes
from the 6
o
/c position on the bottom of the pipe (note that the angular displacement is half of
the total angular movement) :
( ) = Rc Rp y

Equation 2-16
Where:
y: amplitude of the sine wave [m]
Rp: internal radius of the pipe [m]
Rc: external radius of the cable [m]
: angle described in the pipe cross section by the rotation of sine wave [radians]
The horizontal axis in Figure 2-5 is:

L
y
x 100 [%]
The vertical axis is:

L
L
x 100[%]
Where
L: wavelength of the cable sine pattern [m]
L: cable length absorbed in the wave [m]
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-22
Length of Cable Stored in a Helix
The length of cable stored in a helix is given by:
[ ] ( ) L R R 2 L L
2
c p
2
+ = Equation 2-17
The ratio of the length stored to the wavelength is:

[ ]
1
L
R R 2
1
L
L
2
c p

|
|
.
|

\
|
+ =


Where,
L: length stored [m]
L: wavelength [m]
Rp: inner radius of the pipe [m]
Rc: outer radius of the cable [m]
Side Wall Force and Axial Thrust
Sidewall Force and Axial Thrust in a Bend with No Friction
If cable is thrusting into a bend with no cable to pipe clearance, the sidewall force is given by:
R
F
Fsw = Equation 2-18
Where:
Fsw: outwards force per unit length of cable [Nm
-1
]
F: thermomechanical axial compressive force [N]
R: radius of bend [m]
Axial Thrust in a Bend with Friction
The effect of friction is to produce a non linear distribution of force. The force reduces as the
angle from the bend entrance in Figure 2-8 increases as given in Equation 2-19.
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-23

Figure 2-8
Sidewall force in a bend
Equation 2-19
Where:
F

: Axial force in compression at angle [N]


: Coefficient of friction [per unit]
: Angular distance into the bend [radians]
It can be seen from Equation 2-19 that the reduction in force is dependent on friction and on the
bend angle, but not on the bending radius. Also that the force F

is never reduced to zero.


For the cable to be in static equilibrium in the bend, without sliding, there must always be a
compressive cable axial force F

thrusting into the outlet of the bend to oppose the inwards force
F. The force differential F

around a bend of angle is:


( ) ( ) =

exp 1 F F Equation 2-20


( )

= exp F F
EPRI Licensed Material

Thermomechanical and Gravitational Effects: Basic Theory
2-24
Sidewall Force in a Bend with Friction
From Equation 2-18 and Equation 2-19 the sidewall force per unit length is given by;
Equation 2-21


i
Mechanics of Engineering Materials, PP Benham, RJ Crawford and CG Armstrong, Addison Wesley Longman
Limited, England, Second Edition, 1996.
ii
Mechanical Science for Higher Technicians, DH Bacon and RC Stephens, Heinemann Professional Publishing Ltd,
England,1988, ISBN 0 434 91871 7.
iii
Strength of Materials and Structures, J Case and AH Chilver, Edward Arnold (Publishers) Limited, England,
Second Edition, 1972, ISBN 07131 3244 2.
iv
Mechanics of Materials, SP Timoshenko and JM Gere, Van Nostrand Reinhold Company Ltd, London, 1973,
ISBN 0 442 29996 6.
( )
R
exp F
Fsw

=
EPRI Licensed Material
3-1
3
RESULTS OF TESTS ON THE MECHANICAL
PERFORMANCE OF CABLES AND OF CONSTRAINTS
IN JOINT CHAMBERS
This chapter describes a) tests on samples of 138kV and 230kV XLPE cables, b) tests on
methods of cable constraint and c) the parameters that were derived for the finite element
modelling and design work in the main body of the project. These tests were undertaken over a
twelve month period at the EPRIsolutions laboratory at Haslet TX.
The test program is a key part of the project as a) representative parameters are essential for the
accuracy of the modelling work and b) the literature survey showed that published values were
few and inadequate. The measurement of cable mechanical parameters has significant problems
as a) the cable is designed to be flexible and therefore samples are difficult to support in the test
rig, b) XLPE cable samples have pre-set curvature, c) there are significant end effects which
require that where possible, long lengths be tested and d) the XLPE insulation and polyethylene
jacketing materials have highly temperature dependent properties which require that the cable
parameters be measured over a range of temperatures.
Selection of Cable Samples for Test
A list of cable samples available in North America, primarily from EPRI members, was
compiled, Table 3-1.
Table 3-1 gives a list of available cable samples, with brief details where available of their rated
voltage, conductor size, insulation thickness, sheath type and overall diameter.
For the purpose of the project it was considered that cables throughout the range of stiffnesses
should be tested. The samples were therefore arranged in ascending order of bending stiffness.
The stiffest was the LADWP 230kV cable and the most flexible was the BC Hydro 69kV cable.
The Con Edison 138kV cable was also included as being in the mid range of stiffness although
towards the lower end. These three cables were initially selected for the test program. This
permitted quotations to be obtained from different test laboratories and the required cable lengths
to be determined. In consequence the short list was reduced to the 138kV 1500kcmil cable from
Consolidated Edison and to the 230kV 2500kcmil cable from LADWP.

EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-2
Table 3-1
List of available cables
Source Rated
Voltage
[kV]
Conductor
Area
[kcmil]
Type
XLPE
Insulation
Thickness
[mils]
Metallic
Shield/Sheath
Type
Thickness
[mils]
Jacket
Material
Thickness
[mils]
OD
[inches]
CenterPoint
Energy
138 2,400
Segmental
650 Lead alloy 169 LDPE
110
4.13
Con Edison 138 1500
Compact
round
630 Cu tape
Helically 50%
overlapped
2 tapes 10
mils +
aluminum
laminate
LDPE
110
3.30
Con Edison 138 1500
Compact
round
water
sealed
630 Cu wires
71mm
2
and
0.2 mm
Aluminum
laminate

110 3.27
Exelon 138 3200 650 Cu wires Copper
laminate
HDPE -
Exelon - 1600 650 Cu wires Copper
laminate
HDPE -
LADWP 230 2500
Segmental
(5), water
sealed
1060 Lead 117 LDPE
130
5.08
ABB 345 2250
Trapezoidal
wires water
sealed
1024 Cu wires
aluminum
laminate
95mm
2

0.2mm
MDPE 4.25
BC Hydro 69 2000
Concentric
472 Corrugated
aluminum
100 LDPE 3.94
BC Hydro 69 500
Concentric
650 Lead 67 LDPE 2.82


EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-3
The samples were tested with all cable layers complete and also with the outer layers removed,
in the core-only stage. This would enable values for the mechanical parameters to be derived
for the three parts of the cable construction used in the modeling. These being a) the conductor,
b) the extruded XLPE insulation and semiconducting shields and c) the combination of the
sheath, ground conductor and jacket. The sheath parameters could be derived directly from the
test results. The values for the conductor and insulation could be separated using results from
other work carried out by EPRI
[i]
. It was not practical to measure the parameters of the
conductor alone, as a) the characteristics would change markedly when the insulation was
removed and b) the sample would have suffered damage.
Two test schedules were produced, one for 2002 and a second for 2003. The 2002 test schedule
was sent to three test laboratories for quotation. The EPRIsolutions mechanical test laboratory at
Haslet in Texas was chosen by EPRI to carry out the tests.
Samples of cable for the first series of tests were received at Haslet towards the end of
September 2002 and preparation for testing started immediately. The first test series was
completed in May 2003 and the second test series in September 2003.
Tests on Cable Samples to Derive Mechanical Parameters.
The samples of cable were tested in the test laboratory at Haslet. All tests in the first series
except the axial stiffness under compression test were carried out in a modular heavy duty steel
frame located indoors. This frame was capable of being readily altered to suit the test in progress.
Before testing, the samples of cable were straightened by a) reverse bending and b) by heating
and reverse bending. This did not completely remove all the curvature as XLPE cable insulation
has mechanical memory of the manufacturing process. The crosslinked material tends to revert
to the bend inherent in the insulation caused by bending the core on a take-up reel after extrusion
whilst still warm. The presence of a slight bend had to be taken into account in some of the tests.
For the measurements at elevated temperature, a DC current at low voltage was passed through
the cable conductor at a level to produce the required conductor temperature. A temperature drop
across the sample, between the conductor and the outer surface, had been specified in the test
schedule. In order to achieve this, electric heating tapes were wound around the outside of the
sample which was then suitably thermally insulated. Thermocouples were installed onto the
conductor and outer surface of the sample for temperature measurement. The temperatures were
recorded on a data logger. Temperature control was carried out using a control thermocouple on
the sample conductor linked to a conventional temperature controller.
Forces were imparted to the samples using a hydraulic cylinder. The operation of this was
controlled via a PC (personal computer). Results were visible real time on the PC monitors
screen and were immediately available in spreadsheet format.
Forces were measured using a load cell, installed between the hydraulic cylinder and the sample.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-4
Displacements were measured using laser displacement transducers, or string potentiometers, as
appropriate.
Where the testing procedure allowed, mechanical tests were repeated for three loading cycles to
demonstrate repeatability and consistency.
The results of the tests were recorded using the imperial system of units and therefore the
measurements in this section are given in these units. The parameters used for the modeling were
derived in metric (SI) units.
Measurement of Transverse Bending Stiffness
The transverse bending stiffness of a beam Figure 3-1 is a measurement of the ability of the
beam to resist lateral bending at right angles to the beams longitudinal axis. This value is
analytically represented by EI where:
E: Youngs modulus of elasticity of the material [Nm
-2
]
I: second moment of area of the cross section of the cable in the beam [m
4
]
The bending stiffness EI was measured in the laboratory Figure 3-2 by setting up the sample
initially such that it was horizontal and flat. The distance between pivots was 3.05m for the
138kV cable and 2.75m for the 230kV cable. The ends were supported in close fitting, loose
blocks such that they could a) rotate freely in the plane of bending and b) allow longitudinal
movement of the ends of the sample. The hydraulic cylinder was located over the center point of
the cable. The center of the sample was raised and lowered vertically a nominal distance of
150mm (6ins) off the horizontal supports. The associated force, displacement and temperature
were measured. In order to allow for the inherent bend in the sample due to its mechanical
memory four tests were undertaken with the sample positioned at angular position of 0
o
, 180
o
,
90
o
, and 270
o
.

Figure 3-1
Diagram of the beam method of measuring bending stiffness
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-5

Figure 3-2
Photograph of the bending stiffness test rig
Tests were performed at ambient temperature (8-20
o
C), 60
o
C, 90
o
C and 105
o
C. In order to
determine EI, the measured force was plotted against the measured displacement. The gradient of
the resulting graph was used to calculate a value of EI according to the following formula:
EI = FL
3
/48d Equation 3-1
Where:
EI: bending stiffness [Nm
2
]
F: load applied [N]
L: beam length [m]
d : displacement [m]
It was found that when the displacement was plotted against the force by the PC the relationship
was not a straight line and that a substantial hysteresis effect existed Figure 3-3.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-6

Figure 3-3
Displacement-force output for 138kV cable at ambient temperature
The hysteresis effect is attributed to a) plastic deformation of the copper wires, b) friction
between the wires, c) friction between the core and jacket and d) plastic deformation of the lead
sheath, if present. To obtain a single value of EI for use in the modelling study, an average value
was taken by measuring the gradient of a diagonal line drawn between the opposite maximum
and minimum points on the characteristic.
Once the tests were complete a value of EI was calculated for each measurement taken. An
average was taken of the values at each of the four angular rotations as the final results for the
cable in a) the complete and b) the core-only stage. The results for the two cables tested are
shown in figures Figure 3-4 and Figure 3-5.

EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-7

Figure 3-4
Bending stiffness results for 138kV cable

Figure 3-5
Bending stiffness for the 230kV cable
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-8
The results of the measurements at each temperature in the complete cable stage showed
increasing consistency with temperature. Thus at higher temperature the rotational position of the
sample has less of an effect. This is attributed to a) the memory effect of the XLPE material
becoming less pronounced as the temperature increased and b) the increase in elasticity. The
results of the cable in the core-only stage, with the lead sheath removed, also show the marked
effect of the softening of the XLPE core with temperature. Calculations indicate that the
predominant structural component in the cable is the stiffness of the conductor. The cable had a
higher bending stiffness than expected at ambient temperature, this being attributed to the
compressive grip of the XLPE insulation inhibiting the wires from moving. At 60
o
C and above,
the conductor became more flexible as the XLPE softened.
The results of the measurements on the 138kV cable show some interesting features. The values
for the complete cable sample between ambient and 60
o
C show a decrease in stiffness, the values
are then approximately constant at 90
o
C, then a marked rise is present at 105
o
C. The rise is
attributed to the restriction formed by the double copper tapes to the radial expansion of the
XLPE insulation and shields. This restriction increases the compression and friction between the
layers, preventing slip, thus increasing the stiffness of the cable. The results of the measurements
in the core stage show the expected reduction with temperature, because the XLPE softens with
temperature and loosens its grip on the conductor wires.
Finite Element Modelling of the Bending Stiffness Test Rig
After the analysis of the test results had been completed, a finite element model of the rig was
constructed to confirm that the modeling technique employed for the studies on real cable routes
was accurate.
A model was constructed of the test rig to measure the EI value of the 138kV cable sample in the
core-only stage at 60
o
C. A model of two parallel beams was constructed to represent the bending
parameters for the conductor and insulation. The cable in the model was lifted against gravity to
the same height as the tests at Haslet. Figure 3-6 shows the calculated curvature of the cable (the
inverse of bending radius) plotted against bending moment at the center of the beam of cable.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-9

Figure 3-6
Output from the finite element model of the bending stiffness rig on 138kV cable
The output of the model Figure 3-6 was analysed using the fundamental relationship for bending:
1/R = M/EI Equation 3-2
Where
R: cable bending radius [m]
M: bending moment [Nm]
EI: bending stiffness [Nm
2
]
Thus the reciprocal of the slope of curvature (1/R) against bending moment gives the value of the
transverse stiffness EI. The value of EI measured from the graph gave 665Nm
2
. This is very
close to the value used as input to the model of 661N/m
2
and therefore gave added confidence in
both the test and modelling techniques.
Measurement of Torsional Stiffness
The torsional stiffness of a beam is a measurement of the ability of the beam to resist twist along
the longitudinal axis of the beam. This value is analytically represented by GJ where:
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-10
G: shearing modulus of the cable [Nm
-2
]
J: polar moment of area of the cross section of the cable [m
4
]
GJ: torsional stiffness of the cable [Nm
2
]
Angular stiffness was measured in the laboratory by supporting the sample horizontally and flat.
The active length of the sample was 3.07m for the 138kV cable. The test was not performed for
the 230kV cable. One end was fixed in a tightened block. A torque was applied to the other end
of the sample from the hydraulic cylinder via a system of simple levers. This end of the sample
was rotated clockwise and anti clockwise from the initial position of rest. The associated torsion,
angular displacement and temperatures were measured. Figure 3-7 shows a schematic of the test.
Figure 3-8 shows one of the tests in progress.

Figure 3-7
Diagram of the torsional stiffness test
In order to determine GJ, the measured force was plotted against the measured angular
displacement and the gradient of the resulting graph was used to calculate a value of GJ
according to the following formula:
GJ = TL/ Equation 3-3
Where:
GJ: torsional stiffness [Nm
2
]
T: torque applied [Nm]
L: gauge length [m]
: angular displacement [radians]
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-11

Figure 3-8
Photograph of the torsional stiffness test rig
The figure shows a slight bend in the sample, this was present in the untwisted state and is an
example of the memory of the cable insulation.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-12

Figure 3-9
Output of angular rotation with torque for the 138kV cable
Again it was found that when the torsion was plotted against the angular displacement the
relationship was not a straight line and that the substantial hysteresis effect existed Figure 3-9.
In order to obtain a single value at the particular temperature for use in the modelling part of the
project, an average value was taken by measuring the gradient of the diagonal line drawn parallel
to the characteristic as it decreased in torsional force. The shape of the graph indicates a sharp
rise in the characteristic as the test approached its maximum rotational position. This was an
artifact of the test technique and was ignored for the purposes of the measurement.

Figure 3-10
Results of the torsional stiffness tests for the 138kV cable
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-13
Once the tests were complete a value of GJ was calculated for each measurement taken. A
simple average of these at each temperature was taken as the final result for the cable in the
complete and core-only stage. The analysis of the results for the 138kV cable is shown in Figure
3-10.
The results show a progressive reduction in torsional stiffness with temperature. There appears to
be no marked increase with temperature in the complete cable stage, as was the case with the
transverse bending stiffness measurement. The results tend towards a minimum stable value with
temperature.
Overall, it was found that the torsional stiffness of the 138kV XLPE cable was low. This would
not be expected to play a significant part in the thermomechanical behavior of the cable during
the modelling study.
Measurement of Axial Stiffness
The axial stiffness of a beam is a measurement of the ability of the beam to resist longitudinal
forces and of its ability to exert longitudinal forces, for example when heated. This value is
analytically represented by EA where:
E: Youngs modulus of elasticity of the cable [Nm
-2
]
A: the area of the cross section of the beam [m
2
]
In a complicated composite beam, such as the cables under study, the value of EA is likely to
vary substantially depending on the forces applied to it. For this reason the value of EA was
measured under a) tension, b) under compression and c) by thermal methods.
Measurement of Axial stiffness Under Tension
A value of EA under tension was measured in the laboratory using a cable sample 4m in length
and supporting it horizontally and flat. The ends were held in clamps and one end was fixed. The
hydraulic cylinder was fixed to the other end of the sample with the load cell between. A tensile
force was applied to the sample via the hydraulic cylinder and load cell and the extension of the
sample over a gauge length measured. Figure 3-11 shows a diagram of the test rig.

Figure 3-11
Diagram of measurement of axial stiffness under tension
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-14
In order to determine EA, the force per unit length was plotted against the applied extension (the
strain). The gradient of the resulting graph gave a direct value of EA.
F = EA /(L/e) Equation 3-4
F: load applied [N]
EA: axial stiffness [Nm
2
]
L: gauge length [m]
e: extension [m]
It was again found that when the tensile force was plotted against the strain that the relationship
was not a straight line, Figure 3-12. However although the characteristic was slightly different
between that obtained when raising the force and that obtained when lowering the force, there
was no hysteresis effect as found in other tests. As difficulty was experienced in preparing the
sample such that it was completely straight at the start of the test, an initial pretension of 4500N
(1000lbf) was applied to overcome this. The force strain characteristic therefore has this as its
minimum applied force. An average value of EA was calculated by measuring the gradient of a
straight line drawn between the top and bottom points of the curves of the second and third
measurement cycles.

Figure 3-12
Output of axial stiffness under tension measurements for the 138kV cable
Once the tests were complete a value of EA was calculated for each measurement taken. The
results for the cable tested are shown in Figure 3-13.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-15

Figure 3-13
Derived values of axial stiffness under tension for the 138kV cable
The results show little change in the values of axial stiffness with a) temperature and b) presence
or absence of the sheath/jacket. An average value of EA can be taken. At the elevated
temperature the core-only value is very similar to the complete cable value, indicating that the
polymeric jacket is contributing little at this temperature.

Figure 3-14
Photograph of axial stiffness under compression rig
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-16
Measurement of Axial Stiffness under Compression.
For this test short 0.3m long samples were tested in the EPRIsolutions MTS machine, Figure
3-14. The machine was located in an air conditioned room and the samples tested at ambient
temperature were stored in the room prior to test. Samples tested at elevated temperature were
conditioned in an oven at 90
o
C as required by the test schedule and the test carried out as quickly
as possible after withdrawal from the oven.
Measured values of applied force and compression strain were plotted, Figure 3-15. It was found
that the results were not a straight line. The characteristic followed closely that found when this
parameter was measured under tension. After the initial application of compressive force, good
repeatability was found between the three cycles of the test. An average value of EA was
calculated by measuring the gradient of a straight line drawn between the top and bottom points
of the curves of the second and third measurement cycles.

Figure 3-15
Output from axial stiffness test on 138kV cable
Once the tests were complete a value of EA was calculated for each measurement taken. The
results for the cables tested are shown in Figure 3-16.
Some difficulty was experienced, especially with the tests at elevated temperature, in ensuring
that the force was applied to all layers of the samples evenly, owing to longitudinal shrinkage of
the components and the resulting uneven end surfaces. This may account for the variation in the
results.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-17

Figure 3-16
Derived values of axial stiffness under compression for 138kV and 230kV cables
The axial stiffness in compression tests show the importance of the role of the XLPE insulation
in influencing the compressive stiffness of the conductor. Firstly, the results show that the
conductor is the main structural member, as the 230kV cable with the larger 2500kcmil
conductor area, is stiffer than the 138kV 1500kcmil cable. Secondly, both cables show a
reduction in stiffness at 90
o
C. This is attributed to the grip that the XLPE insulation has at
ambient temperature relaxing at 90
o
C, such that the individual strands can spread radially under
compressive load, thus reducing the stiffness of the sample. Both cables showed a small
reduction in stiffness when the sheath/jacket layer was removed. This was slight for the 230kV
cable, but more pronounced for the 138kV cable, especially at 90
o
C. This may again be
indicative of the role that the copper tapes play in holding the insulation and conductor in
compression at the higher temperatures.
Thermal Tests
Measurement of Axial Stiffness
The value of EA of the cable was also measured thermally. A sample of 138kV cable, 5m in
length, was placed in a close fitting steel tube as shown in Figure 3-17. The bore of the tube was
well lubricated. Both ends of the cable were fixed such that no elongation was possible. The steel
tube was intended to prevent lateral bowing of the sample. The cable sample was heated to
125
o
C and the force developed at one end measured.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-18

Figure 3-17
Measurement of thrust and axial stiffness in fully constrained heated cable
A value of EA was calculated from the following formula:
F = EA Equation 3-5
Where;
F: measured force [N]
EA: axial stiffness [N]
: coefficient of thermal expansion [
o
C
-1
]
: rise in temperature [
o
C]
A graph of applied force against temperature was plotted and a value of EA calculated from the
gradient. In order to calculate EA using this formula, a value of , the coefficient of expansion
also needs to be known, and this was measured in the next test.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-19

Figure 3-18
Output of force with temperature from constrained thermal test
The characteristic for the 138kV cable shows an initial period up to 30
o
C where the relationship
is linear. Thereafter the force increases with temperature, but at a decreasing rate. This is
attributed to movements of the strands in the conductor becoming easier, as the XLPE insulation
loosens its grip due to its becoming softer with temperature. There is no peak to the curve over
the temperature range measured which might indicate buckling of the conductor wires. The
return to ambient temperature shows a linear relationship with force exerted.
The peak force for the complete cable is 36,743N (8,257 lbs) at a temperature rise of 75.5
o
C
from 31.5
o
C to 107
o
C. This gives an EA value of 26.3 MN for a cable coefficient of thermal
expansion of 18.5 x 10
-6

o
C
-1
. If an ambient temperature of 15
o
C is taken, the force at 107
o
C is
51,532N and the EA value is 30.3MN. For comparison, if the conductor is taken to be a solid rod
of copper with a reference book value of E, then EA is 98.9MN and the force generated by the
same cross-sectional area of copper would be 162,260 N, which is 3.2 times greater than that
measured. The measured EA for the cable at 107
o
C is 26.6% of the solid rod value. The
reduction in EA is primarily attributed to the stranding effect of the conductor.
The slope of the characteristic at the lower linear end, up to 54
o
C (17,800N) was measured,
giving an axial stiffness of 41.7MN. This is 42% of the value for a solid rod of copper of
98.9MN. The EA value of the cable is shown to be highly temperature dependent, reducing from
41.7MN at 54
o
C to 26.3MN at 107
o
C. This underlines the importance of measuring the
mechanical parameters of an XLPE cable.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-20
Measurement of Coefficient of Thermal Expansion
In order to measure , the coefficient of thermal expansion, a sample of cable, 5m in length was
placed inside a lubricated steel pipe. One end was fixed and the other end was allowed to expand
freely Figure 3-19. The cable and pipe were then heated to 125C. The temperature and
extension of the free end of the sample were recorded.

Figure 3-19
Photograph of the thermal expansion test on 138kV cable
The value of the coefficient of thermal expansion was calculated using the following formula:
L
2
= L
1
[1+ (T
2
T
1
)] Equation 3-6
Where:
L
2
: sample length at temperature T
2
[m]
L
1
: sample length at temperature T
1
[m]
: coefficient of thermal expansion [
o
C
-1
]
T
1
: temperature associated with L
1
[
o
C]
T
2
: temperature associated with L
2
[
o
C]
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-21
The results, Figure 3-20, showed that the 138kV cable had a coefficient of thermal expansion of
18.2 x 10
-6

o
C
-1
in the core stage and 18.7 x 10
-6

o
C
-1
complete. These results indicate that for the
138kV cable it is the copper conductor which predominates.

Figure 3-20
Output of thermal extension with temperature for the 138kV cable
The 230 kV cable had a coefficient of thermal expansion of 11.8 x10
-6 0
C
-1
as a complete cable.
The characteristic of the elongation against temperature chart had a distinctive saw tooth shape.
This is attributed to the friction between the cable and pipe and is believed to be the cause of the
low value of coefficient of thermal expansion. The value for the complete cable should therefore
be discarded. The results of the core-only measurement gave a more realistic value of 21.4
x10
-6

0
C
-1
. This value, although greater than those for the 138kV cable, agrees with published
values on XLPE cables with thick insulation
[ii]

[iii]
.
Measurement of Thermal Relaxation
When a cable conductor is heated as a component part of a buried cable it is effectively
restrained and unable to expand. At low temperature rises it exhibits a substantially linear
relationship between temperature and exerted force. When the temperature increases the rise in
the force exerted decreases and may reach a peak after which the conductor exerts less force with
rising temperature
[iv]
, this is termed conductor relaxation. When the conductor cools to its
original temperature it exhibits a tension. The phenomenon is attributed to a) the creep of the
copper wires beyond their elastic limit and b) the physical movement of the wires in relation to
each other.
To measure thermal relaxation, a 5m length of 138kV 1500kcmil cable was inserted into a well-
lubricated close fitting steel pipe. Both ends were fixed. The temperature was raised to 90
o
C and
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-22
maintained at this temperature for 72hrs. For the core-only test the temperature was maintained
for twice this time. A record of exerted force and temperature with time was made, Figure 3-21.
This graph shows a peak exerted force of 21,150N (4700 lbs) at the start of the test, this is
gradually decreasing as the test progresses to a value of 18,000N (4000lbs). When the heating
equipment is switched off the compressive force rapidly falls with temperature and changes to a
tensile force of 12,600 (2800lbs). This tension is more than the fall in compressive force during
the 72hour period and is indicative of the loss of force due to conductor strand movement during
the initial heating phase.

Figure 3-21
Output of thermal relaxation test on 138kV cable
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-23
List of Parameters Derived from the Mechanical Tests on the 138kV 1500kcmil
Cable

Table 3-2
Parameters derived from tests on a 138kV 1500kcmil XLPE cable

Parameter


Complete Cable

Core-only
Bending Stiffness, EI EI [Nm
2
] EI [Nm
2
]
Ambient temperature 2682 1850
60
o
C 1189 661
90
o
C 1141 379
105
o
C 2450 297

Axial Stiffness under Tension, EA EA [MN] EA [MN]
Ambient temperature 22.9 27.9
60
o
C 26.5 21.8
90
o
C 27.5 28.1
105
o
C 26.9 27.0

Axial Stiffness under Compression, EA (0-
22,500N)
EA [MN] EA [MN]
Ambient temperature 11.4 9.5
90
o
C 6.5 4.2

Axial stiffness from Constrained Thermal
Test, EA
EA [MN] EA [MN]
54
o
C, 30
o
C, (0 20,000N) 41.7 67.5
107
o
C, 114
o
C, (0 37,000N) 26.3 21.5

Torsional Stiffness, GJ GJ [Nm2] GJ [Nm2]
Ambient temperature 1671 1246
60
o
C 1513 902
90
o
C 1275 683
105
o
C 1237 678

Coefficient of Thermal Expansion, [
o
C
-1
] [
o
C
-1
]
18.7 x 10
-6
18.2 x 10
-6

EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-24
List of Parameters Derived from the Mechanical Tests on the 230kV 2500kcmil
Cable

Table 3-3
Parameters derived from tests on a 230kV 2500kcmil XLPE cable

Parameter


Complete Cable

Core-only
Bending Stiffness, EI EI [Nm
2
] EI [Nm
2
]
Ambient temperature 7739 3677
60
o
C 5684 1704
90
o
C 5613 1056
105
o
C 5366 1025

Axial Stiffness under Compression, EA EA [MN] EA [MN]
Ambient temperature 12.5 12.3
90
o
C 10.6 9.0

Coefficient of Thermal Expansion, [
o
C
-1
] [
o
C
-1
]
11.8 x10
-6
21.4 x10
-6


Tests on Types of Cable Constraints
Measurement of Internal Friction between Core and Outer Layers.
In the construction of the cable the XLPE core is separated from the metal or laminate sheath by
a series of cushioning tape layers, sometimes including copper wires or tapes. These layers form
a position in the construction where slip between the two may be possible. This may be
important when designing a cleated installation system, as most of the thrust is generated by the
conductor, but all the restraint is between the jacket and fixed cleats. It was therefore important
to measure the coefficient of friction between the two components to assist in the preparation of
the design guide.
This test was carried out in the open-air test facility at the EPRIsolutions test facility, Figure
3-22. The cable sample was pre-treated by heating to 90
o
C, this caused expansion of the XLPE
insulation, which in turn caused permanent radial extension of the lead sheath. The gap between
sheath and core would thus be as found in service for the test regime. The cable sample was bent
through a right angle with a bend radius of 15D (where D is the diameter of the cable) and fixed
to a rigid steel plate using cleats spaced sufficiently close together to prevent the cable sample
buckling. 15D was selected as a typical minimum bending radius for an XLPE cable held in a
controlled curve by either a template or by cleats at close spacing. Initially the cleats were
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-25
loosely tightened so as not to distort the sheath. A back tension of 1125N (250lb) was applied by
hanging a weight on the conductor of the cable. A rigid anchor plate was fitted to the sheath at
the other end of the sample to prevent movement of the sheath and jacket relative to the
conductor.
A tensile force was then applied to the conductor and raised until the conductor and core moved
in-side the sheath. Initially it was proposed to measure this property with the cable under
compression, as this would be the condition predominating in service, however EPRIsolutions
believed this would not be practical within the constraints of the project.

Figure 3-22
Photograph of the internal friction test rig, 230kV cable
Once a value of friction was measured with the cleats loosely tightened, the cleats were tightened
to a higher value, to simulate the condition in service. The measurement of the coefficient of
friction was then repeated.
A value of , the coefficient of friction, was calculated from the following formula:
F
out
= F
in
e
()
Equation 3-7
Where:
F
out
: outwards or motive force [N]
F
in
: inwards or resistive force applied by the back tension [N]
: coefficient of friction [per unit]
: angle through which the direction of the cable is changed [radians]
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-26
For the 138kV cable the values of coefficient of friction for both tests were greater than unity,
indicating that the cleats did have a clamping effect on the XLPE core. The coefficient of friction
of 1.6 at the higher tightening torque of 68Nm (50lbft) was greater than the 1.4 value at 34Nm
(25lbft) showing the crushing effect on the jacket layers of tightening the cleats. The 230kV
cable had a higher coefficient of frictional of 2.1 at a torque of 68Nm (50lbft).
Measurement of Pull-through Strength of Cable Cleats
One method of installing all or part of a cable system is to provide support using cable cleats.
Often cable cleats provide support, but little or no longitudinal restraint. For example in a
Holttum sagged
[v]
system, the weight of cable is equal on both sides of the support, and there
is little or no longitudinal force on the supports. For the joint chambers studied in this project it
was proposed to improve the performance of the constraints adjacent to the joints by rigidly
cleating the cables against longitudinal forces, thus preventing uncontrolled movement. Such
movement could cause preferential bending. Under load cycling conditions, the possibility of
excessive fatigue to metallic sheaths and foil water barriers existed. It was necessary to know the
withstand strength of a cable cleat to produce designs of cable constraining systems.
The pull-through strength of the cable cleat was measured using the indoor facility at
EPRIsolutions. After initial straightening and cutting the cable sample to length, the layers
within the cable construction were locked together at both ends by a) bolting through all layers,
b) pinning the outer layers to the inner layers with steel nails and c) and applying an epoxy
resin/glass fiber bandage to all the layers at both ends of the sample. Two-part aluminum alloy
cleats were designed and manufactured. These were 100mm long and with a 4.5mm thick
Neoprene rubber liner of 60
o
Shore hardness glued to the inside. The two halves of the cleat were
held together by two high tensile steel bolts seated on flat steel washers. The cleat was bolted to
the cable. The cable was then fitted to a steel plate such that one end protruded. A tensile force
was applied to the cable conductor and the force and movement measured.
The testing was carried out in three parts. Firstly an ultimate pull-through strength was
established. Then a withstand test was performed to verify the long term holding strength of the
cleat. Finally a further ultimate strength test was carried out using two cleats fitted in series with
each other. This test was intended to check whether the strength of two cleats was more than
twice the strength of one cleat.
The tests were carried out at two temperatures a) at ambient temperature and b) the sample
heated such that the temperature of the jacket reached 70
o
C. This temperature was chosen to
simulate the temperature of the jacket when the conductor is at a full load temperature of 90
o
C.
Cleat Ultimate Strength Test
Figure 3-23 shows a test being conducted at ambient temperature on a sample of 230kV cable. In
the photograph a webbing strap is shown supporting the sample to hold it horizontal until the
tensile force was applied.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-27
For this test the tensile force was raised in 900N (200lbf) steps. Each step was held for 5minutes
before being raised again, Figure 3-24. The ultimate strength was reached when a) substantial
movement through the cleat occurred, b) the clamping strength of the cleat decreased, or c) when
the cable moved 25mm through the cleat.

Figure 3-23
Photograph of the cleat pull-through test rig on the 230kV cable
When one of the 138kV cable samples was pulled through the cleat at the 70
o
C test condition the
jacket was found to have suffered mechanical and thermal damage. The jacket exhibited parallel,
saw tooth shaped ridges spaced 7mm apart over a circumferential width of 53mm. The jacket
appeared to have been a) softened significantly by the applied test temperature and b) stretched
when it was dragged incrementally through the cleat. The ridges were up to 2mm deep being
both above and within the normal jacket radial thickness of 3-4.5mm. There was no sign that the
jacket had been ruptured.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-28

Figure 3-24
Output of the cleat pull-through test on the 138kV cable
Cleat Withstand Test
When the movement of the cleats at the 5 minute hold points were plotted against force, Figure
3-25, it could be seen at which tension the movement through the cleat started to accelerate. A
force at which the movement appeared to be low and stable was chosen and this force was then
applied to the cable for a 24 hour period. The movement through the cleat was measured with
time. A force of 6222N (1400lbf) was applied to the 138kV cable and a force of 2700N (600lbs)
was applied to the 230kV cable.
Cleat Withstand Test Result
The results for the cleat on the 230kV cable showed an initial 4mm cable movement, followed
by a smaller movement of 0.7mm over 24 hours Figure 3-26. The initial movement was observed
to be the elastic stretch of the rubber cleat liner when the test load of 2,700N (600lbf) was
applied.
Multiple Cleat Test
For this test two cleats were fitted in series, as close as possible to each other, and then the
ultimate strength test was repeated. It had been intended that the cleats should be touching at the
start of the measurement, but a small gap was present between the cleats due a) to some of the
rubber liner being squeezed out from between the cleat and the cable and b) the natural curvature
in the cable. For the test conducted on the 138kV cable the two cleats came into contact when a
force of 3,600N (800lbf) was applied.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-29

Figure 3-25
Output of the cleat pull through test on 138 and 230kV cables movement at hold points

Figure 3-26
Output of cleat withstand test for 230kV cable

EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-30
The cleat tests showed that the presence of the rubber liner between the cleat and the cable jacket
had to be taken into account in the interpretation of results. When load was applied the rubber
deformed, giving a significant movement of the cable sample. As well as the elastic stretch of the
rubber liner, the cable was also slowly moving through the cleat. The results shown are thus a
combination of these two occurrences. When load was released the elastic nature of the
deformation meant that the cable sample retracted.
List of Parameters Derived from Tests on Constraining Methods for 138kV XLPE
Cable

Table 3-4
Parameters derived from tests on types of constraint on 138kV 1500kcmil XLPE cable

Constraint Tests of 138kV 1500kcmil XLPE Cable


Parameter
Coefficient of internal friction test,
With cleat bolts tightened to 34Nm (25lbf ft) 1.4
With cleat bolts tightened to 68Nm (50lbf ft) 1.6

Cleat ultimate pull-through force, F
pt,
at ambient
temperature
F
pt
[N]
Cleat bolts at 34Nm (25lbf ft), three tests 8100, 8100, 9900
Cleat bolts at 68Nm (50lbf ft), two tests 9900, 9900
Cleat bolts at 102Nm (75lbf ft) two tests 8100, 11700

Cleat ultimate pull-through force, F
pt,
at 70
o
C F
pt
[N]
Cleat bolts at 34Nm (25lbf ft), two tests 5400, 7200
Cleat bolts at 68Nm (50lbf ft), two tests 7200, 8100
Cleat bolts at 102Nm (75lbf ft), two tests 9000, 10800

Double cleat ultimate pull-through force at 68Nm (50lbf
ft) and ambient temperature
22500

Cleat 24 hour withstand test at ambient temperature
Cable permanent creep movement
Single cleat tightened to 68Nm (50lbf ft)
Applied force 6300N
2 mm

EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-31
List of Parameters Derived from Tests on Constraining Methods for 230 KV XLPE
Cable

Table 3-5
Parameters derived from types of constraint on 230kV 2500kcmil XLPE cable
Constraint Tests on 230kV 2500kcmil XLPE Cable Parameter
Coefficient of internal friction test,
With cleat bolts tightened to 34Nm (25lbf ft) (no movement at
13500Nm)
With cleat bolts tightened to 68Nm (50lbf ft) 2.1

Cleat pull-through force, F
pt
, at ambient
temperature
F
pt
[N]
Cleat bolts at 34Nm (25lbf ft), three tests 3600, 5400, 2700
Cleat bolts at 68Nm (50lbf ft), two tests 3600, 3600
Cleat bolts at 102Nm (75lbf ft), two tests 4500, 5400

Cleat pull-through force, F
pt
, at 70
o
C F
pt
[N]
Cleat bolts at 34Nm (25lbf ft), two tests 2700, 4500
Cleat bolts at 68Nm (50lbf ft), two tests 2700, 4500
Cleat bolts at 102Nm (75lbf ft), two tests 4500, 4050

Double cleat ultimate pull-through force at 68Nm
(50lbf ft) and ambient temperature
20700

Cleat 24 hour withstand test, ambient temperature
Cable permanent creep
Single cleat tightened to 68Nm (50lbf ft)
Applied force 2700N
0.8 [mm]
Discussion
The bending stiffness was found to be of larger magnitude than expected from work performed
on paper insulated cables and from the bending theory of conductors without coverings. It was
also found that the conductor stiffness reduced with temperature. It was concluded that the
covering layer of XLPE insulation played a significant role in increasing the stiffness of the
conductor. The XLPE insulation and semiconducting shields grip the conductor as a result of the
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-32
differential thermal contraction strain that is formed at the end of the factory crosslinking process
when the cable is cooled from 200
o
C to room temperature. This increases the stiffness of the
conductor by restricting the individual movement of each wire and by increasing the inter wire
friction during bending. XLPE is a comparatively rigid semi crystalline material up to 60C, but
above that temperature it progressively softens as the crystallinity reduces until it exhibits
elastomeric properties due to the attainment of a non crystalline amorphous structure above
103C. Thus the grip of the insulation on the conductor reduces above 60C and the conductor
progressively becomes more flexible.
The test results were analysed and the XLPE insulation and shields were determined to exhibit a
significantly smaller bending stiffness compared to the conductor.
The sheath for the purpose of the tests was considered to be comprised of all the layers that
were applied onto the extruded core. The performance of the sheath could be deduced as cable
samples were tested as a) complete cables and then b) as extruded cores with all of the covering
layers removed. It was shown that the bending stiffness of the 230kV cable reduced with
temperature both in the complete and in the core-only stages. This demonstrated that the
polyethylene jacket and the extruded lead sheath did not significantly restrict the bending
performance. However in the 138kV cable, the combination of the layers of copper tape,
aluminum foil co-polymer and bonded polyethylene jacket started progressively to increase the
bending the stiffness of the complete cable when the temperature rose above 60
o
C. This was
attributed to the restriction by the copper tapes and aluminum foil of the radial expansion of the
XLPE insulation and shields.
The primary function of most cleats in service installations is to control the bending radius of the
cable rather than alone to withstand longitudinal force. By holding the cable in a curved offset
shape it is possible for the geometry and cable stiffness in combination to transfer load to the
cleats, but in a transverse direction.
The cleat designs in this project would however be required to withstand longitudinal forces as
the cables will exert a force at the joint positions. The tests on the cleats showed that the rubber
liner played an unexpected role as it readily extended longitudinally when tension was applied to
the cable. This gave some elasticity to the clamping effect of the cleat. Once this effect was
allowed for in the results, the cable showed a stable and predictable movement through the cleat.
Tests on two cleats in series showed they doubled the pull-through force.
Values for the internal coefficient of friction of the XLPE core within the sheath/jacket have not
previously been found in published literature. The measurements of internal friction between the
cable and core were thus essential for the design of constraining systems in joint chambers.
Increasing the tension of the bolts on the cleats showed that a proportion of the clamping force
was transferred to the extruded XLPE cable core beneath the jacket. However it was clearly
shown that, when the thermomechanical force is high enough, the cable core will slide inside the
sheath and apply thrust directly to the cable joint or termination.
EPRI Licensed Material

Results of Tests on the Mechanical Performance of Cables and of Constraints in Joint Chambers
3-33
After the test results had been analysed the derived parameters were applied to a finite element
model of the bending stiffness measuring rig. It was verified that the measuring and modelling
techniques were sound and gave closely similar answers.
Conclusions
Tests to determine the mechanical properties of the two XLPE cables to be used as a basis for the
thermomechanical project were successfully carried out and yielded essential values for the
engineering models.
Mechanical tests on cables are difficult to perform and there were no published values for some
of the XLPE cable properties prior to the test program. Test methods were developed to measure
them and although not perfect will provide the foundation for future measurements.
XLPE cables were found to behave differently to paper cables, in that XLPE cables are
significantly stiffer. As with paper cables it was found that the conductor is the key structural
component dictating the cable stiffness. The analysis of the test results showed that the properties
of the XLPE insulation had very significant effects on the cable stiffnesses in that a) the bending
and compressive axial stiffnesses of cable in the core-only stage were increased at temperatures
below 60
o
C and b) the stiffnesses of the complete 138kV cable was increased at 60
o
C and above.
The first effect was attributed to the compressive grip of the XLPE insulation formed during
thermal contraction in manufacture. The second effect was attributed to the constraining effect of
layers of copper shielding tapes. Theses findings show that cables with different designs and
different manufacturing processes are likely to exhibit different stiffnesses and different
thermomechanical performances in service. Thus it is recommended that cable samples be
submitted to mechanical test if the design is different, or if the cable is provided by a different
supplier/manufacturing route.



i
Research to determine the acceptable emergency operating temperatures for extruded dielectric cables. EPRI
Report EL-938 November 1978.
ii
Thermomechanical behavior of 400kV synthetic cables. Brincourt, Dorison, EDF. Jicable 1999 Paper A4.2
iii
Thermomechanical modeling of 345kV XLPE cables in duct Tarnowski, Iordanescu, Awad, Royer. Hydro
Quebec. Jicable 1999 Paper A4.1
iv
Thermomechanical behavior of large conductor cables. Holdup, Occhini, Skipper IEEE Paper 81 TP 67-478
July 1967.
v
The installation of metal sheathed cables on spaced supports. Holttum. W. (1955) Proc IEE A 102, 729 742.
EPRI Licensed Material
4-1
4
PARAMETERS LIMITING SYSTEM PERFORMANCE
Introduction
This chapter records the limiting parameters that may inhibit service reliability and reduce the
specified service life of cable circuits.
At transmission voltages manufacturers do not usually publish design data nor are there
specifications or standards that cover limiting design parameters. One reason is that at these
voltages cables are designed to suit each particular application, thus standard designs do not
exist. Another reason is that some suppliers prefer to supply a turn-key service covering system
design, cable and accessory manufacture and system installation; thus to enhance their
competitive position they prefer to keep their in-house design data confidential.
Purchasers of cable and cable systems are advised to obtain the values of the limiting parameters
at the project tendering and bidding stages by requiring the prospective suppliers to state the
limiting values for both the particular cable construction and the particular application. The
following sections recommend limiting parameters that were compiled from a) existing
knowledge, b) analysis of the literature search and c) discussions with the utilities who supported
the project.
Limiting Values of Cable Bending Radii
The bending radius is one of the most important limiting mechanical parameters of an electric
power cable. All layers of the cable construction are sensitive to being bent. The most critical
component is the metallic sheath, whether it is formed from an extruded tube or from a foil
laminate bonded to the polymeric jacket. Lead sheathed cables are prone to bending damage as
too small a bending radius can overstretch the extruded material on the outside of the bend.
Aluminum sheathed cables require more force in order to bend them and for this reason extruded
aluminum sheaths are corrugated to facilitate bending. Although corrugated aluminum sheath is
the most robust type, cracks can be formed in the trough of the corrugations by overbending,
particularly on large diameter transmission class cables. Cables with a foil laminate sheath have
no extruded metal sheath, but over-bending can lead to the foil splitting on the outside of the
bend and to knuckling on the inside. The purpose of bonding the foil to the polymeric jacket is
to a) uniformly distribute bending strain and b) protect the foil from the possibility of
concentrated local bending. In transmission class cables the macro bending characteristics of the
cable will be determined primarily by the conductor, which is the stiffest structural element, and
not by the foil/jacket laminate.
EPRI Licensed Material

Parameters Limiting System Performance
4-2
The importance of the bending radius is underlined by it being one of the few mechanical
parameters relevant to the whole cable for which a test is invariably specified in qualification test
specifications. Examples are AEIC CS7
[i]
and IEC 62067
[ii]
. Additionally minimum installation
bending radii are also specified in some documents such as AEIC CG5 2001
[iii]
. The minimum
bending radii for XLPE transmission class cables abstracted from publications are given in Table
4-1.
Table 4-1
Cable minimum bending radii
Source Application Situation Bend Radius
[m]
AEIC CS7
1993
XLPE Cable 69-138kV
Lead, copper, aluminum,
metallic foil tapes
Qualification bend test
3 cycles
12.5 (D+d)
IEC 62067
2001
XLPE Cables 150-500kV
Lead, lead alloy, corrugated
aluminum, longitudinally applied
metal foils bonded to oversheath
Qualification bend test
3 cycles
12.5 (D+d)
IEC
60840
[xi]

1999
XLPE cables 30-150kV
Lead, lead alloy, corrugated
aluminum, longitudinally applied
metal foils bonded to oversheath
Qualification bend test
3 cycles
12.5 (D+d)
Electra
141
[x]

April 1992
Guidelines for tests on HV
extruded cable with laminated
protective coverings
As IEC 60840 bend test, 3 cycles
Sidewall pressure test, 7,358Nm
-1
, 1
cycle around 180
o
pattern
12.5 (D+d)
12.5 (D+d)
Jicable 03
Paper
A.1.5
[iv]

275 XLPE 3000kcmil, 122.6mm
OD cable development in Japan.
Special four layer aluminum foil
laminate bonded to PVC jacket.
Foil thickness: 0.05mm, laminate
thickness: 0.3mm.
Development bend test, 5 cycles
Development bend test, 30 cycles
5D
10D
Lead sheathed cable.

12D minimum
14D preferred
AEIC CG
5-2001
[iii]

Extruded power cable pulling
guide 46-138kV.
Recommended installation
bending radii:

Corrugated metal tape sheath,
longitudinally applied and folded
18D minimum
20D preferred
EPRI Licensed Material

Parameters Limiting System Performance
4-3
Source Application Situation Bend Radius
[m]
On pattern adjacent to accessories: 12D (15D)
Direct buried: adjacent to
accessories:
12D (20D)
Direct buried: main cable span: 15D (20D)
XLPE cables 33-200kV
Lead alloy, corrugated aluminum
and foil laminate.

In duct: main cable span: 25D (30D)
Adjacent to accessories: on former 15D
Adjacent to accessories: direct buried 20D
Main cable span: direct buried 30D
Electric
Cables
Handbook
1997
[v]

XLPE cables 200-400kV
Lead alloy and corrugated
aluminum
(Foil laminate radii were not
listed) Main cable span: in ducts 35D
Notes;
Where applicable, foil laminate bending radii are given in (brackets)
D: diameter over the jacket
d: diameter over the conductor

The bending radii listed in Table 4-1 have been reviewed and simplified to give the
recommended design values in Table 4-2 and Table 4-3.

Table 4-2
Minimum bend radii for cable inside preformed pipe and duct bends
Minimum Bending Radius [m] Cable
Thick Wall
(1)
Sheaths Thin Wall Sheaths
69-161kV 25D 30D
220-345kV 30D 35D
(1): 2mm and greater radial thickness of metallic cable component


EPRI Licensed Material

Parameters Limiting System Performance
4-4
Table 4-3
Bend radii for design of joint manholes
Minimum Bending Radius:
Cable in Close Cleated Bend [m]
Cable
Thick Wall Sheaths Thin Wall Sheaths
69-161kV 10D 15D
220-345kV 12D 19D

The limiting values that were calculated for the minimum bending radii for the 138kV and
230kV XLPE cables in this project are given in Table 4-4 and Table 4-5.
Table 4-4
Limiting value of bend radii for the main run of 138kV and 230kV XLPE cables
Minimum Bending Radius [m] Cable Design Cable Diameter [mm]
Criterion Value
138kV 1500kcmil,
aluminum foil sheath
83 30D 2.49
230kV 2500kcmil lead
sheath
129 30D 3.87

Table 4-5
Limiting bend radii in joint chambers for 138kV and 230kV XLPE cables
Minimum Bending Radius:
Cable in Controlled Bend [m]
Cable Design Cable Diameter [mm]
Criterion Value
138kV 1500kcmil,
aluminum foil sheath
83 15D 1.25
230kV 2500kcmil, lead
sheath
129 12D 1.55

Limiting Value of Tensile Force
Tensile forces are generated during a) the pulling-in process, b) the combined application of
gravitational and thermomechanical loading, usually acting at the top end of a cable span and c)
EPRI Licensed Material

Parameters Limiting System Performance
4-5
thermal contraction when the cable cools down at the end of a loading cycle. The loads are
experienced predominantly by the conductor and to lesser extent by the metallic sheath/ground
wire combination. The majority of transmission class cables have a sufficiently large conductor
cross sectional area to be able to withstand the load without damage and without excessive
extension. The latter is important as excessive conductor extension or uneven load sharing will
place a strain on the metallic sheath combination, which is not designed to carry a tensile load
and is the least capable of withstanding it. Designs of metallic sheaths in decreasing order of
robustness are a) extruded corrugated aluminum and lead sheaths, b) thin welded/brazed sheaths
of stainless steel and copper, usually all corrugated, and c) thin copper or aluminum foil
copolymer bonded to the polymeric jacket.
A cable route consists of straight sections and bends. Significant tensile force is generated in the
cable during the pulling-in operation in overcoming a) cumulative frictional forces and b) the
tension multiplication effect around bends. During pulling-in there will be a strain in each cable
component, which may result in a residual strain remaining afterwards, particularly if the
coefficient of friction is high. During pulling-in it is possible for the conductor pulling bolt to be
broken and for the metallic sheath to be severely extended or broken. Limiting values of
maximum pulling force F are available
[v] [vi]
:
Reference
[v]
: F = (59 x A x 2) N, where A [mm
2
] is the cross sectional area of the copper
conductor. Only two out of three cables are taken to share the load.
Reference
[vi]
: F = (87 x A x 2) N, where A[mm
2
] is the cross sectional area of the copper
conductor. Only two out of three cables are taken to share the load.
Reference
[iii]
: Maximum pulling in tension for the eye: F = 81,000N (18,000lbf), with the proviso
that the conductor is large enough to handle the tension value.
Limiting values for the tensile pulling-in force recommended for the 138kV and 230kV XLPE
cables in the main span are given in Table 4-6. The lower of the two conductor stresses given
above is recommended as being a safer value i.e. 59Nmm
-2
.
Table 4-6
Limiting value of pulling-in force for 138kV and 230kV XLPE cables
Maximum Pulling-in Force [N] Cable Design Conductor Area
A [mm
2
]
Criterion Value per Cable Value per
Installation Group
138kV 1500kcmil 760 (59 x A x 2) 44,840 89,680
230kV 2500kcmil 1266 (59 x A x 1) 74,694 74,694
When cables are load cycled, the connectors in the joints must be able to withstand the
compressive thrust due to thermal expansion and the tensile retraction loads due to conductor
relaxation.
The tensile load that is generated by thermomechanical force when a straight cable cools is:
EPRI Licensed Material

Parameters Limiting System Performance
4-6
EA F = Equation 4-1
Where:
F: tensile force in conductor [N]
E: effective value of Youngs modulus of elasticity of the conductor [Nm
-2
]
A: cross sectional area of conductor [mm
2
]
EA: effective axial stiffness of the conductor [MN]
: coefficient of thermal expansion of the conductor [
o
C
-1
]
: temperature fall of the conductor during a cooling cycle [
o
C]
Suitable measured values of the axial stiffness EA of the cable and core are given in Chapter 3,
which records tests on 138kV and 230kV cables. The method of deriving the EA value for the
conductor is given in Chapter 7.
The maximum thermomechanical force that can be generated with conductor relaxation in a
stranded conductor is:
Ak F = Equation 4-2
Where:
F: maximum thermomechanical tensile force in a stranded copper conductor [N]
A: cross sectional area of conductor [mm
2
]
k: constant giving maximum attainable stress [Nmm
-2
]
In practice the maximum force is limited by the stress relaxation of the conductor, due to
physical movement between the wires and to creep of the copper. The maximum attainable axial
force is calculated from Equation 4-2. For the normal slow rate of temperature rise in a
transmission class cable, the constant k = 30Nmm
-2
for the generation of a compressive force,
(obtained from reference
[vii]
for large conductor sizes up to 2000mm
2
.) The same reference
[vii]

showed that a tensile force was produced on cooling; this is taken to be reciprocal, with the same
value of k.
The maximum attainable long term tensile force with conductor relaxation present for the 138kV
1500kcmil cable is calculated to be 22,800N from Equation 4-2. The maximum short time tensile
load applied during the tests on the 138kV 1500kcmil cable was 53,333N, at which no obvious
deformation was seen.
Limiting Value of Compressive Force
The conductor connectors in the joints must be able to withstand both the compressive thrust and
tensile retraction forces due to cable thermal expansion and contraction.
EPRI Licensed Material

Parameters Limiting System Performance
4-7
The maximum value of compressive force that can be generated by a stranded copper conductor
is the same as that for tensile force and is given by Equation 4-1 and by the limiting value in
Equation 4-2. The results of the compression tests on 138kV and 230kV XLPE cables are
described in Chapter 3. The tests performed were of short duration, i.e. less than 3 hours. The
tests were a) a compression test in which 45,000N was applied to the 138kV 1500kcmil cable
and b) a thermally constrained test in which the cable was heated from 31.5C to 107
o
C,
generating a force of 36,550N. These forces are higher than the limiting value of 22,800N from
Equation 4-2, based on long term conductor relaxation measurements. The difference is
attributed to the short duration of the test. There were no obvious signs of cable deformation.
The maximum thermomechanical force generated in the 72 hour mechanically constrained
relaxation test on a 5m sample heated to 90C was 20,915N, which showed that stress relaxation
had reduced the force to 18,000N. The measured forces are of similar magnitude to the limiting
value of 22,800N from Equation 4-2, based on measurements of conductor relaxation
[vii]
. There
were no obvious signs of damage.
The maximum force applied to the 230kV cable in a short term compression term test was
90,000N, compared to the limiting value of 38,000N with conductor relaxation. A mechanically
constrained heating test was not performed. There were no obvious signs of damage.
Limiting Values of Sheath Strain
When cable is heated within a pipe or duct, wave patterns form adjacent to bends. As the cable is
not a tight fit in the pipe the bending radius is reduced within the deformation patterns. When the
cable is heated and cooled during load cycling, the bending radius increases and decreases. The
reduction in bending radius directly increases the strain in the metallic sheath or foil. The cyclic
fluctuation in sheath strain will continue to take place over the lifetime of the cable. If not
properly controlled, the change in bending strain will lead to sheath fatigue and cracking. The
consequences of a cracked foil are that a) moisture can enter and b) induced voltage may cause
sparking and erosion. An additional consequence with an extruded metallic sheath is that there is
an ineffective return path for short circuit current, leading to severe local heating in the event of a
system fault. Cable installation designers endeavor a) to minimize the cyclic change in strain and
b) to keep the strain below the defined fatigue limit for the particular cable application and
sheath material.
Equation 4-3 calculates the absolute magnitude of strain in the metallic sheath, this being
dependent on a) cable bending radius and b) longitudinal thermal extension. Both of these
parameters are of particular importance in pipe and duct systems as the cable is free within the
constraints of the pipe clearance to bend and stretch in a natural and comparatively uncontrolled
way.
(


+ =
L
L L
R
r
1
Equation 4-3
Where:
EPRI Licensed Material

Parameters Limiting System Performance
4-8
: combined strain in the sheath at radius r [per unit]
r: outer radius of the sheath [m]
R: bend radius at the neutral axis of the cable [m]
L
1
: extended sheath length [m]
L: original sheath length [m]
The extended sheath length (L
1
-L) is due to the predominating expansion of the conductor L
c,
,
less the thermal expansion of the metallic sheath L
s
, where (L=L), thus:
( ) ( ) ( ) [ ]
s c 1
L L L = Equation 4-4
Where:
: product of coefficient of thermal expansion and temperature rise [per unit]
c: refers to the conductor material and temperature rise
s: refers to the sheath material and temperature rise
The temperature rise of the sheath is less than the temperature rise of the conductor, due to the
temperature drop across the XLPE insulation and shields, typically this is 10
o
C. The coefficients
of expansions of the sheathing materials are equal to or greater than that of the conductor,
which is usually copper. Typical values of are given in Table 4-7.
Table 4-7
Properties of sheath materials
Sheath Material Coefficient of Thermal
Expansion,
[
o
C
-1
]
Modulus of Elasticity, E
[Nm
-2
]
Lead Alloy 29 x 10
-6
17 x 10
9

Aluminum 23.5 x 10
-6
70 x 10
9

Copper 17 x 10
-6
130 x 10
9

Stainless Steel 16.3 x 10
-6
201 x 10
9

It will be apparent that it is possible for the longitudinal strain in the sheath to be negative, in that
the sheath attempts to expand more than the conductor. In the majority of cases the conductor
axial stiffness EA will predominate over that of the sheath. This will hold the sheath in
compression (a negative sign convention). Note that the bending strain on the outside of the bend
is positive (extension) and on the inside is negative (compression). In pipe and duct systems
sinusoidal and helical wave patterns are likely to form. These will partially accommodate the
thermal strain released from the adjacent rigid cable. Equation 4-4 is a simplified situation which
assumes that all of the thermal strain generated by the cable is released into the sine wave. For
greater accuracy, the finite element ABAQUS model calculates the distribution of sheath strain
in each cable throughout the length of the route.
The limiting values for the sheathing material are derived from the strain at the ultimate tensile
force, with a factor of safety of 2 applied. However, the common sheathing materials are ductile,
EPRI Licensed Material

Parameters Limiting System Performance
4-9
for example a) lead alloy has a very low elastic limit and in its plastic range will readily extend
when the load is applied quickly and b) sheathing grade aluminum is 99% pure and so does not
have a defined elastic limit. In calculating a limiting value of strain it is good practice to take a
0.1% proof stress and apply a design margin of 2.
Calculation of Sheath Strain from Bending Radius
The sheath strain can be calculated using Equation 4-7 if the radius of curvature of the cable is
known for the design of the particular pipe or duct system. The magnitude of the radius is
obtained from a) the pre-formed pipe bend, b) the manhole offset or s-bend or c) the calculated
curvature of the cable within the wave patterns. The finite element analysis modelling and
normalized span calculation methods described in Chapters 5, 6, 7 and 8 report the magnitude of
strain on the outer surface of the cable jacket. This is converted to the strain on the outer surface
of the metallic sheath (for convenience this simple calculation can be performed by the post
processing program).
From Equation 4-3:
R 2
Ds
= Equation 4-5
j
j
s
s
D
D
= Equation 4-6
Alternatively, if the stress in the sheath is required:
R . 2
E D
s
s
= Equation 4-7
Where:

s
: strain in sheath [per unit]

j
: strain in jacket [per unit]

s
: stress in the sheath [Nm
-2
]
D
s
: outer diameter of metallic sheath or foil [m]
D
j
: outer diameter over jacket [m]
E: elastic modulus of sheath material [Nm
-2
]
R: radius of curvature of cable [m]
Values of E, the modulus of elasticity of the sheathing metals are given in Table 4-7.
EPRI Licensed Material

Parameters Limiting System Performance
4-10
Limiting Value of Sheath Fatigue Strain
Most of the worldwide installations of cable with metal sheaths are in fully constrained
applications where no movement is possible, such as a) direct buried, b) close cleated and b) in
filled ducts. For this reason information on the limiting values of strain necessary to prevent
sheath fatigue are sparse.
The strain limit is based on the number of daily load cycles in the specified design life of the
cable circuit. The higher the number of cycles, the lower is the permissible design strain.
Historically most cables had a lead sheath. Limiting values of strain that have been used for a
typical design life of 40 years (0.146 x 10
5
daily load cycles) with 5 cycles per day are given in
Table 4-8
[viii]
.
Table 4-8
Limiting values of strain for lead sheathing materials
Lead Sheathing Material Limiting Strain [%] Associated Stress [MN m
-2
]
Plain lead 0.016 2.780
Alloy E: (0.4% Sn + 0.2% Sb) 0.037 6.332
Alloy C: (0.2% Sn + 0.075% Cd) 0.024 4.169
Alloy Copper-tellurium: (0.06% Cu + 0.04% Te) 0.048 8.339
The values in Table 4-8 are derived from the levels of strain necessary to achieve a life of 10
7

cycles at room temperature. The limiting values for sheath strain for a life of 5 x 0.146 x 10
5
load
cycles are derived from the stress values using the modulus of elasticity for lead of 17.2 x10
9

Nm
-2
. This value is taken as constant for each of the alloys in the table
[viii]
.
Good service experience has been gained with lead alloy E sheaths on self-contained pressure
assisted cables in many applications where fatigue resistance is important. Alloy E is an
acceptable engineering choice for the installation of XLPE cables in pipes in which a) there is
sufficient clearance in the pipe to permit lateral movement and b) the cable is allowed to flex
within offsets in the joint chamber. Copper-tellurium alloy has a superior fatigue resistance and
has been used both a) in past applications such as cables in unfilled ducts and b) in present
applications such as the 230kV 2500kcmil XLPE duct installation studied in this report.
The preference is to select lead sheaths that have been extruded without stop marks as these
positions introduce the risk of preferential fatigue failure occurring at the interface where the
grains of metal are of increased size. Stop marks occur when the cable movement is stopped to
allow the extrusion press to be recharged with metal. The risk is eliminated by specifying that the
lead sheath be applied by a continuous screw press. The disadvantage is that the choice of lead
alloys available for continuous extrusion is narrower than that available for extrusion in the older
designs of ram press. The type of press may thus restrict the choice of lead alloy and the limiting
fatigue strain.
EPRI Licensed Material

Parameters Limiting System Performance
4-11
Aluminum is much more resistant to fatigue strain than lead alloy and consequently higher
limiting values are permitted. Although corrugated sheaths are more flexible than straight sided
sheaths, tests have shown that the same strain fatigue limits apply. The relationship between
sheath strain and the time to failure is shown in Figure 4-1
[ix]
.

Figure 4-1
Flexing life characteristic for corrugated and smooth sided aluminum sheaths
Figure 4-1 shows that a fatigue strain of 0.25% (change in strain) corresponds to approximately
10
5
cycles. At a rate of five loading cycles per day this corresponds to a life of over 50 years.
The improved fatigue performance of aluminum over that of lead is clear, as the allowable
fatigue strain is approximately ten times that of lead alloy 1/2C. It has been practice to specify
corrugated aluminum sheaths when the cable is required to a) flex cyclically such as in sagged
cleated systems and in semi-controlled positions such as in bridge crossings with expansion
joints and b) withstand significant vibration, for example installation next to a railroad.
There is little published information on the bending fatigue performance of large diameter,
transmission voltage, power cables which employ a radial water barrier formed from a metallic
foil laminated co-polymer. In the absence of published information, it is our engineering
judgment that a thin metallic foil of typically 0.03mm thickness, with the support of a bonded
polymeric jacket with temperature dependent properties, may not exhibit the uniform bending
performance of a self-supporting, thick wall metallic sheath (i.e. equal to or greater than 2mm in
thickness). Prospective users of foil laminate protections for transmission class cables in
applications where cyclic movement is expected to occur throughout the life of the system, such
as in pipes or ducts, are encouraged to require that fatigue tests be performed to confirm
suitability. The following information has been abstracted from similar test protocols to guide the
prospective user. Similarly, a provisional limiting value of strain is given in Table 4-9 for
guidance until test data becomes available.
EPRI Licensed Material

Parameters Limiting System Performance
4-12
Detailed recommendations on a range of tests on laminated coverings are described in CIGRE
Electra 141, published in 1992
[x]
. Specific fatigue tests with limits are not described. The most
relevant tests are a) a thermal cycle test to measure the fatigue resistance during heating and
cooling when submitted to 100 temperature loading cycles to IEC 60840
[xi]
, with a static pre-
formed cable bend of radius 12.5(D+d), where D is the outer diameter and d is the conductor
diameter, and b) a bending test of 3 cycles around a bend radius of 12.5(D+d) and c) a 1 cycle
sidewall pressure test of 7,358Nm
-1
, around a 180
o
pattern of radius 12.5(D+d). None of the tests
in Electra 141 apply a true cyclic strain to the cable for the number of cycles required to simulate
the service life.
Published test results on the higher voltage, larger conductor diameter HV and EHV XLPE foil
cables are rare. It is noted that the CIGRE Electra 141 1992
[x]
recommendation was based on
service and test experience with HV cables containing small diameter conductors, as EHV cables
with foil laminate coverings had yet to be installed.
Development tests in Japan are described
[iv]
on a special design of four layer, aluminum foil
laminate applied to a 275kV XLPE 3000kcmil, 122.6mm OD cable with a PVC jacket. Cable
was formed into a 14m diameter continuous loop, which was then continuously rotated on rollers
to simulate the transportation of the cable within a tunnel for a distance of a) 5km and b) 10km.
Bending fatigue tests were then conducted to evaluate the residual life of the aluminum
laminated tapes (details of the test procedure not given). Graphs are presented of the number of
cycles to failure at different levels of stress, from which it is calculated that, for one cycle per
day for a 40 year life, ie 0.146 x10
5
cycles, the failure strain is 0.6%. In a service application the
polymeric jacket will experience a temperature near to 80
o
C, at which the jacket will be
significantly softer and less able to provide uniform mechanical support to the thin aluminum
foil.
A design factor of 4 has been selected to allow for the expected reduction in mechanical support
from the PVC or PE jacket at 80
o
C. When applied to the strain level at failure of 0.6% this yields
a design value of 0.15%. There is a need for more test data, especially on XLPE cable
constructions intended for pipe and duct applications. It is recommended that cyclic fatigue tests
be performed on foil laminate HV and EHV XLPE cables over the range of temperatures likely
to be encountered in North American applications (e.g. 0
o
C, 20C and 80
o
C).
A third type of sheath exists comprised of thin wall, welded or brazed, sheet of less than 2mm
thickness. The types of sheathing material being stainless steel, copper and aluminum, usually in
a corrugated state to promote flexibility. No information was found on the fatigue performance
of HV or EHV XLPE cables protected with this sheath type. It is recommended that, to
demonstrate suitability for pipe and duct applications, cyclic fatigue tests be performed over the
same temperature ranges recommended for foil coverings i.e. 0
o
C, 20C and 80
o
C.
A summary of the recommended fatigue limits for the different types of sheath is given in Table
4-9.
EPRI Licensed Material

Parameters Limiting System Performance
4-13
Table 4-9
Sheath fatigue limits
Sheath Type Limiting Strain [%]
Category
(1)
Construction Material 0.146 x10
5

cycles
(40 years)
5 x10
5

cycles
Thick wall Corrugated continuous
extrusion
Corrugated aluminum sheath 0.5 0.25
Thin wall Foil laminate +PE Copper + copolymer 0.15
(2,3)
0.01
(2)

Thin wall Foil laminate +PE Aluminum + copolymer 0.15
(2)
0.01
(2)

Thick wall Continuous extrusion Lead alloy: copper-tellurium:
(0.06% Cu + 0.04% Te)
0.048 -
Thick wall Continuous extrusion Lead alloy E: (0.4% Sn + 0.2%
Sb)
0.037 -
Thick wall Continuous extrusion Lead alloy C: (0.2% Sn +
0.075% Cd)
0.024 -
(1): Thick wall is 2mm and greater, thin wall is less than 2mm.
(2): A factor of 4 is taken to extrapolate results from 20
o
C to 80
o
C
(3): Copper expected to have equal or higher fatigue strength than Al depending upon condition
Limiting Value of Sidewall Force
Load current causes the conductor temperature to rise to the operating temperature of 90
o
C. If the
conductor is constrained, longitudinal and lateral forces are exerted. If the cable is unconstrained,
the rise in temperature causes an extension in length. Tensile forces are developed during cooling
resulting from a) conductor relaxation and b) low ambient temperatures. Forces and extensions
were measured experimentally and are listed in Chapter 3 for the 138kV 1500kcmil and 230kV
2500kcmil cables. Pipe and duct-manhole systems are a combination of the two extreme methods
of cable installation in that they are neither unconstrained nor fully constrained. Behavior is
controlled by the pipe clearance such that the cable is free to generate controlled thermal
extension, thereby limiting the axial force to an acceptable design level.
Where the cable in a bend is in contact with the pipe wall the thermomechanical axial force
subjects the cable to a sidewall force. Axial and sidewall forces are generated by the conductor.
When the force is compressive, the cable is pushed to the outside of the pipe wall and when
tensile, is pulled to the inside wall. In both cases the XLPE insulation and PE jacket are in
compression under the action of the sidewall force.
The calculation of sidewall force, F
sw,
is given in Equation 4-8, it is expressed as the force per
unit length between the cable surface and the pipe wall. It is sometimes incorrectly called side-
wall pressure.
EPRI Licensed Material

Parameters Limiting System Performance
4-14
Sidewall pressure is an important calculable design limit for the insulation as the conductor
exerting the force has a defined diameter, D, thus the pressure is Fsw/D.
Sidewall force, F
sw
, is the preferred limiting parameter for the jacket. The term sidewall
pressure has no definable meaning for the jacket as its curved shape is in point contact with the
pipe wall, thereby generating an indeterminate pressure of infinite magnitude. At elevated
temperatures, near the melting point of the jacket, the sidewall force will progressively flatten the
jacket until either the contact area reduces the pressure to a withstandable level, or the metallic
sheath or foil is uncovered. PVC jackets are significantly softer than PE jackets in the normal
operating temperature range. The typical melting temperatures are 105
o
C for LDPE, 115
o
C for
MDPE/HDPE, 130
o
C for HDPE: and 160
o
C for PVC. A jacket of MDPE/HDPE will normally be
selected to give the best resistance to sidewall force.
Equation 4-8
Where:
Fs: sidewall force on the cable in bend [Nm
-1
]
F: axial force in cable [N]
R
b
: radius of bend [m]
Inside the cable, the conductor sidewall load compresses the XLPE insulation and shields and
may result in their permanent deformation. The compressive force may also cause indirect but
equally serious damage, such as the indenting of screening wires into the semiconducting
insulation shield and jacket. Deformation will generally be greater at the cable operating and
emergency rating temperatures as a) thermomechanical forces tend to increase with temperature*
and b) the softness of the XLPE insulation, shields and MDPE jacket increase with temperature.
* The Application Guide in Chapter 13 explains that the selection of an increased pipe or duct
clearance may usefully result in the sidewall force being limited to a maximum at less than the
operating temperature, i.e. in the range of 40-60
o
C.
Limiting values for sidewall force are published to protect XLPE cables during the pullingin
operation. These can be a useful guide in the prevention of crushing and splitting of such
mechanical components as foil laminate sheaths, semi-conducting tapes and cushioning tapes.
These values are not applicable to the polymeric components at operating temperatures, as the
pulling-in operation occurs at ambient temperature when the polymeric components are rigid and
better able to withstand sidewall force.
A limiting value of sidewall force of 29,233Nm
-1
(2,000lbs/ft) is given for pulling-in XLPE
cables
[iii,vi]
at ambient temperature.
Limiting values of sidewall force for XLPE cable and polymeric jackets at operating
temperatures are derived in the next section, with values being given in Table 4-11.
b
sW
R
F
F =
EPRI Licensed Material

Parameters Limiting System Performance
4-15
Limiting Value of Lateral Pressure on the Insulation and Jacket
There are few published values for the limiting lateral pressure for XLPE insulation. Limiting
values are derived later in this section.
The semi-crystalline morphology of XLPE gives it rigidity at ambient temperature, however this
starts to reduce above 60
o
C until it becomes a soft amorphous elastomer at 105
o
C. Melting of
XLPE at 105
o
C is prevented by the presence of crosslinking between the LDPE molecules.
XLPE is capable of being deformed under compressive load by a) permanent compression-set
and b) memory of the elastic deformation, which is locked-in by the formation of the semi-
crystalline structure as the material cools. Elastic memory and deformation can be released by re-
heating the cable, but compression-set cannot be released. The electrical performance of the
cable will be reduced by the total deformation that it suffers, comprising a) the reduction of
insulation thickness and b) the stress raising effect at the insulation-shield interface, due to local
distortion (for example by the impression of an individual shielding wire).
The temperature drop from the conductor to the jacket is typically 10-15
o
C. It is unwise to
operate a PE jacket within 15
o
C of its melting temperature. If an LDPE jacket with a melting
temperature of 100
o
C is selected, then the cable short term emergency temperature would be
limited to 100
o
C (ie 105-15+10). If an MDPE/HDPE grade is selected with a melting
temperature of 115
o
C, then the short term conductor temperature would be limited to 110
o
C.
Guidance on the design temperature for a particular jacket material is obtained by performing a)
compression-set tests and b) VICAT deformation tests on both plaque samples and on cable
samples. HDPE has an attractive elevated melting temperature of 130
o
C. The disadvantages of
HDPE as a jacketing material are that a) the risk of stress cracking is increased and b) the higher
modulus increases the bending stiffness EI, thereby prospectively increasing the axial and
sidewall force.
The lateral pressure limits derived in the following section for XLPE insulation, Table 4-11, are
equally suitable for an MDPE jacket, providing the maximum short term jacket temperature does
not exceed 100
o
C. This effectively limits the conductor emergency operating temperature to
110
o
C.
Table 4-10 gives values from laboratory measurements that were performed for the Central
Electricity Generating Board in 1984 in anticipation of the future installation of transmission
class cables
[xii]
. Two types of tests were performed. A bending rig was designed to test a 14m
length of 132kV, 3,500kcmil, lead sheathed cable with an XLPE insulation thickness of 21mm.
The cable was laid around a semicircular former with a 20D radius of 2.12m. A range of tensions
was applied to the cable such that sidewall loads were generated around the bend, tending to pull
the conductor through the insulation. Tests were performed at temperatures of 90
o
C and 105
o
C.
The tests culminated with the application of 13 load cycles at 105
o
C at an applied insulation
pressure of 8bar. This resulted in an insulation deformation of 2.45mm (11.6%). The cable was
then ac voltage tested whilst still in the bending rig and was demonstrated to be free of partial
discharge at 10pC with a test voltage of 100kV (1.3Uo).
EPRI Licensed Material

Parameters Limiting System Performance
4-16
Table 4-10 also lists some of the results from a series of bench top tests that were performed on
short 150mm long cable samples heated in an oven whilst a lateral force was applied to a rigid,
replacement conductor. The sheath had been replaced by a steel tube. Different sizes of XLPE
cable were tested, including a 275kV 1260kcmil

cable core with an insulation thickness of
25mm. A graph of the insulation deformation with pressure at 92.8
o
C and 105.6
o
C is given in
[xii]
.



Table 4-10
Insulation deformation and pressure measurements on XLPE cables
90
o
C Temperature 105
o
C Temperature Cable Sample
Lateral
pressure [bar]
Insulation
indentation
[%]
Lateral
pressure [bar]
Insulation
indentation
[%]

5.3 2.8 - -
8.0 5.4 - -
- - 8 11.6
Bend test rig, 14m long cable.
132kV, 3500kcmil, 21mm
insulation.

Extrapolated from the test result
for 8bar at 90
o
C.
15 11.6 - -

6 5 1.8 5
12.2 10 4.3 10
Bench top compression rig,
150mm long samples of XLPE
core.
275kV, 1260kcmil, 25mm
insulation.
Interpolated from graph.
14 10 (nom) - -

In both of the test methods the lateral pressure acting on the insulation was calculated by
dividing the sidewall force per unit length by the projected conductor lateral area per unit length :
Equation 4-9
Where:
P: pressure on insulation [Nm
-2
]
F
sw
: sidewall force per unit length acting on cable [Nm
-1
]
R
c
: radius of conductor [m]
c
sw
R 2
F
P =
EPRI Licensed Material

Parameters Limiting System Performance
4-17
The Principal Investigator attended an international workshop held in Paris in 1987 to discuss the
thermomechanical limits of polymeric cables. The workshop was held at the same time as the
1987 Jicable Conference dedicated to the technology of polymeric cables. XLPE cables up to
150kV were being installed at that time. Major utilities were investigating the prospective use of
transmission class cables above 200kV and up to 500kV with an operating temperature of 90
o
C.
The workshop considered that a) deformation of the insulation should be limited to 10% and b)
XLPE cable should not exceed a short term emergency temperature limit of 105-110
o
C. The
latter temperatures were slightly above the transition temperature of 103-105C at which
crosslinked low density polyethylene loses its crystalline structure and behaves as a soft
elastomer.
Since 1987 significant advances have occurred in the application of XLPE cables leading to the
commissioning of the first major transmission class circuits in the late 1990s
[xiii]
. Care has to be
taken when reviewing whether thermomechanical design limits have been truly validated by time
in service. For example, in the first long-length circuits at 500kV in Japan a) a factor of safety of
10% was included in the calculation of the design stress and service life, b) 2mm of extra
insulation thickness was added, c) the operating temperature was reduced by 5
o
C to 85
o
C and d)
three parallel circuits were installed when two would have sufficed, thereby reducing the loading
and reducing the operating temperature.
Technological improvements and commercial competition have led to reductions in the design
margins to permit the insulation electrical design stress to be increased
[xiii]
. For example the
insulation thickness of some 132kV cables has been reduced from 21mm to 14mm and the
insulation thickness of some 220/230kV cables has been reduced from 27mm to 21mm.
Transmission class cable designs have been introduced with wire shield conductors applied over
the extruded core for use with both thin stainless steel sheaths and with metallic foil laminates.
These are in addition to the existing traditional designs of extruded metallic sheaths, which have
smoother inner surfaces facing the insulation.
In recommending safe thermomechanical design limits for transmission class cables the
Investigators consider that uncontrolled indentation of the insulation is now more likely to occur
with a) present designs of XLPE cable and b) present levels of circuit loading. Thus the
deformation limit of 10% reduction in insulation thickness proposed in 1987 is now considered
by the Principal Investigator to be unwise. In combination with deliberate reductions in other
design tolerances, the 10% value introduces an unnecessary risk to service life, particularly as
XLPE transmission class circuits are now being installed and operated, some of which are likely
to be loaded close to their design operating temperature of 90
o
C.
It is recommended by the Principal Investigator that, for the purpose of system design, the
limiting transverse pressure acting on the insulation is the value which limits its reduction by
equal to or less than 5%. This recommendation should be reviewed when cable samples from
future tests and service installations become available for analysis.
Recommended pressure limits were derived from Table 4-10 and are given in Table 4-11,
together with the resulting sidewall force limit at bends calculated from Equation 4-9 for the
138kV and 230kV cable sizes studied in this project. It should be noted that the 230kV cable has
EPRI Licensed Material

Parameters Limiting System Performance
4-18
a low design stress (i.e. thick insulation) and that the 138kV cable has a moderately increased
design stress (i.e. slightly thin insulation), thus in these cases the 5% design limit would be a safe
value.
Table 4-11
Lateral force limits for 138kV and 230kV XLPE cables
Insulation Pressure P and Corresponding Sidewall Force F
sw

Limits for 5% Indentation [Nm
-1
]
Cable Design
2 bar limit at 105
o
C
Emergency Loading
6 bar limit at 90
o
C
Continuous Loading
138kV, 1500kcmil, 18mm XLPE 6,580 19,740
230kV, 2500kcmil, 27mm XLPE 8,840 26,520
Limiting Value of Axial Thermomechanical Force
The generation of thermomechanical wave patterns within the pipe, results in both reduction in
axial thrust and increase in cable curvature. The sidewall force is reduced within the preformed
pipe bends, but increased in the regions within the main pipe span that contains cable
deformation patterns.
The axial force necessary to limit the insulation deformation for a given bend radius is calculated
using Equation 4-10 (derived from Equation 4-8 and Equation 4-9). The limiting pressure P is
obtained from Table 4-11. The limiting bending radii R
b
for preset bends is obtained from Table
4-1 and for the 138kV and 230kV examples from Table 4-4 and Table 4-5. The local bending
radii R
b
for the cable within thermomechanical deformation patterns is obtainable from the finite
element modelling studies given later in this report.
F 2R
b
R
c
P Equation 4-10
Where:
F: limiting value of thermomechanical force [N]
R
b
: radius of the cable bend [m]
R
c
: radius of the cable conductor [m]
P: limiting value of lateral pressure on insulation [Nm
-2
]
Limiting Value of Pipe to Cable Clearance
The cable to pipe clearance is the most important parameter in the thermomechanical design of
pipe and duct-manhole systems. The radial clearance determines the balance between the
presence of deformation patterns and the reduction in thermomechanical axial thrust that results.
However, in itself, pipe clearance is only a limiting factor in a) providing sufficient clearance for
installation and b) preventing jamming during installation and in service.
EPRI Licensed Material

Parameters Limiting System Performance
4-19
Pipe Jam Ratio
If the pipe is sufficiently large in diameter one of the three cables may slip between the other two
causing them to jam against the pipe wall. This will only occur where the sum of the three cable
diameters is close to the internal diameter of the pipe. This effect is termed the jam ratio and is
defined according to Equation 4-11:
J = D/d Equation 4-11
Where:
J: jam ratio [per unit]
D: inside diameter of the pipe [m]
d: outside diameter of the cable [m]
To prevent jamming occurring, the 1992 EPRI Underground Transmission Systems reference
Book
[vi]
recommends

that J should be outside the range of 2.9 to 3.1
.
The equivalent range given
in the more recent publication of AEIC CG5-2001
[iii]
is 2.8-3.0. Variations in the cable diameter,
pipe diameter and ovality in the pipe diameter at bends were taken into account in these limits
[iii]
.
Bibliography
Further background reading is given in references
[xiv,xv,xvi,xvii,xviii]
.


i
AEIC CS7-93 (3rd Edition) June 1993 Specifications for crosslinked polyethylene insulated shielded power
cables rated 69 through 138kV.
ii
IEC 62067 (2001-10) Power cables with extruded insulation and their accessories for rated voltages above
150kV (Um = 170kV) up to 500kV (Um = 550kV) Test methods and requirements, 2001.
iii
AEIC CG5 2001, Underground extruded power cable pulling guide.
iv
Umeda K et al, Development of 275kV XLPE cable with aluminium laminated tape and radial moisture barrier,
Jicable 2003, paper A.1.5.
v
Electric Cables Handbook, Blackwell Science, third edition, 1997, ISBN 0-632 04075-0.

vi
Underground Transmission Systems Reference Book, 1992 Edition, EPRI, Palo Alto CA,TR-101670, Research
Project 7909-01.

vii
Holdup W, Occhini E, Skipper D.J, Thermo-mechanical behavior of large conductor cables, Paper 31 TP 67-478,
IEEE trans, 1967.
viii
Hiscock S.A, Lead and lead alloys for cable sheathing, Ernest Benn Ltd., London 1961.
ix
Galloway S.J, Thermomechanical design of supertension cable systems, IEE Discussion Meeting, London,
October 1994.
x
Guidelines for tests on high voltage cables with extruded insulation and laminated protective coverings, CIGRE
Electra 141, April 1992.
EPRI Licensed Material

Parameters Limiting System Performance
4-20

xi
IEC60840, 1999, Power cables with extruded insulation and their accessories for rated voltages above 30kV (Um
= 36kV) up to 150kV (Um = 170kV)- test methods and requirements.
xii
Head, Crockett, Taylor, Wilson, Thermomechanical characteristics of XLPE HV cable insulation, 5
th
International
Conference on Dielectric Materials, Measurements and Applications, IEE, 1988
xiii
Cable system technology review of XLPE EHV cables, 220kV to 500kV, EPRI, Palo Alto, CA: 2002. 1001846.
xiv
Anelli P, Donazzi F, Lawson W.G, The fatigue life of lead alloy E as a sheathing material for submarine power
cables. IEE Trans on Power delivery, Vol 3 No.1 1988.
xv
Holttum. W. The installation of metal sheathed cables on spaced supports. Proc IEE A 102, 729 742. (1955).
xvi
Harvard D.G. Selection of cable lead sheath alloys for fatigue resistance. IEEE PAS-96 No.1 Jan/Feb 1977.
xvii
Brincourt T, Dorrison E, Thermomechanical behavior of 400kV synthetic cables, Jicable 1999, paper A 4.2.
xviii
Reliability parameters of XLPE UT systems based on EDF experience, Part 2 Thermomechanical stresses on
cables and accessories, Unattributed publication.

EPRI Licensed Material
5-1
5
CABLE AND ROUTE MODELLING: METHODS AND
ANALYSIS
Summary
Prior to this project, mathematical modeling of the thermomechanical performance of cables and
joints in pipe systems had not been possible because of the complexity of the interaction of the
thermomechanical phenomena, the multi-directional route geometry, the long route lengths and
the dynamic nature of frictional contact and pattern formation between the three cables and the
pipe wall.
Finite element analysis using the traditional implicit method had been successfully applied in
recent years to solve cable thermomechanical problems in which the cable was either rigidly
constrained (e.g. close cleated or buried direct), or allowed to move in simple semi-constrained
situations ( e.g. uniformly sagged from cleats, controlled bends underneath cable terminations
and controlled crossings of bridge expansion joints). However, neither of these cases apply to the
behavior of cables in ducts and pipes. The implicit method has been successfully applied to the
simple cases of one cable in a duct in an elementary route
[i] [ii]
, this work used the ABAQUS-
Standard software. As described later, the implicit method is unsuited to model the dynamic
behavior of long complex routes and of three cables within one pipe.
Exploratory work before this project demonstrated the potential of a different solving technique
using the explicit finite element method, originally developed to solve dynamic problems. A
pipe containing a 30m length of cable had been successfully modeled in this way. The main
thrust of this project would to a) understand and overcome any limitations in the explicit
technique, b) apply the technique to long length models of real pipe and duct routes containing
cables and joints and c) analyze cable and joint thermomechanical performance. Additionally the
technique would model the effects of traffic vibration and short circuits. The understanding of
the thermomechanical performance would then be used as the basis for the preparation of the
design guide described later in Chapter 13.
During the project the explicit modelling technique was successfully extended from 30m to
1,200m and was applied to model the behavior of real routes comprising a) one cable in a duct
and b) three cables in a pipe. These models typically comprised two pipe spans containing
between one and three straight joint manholes. When a part of a route is studied a decision has to
be taken on where to terminate the model, this is usually selected to be at a straight joint position,
where for the purposes of modeling, it is treated as an anchor joint. No decision was necessary
for the 230kV duct route as every manhole contained an anchor joint and so the model could be
terminated at any of the existing manhole positions. In total the technique was successfully
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-2
applied to model a) four double spans in a 138kV pipe route and b) four spans in a 230kV duct-
manhole route containing three manhole positions.
At the end of the project an experimental solution was successfully obtained on a single beam
trial model of 16,000m length. This points to the future capability of studying a route of
approximately 16km length (10miles) comprising 20-30 spans in one model. However, there is
no true engineering benefit in solving one large model of a long length cable circuit, providing
sufficient spans can be modeled a) to permit cumulative sliding to occur from one span to
another and b) to eliminate possible end effects of introducing model anchor joints.
During the project the workers faced a number of problems. There was a choice of engineering
facilities in the software, few of which had previously been applied to modelling the dynamic
behavior of cables in pipes. It was necessary to select and prove each particular facility and when
necessary to either adapt, or discard it. In parallel with the modelling work the software house
was continuously developing and issuing updates to their finite element program. This gave the
opportunity to feed back adverse experience and also to request enhancements. In this way the
ability to model long length cable routes in pipes was successfully developed.
The modeling method is now a proven technique that can be applied with confidence to study the
thermomechanical performance of both planned and existing routes. In terms of the Guide to the
Design of Duct and Pipe Systems in Chapter 13, the explicit finite element method is the first
choice as it models performance as closely to real life as possible and requires fewer engineering
assumptions. However significant resources and expertise is required in a) preparing the models,
b) checking that the solution is satisfactory and c) analyzing the output data. Five engineers (two
cable and three FEA engineers) were involved at the most intensive parts of the project. At the
present time two engineers are required to model and analyze the thermomechanical behavior of
a route.
The initial modelling output showed that a) cable deformation patterns formed adjacent to the
positions of bends and joints, b) each of the three cables in a pipe systems were able to move to
form patterns with more freedom than expected and c) that joints in pipe systems deflected
within their casings. The formation of cable deformation patterns within pipes and ducts is
studied in detail in Chapter 6 and Chapter 7. The deflection of joints in pipe systems is studied in
Chapter 11.
The question at the end of the work is does the technique model the real world? The work on
modelling described in this chapter and on cable parameter measurements described in Chapter 3
has given convincing answers to a) what happens in pipe and duct systems and b) the importance
of the input parameters. It was possible to validate the explicit software a) by comparison with
the industry standard implicit software and b) by modelling the beam deflection experiment.
There was no experimental information available on the behavior of XLPE cables in pipes.
However photographic evidence was available of the thermomechanical displacement of joints
within their casings in a 138kV XLPE pipe system. A comparison of the modeled deflection with
the actual joint deflections demonstrated the importance of a) closely modelling the joint and
cable constraints present within the real casing and b) the values for the key variables of
coefficient of pipe wall friction () and of the bending stiffness (EI) of the cable. The comparison
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-3
showed that the thermomechanical cable forces present were larger then those initially modeled
and in consequence the values of EI and needed to be higher than those originally selected for
the studies on the 138kV and 230kV cables. The sensitivity study in Chapter 7 quantifies the
effect of varying each cable parameter and will permit the work in this report to remain relevant
when future information inevitably becomes available from circuit experience and full size
experiments.
Contents
This chapter starts with the objectives set for the modelling technique.
An explanation is made of the model cable route. This starts with the cable components and ends
with the geometry of the pipe and joint casing. In particular this section explains how the
temperature dependence of the XLPE insulation is dealt with.
The preparation of the loading cases that are applied to the model is described.
The step by step processes are explained of a) preparing the model data for the cable, joints and
pipe route, b) preparing and checking the solution and c) performing the engineering analysis
and interpretation of the results.
The computing requirements are listed and the principle of the explicit finite element analysis
method is explained.
The validation of the explicit FEA cable modelling method is described.
Finally, areas for possible future work are discussed.
Modeling: Objectives
The prime objective was to produce a method of simulating the behavior of power cables
installed in ducts with particular emphasis on modeling real installations that include typical duct
lengths between manholes of 600m (2000ft). While previous work had concentrated on modeling
behavior in small test installations that include flat, straight pipes with one or two curves, the
analytical objective of this task was to produce a method in which long pipe lengths containing
many straight sections as well as changes in elevation could be investigated. Additionally, the
method had to cater for three cables per duct as well as single cables and allow for the inclusion
of joint manholes between the duct runs.
To calculate the effect of the deformation of hundreds of meters of cable on a global scale, it was
unnecessary to model the constituent parts of the cable in micro detail. It would be necessary to
model the cable in reduced complexity whilst maintaining the quality of the results. The
modeling details are:
3 cables in a pipe route
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-4
1 cable in a duct-manhole route
Representation of the cable performance by division into three components i) the conductor,
ii) the extruded core and iii) the metallic sheath/foil/jacket; using three parallel beam
elements
Each cable component to be represented by axial stiffness, lateral bending stiffness and
torsional stiffness
The conductor to be represented by a stress strain characteristic typical of stranded
conductors
The effective modulus of the XLPE core to be derived from a) a radial temperature gradient
across insulation, b) non-linear dependence of XLPE elastic modulus with temperature and c)
difference in XLPE elastic modulus in compression and tension (i.e. either side of the neutral
axis in bending)
Cable within bends; using curved beam elements
Friction between each cable and the pipe wall
Friction between individual cables
Representation of the joint casing in the pipe system, comprising steps between the
telescoped pipes
Representation of the one-piece pre-molded joint in the 138kV pipe system
Representation of the composite anchor joint,
cable offsets and supports/cleats in the 230kV duct-manhole system
Reporting sidewall pressure between the cable and the pipe/duct wall
Gravity
Heating through a defined temperature rise e.g. 15C to 90C
Temperature cycles
Method to represent traffic induced vibration
Method to represent impulse loads
Modeling the Cable
An effective way of modeling a power cable inside a pipe is as a beam; this represents both its
effective overall geometry and its predominant mode of structural behavior i.e. bending. The
finite element method is a) computationally powerful in analysing a multi-variable problem and
b) flexible in adaptation to different conductor sizes, transmission voltages (insulation
thicknesses) and sheath/jacket finishes.
The XLPE cable is represented as three parallel beams comprised of equal length longitudinal
elements. The ABAQUS facility called general contact was selected to model friction, this
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-5
required that the minimum length of an element be greater than 1.66 x the contact diameter. The
diameter of the larger of the two cables was 125mm. The smallest cable element length was set
at 300mm in the main pipe span, but smaller when modelling some parts of the joint geometry in
detail (in which case general contact was not applied).
Each of the three beam elements was given the property of an annular cylinder:
The first beam element models the conductor.
The second beam element models the extruded XLPE insulation and shields.
The third beam element models the remaining layers, comprising various combinations of a)
shield wires or tapes, b) metallic sheath, c) aluminum or copper foil laminate and c)
polymeric jacket.
Each beam element has a node point at either end with one integration point at the center. Each
element has the mathematical stress-strain behavior of a perfect beam. The geometry of each
beam is a straight, annular cylinder, with each cylindrical layer covering those beneath. A full set
of mechanical properties is specified for each beam:
Axial stiffness, EA
Bending stiffness in two different planes, EI
Torsional stiffness in three different planes, GJ
The stiffness of each of the elements was represented as being temperature dependent. The
values were obtained from the analysis of experimental results from a) 138kV and 230kV cable
samples at different temperatures and b) from a previous EPRI project recording the results of
bench top measurements on plaque and cable samples of XLPE
[iii]
.
Each beam was given the same value of coefficient of thermal expansion () this was found to
be necessary to obtain satisfactory solutions to long length cable route models. It was confirmed
that this single value method gave the same mechanical performance, providing it was the
effective value of for the cable as a whole as that of a model with a different value for each
layer. This was verified by constructing two short length conventional implicit finite element
models with detailed cylindrical symmetry, one with a constant coefficient of expansion and the
other with a different value for each layer. Effective values of for the whole cable were
initially derived from these models and later in the project from measurements made on the
138kV and 230kV cable samples.
The stiffness characteristics were specified to be linear, i.e. a constant value of stiffness
representing a linear relationship between stress and strain. At the start of the project it was
found that satisfactory solutions for a particular cable route could only be obtained with linear
properties. As a result of modeling and software improvements it is considered that, for future
work, non-linear parameters could be represented.
At the start of the project the first approximate value of bending stiffness (EI) for the XLPE
insulation was taken to be 10% of the axial stiffness (EA). For later work the effective value of
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-6
EA of the XLPE insulation was derived from a separate detailed implicit FEA model in which
the insulation was divided into ten layers and each ascribed values for Youngs modulus of
elasticity in compression (on the inside of the bent insulation) and in tension (on the outside of
the bend) at different temperatures. A radial temperature gradient was applied across the
insulation according to the rating of typically 10
o
C.
For the final modeling work (the sensitivity study and the studies on the 138kV and 230kV
routes), the effective EI was abstracted from the cable measurements recorded in Chapter 3.
The value of EI for the sheath/jacket layer was obtained by subtracting the measured EI for
the cable core (XLPE insulation and conductor) from the measured EI for the complete cable
at each temperature.
The value of EI for the XLPE insulation was derived using the calculated second moment of
area for the cylindrical layer and the effective modulus of elasticity from Table 5-3 and Table
5-4. The values of the modulii of elasticity are given in the following section.
The value of EI for the conductor was obtained by subtracting the calculated EI for the
insulation from the measured value for the core (XLPE insulation and conductor). This
approach was valid as the EI of the conductor is significantly greater than that of the
insulation. (An approximate value of the conductor is revealed by the 105
o
C measurements,
at which the soft XLPE insulation contributes little to EI).
Calculation of the Effective Axial and Bending Modulii of XLPE Insulation at
Different Temperatures
The properties of the Youngs modulus of elasticity (E) of XLPE are complex. E has different
properties in a) tension as shown in Figure 5-1 and b) compression as shown in Figure 5-2. E
also reduces very significantly with temperature when the morphology changes from a semi-
crystalline rigid solid to a soft elastomer at 105C. The increase in softness and elasticity become
pronounced at temperatures above 60C. The flow of heat outwards from the conductor through
the insulation produces a temperature gradient such that the outer surface may be 5C to 15C
cooler than the conductor (depending on the current rating and the thickness and diameter of the
particular cable).
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-7

Figure 5-1
Elastic modulus of XLPE in tension (linear/linear scale)

Table 5-1
Modulus of XLPE in tension at different temperatures
Modulus of Elasticity of XLPE in Tension
Temperature [C] Young's Modulus [p.s.i.] Young's Modulus [N/m
2
]
10 10430 71.91 x 10
6

20 7825 53.95 x 10
6

30 5550 38.27 x 10
6

40 3800 26.20 x 10
6

50 2600 17.93 x 10
6

60 1750 12.07 x 10
6

70 1200 8.27 x 10
6

80 775 5.34 x 10
6

90 500 3.45 x 10
6

95 350 2.41 x 10
6

100 180 1.24 x 10
6

105 90 0.621 x 10
6

110 80 0.552 x 10
6

120 72 0.496 x 10
6

130 72 0.496 x 10
6

EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-8

Figure 5-2
Elastic modulus of XLPE in compression versus temperature (linear/linear scale)

Table 5-2
Elastic modulus of XLPE in compression
Temperature [
o
C] Young's Modulus [p.s.i.] Young's Modulus [N/m
2
]
10 37000 255.11 x 10
6

20 27500 189.61 x 10
6

30 17500 120.66 x 10
6

40 11250 77.57 x 10
6

50 7500 51.71 x 10
6

60 5800 39.99 x 10
6

70 4500 31.03 x 10
6

80 3150 21.72 x 10
6

90 2100 14.48 x 10
6

95 1300 8.96 x 10
6

100 700 4.83 x 10
6

105 280 1.93 x 10
6

110 162.5 1.12 x 10
6

120 131.25 0.905 x 10
6

130 131.25 0.905 x 10
6

EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-9
The axial stiffness EA is predominantly influenced by the cooler outer insulation because EA is
proportional to diameter. The bending stiffness EI is very dependent upon the outer layer
because the second moment of area I is proportional to the 4th power of diameter.
It is necessary to calculate the effective values of EA and EI for the geometry and ampacity of
each cable as the selected beam element model of the cable permits only one set of values to be
entered. The method of calculation was to construct axi-symmetric FEA models of the insulation
dimensions for the 138kV Consolidated Edison and 230kV LADWP cables, each being divided
into 10 cylinders. The ampacity ratings of each cable were used to calculate the temperature
gradients across the insulation. The 138kV gradient was 8C at the 90C conductor temperature
and the 230kV gradient was 16C. Temperatures were applied to each of the cylinders and values
at each temperature were abstracted from Figure 5-1for the tensile modulus and Figure 5-2 for
the compressive modulus. These graphs had been redrawn from the experimental results in EPRI
report EL-938
[iii]
. Corresponding tabular values are recorded in Table 5-1 and Table 5-2. The
models were first a) deformed in axial compression, then b) deformed in axial tension and then
c) bent. The deformation results were then compared with a single beam model of the cylindrical
insulation and, by iteration, single values of EA and EI were derived; these are shown in Figure
5-3 and Figure 5-4 and recorded in Table 5-3 and Table 5-4.

Figure 5-3
Effective axial and bending stiffnesses for XLPE insulation in a 138kV 1500kcmilcable

EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-10
Table 5-3
Effective axial and bending stiffnesses for XLPE insulation in a 138kV 1500kcmil cable
Conductor
Temperature
[C]
Temperature
Gradient
[C]
Axial Stiffness
EA

[N]
Bending Stiffness
EI
[Nm
2
]

15 0 8.01 x 10
5
204.99
20 1 6.96 x 10
5
179.04
30 2 4.63 x 10
5
122.34
40 3 3.06 x 10
5
82.57
50 4 2.07 x 10
5
56.66
60 5 1.56 x 10
5
41.68
70 6 1.23 x 10
5
31.97
80 7 0.92 x 10
5
23.62
90 8 0.64 x 10
5
16.45
100 9 0.34 x 10
5
9.29
110 10 0.096 x 10
5
2.91


EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-11

Figure 5-4
Effective axial and bending stiffnesses for XLPE insulation in a 230kV 2500kcmil cable

Table 5-4
Effective axial and bending stiffnesses for XLPE insulation in a 230kV 2500kcmil cable
Conductor
Temperature
[C]
Temperature
Gradient
[C]
Axial Stiffness
EA x 10
-5
[Nm]
Bending Stiffness
EI
[Nm
2
]
15 0 16.5 x 10
5
824.39
20 2 14.6 x 10
5
736.37
30 4 10.1 x 10
5
525.33
40 6 6.83 x 10
5
364.50
50 8 4.7 x 10
5
254.6
60 10 3.45 x 10
5
184.85
70 12 2.75 x 10
5
144.03
80 14 2.15 x 10
5
111.94
90 16 1.58 x 10
5
82.36
100 18 1.05 x 10
5
56.52
110 20 0.50 x 10
5
30.43
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-12
Modeling the Pipe or Duct
The pipe is modeled as a rigid cylindrical surface that can neither move nor deform, thus
material properties for different types of pipe and duct are not required. The cylindrical surface
was comprised of 24 elements. In the straight sections of the pipe route, the longest element that
was used was 100m. In pipe bends the element lengths were the same as used to model the cable
i.e. 300mm minimum. Curved elements were used at bends. The economical specification of
rigid elements for the pipe gave significantly faster solution run times.
Changes in route direction and inclination were modeled in the [x,y,z] planes. The study of
routes containing vertical sections was excluded from the scope of the study. This was partly
because a) the 138kV and 230kV cable routes studied had only shallow slopes and partly
because b) of concern that this would inhibit the application of the modelling technique to long
route lengths. It is now considered that vertical section of route could be modeled.
Modeling the Joint and Casing
The three beam element model of the cable span was also used to model a) the cable geometry
within the manhole and b) the joint geometry. In the models of long pipe spans the joints were
represented as having simple cylindrical geometry.
In the detailed and separate model of the pipe system manhole it was necessary to select short
lengths of a) beam elements and b) transition elements to be able to represent the complex
geometry of the joint. The transition elements modeled the conical end shapes of the joint bodies
and were necessary to connect beams representing large differences in diameter. The ABAQUS
general contact facility was not used to represent friction with transition elements.
The casing which houses the three joints in a pipe system was modeled using rigid elements to
form a geometrically accurate surface, representing five different pipe diameters welded
together, Figure 5-5. The rigid surface was a continuation of that modeling the pipe surrounding
the main spans. Thus the cables and joints were free to a) slide longitudinally through the pipe
and casing until the joints contacted the smaller diameter pipes, or b) to deflect laterally, twist
and rotate until they contacted the casing surface.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-13

Figure 5-5
View of modeled joint casing in an in-line 138kV pipe system
It was not necessary to model the joint in the 230kV duct system in detail as an anchor joint was
employed together with a cable cleat on either side, thus the joint was prevented from movement
and deflection. The cable offsets and cleats within the manhole were modeled in detail.
Different types of constraint were applied to the joint models at the element nodes. The anchor
joint in the duct-manhole system was simply pinned to a rigid anchor at its center line. The three
joints within the casing of the pipe system were generally not constrained to the casing in the
longitudinal direction, but were constrained together a) at the initial stage of the project using a)
radial thrust connectors, b) triangular elements with connected elasticity to model binding tapes
and c) at the end of the project using rigid connectors to model the role of the central aluminum
spider in holding the joint bodies parallel, whilst giving freedom to rotate, twist and move
longitudinally.
Different types of constraining cleats were applied to the cable model at the element nodes both
within the manhole and within the joint casing to represent different types of cleat, for example
a) guide cleats which positioned the cable in two geometric planes, but permitted sliding in the
axial direction, b) clamp cleats that applied a longitudinal constraining force, c) ratchet cleats
that permitted a clamping cleat to travel under the constraint of a longitudinal spring and d) cable
stockings that limited movement in one direction, but not in the reverse direction.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-14
Modeling Friction
This was a key part of the modeling. The cable beam senses the surface of the pipe and
registers that contact has occurred. It registers the coincidence of the geometric co-ordinates of a
single line beam (representing the centerline of each cable) with those of the inner surface of a
cylinder of diameter (Dp-Dc). The latter is the locus of the cable centerline when its outer
surface of diameter Dc contacts the inner surface of the pipe of diameter Dp. The use of this
surface-based algorithm is more accurate than a contact-based approach. When surface co-
incidence occurs, the ABAQUS general contact facility calculates the values of contact
pressure and frictional constraint as though the surface of the cable beam was in contact with the
pipe. Frictional constraint was calculated using a shear force definition. The same algorithm was
used to calculate friction between each of the three cables.
Loading Conditions: Heating and Load Cycling
Gravitational force and thermomechanical expansion strain were applied in three or more
computational steps. Each is sub-divided into the number of computational time steps at which
output reports are required. For example, the specification of 30 time steps was used when
heating from 15
o
C to 90
o
C, thereby giving full output of force and movement at each 2.5C
increment in temperature. (If too may time intervals are specified the solution run time becomes
excessively long, ie longer than 16 hours, particularly if load cycling is applied.)
First Computational Step: Dropping under Gravity
The cables were first positioned 2mm above the bottom surface of the pipe. The three cables in
the pipe system were positioned in triangular trefoil formation, however they could equally have
been positioned in, for example, an inverted cradle formation. They were then dropped under
gravity, such that they fell in a natural position according to the geometry of the route. The end
boundaries of the cable were then fixed. In the routes that were studied the angles of pipe
inclination were less than the critical sliding angle of 5.6
o
, for the coefficient of friction of 0.1,
thus the cable did not slide when it was dropped. If a greater angle is required to be modeled then
it would be necessary to fix the end of the cable prior to it being dropped in the same way that a
cable in the real world would have to be gripped to prevent it sliding downhill after installation
and during jointing.
At the ends of the cable spans, boundary conditions were selected that did not apply artificial
curvature to the cable. This avoided the risk that rigid pinning could initiate thermomechanical
deformation patterns. The end of the cable has to be pinned longitudinally, but requires freedom
to translate, rotate and twist across the plane of the pipe cross section, thereby giving the cable
freedom to move at a later stage of temperature rise. This type of boundary would act as though
there was another length of cable on the opposite side of it. The stipulation that the end has to
remain at 90
o
to the plane would simulate a mirror of symmetry as though the model continued
but with a reverse image of the same model shape.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-15
Second Computational Step: Heating
The thermal expansion facility of the software is applied. Typically the cable was heated through
75
o
C from an ambient starting temperature (15
o
C) to the cable design temperature (90
o
C). The
effect is of the cable being switched onto full rated load, but with a slow, natural rate of rise of
temperature. This is consistent with the long thermal time constant of a transmission class cable
and low rate of temperature rise. Thus, it is not necessary to model transient heating effects by
the inclusion of thermal capacity. The heating step is sub-divided into computational time steps
to give output at the required temperature increments, as described previously.
Third Computational Step: Constant Temperature
The third time step is specified to represent a nominal period at design temperature (note that real
time has no meaning). This step is simply for the benefit of the observer to separate the
thermomechanical activity during the heating and cooling phases, as no activity occurs at
constant temperature.
Fourth Computational Time Step: Cooling and Completion of Load Cycle
The temperature is specified to fall either to ambient temperature (a free choice exists, higher or
lower than the original ambient temperature), or to an intermediate load cycling temperature of
the level that would occur if the cable was experiencing daily load fluctuations. In this project
the cable was load cycled back to 15
o
C to permit the maximum contraction movement and strain
levels to be studied.
By the repetition of the number of computational process steps the required number of load
cycles can be applied. At the start of the project it was possible to run three loading cycle, with
computer run times in excess of 22 hours. During the project significant efficiencies in run time
were achieved. At the end of the project the processing run time of one load cycle on the 200m
long sensitivity model had been reduced to 10 minutes and the time of a double span model on a
real 138kV route had been reduced to 2hours 18mins. These short run times mean that a number
of runs can be chained to run consecutively overnight. A maximum number of 20 load cycles
was successfully applied to the sensitivity study route, with the run terminating after 23 cycles
due to insufficient hard disk space. This limitation is easily overcome by reducing the request for
output time steps and by making more disk space available.
Loading Conditions: Vibration and Short Circuit
These load conditions are described in detail in Chapter 9 and Chapter 11. In both cases the new
load case was applied in an additional computational step, inserted after the cable model had
been heated to its operating temperature, such as in computational step three.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-16
Computational Step: Vibration
To simulate road traffic vibration a second, mobile pipe surface of defined length (10.7m) was
defined to have the same surface diameter (Rp-Rc). The mobile pipe surface was termed a
shaker. The cable could sense both the static pipe surface and the mobile surfaces even though
they were exactly coincident, but the pipe surfaces could not sense or interfere with each other.
The required amplitude and frequency of vibration were applied to the mobile pipe. It was then
moved along the pipe route at a speed of 30mph simulating vibrations generated by the passage
of a truck along a rough road surface. The vibration was restricted to the straight sections of the
route.
Computational Step: Short Circuit
To model the application of an electrical short circuit, the ABAQUS facility of a right angled
radial thrust connector was connected between the beam element nodes in each pair of the phase
cables (e.g. between phase cables A and B). Thus each cable was connected to the other two
cables. The thrust connector was programmed to apply the appropriate force vector generated by
the specified short circuit current level at the appropriate supply frequency (60Hz). The thrust
connector applied a time varying force to the cables, whilst allowing movement of the cables
within the pipes and ducts in all three planes [x,y,z].
Process of Modeling a Cable Route
Firstly the [x,y,z] co-ordinates have to be abstracted from the route records by hand. This is a
very labor intensive task and has to be performed by a combination of a) transcribing the co-
ordinates from the statements on the drawing and b) direct measurement and scaling off the
drawing. An alternative is to measure and scale-off a scanned electronic drawing using an on-
screen measuring technique. The co-ordinates have to be registered at every change in slope,
direction and curvature. Generally it was found that the route records had anomalies as a) the co-
ordinates of different points on a drawing didnt match with sufficient accuracy for modelling
and b) the radii of pipe bends were often inadequately specified. For the purpose of the
computation it is essential that each of the inputted changes in pipe section geometry coincide
with absolute geometric precision. It is anticipated that new cable routes will be surveyed using
GPS methods such that computer files of co-ordinates will be available. However it is still
anticipated that human judgment will be required to check the dimensions and correct the
inevitable anomalies. The dimensions of the cable offsets within the manhole and joint casing
would still require to be hand calculated.
An input sheet Table 5-5was designed together with the appropriate links to the pre-processor
code that would permit each straight and curved section of the pipe to be characterized by its
position in terms of [x, y, z] co-ordinates, together with a description of whether a section is
straight or curved. Table 5-5 is the [x,y,z] definition sheet of an example route fabricated from
two actual spans in a 138kV 1500kcmil pipe route. The three cables are in an 8 steel pipe (8
1/8 internal diameter). The first span in the model is identical to an existing pipe span. The
second span has had its angles of inclination reversed, such that the slope of the model is all
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-17
down hill either side of a central manhole. Each section of straight and curved pipe is numbered,
described and has a set of Cartesian co-ordinates. The dimensions are abstracted from the route
drawings in dimensions of feet and are then converted to meters. The x-z plane gives a plan view
of the route and the y-axis gives the vertical elevation. It will be seen that 61 sets of co-ordinates
have characterised the multiple changes in direction and elevation in the 3,772ft model, which
includes 9 pipe bends and 3 joint manholes. The straight sections are defined simply by entering
the end co-ordinates of the particular section. If the section is curved then the radius of the curve
also needs to be entered, this is omitted from the table for economy of space.

Table 5-5
Example route geometry sheet for 138kV pipe span
Section
Number
Section
Geometry
Section
Length
x-comp [ft]
Section
Vertical
y-comp [ft]
Running
Length
z-comp [ft]
1 - 0 0 0
2 straight 3.281 0 0
3 joint 16.114333 0 0
4 straight 48.922333 0 0
5 curved 93.729945 0 -40.843324
6 straight 97.655695 0 -83.161623
7 straight 105.853584 1.727 -171.53219
8 straight 110.472114 2.7 -221.318425
9 straight 119.709172 2.1 -320.890895
10 straight 128.946231 1 -420.463365
11 straight 138.18329 -2.9 -520.035835
12 straight 147.20349 -4.9 -619.608304
13 straight 156.657407 -6.9 -719.180774
14 straight 165.894466 -10.9 -818.753244
15 straight 168.276928 -11.8 -837.351265
16 straight 173.248967 -13.2 -869.468687
17 straight 180.611411 -14.9 -917.027178
18 straight 195.909994 -18 -1015.85002
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-18
Section
Number
Section
Geometry
Section
Length
x-comp [ft]
Section
Vertical
y-comp [ft]
Running
Length
z-comp [ft]
19 straight 211.208578 -20 -1114.67285
20 straight 226.507162 -23.8 -1213.49569
21 straight 241.805745 -27.1 -1312.31853
22 straight 257.104329 -29 -1411.14137
23 straight 272.402912 -32.9 -1509.96421
24 straight 287.701496 -35 -1608.78705
25 straight 301.470221 -37 -1697.7276
26 curved 303.384839 -37 -1715.06999
27 straight 307.029903 -37.9 -1769.94907
28 curved 324.309542 -37.9 -1838.99716
29 straight 347.971617 -37.9 -1891.40285
30 straight 356.901479 -41.05 -1911.1803
31 straight 358.259476 -41.05 -1914.18793
32 joint 363.540577 -41.05 -1925.88427
33 straight 364.897751 -41.05 -1928.89008
34 straight 398.519297 -42.5 -2003.35355
35 curved 420.028581 -41.3 -2040.01524
36 straight 449.026191 -39.9 -2079.20326
37 curved 478.849401 -40.7 -2111.40939
38 straight 519.476525 -41.9 -2146.60171
39 straight 596.19304 -42.8 -2210.74668
40 straight 672.909555 -44.2 -2274.89164
41 straight 751.387192 -44.9 -2336.86954
42 straight 790.626011 -45.8 -2367.85849
43 curved 807.100815 -46.4 -2382.40856
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-19
Section
Number
Section
Geometry
Section
Length
x-comp [ft]
Section
Vertical
y-comp [ft]
Running
Length
z-comp [ft]
44 straight 835.581857 -47 -2410.49468
45 curved 854.343278 -47.6 -2426.82387
46 straight 903.947777 -48.2 -2464.84549
47 straight 983.314975 -49.3 -2525.68008
48 straight 1062.68217 -51.7 -2586.51468
49 straight 1142.04937 -52.4 -2647.34927
50 straight 1221.41657 -53 -2708.18387
51 straight 1300.78377 -54.6 -2769.01846
52 straight 1329.55438 -54.7 -2791.071
53 curved 1344.14664 -54.7 -2801.25828
54 straight 1386.37706 -55 -2828.02751
55 curved 1401.24618 -55.2 -2838.42739
56 straight 1468.61794 -55.9 -2890.25446
57 straight 1635.76016 -58.6 -3010.66824
58 straight 1668.58003 -59.3 -3034.31251
59 straight 1671.2576 -59.3 -3036.24155
60 joint 1681.67018 -59.3 -3043.74306
61 straight 1684.33228 -59.3 -3045.66091

An example of the input sheet for the joint casing is shown in Table 5-6. As with the cable pipe
sections the joint longitudinal sections are defined by entering the section end co-ordinates.
Additional data is required to define the casing profile. This data is entered in the form of a series
of pairs representing lengths and diameters. The dimensions are finally converted from feet to
meters. The modeled profile of the joint casing is shown in Figure 5-5 and Figure 5-6.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-20

Figure 5-6
Example profile of a joint casing in a 138kV pipe system
The shape is from the Utilitys Drawing No 301680 of 24.01.95, with the exception of the ends,
which were angled for modeling reasons. It is now possible to model vertical ends.
Table 5-6
Example input sheet for the joint casing in a 138kV pipe system
Length (feet) Diameter (feet)
0 8.125
2 7 8.125
3 1 4
3 7 1 4
4 7 1 4
4 7 1 6
7 1 1 6
7 1 1 8
10 11 1 8
10 11 1 6
13 5 1 6
13 5 1 4
14 5 1 4
15 1 4
15 4 8.125
18 8.125
Solving the Thermomechanical Behavior of the Model Route
Inputting Information into the Solver
The route input sheets are submitted to the ABAQUS pre-processor software. A script was
written within ABAQUS to generate the nodes and elements for the cable in pipe
application. A second script calculates the radii of curvature at each change in pipe route, if
this had not already been pre-specified (for example if the radius is less than 10m). This
script is necessary because the geometric information is in the form of a series of short
straight pipe sections with angular changes in direction. A script was also written to show the
cable route in three dimensions, but mounted on a series of stilts, which drop down onto a
horizontal plane Figure 5-7 is the stilt diagram of two 138kV pipe spans totaling 763m in
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-21
length. The stilt diagram shows the operator the orientation of the route and permits the
vertical and horizontal bends and changes in elevation to be separated and identified. Without
this facility it is easy to lose orientation when the image of the route is being rotated on the
screen, particularly as routes are long and the angles of slope are small. This file is then
checked and corrected by the operator before being submitted to the solver software.
The mechanical characteristics of each of the three cable beams are entered into ABAQUS.
These will be different for each cable size. At the moment the characteristics are hand
calculated, being part derived from theoretical formulae and part from measured values. It is
anticipated that it will be possible to incorporate the calculation into software.
The required load cases are then specified. The solution method is dynamic and requires the
specification of a pressure wave, which travels at the speed of sound through the cable beam
elements, together with the specification of appropriate damping factors. This process
requires an experienced operator.
The intermediate time steps and the types of required output are specified.
The model is then submitted to the solver software and the run is commenced.

Figure 5-7
Three dimensional stilt diagram of two 138kV pipe spans totaling 763m
Checking the Success of the Run
The following points are checked:
The computation went to completion successfully
The warning messages, if any, are understood and are judged to be insignificant
The time steps were reasonable.
The kinetic energy in the solution is negligible compared with a) the elastic energy and b) the
frictional dissipation. (The solution is a valid quasi-static solution and thus must not exhibit
kinetic energy).
The history output charts of reaction forces against computation time steps, at each end of the
cable route, are sound.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-22
The output looks to be correct and has no anomalies, from the engineering view point (e.g.
the output charts of axial force, the bending radii the bending moments).
Engineering Analysis of the Model Output
The output is then viewed in detail to analyze the thermomechanical behavior of the cables and
the joints. Appropriate charts are printed using the ABAQUS Viewer software:
The cable deformation diagrams are analysed at different temperatures showing
thermomechanical patterns along the route in each of the cables, in terms of their position,
length, type (e.g. small cylindrical sine waves, large cylindrical sine waves/loops and helices)
and temperature of formation. (These will require to be plotted in plan and elevation view to
obtain the correct interpretation).
The whole route is searched for the unwanted presence and location of extreme pattern
formations (e.g. large waves, helices, short wave slopes and cables in potential jamming
formation i.e. with axes in straight alignment)).
History plots are analysed of axial force (section force) with time step during heating at
selected location along the route. To identify the values of temperature and force at the
significant events e.g. a) the temperature at which axial force starts to rise, b) the turn-over
temperature at which the force peaks, c) the force at the operating temperature of 90
o
C and d)
the force at the emergency operating temperature.
Check the slope of the force/temperature plot and compare with the maximum theoretical
value. Check the deformed length of the region of cable containing fully developed
thermomechanical patterns and compare with the maximum possible unconstrained
expansion length.
Compare the output with the limiting parameters for the particular cable, as calculated from
Chapter 4. For example in the cable span sections of the route, record the extreme values of
axial force, side wall force, the minimum bending radius, the strain at the surface of the
cable, the change in strain (fatigue strain).
Within the joint casings in pipe systems, record the maximum axial force at each end, the
differential axial force, longitudinal movement at each end, differential movement, minimum
bending radius of the adjacent cable, the absence of excessive local thermomechanical
deformation of the joints and adjacent cable (i.e. lateral deflection, twist and rotation). Ensure
that the joints and cable have not slid longitudinally and are being constrained by contact
with the casing or pipe. If one-piece joint moldings are being used in the role of semi-anchor
joints, check that the magnitude of the differential thrust does not exceed the cable pull-
through strength of the joint.
In the joint chamber in duct-manhole systems check for the absence of excessive values of a)
cable slip through clamping cleats, b) lateral deformation between cleats, c) strain at the
surface of the cable, d) fatigue strain of the cable in the offsets, e) conductor/core thrust at the
center of the straight/anchor joint.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-23
Charts and Videos
At the end of the solution the three dimensional image of the deformed cables along the
complete route is viewed on the screen in plan and elevation, to give a quick impression of
thermomechanical activity. The ABAQUS Viewer software is used. Each cable is seen as a line
which follows the axis of the deformed cable. The output from the solution is converted from
line contour plots of the type shown in Figure 5-11, to show the three dimensional geometry of
the cables, pipe, joints and casings using a software script called Cable View to give a three
dimensional visualization of the cables, pipes, joints and casing. This gives an image as close to
real life as possible. The output can be viewed as either discrete images, or as a video clip.
The amplitude of the lateral deflection of the cable wave patterns is determined by the cable to
pipe clearance, which is negligibly small in comparison to the length of a double span of pipe
route (0.123m compared to 2,000m). The presence of wave patterns can just be seen in the plan
view in Figure 5-8, however details and dimensions cannot be discerned.

Figure 5-8
Plan view of the 200m normalized reference span at 90
o
C with wave patterns adjacent to
the bend at B
Scale 1:1
The details of the cable deformation patterns are obtained from the following views:
Three dimensional rotation of the complete span, to give a foreshortened, but correctly scaled
view, as shown in Figure 5-9.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-24

Figure 5-9
Rotated perspective view of 200m normalized pipe span at 90
o
C
Scale 1:1
Increasing the scale factor in the lateral direction by x 10-20. This is an artificial view, but
permits the whole span to be seen and printed out in plan/elevation view, so that the regions
of deformation patterns can be located. Figure 5-10 is the plan view of a double span from a
138kV pipe system of 763m length, showing the position and growth of the wave patterns
from the starting condition at the 15C ambient, then at 30
o
C, 45
o
C, 60
o
C, 75
o
C and 90
o
C.

Figure 5-10
Thermomechanical pattern distributions in a double span 138kV 1500kcmil pipe system at
15
o
C, 30
o
C, 45
o
C, 60
o
C, 75
o
C and 90
o
C
Plan view, with 10:1 lateral distortion
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-25
Increasing the scale factor in the lateral direction and three dimensionally rotating the model.
Figure 5-11 is a line deformation plot of a 200m span model that is being used to study the
effect of traffic vibration on a cable that has already developed thermomechanical wave
patterns. The vibrations can be seen to have a) increased the length of the pattern region from
47.3m to 70m and b) increased the number of helical waves from 2 to 3. Details of the traffic
vibration study are given in Chapter 9.

Figure 5-11
Deformation line plot with 200m duct model rotated and lateral scale increase (before and
during journey of vibrating truck over duct)
Projection view, with 10:1 lateral distortion
Zooming in to one part of the span. Figure 5-12 shows the first 25m length of the 200m pipe
span adjacent to the bend in correctly 1:1 scaled plan view. Individual wavelengths can be
seen and measured.

Figure 5-12
Close-up view of the end 25m of the 200m pipe span next to bend B at 90
o
C, to observe
individual wave patterns
Plan view, scale 1:1
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-26
Zooming in and rotating the model. Figure 5-13 is a view of the same 25m region as shown
in Figure 5-12, but at a higher temperature. The model has been rotated to give a
foreshortened perspective view of the waves as they change formation to enter the bend in
vertical alignment. The view is still a true 1:1 scale.

Figure 5-13
Rotated close-up view of the end 25m of the 200m pipe span at 125
o
C to show change in
wave shape at bend entrance B
Projection view, 1:1 scale
Cutting out one part of the route and zooming-in, for example to look up the inside of the
duct. Figure 5-14 is the view looking up the complete 200m length of the normalized pipe
span in Figure 5-8 from A to B for three ducts in triangular formation. The cables have been
heated to 90
o
C. The first part of the cable at end A is straight. The region close to the bend
has wave patterns. The cables in the bend from B to C are straight. A 50kA, 60Hz 0.1s short
circuit has been applied. The cables have been repelled away from each other and are held
against the walls of the duct, permitting a clear view from end A to end B. Details of the
short circuit study are given in Chapter 10.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-27

Figure 5-14
End view along ducts in trefoil at 90
o
C during 50kA short circuit
End view, scale 1:1
The deformed cable is then viewed dynamically to give a video effect comprised of the
solutions at each of the computational time steps. This shows when, where and how the cable
deformation patterns start. The video effect can be run forward and backwards at different
speeds.
The deformed cable view can be probed to find the node numbers and hence the exact location
of thermomechanical activity.
The forces are visualized by field plots, which superimpose colour onto the deformed cable to
illustrate the distribution and magnitude of force along the route.
Charts are produced in two forms a) a history plot of the change in force with temperature at
one specific location, as shown in Figure 5-15 and b) a path plot which shows the variation of
force (y axis) along the route distance (x axis) at one computational time step/temperature as
shown in Figure 5-16.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-28

Figure 5-15
History plot of axial force with temperature in the three cables in the 200m normalized
route at the end of the straight span at A and adjacent to the bend at B
Figure 5-15 gives the force-temperature characteristics for the three 138kV cables in the 200m
normalized span pipe model. The characteristics are shown at end A at the start of the
normalized span and at B, the entrance to the bend. If required, the characteristics can be studied
at each of the element positions, these being spaced at 0.3m intervals. Small output temperature
steps of 2.5
o
C have been chosen. The software always solves the problem in very small
temperature steps, however a sensible compromise is required in the specification of output
temperature steps as this will very significantly increase the disk storage space required for long
pipe routes.

Figure 5-16
Path plot of axial conductor force along a double span of a pipe system at 90
o
C
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-29
Figure 5-16 is at the specified temperature of 90
o
C and is the distribution of axial force along
each of the three conductors in a 1,187m, 138kV double span pipe system. The force is
calculated at each of the element positions, which in this model are 0.5m long in the main spans
and shorter in the manhole. The variations in force are associated with the positions of bends and
regions containing wave patterns. The manhole position is revealed to be at 650m by a vertical
discontinuity in the curve, this being the force differential across the joints.
The force-temperature charts can be overlaid on top of each other to produce families of curves
giving the development of force with temperature at each position along the route.
In addition, the axial force distributions can also be plotted of bending radius, sidewall force,
bending moment, strain and axial movement. Different variables can be plotted against each
other, for example the variation of axial force and axial movement.
The use of colour is essential in distinguishing different superimposed plots, for example to show
a) the intertwining deformation patterns of three cables in one pipe and b) the development of
force with temperature at one location. The majority of diagrams given in this report are in
colour. The reader is advised to obtain a copy of the report on CD to get the full benefit from the
diagrams.
Computing Requirements
The pre-processing, solving and post processing can be undertaken on one work station
computer, however two computers were used in parallel at various stages of the project, this was
to double the number of models that could be analysed. Each computer contained 3GHz twin
Xeon 32 bit Intel processors, these having large cache memories. The machines had 3GBytes of
onboard RAM, however RAM was not a limitation as the explicit finite element method is
memory efficient. The requirement for hard disk space dedicated to the individual model
solutions was generally less than 5GByte, however when applying a thermomechanical loading
case of greater than 20 load cycles this was exceeded. Analysis of the post-processed output data
was undertaken using the workstations and standard PCs, although even with fast processors
these were distinctly slower in processing data.
Description of the Finite Element Analysis Modeling Methods:
ABAQUS is a proprietary, finite element analysis software system that contains sophisticated
features that can be adapted to model pipe and duct cable systems. A cable subjected to thermal
expansion in a pipe is a complex system that requires the resolution of the 3-dimensional
deformation of a very long, slender, beam-like structure following a devious route, with a non-
trivial physical behavior. The deformation is influenced by intermittent and arbitrary contacts
between the cable surface and the pipe wall, including frictional effects. If three cables are
installed in the same pipe then the behavior is complicated still further by inter-cable contact.
ABAQUS is a non-linear analysis software code that includes efficient routines for solving
complex geometric problems with a facility to model contact between surfaces. It includes two
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-30
methods for solving the equations which govern the behavior of the cable; namely the implicit
and explicit methods of analysis. They are very different in nature with strengths and
limitations depending on the type of problem and the structure that is to be analyzed. Implicit
methods can be applied to solve certain static and dynamic problems, whereas explicit methods
are primarily applied to a range of dynamic problems. Explicit methods can also be used to
analyze quasi-static events as they are a special case dynamic problem. This enables key benefits
of the explicit method to be applied to static problems.
The key advantage of the explicit method to the modeling of cables in a pipe is the ability to
handle complex contact phenomena with great efficiency.
The implicit and explicit methods are described below.
Modeling: Implicit Analysis Method
The implicit finite element method is the standard technique used to solve complex static and
some simple quasi-static structural problems. The implicit method of analysis is provided by
the ABAQUS/Standard program. In the implicit method, the structural model is discretised i.e.
divided into discrete finite elements connected by nodes that describe the geometric form of the
structure. For a model to be in static equilibrium, the net force acting on each node in the
structure must be zero. Based on this fact, a system of equations is formed which relate the
applied external forces to the displacements at each node. The nodal displacements and external
forces are related by the element stiffness, with potentially a different stiffness from each
element in the model. The individual element stiffnesses are assembled into a global stiffness
matrix and this set of equations is then solved simultaneously to find the displacements at each
node. Element strains can then be calculated and, through the material constitutive law, the stress
in each element can be found.
A key point is that to solve for nodal displacements, the stiffness matrix has to be inverted.
Matrix inversion is a computationally intensive operation and for large models can take long
processor times. For simple linear elastic structures, not involving contact or large deformations,
this operation need only be done once in order to obtain a solution. ABAQUS/Standard also
incorporates fast solvers to reduce the time taken to process the matrices. For structures that
involve contact with geometric and material non-linearities, matrix inversion has to take place
every time there is a change in the global stiffness matrix. ABAQUS/Standard has very efficient
iterative solution schemes for resolving contact and non-linearity in general, but a complex
analysis will involve much iteration before a stable solution is achieved. In particular,
intermittent and arbitrary contact between bodies can cause an excessive number of iterations
that may result in inefficiency with no final solution being obtained.
For static analysis, the mass of the structure is unimportant as inertia forces are ignored and the
solution depends entirely on the stiffness of the structure.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-31
Modeling: Explicit Analysis Method
The explicit method of finite element analysis is used to solve dynamic problems, some of
which can occur at very high speed. The explicit method of analysis is provided by the
ABAQUS/Explicit program. As in the implicit method, the structure is discretised into nodes
and elements, but the structural response is solved by the method of explicit dynamics.
Explicit dynamics is a mathematical technique for integrating the equations of motion through
time. ABAQUS/Explicit uses a central difference rule to integrate the equations of motion
through time, using the kinematic conditions at one increment to calculate the kinematic
conditions at the next increment. At the beginning of the increment the program solves for
dynamic equilibrium, which states that the nodal mass matrix, multiplied by the nodal
accelerations, are equal to the total nodal forces (the difference between the external applied
forces and the internal element forces).
I P a . M =
where:
M is the nodal mass matrix [kg]
a is the nodal acceleration matrix [m.s
-2
]
P is the external applied forces matrix [N]
I is the element internal forces matrix [N]
The accelerations at the beginning of the current time increment (time t) are calculated as:
( )
t
1
t
I P M a =


Equation 5-1
Since the explicit procedure uses a diagonal, or lumped mass matrix, solving for the
accelerations is trivial; as there are no simultaneous equations to solve. The acceleration of any
node is determined completely by its mass and the net forces acting upon it.
The accelerations are integrated through time using the central difference rule, which calculates
the change in velocity, assuming that the acceleration is constant. This change in velocity is
added to the velocity from the middle of the previous increment to determine the velocities at the
middle of the current increment.
The velocities are integrated through time and added to the displacements at the beginning of the
increment to determine the displacements at the end of the increment. Hence, satisfying dynamic
equilibrium at the beginning of the increment provides the accelerations and, knowing the
accelerations, the velocities and displacements are advanced explicitly through time. As in the
implicit method, the displacements are used to calculate elemental strains which in turn are
used with material constitutive relationships to calculate stresses and, consequently, internal
forces.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-32
For the method to produce accurate results, the time increments must be quite small so that the
accelerations are nearly constant during an increment. Typically the time increment used must be
less than the time taken for a stress wave to traverse an element. This results in incremental time
steps ranging from micro to milliseconds with a typical analysis taking many thousands of
increments. However, the operations described above can be carried out rapidly and
ABAQUS/Explicit is designed to complete these operations as efficiently as possible.
The selection and adaptation of the explicit finite element method to solve the cables in pipe
problem was the major technological breakthrough that made the advances in knowledge
described in this report possible.
Validation of the Explicit Finite Element Analysis Modelling Technique
The dynamic explicit FEA method selected as the basis for the thermomechanical project was
validated by comparison with the industry standard implicit method used to solve static
structural problems. At the start of the project a cable in pipe model was prepared and submitted
to the ABAQUS/Explicit and to the ABAQUS/Standard proprietary programs. There is much
experience and confidence in the use ABAQUS/Standard technique, hence its use in this
application as the reference standard. The application of the ABAQUS/Standard technique to
cable in duct problems with comparatively simple geometry is recorded in 1994
[i] [ii]
.
The model selected was a short section of straight duct containing one cable as this was simple
enough to be solved by the standard structural implicit code. The cable selected for the
comparison was a two beam approximation of the 138kV 1500kcmil cable, however only one
cable was inserted in the 206mm internal diameter where normally there would be three. The
analysis consisted of two steps. The dimensions of the model were:
Duct:
Length: 30.0 m (98.4ft)
Diameter: 0.206 m (8.125)
Cable:
Length: 30.0 m (98.4ft)
Conductor diameter: 0.0335 m (1.32)
Insulation diameter: 0.083 m (3.27)
At the beginning of the analysis the cable was placed at the centre of the duct. The first
computational step allowed the cable to drop to the bottom of the duct under gravity. A small
perturbation was formed in the centre of the cable length by displacing three nodes laterally. The
perturbation was necessary to prevent the cable behaving as a geometrically perfect cylindrical
beam, thereby remaining perfectly straight throughout the analysis.
At the beginning of the second computational step, the ends of the cable were fixed and the cable
was then subjected to a temperature increase. For purposes of validation the cable was heated
through the exaggerated temperature range from 15C to 180C, this being large enough to
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-33
ensure that the model cable would generate thermomechanical deformation patterns comprising
both pronounced sinusoidal shapes and helical shapes.
The cable material properties were indicative only and were selected before values of cable
components became available from either the literature survey and from the tests program on
138kV and 230kV cables. The indicative values are recorded in the Interim Report for the project
[iv]
.
Validation Results
Description of thermomechanical phenomena
Figure 5-17 is a 3D visualization of the cable deformation pattern at the end of the load step at
180
o
C, it was viewed in conjunction with a video animation of the formation of the cable pattern.

Figure 5-17
Cable thermomechanical patterns in the 30m duct validation model
Figure 5-18 shows the reaction forces experienced by each end of the cable at the points of fixing
at the boundaries. The sign convention is that forces acting in the x direction, from left to right,
are positive, thus the forces on the left hand of the cable are positive and on the right hand end
are negative. To obey the equations of motion, the forces have to be of equal magnitude and act
in opposite directions.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-34
It was shown that from 15C to 23C that the cable appeared to behave as a straight rod
developing high compressive loads up to 4,500N, however small deformation patterns were
already growing adjacent to the central perturbation. At 23C the small deformation patterns
developed into lateral sine waves. The thermal expansion strain within the rigid cable was
released into the sine waves and the force fell slightly to less than 4,000N. As the temperature
continues to rise the sine waves absorb much of the thermal expansion strain and grow to the
ends of the pipe, thereby resulting in a much reduced rate of increases of force with temperature.

Figure 5-18
Reaction forces at each end of the 30m length in the validation model
The length of the cable within the patterns is shown in Figure 5-19. At 90C the cable sine
patterns have absorbed 40mm of free cable length. Figure 5-18 shows that at 105C the end
thrust developed is 5,000N. Helical loops start to form at each end of the cable, in opposite
helical rotation to each other. The helices are able to absorb the cable expansion more efficiently
and this reduces the cable end thrust, falling from 5,000N to a minimum of 3,500N at 128C.
Thereafter the cable patterns exhibit higher stiffness and are not able to absorb thermal strain as
easily, so force rises at an increased rate. The solution was stopped at an arbitrary temperature of
180C with the cable having developed 6,000N thrust and the pipe having absorbed 90mm of
cable expansion.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-35

Figure 5-19
Length of cable inside the 30m duct with temperature for the validation model
The wavelength of the horizontally snaked sine waves was 4.9m (16ft) and the length of the
helical loops at each end was 3.7m (12ft). To form the horizontal sine waves, the cable slid
uniformly up the walls of the pipe against the force of gravity until it became possible for loops
to pass right over the top of the pipe into a helical loop. As this is a uniform process without
sudden buckling, the cable length graphs are nearly linear with temperature. However, the force
graph is almost a constant value, which shows that the cable patterns are efficiently absorbing the
expansion.
Comparison of explicit and implicit solutions
Figure 5-20 shows correctly scaled orthogonal views of the patterns exhibited by the centre line
axis of the cable heated to 180C within the 30m pipe, to compare the geometry of the explicit
and implicit cable deformation patterns. The orthogonal view is essential to be able to view the
patterns. (The orthogonal views are shown without the perspective of distance being applied).
The cable patterns seen in plan view over the 30m length would be too small to print on one
sheet of paper, (the ratio of the pipe internal diameter to length is 1:150). The geometric
comparison confirmed that the explicit methods produced closely identical cable deformation
patterns.
A comparison of the distribution of bending radii along the 30m length was made, Figure 5-21. It
was confirmed that the implicit method gave a closely similar distribution. This also confirms
that the bending moments (that generate the patterns) are the same, because the reciprocal of
radius is directly proportional to the bending moment.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-36

Figure 5-20
Comparison of 'explicit' and 'implict' cable deformation patterns for the validation model

Figure 5-21
Comparison of cable bending radii in the explicit and explicit FEA models
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-37
The explicit and implicit solutions were also compared by printing out a number of load cases
and these confirmed close coincidence, for example the longitudinal force distributions agreed
within 2.5% Figure 5-22.

Figure 5-22
Comparison of axial force distribution in 'explicit' and implicit FEA validation model
Comparison with In-Service Results
Feedback from in-service XLPE cable systems is limited, as the first 230kV duct-manhole and
138kV pipe systems containing joints have only recently been installed.
No published experimental information was found on the performance of test rigs containing
significant lengths of XLPE cables in either pipes or ducts. Some recent published information
was available
[i] [ii]
on XLPE cable. Older information was available from experiments on HPFF
pipe type paper cable and on self-contained paper cable. This information showed that cables do
form sinusoidal and helical patterns similar to those produced by the model.
The information from service experience to date is that joints within a 138kV XLPE pipe system
were reported to have deflected laterally within a 1-2 year period in service. The adjacent cables
within the casing had deflected laterally and had rotated axially.
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-38
This information was extremely useful in checking the cable mechanical parameters used in the
model. The prototype pipe system models studied at the start of the project were based on a)
assumed cable parameters and b) a coefficient of friction of 0.1, a typical value measured on a
well lubricated cable during installation. The cable parameters produced deflection , but not as
much as recorded photographically. Following the measurement of the 138kV 1500kcmil cable
parameters at EPRISolutions, the model was modified to include the measured values of axial
stiffness EA, bending stiffness EI and the coefficient of thermal expansion , together with a
value of coefficient of friction of 0.3
[v]
, this being more representative of the service condition
after the lubricant had dried and dispersed. The degree of deflection produced by the model
joints was closely similar to that reported Figure 5-23.
At the end of the project three double span manhole positions from the 138kV service route were
modeled, each containing one manhole position. The three joints in each of the manholes were
found to start deflecting at a temperature rise of less than 15
o
C above ambient temperature.

Figure 5-23
Deflection of 138kV joints, top is at 90
o
C, bottom is at 15
o
C at end of 3
rd
cycle
Perspective view, scale 1:1
EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-39

Figure 5-24
Deflection of 138kV joints, top is at 90
o
C, bottom is at 15
o
C at end of 3
rd
cycle
Elevation view, scale 1:1
The shape of the deflection patterns as shown in Figure 5-23 (top), was also found to be closely
similar to that a) seen by the Principal Investigator during the opening of one of the manhole
positions in the service installation and b) in photographs of another manhole position. Figure
5-23 is orientated to give the same perspective view as photographically recorded. Figure 5-24
shows the same deflection at 90
o
C and back at 15
o
C in elevation view.
One difference was that the joints in the service installation exhibited significant deflection at
ambient temperature and thus had not straightened themselves. The model had predicted that the
joints should partially straighten, as shown in the bottom halves of Figure 5-23 and Figure 5-24.
The difference is attributed to a) the number of modeled load cycles being less than 5 and b) the
cable stiffnesses being modeled as elastic, whereas the measured cable parameters revealed the
presence of a significant non-elastic component. During the project, successful advances were
made in the modelling technique to represent a) complex geometries within long span lengths, b)
sliding contact between the three cables and pipe, c) temperature dependent properties and d)
non-linear elastic properties. However the representation of inelastic properties was not possible.
Technological advances in the software have since shown that a) the additional representation of
inelasticity is now possible and b) the number of load cycles can be significantly increased. It is
recommended that these be included in future modelling work.
Overall it was concluded that the modelling technique had been satisfactorily validated by
calculation and by the close similarity with the thermomechanical deflections seen in the service
pipe installation.

EPRI Licensed Material

Cable and Route Modelling: Methods and Analysis
5-40


i
Tarnowski.J, Iordanescu.M, Awad.R, Royer.C, Thermomechanical modelling of 345kV XLPE cables in duct,
paper A4.1, Jicable 99, Versailles.
ii
Tarnowski J, Iordanescu M, Hardy C, Chaaban M, Thermomechanical modelling of extruded insulation HV cables
in ducts, paper 21-101, CIGRE Conf, 1994, Paris.
iii
Research to determine the acceptable emergency operating temperatures for extruded dielectric cables, EPRI Palo
Alto CA, Report EL-938, Project 933-1,November 1978.
iv
Mechanical effects on extruded dielectric cables and joints installed in underground transmission systems in North
America, Technical Update Report, August 2002, EPRI, Palo Alto, CA:2002: 1001848
v
Underground transmission systems reference book,1992 Edition, EPRI, Palo Alto:CA: 1992,TR-101670, project
7909-01.
EPRI Licensed Material
6-1
6
THERMOMECHANICAL PERFORMANCE OF CABLES
IN A STRAIGHT ROUTE
Introduction
This chapter studies a straight route and explains, using model output, how temperature rise
causes thermomechanical deformation patterns to grow and cable axial forces to increase.
The initial studies of real cable routes in this project
[i]
showed that thermomechanical
deformation patterns occurred next to bends in the route and next to joints, with sections of
straight cable in between. The investigation then divided into two parts, one as described in this
chapter, to study whether patterns could form in straight pipes and the second as dealt with in
Chapter 7, to study how the geometry of bends initiates patterns. The understanding of initiating
causes in straight pipes is important because this situation was not included in the models of real
routes, thus it needed to be known whether forces and patterns would be increased or decreased.
It was found that the initiating cause in both cases was the same, ie the presence of curvature in
the cable or pipe, no matter how mild. The study of straight routes and ones containing bends as
described in this project report is intended to give a complete understanding of
thermomechanical phenomena.
In the simplest route the pipe or duct is everywhere straight, containing a) no variations in the
pipe inner diameter, b) no bends and c) no slopes. In practice this perfect state cannot exist as a)
the pipe will be of increased diameter at each accessory, b) in the case of a duct installation the
cable will experience bends and a different method of constraint in the manhole, where it
emerges from the duct, c) the pipe will exhibit perceptible changes in alignment at pipe joins and
d) the cable as supplied will not be perfectly straight.
XLPE cables are designed to be flexible such that they can be wound onto reels and pulled into
pipes and ducts without damage. The cable is not perfectly straight. Residual curvature is formed
by a) the factory take-up reel at the end of the extrusion process when the cable is still warm, b)
the factory shipping reel during the installation haul-off process, in which the diameter and
change of lay direction will impress a curvature to the cable and c) sidewall forces during
installation, for example when guiding the cable into the pipe or pulling it around a bend. In
Chapter 3 it was described that significant difficulty was experienced in testing both the 138kV
and 230kV XLPE cables because of their preformed curvature and high stiffness. The Test
House attempted to straighten the XLPE cable samples, but as cable jointers experience, the best
that can be achieved by the use of heat and bending is a partial straightening. The elastic memory
of the cable typically reasserts the original curvature within 24 hours.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-2
Three cases of straight route models are considered:
Theoretically Perfect Straight Cable
The cable and pipe are perfectly straight with each end of the cable anchored to prevent
longitudinal expansion. Upon heating, cable will generate an internal thermomechanical
compressive force and, because the theoretical cable is perfectly straight, it will behave as a
perfect strut under column load. It will neither buckle into sideways deflection, nor form
thermomechanical patterns within the pipe. The computer model of the route will behave in
exactly the same theoretical way because the cable is modeled with perfect axial symmetry along
the perfectly straight pipe route. The force generated is given from Chapter 2 as:
EA F = Equation 6-1
Where:
F: compressive axial force [N]
EA: effective axial stiffness of the cable in compression [N]
E: effective modulus of elasticity in compression [Nm
-2
]
A: effective cross sectional area of the cable [m
2
]
: effective coefficient of thermal expansion of the cable [
o
C
-1
]
: temperature rise of the cable (referred to conductor) [
o
C]
This is the simplest model case in which the magnitude of force is directly proportional to the
temperature rise, with the slope of the F/ graph being constant at EA for all temperatures.
It is seen that force is independent of route length L. The cable in a short route will generate the
same force at the anchors (joints or terminations) as in a longer route. However the locked-in
thermal strain is proportional to the length L. Thus if one of the anchors is allowed to move
lengthwise, the axial force may fall to near zero in a short span, but will be maintained at a
higher value in a long span.
Equation 6-1 is the maximum axial force that can be generated in any horizontal route with or
without bends. This force F is a useful absolute point of reference with which to judge the force
reducing capabilities of routes containing bends, different diameter pipes and different cable
characteristics.
An example of the calculation of the maximum force is given in Table 6-1 based on parameters
similar to those in the cable model (which have more complex temperature dependent
properties). The values of EA were derived from the compression tests on cable samples
described in Chapter 3. The maximum theoretical force at 90
o
C is 24,540N.
(Note: Subsequent analysis of the results from the constrained thermal test in Chapter 3 and a
comparison with published results for large conductors has indicated that a more realistic
average value for the 20-90
o
C range is 47.3MN. This is 2.5 x the average value of 19.3MN
given in Table 6-1. The maximum theoretical force is increased from 24,540N to 60,300N. The
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-3
effect of increasing the value of EA on cable thermomechanical performance is given in the
sensitivity study in Chapter 7.)
Table 6-1
Example calculation of maximum theoretical cable thrust
138kV 1500kcm XLPE Cable Parameter
20
o
C 90
o
C
EA [MN] 18.2 20.3
[
o
C
-1
] 17 x 10
-6
17 x 10
-6

[
o
C] 75 75
Slope of F/ [N.
o
C
-1
] 309.4 345
Average Slope of F/ [N.
o
C
-1
] 327
F
max
[N] 23,205 25,875
F
max
[N] 24,540
Straight Route Containing a Single Kink
The model pipe route was selected to have the same straight length of 200m as contained in the
hockey-stick shaped model used in the Chapter 7 sensitivity study. 200m is less than a typical
span length of 500-800m, but is longer than straight sections of pipe found in real spans. The
model contained one 138kV 1500kcm cable. The outer diameter of the cable was 83mm and the
diametral clearance to the pipe was 150mm, hence the internal diameter of the pipe was 233 mm.
The cable and pipe dimensions are those of the reference case in the sensitivity study. As
described in Chapter 7 the models are normalized on the dimension of the clearance between the
cable and pipe. The dimensions of the model are similar to those of a real 138kV 1500kcm cable
installation inside an 8.125ins pipe with a diametral clearance of 123mm (4.85ins).
Two models were studied using the same pipe diameter, one containing a single 138kVcable (a
duct system) and the second containing three 138kV cables (a pipe system).
Single Cable Model
The model cable was given a gentle preformed inelastic lateral deformation of 20mm over a
length of 600mm, this being equivalent to an angular deviation in the cable axis of 3.8
o
. The
scientific name for the deformation is a perturbation, but for brevity the name kink was
selected.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-4
Single Kink- Deformation Patterns
The small kink initiated the formation of long thermomechanical deformation patterns. A 3D
visualization of the pattern at 50C is shown in Figure 6-1. The model was heated from a 15C
ambient temperature through a rise of 35 C. The kink had initiated a 33m long pattern
containing two helical loops and a number of lateral sine waves. Figure 6-1 is a correctly scaled
1:1 view looking down and along the cable in plan view. This has the effect of naturally
foreshortening the length and emphasizing the deflection patterns. The purpose was to fit the
pattern into the page width. The kink is still present and its position is marked in the figure, but is
too small to be seen clearly in this particular view.

Figure 6-1
Thermomechanical pattern at 50
o
C initiated by one kink.
Plan perspective view, 1:1 scale
Figure 6-2 is a plot of the central axis of the cable showing the nodes between the 300mm long
elements. The kink at the starting temperature of 15
o
C was formed by taking two 300mm
elements and displacing the center node transversely by 20mm. The shape was formed before the
cable was dropped 1mm under gravity onto the pipe floor. This is the neutral shape of the model
cable, free of elastic strain. The nodes of the kink form an isosceles triangle, shown in Figure
6-2, with a base length of 600mm, height of 20mm and angle to the duct axis of 3.8
o
. There is
already 1.3mm of additional cable length in the shape of the deformation.
In Figure 6-2 at 35
o
C (20
o
C rise), the thermal deformation has extended from 600mm to
1200mm.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-5

Figure 6-2
Initial deformation at 15 C and at 35 C-single kink case
Plan view, 1:1 scale
The scale in Figure 6-3 has been reduced to increases the length of cable in view from 2.1m to
20m. The 300mm node markers have been left on. To permit the deformation to be seen the
lateral scaling factor has been increased to 10:1.
At 42.5
o
C the overall length of the pattern has grown to 13.8m. The wavelength is 4.2m. The
peak to peak amplitude in plan projection (not circumferential projection) is 93.2mm and in
elevation projection is 30mm. The axial force peaks at this temperature. A sine wave pattern of
larger amplitude is about to be formed. The middle half wave containing the original kink has
started to lift off the bottom of the pipe under the influence of the high thermomechanical thrust
and appears to have contacted the opposite pipe wall. The two adjacent waves are starting to lift
to form two circular loops. Loop formation only occurs in this single kink case and is attributed
to the axial thrust being higher than in the cases studied containing multiple kinks.

Figure 6-3
Thermomechanical deformation at 42.5
o
C and 45
o
C-single kink case
Plan view, 10:1 scale
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-6
At 45
o
C, sine waves and two helical loops have formed and the axial force is falling. The pattern
is now absorbing thermal strain from the adjacent cable at a faster rate than the temperature rise
is generating it. In Chapter 2 on Basic Theory, it was shown that the cable absorption in a sine
wave increases as a power law function of the ratio of amplitude to wavelength. The length of
the pattern has grown to 21m with 4.4-4.8m wavelengths. The peak to peak amplitude in plan
projection is now 137mm and the elevation projection is 47.7mm. The maximum circle that the
axis of the cable can describe is the difference of the cable and pipe diameters, this being
150mm. This indicates that the axis of the wave has lifted up the pipe wall by 47.7mm compared
to the radius of 75mm. Thus the cable is rising towards the 3
o
/c -9
o
/c axis. The lifting of the two
adjacent waves off the floor of the pipe continues.
Table 6-2
Dimensions of the incipient thermomechanical pattern
Temperature
[
o
C]
Pattern
Type
Half
Waves
Length of
Pattern [m]
Peak to Peak
Amplitude[mm]
Height
[mm]
Min bend
radius [m]
15 kink 1 0.6 - - -
20 oscillating
decrement
3 2.4 19.5 - 450
25 3 2.4 20.8 - 74
30 3 3.6 22.8 3.7 32
35 3 4.8 27.1 4.8 18
40 5 9.6 47.8 12 -
42.5
(force peaks)
7 13.8 93.2 30 6.1
45 oscillating
sine wave
9 21 137 47.7 4.9
Table 6-2 records the dimensions of the patterns that were initiated by the kink. When the cable
was dropped under gravity at the starting temperature of 15
o
C the kink relaxed and had spread
further than the preset 600mm length. The temperature rise to 20
o
C produced a train of small
waves that oscillated either side of the cable central axis and decreased in magnitude. The length
recorded in the table was measured over the 3 half waves that were visible at the selected lateral
scaling factor. The peak to peak amplitude is a measure of the lateral displacement from the plan
view. It should be noted that the amplitude is the chord length of the circle of diameter (Dp-Dc),
where Dp and Dc are the diameters of the pipe and cable. As the temperature rises to 42.5
o
C,
increases occur in a) the number of half waves, b) pattern length, c) amplitude (width) and d)
height. The minimum bending radius reduces, this being in inverse proportion to the magnitude
of the bending moment, Equation 6-2. The bending radius at the transition between the
oscillating decrement pattern and the sine wave patterns is 6.1m. The original kink position
remained at the peak of the central half wave. At 45
o
C the pattern changed to two equal
amplitude sine waves still symmetrically positioned either side of the kink.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-7
M
EI
R = Equation 6-2
Where:
R: radius of curvature in cable [m]
M: bending moment [Nm]
EI: effective axial stiffness of the cable in compression [N]

Figure 6-4
Cable thermomechanical patterns at 45
o
C, 50
o
C and 55
o
C-single kink
Plan view, 10:1 scale
The width of the pattern in Figure 6-4 has increased from 20m to 54m showing that the pattern
length has doubled between 45
o
C to 55
o
C. At first sight the patterns appears to be a train of sine
waves, but at 50
o
C and 55
o
C the two waves either side of the centre can be seen to have lower
amplitudes. These are not sine waves, but circular loops that are expanding radially outwards and
have not yet completely contacted the pipe wall. These partly formed loops can be seen in the 3D
visualization in Figure 6-1. The axial thrust is now falling at a lower rate, reaching a minimum at
65
o
C.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-8

Figure 6-5
Growth of thermomechanical patterns between 15
o
C and 125
o
C-single kink
Plan view, 10:1 scale
The 200m route length shown in Figure 6-5. Although the first deformation can be seen at 35
o
C
in Figure 6-2, it is too small to be seen in Figure 6-5 even at the 10:1 transverse scaling factor. At
the XLPE cable operating temperature of 90
o
C the patterns extend to 72% of the cable length.
The patterns reach the end of the cable at 112.5
o
C.
Figure 6-6 shows the cable climbing around the circumference of the pipe. The view is the locus
of the cable axis looking along the pipe from the kink at the centre of the route. The largest
diameter that the locus can form is 150mm. (The dotted circle was added to illustrate the pipe
wall position and is not to scale, the internal diameter of the model pipe is 233mm).
At 42.5
o
C the cable has reached its peak thrust, one wave is sliding up the pipe wall and
another is starting to lift vertically.
At 45
o
C the half wave containing the kink has contacted the opposite pipe wall.
At 47.5
o
C two loops can be seen to be lifting.
At 50
o
C one loop has almost formed a circle, this is shown in the 3D visualization in Figure
6-1.
At 55
o
C the loop has formed a complete circle with the jacket of the cable contacting the
pipe wall. At the bottom of the pipe the sine waves can be seen end on. The higher amplitude
waves are not laying perfectly on the bottom of the pipe.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-9

Figure 6-6
Lengthwise view inside the pipe from the 100m central position
Single Kink Force Distribution Along the Route
The force distribution along the 200m route has four distinct phases A, B, C and D. In Chapter 2,
Basic Theory, it was explained that the coefficient of friction, in combination with the cable
weight prevented all of the locked-in thermal expansion strain from the 200m cable length being
fed into the deformation pattern. A given length of cable can withstand a certain difference in
force F along the length of the route, Equation 6-3. As the temperature continues to rise, the
force in the straight rigid cable at the end of the route attempts to increases in a linear way
according to Equation 6-1. The frictional constraint force F is continuously being exceeded.
So thermal strain is being fed into the deformation pattern permitting it to grow in length. At the
same time the front of reducing force travels along the constrained straight cable, releasing a
proportion of the locked-in thermal strain until it reaches the cable ends.
LWg F = Equation 6-3
Where:
F: force difference supported by frictional constraint [N]
L: length of cable supporting F [m]
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-10
W: unit mass of cable [kg m
-1
]
g: acceleration due to gravity, 9.81 [m s
-2
]
: coefficient of friction
Phase A, 15
o
C-42.5
o
C
Figure 6-7 plots the axial force along the 200m span. The sign convention is that compressive
forces are negative, thus increasing compressive force is plotted in the downward direction.
The force increases in a near linear way with temperature to reach a maximum at 42.5
o
C.
The force is higher at the ends of the cable as these are insulated by frictional constraint.
The force fronts spread out and reduce the magnitude of force over 75% of the 200m route
length.
A small reduction in force can be seen to have occurred at the kink at 20
o
C, showing that the
kink is already growing and locally absorbing some cable expansion. At 35
o
C, the force
front has reduced the force 10m away, even though Figure 6-2 shows that the deformation
pattern is still small and has only extended its overall length from 600 to 1200mm. At 40
o
C
(not shown on the graph) the force front has reduced the force 40m away and this length is
sustaining a force differential of 460N, this being in agreement with the value calculated by
the static frictional constraint of the cable with a coefficient of friction of 0.1. Agreement is
also obtained at 45
o
C in which a cable length of 83m sustains a force difference of 990N. At
42.5
o
C the force is clipped to its maximum value by the transition from a small amplitude
deformation to a longer pattern of larger amplitude sine waves.

Figure 6-7
Force distribution along route, Phase A, 20
o
C-42.5
o
C
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-11
The lengths of the thermomechanical patterns and the lengths of cable experiencing reduced
axial force either side of it are given in Table 6-3.
Table 6-3
Length of route experiencing reduced force-single kink
Temperature
[
o
C]
Length of Route
Experiencing Drop in
Force [m]
Length of
Pattern [m]
Length of Force
Wave Front [m]
15 - 3.6 -
20 2.7 2.4 0.15
25 21.9 2.4 9.8
30 27.6 3.6 12
35 42.9 4.8 19.1
40 85.9 9.6 38.2
42.5 152 13.8 69.1
Phase B, 42.5
o
C-65
o
C
Figure 6-8 shows that:
The force at the kink falls by 45% from 7,285N to 4,000N.
At the ends of the cable the force is higher than at the center, but both are falling.
The thermomechanical pattern increases in length from 7% to 75% of the route length. The
growth of the sine wave and loop patterns efficiently absorb a proportion of the locked-in
thermal strain from all of the route length.

Figure 6-8
Force distribution along route, Phase B, 42.5 65
o
C
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-12
Phase C, 65
o
C -90
o
C
Figure 6-9 shows that:
The force is nearly constant over the route and changes with temperature are small.
The force at the centre starts to rise. The thermal strain is being generated at a higher rate
than can be absorbed by the thermomechanical patterns.
The force at the ends continues to fall slightly. The patterns are still growing in length and
are absorbing thermal strain as they approach.

Figure 6-9
Force distribution along route, Phase C, 65-90C
Phase D, 90
o
C-125
o
C
Figure 6-10 shows that:
The forces at both ends of the route are increasing with temperature.
The force at the center is now higher than at the ends.
The thermomechanical patterns reach the ends of the cable at 112.5
o
C.
From 112.5
o
C to 125
o
C the thermal strain throughout the route is being generated at a higher
rate than the thermomechanical patterns could absorb. However the absorption is still
efficient and the rate of increase of force is much lower than in phase A.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-13

Figure 6-10
Force distribution along route, Phase D, 90-125C
Single Kink - Cable Length Absorption With Temperature
Figure 6-11 plots the increasing length of free cable stored in the patterns (extension) as
temperature rises. It is shown that:
The patterns absorb negligible cable length up to 40
o
C.
Above 40
o
C the patterns very efficiently release locked-in thermal strain from the cable and
store it at a lower axial force.
From 15
o
C -125
o
C the patterns have stored 325mm of thermal expansion, this is 87% of the
total prospective free expansion length of 374mm. Thus only 13% of the prospective locked-
in thermal strain remains in the cable. This is equivalent to the coefficient of thermal
expansion having been reduced from the model value of 17 x 10
-6

o
C
-1
to an effective value
of 2.2 10
-6

o
C
-1
. The extension/temperature slope is linear above 50
o
C.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-14

Figure 6-11
Variation of cable absorption length (extension) with temperature
Single Kink - Force Increase With Temperature
Figure 6-12 shows the force increasing with temperature at the ends of the model route. It is
shown that:
From 15
o
C to 42.5
o
C the force rises rapidly with temperature reaching a peak value of
8,942N from which the effective value of axial stiffness EA of 19.1MN can be calculated.
This is within the 18.2MN-20.3MN range of temperature dependent EA values set in the
model. In this temperature range it is confirmed that the cable has behaved as a straight rod in
compression containing the maximum theoretical value of locked-in thermal strain.
The force falls to a minimum at 65
o
C.
The forces at the ends of the cable rise to 4,650N at 125
o
C, from which it can be calculated
that the effective axial stiffness over the temperature rise of 110
o
C has been reduced to 13%,
ie from 19.1MN to 2.49MN.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-15

Figure 6-12
Increase of cable axial force at the ends of the route with temperature
Figure 6-13 at the kink position at the center of the span shows that:
From 15
o
C to 42.5
o
C the force rises rapidly and is linear with temperature. The slope of the
F/ curve is linear to 40
o
C and is coincident with that in Figure 6-12, giving the same value
of effective axial stiffness EA of 19.1 MN. The force reaches a peak value of 7,625N at 42.5

o
C, but as the peak has a rounded shape this reduces the effective EA value to 16.3MN. The
cable at the kink position has developed 85% of the maximum theoretical value, the reason is
that some of the thermal strain has been released to form an enlargement of the kink. The
sine wave patterns form at 42.5-45
o
C.
The force drops to a minimum value at 65
o
C and rises progressively to a value of 5,375N at
125
o
C. It is interesting to note that this is 16% higher than that at the ends of the cable.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-16

Figure 6-13
Increase of cable axial force at the central kink with temperature
Multiple Kinks - Deformation Patterns
The model was re-run with the number of kinks increased progressively from 1 to 2,3,4,5 and 29.
The kinks were equispaced along the 200m length. The spacing between the 29 kinks was 6.9m.
This was selected as being representative of the kinks expected to be present in duct and pipe
routes resulting from either a) pipe joints every 6.9m (in practice between 6m and 12m) or b) the
average circumferential length of one turn of cable on a shipping reel with a diameter of 26Dc
(Dc: cable diameter).
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-17

Figure 6-14
Growth of deformation patterns-29 kinks
Plan view, 10:1 scale
For 2,3,4 and 5 kinks, the deformation patterns were initiated in exactly the same way as for a
single kink.
The wavelengths of the sine wave patterns increased as the number of kinks increased, as shown
in Table 6-4. The progressive lengthening is attributed to the reduction in force caused by the
increasing number of kinks. For the 2 to 5 kink cases, the kink positions did not coincide with the
node points in the wave patterns. In the 29 kink case, the kinks coincided with the crest of each
wave and enforced a wavelength of 6.9m, as shown in Figure 6-14.
Table 6-4
Wavelengths at 90
o
C
Number of kinks Wavelength at 90
o
C
[m]
1 5.0
2 5.2
3 5.7
4 6.0
5 6.6
29 6.9
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-18
Figure 6-11 shows that there is little difference to be gained in the absorption of thermal strain by
increasing the number of kinks. The characteristic for 1 and 2 kinks is almost identical. The
characteristic for 29 kinks has an offset at the starting temperature of 15
o
C of 37mm. This is the
cumulative sum of the increase in cable length in each kink, before the temperature increase was
applied. There is a 1.3mm increase in length at each kink. If the offset of 37mm is removed then
the curves are nearly coincident. Thus increasing the number of kinks does not increase the
absorption of thermal strain, but does distribute the thermomechanical patterns more uniformly.
This would reduce the longitudinal movement of the cable.
Multiple Kinks Force Distribution Along the Route
The force distribution along the route is altered by the presence of multiple kinks. The force
fronts formed by release of the thermal strain in overcoming friction overlap at a progressively
lower temperature, as shown in Table 6-3. For example, the fronts from one kink have to travel
100m to reach the end of the cable, those from two kinks have to travel 33.3m before they
overlap and those from three kinks have to travel 25m. With more than one kink in a 200m route,
the fronts overlap and reduce the temperature at which the patterns fully occupy the cable span,
as shown in Table 6-5.
Table 6-5
Temperatures at which thermomechanical patterns reach the ends of the span
Number of Kinks Spacing of Kinks
[m]
Temperature at which
Patterns Reach Ends of
Cable [
o
C]
1 100 112.5
2 66.6 90
3 50 82.5
4 40 80
5 33.5 77.5
29 6.9/3.3* 42.5
* In the case of 29 kinks only, the spacing between the last kink and the cable
end was reduced.
Multiple Kinks - Force Increase With Temperature
Figure 6-12 shows the family of axial force/temperature curves for an increase in the number of
kinks. The force is plotted at the end of the cable span. Figure 6-13 is the equivalent curve at the
kink closest to the centre of the span. Increasing the number of kinks does not substantially alter
the shape of the curves, although there is a reduction in force. Table 6-6 and
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-19
Table 6-7 give a) the peak axial force and b) the force at 90
o
C for the cable at the ends of the
route.
Table 6-6 shows that an increase in the number of kinks from 1 to 5 reduces the peak force by
17%. A worthwhile reduction in force would be gained in a manhole at the end of the cable span.
The peak force for 1 and 2 kinks occurs at 42.5
o
C, but for the other cases occurs at 40
o
C. The
smallest temperature step in the computations was 2.5
o
C. Only the single kink case produced two
helical loop patterns, all of the others were sine waves, this is attributed to the higher force. The
F/ slopes for 1-5 kinks are identical. The characteristic for the 29 kink case has a progressively
reducing F/ slope with temperature, this is attributed to the cumulative absorption of thermal
strain by the 29 patterns.
Table 6-7 compares the force at 90
o
C at the ends of the cable span. From 90
o
C up to 125
o
C
increasing the number of kinks has a minimal effect on force. The absorption effect of the
thermomechanical patterns now predominates and reduces the force to a similar low level in each
case.
Table 6-6
Axial force at the ends of the cable span-multiple kinks.
Number of Kinks Peak Force [N] Temperature at
Peak Force [
o
C]
1 8,942 42.5
2 8,185 42.5
3 8,010 40
4 7,757 40
5 7,411 40
29 6,026 40

Table 6-7
Axial Force at 90
o
C at the ends of the cable span-multiple kinks.
Number of kinks Force at 90
o
C [N]
1 4,434
2 3,882
3 3,974
4 4,273
5 4,333
29 4,464

EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-20
Three Cable Model
The same procedure was followed for the modeling of three cables in a pipe. The 138kV
1500kcm model cable had a diameter of 83mm, pipe internal diameter of 233mm and clearance
of 150mm. Similar results to the single cable case were obtained. The salient points are
summarized below.
Three Cable Case: Single Kink-Deformation Patterns
At the start of the load case, the three cables were positioned in a trefoil formation. The same
20mm inelastic deformation was applied at the central node between two 300mm beam elements,
as shown in Figure 5-15. As in the single cable case, a deformation pattern formed immediately
the temperature was increased. Figure 5-15 shows that at 35
o
C the deformation length in each of
the three cables had extended to 1.2m.

Figure 6-15
Deformation patterns in three cables at 15
o
C and 35
o
C-single kink
Plan view, 1:1 scale
The cables kept their straight trefoil formation up to the peak temperature of 45
o
C. The cables
then formed a pattern of intertwining sine waves. Figure 6-16 shows that the pattern at 50
o
C has
a length of 33.6m, which is close to the 33m length for the single cable, Figure 6-1. The helical
loop patterns seen in the single cable case for one kink did not form. The absence of circular
loops is common to each of the three cable cases modeled in this project and is attributed to the
symmetrical constraint imposed by the adjacent cables.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-21

Figure 6-16
Thermomechanical pattern in three cables at 50
o
C-single kink
Plan projection view: 1:1 scale
Figure 6-17 shows the full 200m route at 90
o
C in plan view. The lateral to longitudinal scale
factor has been increased to 20:1 to permit the lateral deformation patterns to be seen. The
pattern has grown to 126m compared to 144m for the single cable case in Figure 6-5. The
patterns in the three individual cables are shown in Figure 6-17 and are seen to be the same. It is
shown that two regions exist along the route where the amplitudes and wavelengths of the sine
waves are larger. The wavelength in this region is 6.9m, compared to 4.7m elsewhere. It is noted
that the two regions are not completely symmetrical either side of the 100m centerline. Figure
6-17 shows what appears to be depressions on the peaks of the larger sine waves, however this is
an optical effect caused by the peak of the cylindrical sine wave passing around the inside of the
pipe.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-22

Figure 6-17
Deformation patterns in the three cables at 90
o
C-single kink
Plan view, 20:1 scale
Three Cable Case: Single and Multiple Kinks-Force Increase with Temperature
Figure 6-18 shows the force/temperature characteristic at the end of the 200m cable span for the
cases of 1 and 29 kinks. Table 6-8 lists the magnitudes of force. In summary, the characteristics
for the three cable and single cable cases are closely similar:
All three cables in the single kink case generate similar peak forces of 9810N average, this is
close to the level of 8,942N (10% less), generated in the single cable case.
The temperature at which the force peaks is slightly higher at 45
o
C compared to 40-42.5
o
C.
The gradient F/ of the line to the peak force for the single kink case gives an effective axial
stiffness EA value of 19.2MN. This is close to 19.1MN for the single cable case and within
the18.2MN - 20.3MN range of EA values set in the model.
In the 29 kink case the peaks of the three cables have a greater variation. The average value
is 6,678N, which is 11% higher than the value of 6,026N in the single cable case.
The peak forces in the three cable 29 kink case are 32% less than the single kink case. This is
close to the 33% reduction in the single cable case.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-23
The gradient F/ of the line to the peak force for the 29 kink case has a similar curved shape
to the single cable case.
At 90
o
C there is a significant variation in the values of force in the three cables in both the
single and 29 kink cases. The 29 kink average is 2400N, this being 48% below the single
kink value of 4651N. The equivalent single cable values were 4,464N and 4,434N. The
single kink cases for three cables and one cable are close (within 4%), this is not so for the 29
kink case.

Figure 6-18
Increase of axial force at the ends of three cables-1 and 29 kinks

Table 6-8
Axial forces at the span ends, three cables with multiple kinks
Number of
Kinks
Cable
Number
Peak
force [N]
Temperature
at Peak
Force [
o
C]
Force at 90
o
C
[N]
Force at
125
o
C [N]
1 (top) 9,837 45 2752 2426
2 9,797 45 5600 4614
1 kink
3 9,798 45 5601 4614
1 (top) 7,124 42.5 550 2512
2 6,237 45 2837 3378
29 kinks
3 6,672 45 3840 4373

EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-24
Conclusions
The following conclusions are general to all applications of cables in pipes and ducts. Some
points have been illustrated with values from the study of the 138kV 1500kcm XLPE reference
cable with a pipe clearance of 150mm in a 200m span. Variations in cable parameters and pipe
geometry will affect the magnitudes of thermomechanical forces. For example a cable in a tight
fitting duct is unlikely to generate thermomechanical patterns and so will obtain little benefit in
force reduction. Variations in cable parameters and route geometry are covered in the sensitivity
study in Chapter 7.
From the study of the straight cable models it was concluded that:
Small perturbations (kinks) in the cable and small deviations in the pipe route initiate
thermomechanical patterns in the same way as large preformed bends in the pipe route. The
patterns are the same. One difference is that a pattern from a kink grows in both directions,
whereas a pattern from a bend grows in one direction only (it does not grow around the
bend).
Thermomechanical patterns can be expected to form in all nominally straight pipe routes,
because practical pipes and cables will possess some deviation and curvature.
Three cables in a pipe system behave in a closely similar way to one cable in a duct system,
providing they have the same clearance between one of the individual cables and the pipe.
Differences were that a) the peak force at the turn-over temperature was 10% higher in the
three cable case and b) a three cables in a pipe system can only produce sine wave patterns,
whereas one cable in a duct can also produce helical loop patterns.
The sine wave thermomechanical patterns in both pipe and duct systems have a force
sensitive, non-linear characteristic that can absorb thermal strain over a wide range of
temperature with only a small increase in force.
Thermomechanical patterns are comparatively mild geometric shapes with a wavelength to
amplitude ratio of more than 20:1. For example they have typical wavelengths of 5-7m,
compared to a circumferential peak to peak amplitude of less than 0.47m (this is one
circumference for the locus of the cable axis).
The presence of the thermomechanical pattern has a very beneficial effect in reducing the
maximum cable thrust, for example by 64% in the single cable case.
The peak force in the theoretical case of a perfectly straight cable in a tight fitting pipe occurs
at the design temperature of 90
o
C. However the presence of a thermomechanical pattern
limits the force to the value reached at the turn-over temperature of 40-45
o
C at which a sine
wave pattern is formed. The pattern absorbs locked-in thermal strain from the cable and
further reduces the force. The force at the 90
o
C continuous operating temperature for an
XLPE cable was reduced by 82% (single cable case).
Possible disadvantages of thermomechanical patterns are that a) bending movement is
increased, thereby introducing the risk of fatigue failure and b) the reduction in bend radius
locally increases sidewall load on the cable components. These subjects are considered in
later chapters.
EPRI Licensed Material

Thermomechanical Performance of Cables in a Straight Route
6-25
The cable in the vicinity of a kink will benefit by greater reduction in peak force, for example
in the single kink case, the peak force is reduced by 69%, whereas the force 100m further
away at the ends of the circuit is reduced by 64%.
Increasing the number of kinks is beneficial in further reducing the force at the ends of the
route, however the percentage improvement is small. For example in the single cable case,
one kink reduced the peak force by 64%, 5 by 70% and 29 by 75%. Additional kinks also
produce a) uniform distribution of force, b) reduced force at nearby manhole positions and c)
reduced longitudinal thermomechanical movement.
Transition from the incipient oscillating decrement pattern to the sine wave pattern, from a
geometrical point of view, is continuous and smooth. The half waves of a decrement pattern
grow in magnitude until they become large amplitude sine waves. However the transition
occurs rapidly in less than the minimum temperature step of 2.5
o
C used in the study.
Thermomechanical patterns reduce the force along the cable for distances well in excess of
their own length. For example at 45
o
C in the single cable case the pattern was 21m wide, but
the length of reduction in force had extended to the full 200m cable span. The length of cable
affected is determined by the frictional constraint (weight of cable and the coefficient of
friction). The length benefiting from a reduction in force is reduced for a) heavy cables and
b) pipes with a high coefficient of friction.
The study of real route models containing preformed bends and straight sections is not
invalidated by the work in this chapter on kinks. The cable curvature in the bends and in the
joint chambers will initiate thermomechanical patterns. The presence of additional pattern
initiating sources will result in comparatively small further reductions in peak force.
A possible practical application for further study is the inclusion in pipe and duct routes of
small deviations in the pipe, or of deliberate kinks in the cable. Deviations could be
positioned in the vicinity of manholes particularly in straight spans, with the objectives of a)
reducing the thrust acting on the joints and cleats and b) reducing longitudinal cable
movement.





i
Mechanical Effects on Extruded Dielectric Cables and Joints Installed in Underground Transmission Systems in
North America, Technical Update, August 2002, EPRI, Palo Alto: CA, 1001848.
EPRI Licensed Material
7-1
7
THERMOMECHANICAL PERFORMANCE OF CABLE IN
A ROUTE CONTAINING BENDS AND SLOPES:
SENSITIVITY STUDY
Introduction
At the start of the project, models were constructed from representative parts of a 138kV and a
230kV duct route. These showed that thermomechanical patterns formed in the pipe system close
to bends and to joint positions
[i]
. The patterns appeared to form adjacent to bends and some joint
positions. There were no patterns present in the 230kV duct system, but cable movements
occurred into bends and into manholes. A number of interactions were identified, although the
routes had too many geometric features to be able to isolate cause and effect. The work in this
chapter was designed to identity and quantify the interactions in a simple route. The knowledge
gained could then be applied to the analysis of real duct and pipe routes, as studied in later
chapters.
The sensitivity study was designed to investigate:
Thermomechanical and gravitational effects
Effect of varying the cable parameters
Effect of varying pipe and duct system route geometries
Predictability of the thermomechanical effects
Usefulness of the analysis as the basis of a design guide
A route model was selected with a) the basic elements present in real routes and b) geometries
that would permit a direct comparison of a duct and pipe route. The route selected contained one
straight section and one bend. The purpose of the straight section is a) to generate locked-in
thermal strain and axial thrust and b) to generate frictional constraint to cable movement. The
purpose of the bend is a) to absorb thermal expansion strain and b) to generate thermomechanical
patterns in a natural and repeatable way.
From the study it was found that patterns were initiated by bending moments formed from the
combination of a) cable curvature and b) thermomechanical axial thrust. These are described in
Chapter 2, Basic Theory. Cable curvature has to be present in the route before load is applied. In
the case of the hockey stick model it is provided by the pre-formed bend. In the case of the
straight section studied in Chapter 6 it is provided by curvature at the cable kink inserted into
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-2
the model. The work in this chapter shows that the initiation of thermomechanical patterns is
closely identical in both cases.
The reference model was later used to investigate the effects of vibrations (Chapter 9) and short
circuits (Chapter10). The figures and tables in Chapter 10 are a useful source of supplementary
information as they show, quantify and compare thermomechanical patterns at 90
o
C in pipe and
duct systems.
This is a self-contained chapter, the work is recorded in the following stages:
Selection of the route model
Selection of the key variables and thence a method of normalizing the geometry
Identification of the key parameters to be varied and the ranges of their variation
Investigation of pattern initiation at a 5
o
deviation in the route
Investigation of pattern initiation at the 90
o
bend in the reference route
Ranking the results of the sensitivity study
Magnitude of the parameters that limit service performance
Discussion of results
Conclusions
Further work
The sensitivity study bar charts and graphs are included at the end of the chapter.
Reference Model
The geometry of the reference model is shown diagrammatically in Figure 7-1. It comprises a
horizontal 200m hockey stick shaped span, with a 90
o
bend of 35Dc radius, where Dc is the
diameter of the cable. The 200m length of straight section was selected as being longer than the
typical straight sections present in real routes, but shorter than typical span lengths of 500-800m.
The 35Dc radius is selected a) as being the smallest present in the 138kV pipe route studied and
b) at the upper end of the 25-35D range of minimum duct radii given in Chapter 4 for HV and
EHV XLPE cables. The scaled print-out in Figure 7-2 shows that the bend is a minor part of the
route.
The reference model comprises either one duct containing a single cable or an identical pipe
containing three cables in trefoil formation.
An element length of 300mm was selected as the shortest possible, consistent with the software
rules.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-3

Figure 7-1
Dimensions of reference model
Not to scale


Figure 7-2
Print-out of reference model (200m straight section and 2.9m radius bend)
Plan view, scale 1:1
Normalization of the Sensitivity Study
A method of normalizing the cable and pipe geometry is required to a) aid the understanding of
the phenomena and b) reduce the number of models that require to be solved. Without a
normalized geometry it would be necessary to run the model for every conductor size, insulation
thickness, sheath type, cable diameter and pipe diameter within the 69-345kV range. This is
clearly impractical and would be of limited use as a design guide, when the dimensions of a
future cable design differed.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-4
A normalization study was performed to test the suitability of different geometric ratios. It was
concluded that the model should be normalized on the pipe to cable clearance as given in
Equation 7-1:
D=(Dp-Dc) Equation 7-1
Where:
Dp: internal diameter of pipe of duct [mm]
Dc: outer diameter of the cable [mm]
The reason for the choice of (Dp-Dc) is that it is the key parameter determining the amount of
thermal strain that can be absorbed by the different cable geometries present in the route
(Chapter 2, Basic Theory):
Preset bends
Cylindrical sinusoids (abbreviated as sine waves)
Helical loops
The useful consequence is that two cables of different diameters, but the same pipe clearance
D, will have the same thermomechanical performance, if all other parameters are the same
(stiffnesses, weights, coefficients of friction etc). Thus one model cable diameter Dc will provide
results that can be used across the complete voltage and conductor size ranges.
Verification of Normalization on Pipe Clearance
The principle of normalization was successfully validated by modelling performance with
extreme differences in cable diameter over a wide range of temperatures and force as described
below.
The test parameters were chosen to be at the ends of the transmission class XLPE cable range,
Table 7-1. The 345kV 3000kcm cable has nearly twice the diameter of the 69kV 1000kcm cable.
All of the other parameters of the two cables had to be identical for the comparison.
The pipe clearance of 150mm was chosen by taking the pipe to diameter ratio of 2.48 from a true
pipe system with an internal diameter of 206mm containing a 138kV 1500kcmil cable with
diameter of 83mm and then applying the ratio to the 69kV and 345kV cable diameters. This gave
pipe clearances of 114mm and 200mm, which when averaged and rounded gave a 150mm
clearance. The coefficient of friction of 0.2 was calculated from recorded pulling tensions for the
138kV cable. The other parameters were those of the 230kV 2500kcm cable that had been a) part
measured from initial test results (Chapter 3) and b) calculated for each of the three layers
(conductor, insulation and sheath/jacket). The weight of 34 kgm
-1
is that of the 230kV 2500kcmil
cable. The 230kV cable was chosen because the 27mm insulation thickness is equivalent to that
of a 345kV cable.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-5
The parameters to be modeled for the 345kV test cable were close to reality, although it would
be unusual to install a cable in a duct system with a clearance of 150mm. The weight and
stiffnesses were applied to the 69kV test cable and are significantly higher than would be
encountered in reality.
Table 7-1
Test parameters to validate normalization
Parameter 69kV 1000kcm 345kV 3000kcm
Outer diameter [mm] 77 133
Pipe ID [mm] 227 283
Pipe clearance D [mm] 150
Straight section [m] 200
Bend radius [m] 2.69 (35D)
Weight [kgm
-1
] 34.2
Coefficient of friction with pipe, 0.2
Coefficient of thermal expansion, [
o
C
-1
] 17 x 10
-6

Axial stiffness, EA [MN] 28
Bending stiffness, EI [Nm
-2
] 7,740
Torsional stiffness, GJ [Nm
-2
] 9,144
The load case that was applied to the horizontal test model was a) to place the cables on the pipe
floor in a neutral condition by dropping them under gravity and b) to heat them from 15
o
C to
215
o
C. The 200
o
C temperature increase was chosen to ensure that thermomechanical patterns
would form and fill the pipe.
The force temperature characteristics are shown in Figure 7-3 for position B at the entrance to
the bend. Three heating phases were applied to the model, each with different temperature
intervals, thus the calibration in Table 7-2 is needed to relate the process time scale to
temperature.

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-6

Figure 7-3
Axial force/temperature characteristics for the 69kV cable (left) and 345kV cable (right)
Note: computation time scale is non-linear with temperature

Table 7-2
Calibration between time and temperature
Process
Time [s]
Temperature
[
o
C]
Temp Step
Size [
o
C]
Number of
Temperature
Steps
Process
Step No
Comment
0-1 15 - - 0-1 Drop cable
1-2.2 15-45 6 5 1-2 Heat to 45
o
C
2.2-4 45-90 3 15 2-3 Heat to 90
o
C
4-10 90-215 5.2 24 3-4 Heat to 215
o
C

Table 7-2 compares the axial forces at the five sample positions shown in Figure 7-3. The forces
agree over a temperature range of 65-152
o
C to within 0.88%.


EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-7
Table 7-3
Test case: comparison of axial forces
Axial Force [N] Point
on
Graph
Time
[s]
Temp
[
o
C]
69kV
1000kcm
345kV
2500kcm
Difference
[%]
Comment
1 3 65 4,083 4,119 +0.88 Rate of force increasing
2 4 90 14,364 14,310 -0.38 Cable operating temperature
3 5.25 116 25,599 25,605 +0.02 Force turn-over point
4 6 131.6 26,229 26,277 +0.18 6s marker on time scale
5 7 152.4 10,120 10,158 +0.38 7s marker on time scale


Figure 7-4
Identical thermomechanical pattern formed at the turn-over temperature/peak force
position for the 69kV and 345kV duct system models
Scale, 10:1
Figure 7-4 is a plan view of the deformation plot at the turn-over force position at which the
sinusoidal and helical patterns become predominant at position 3, 116
o
C. showing the 32m cable
length closest to position B at the start of the bend. The plots for the 69kV and 345kV cables
were superimposed on each other and were confirmed to be identical. A three dimensional view
showed that the pattern was comprised of sine waves with one helical loop at exactly the same
stage of formation in both plots. At position number 5, 152
o
C, the deformation pattern had
increased in length to 98m. The plots were superimposed and were found to be closely similar
with a slight difference developing in the wave shapes remote from the bend, however the axial
forces still agreed within 0.38%.
Having validated the principle of normalization on pipe to cable clearance the sensitivity study
could proceed.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-8
Sensitivity Study: Variations in Geometry of Model Route
Clearance Between Cable and Pipe
Table 7-4 covers the extreme values of pipe clearances. The maximum and minimum values are
selected for study and not recommended for installation design purposes.
The range of clearances was selected by applying the jam ratios for a pipe system to each of the
selected cable diameters, the largest clearance being 250mm. The reference cable clearance is
selected to be 150mm for three cables in the pipe, for the reason given in the previous section.
The minimum clearance between the exscribed circle encompassing the three cables in trefoil
(95mm OD) and the pipe (100mm ID) is selected to be 5mm.
For one cable in a duct, the minimum clearance is selected to be 10mm. The maximum clearance
of 250mm is the same as the pipe clearance. This was to compare the differences in performance
between one and three cables in the same pipe diameter.
The clearances present in the real cable installations studied in later chapters are:
138kV 1500kcm pipe system: 123mm clearance (27.5mm installation clearance on exscribed
circle around trefoil group of 179mm)
230kV 2500kcm duct system: 37.3mm clearance

Table 7-4
Cable to Pipe Clearances
Range of Cable to Pipe Clearances (Dp-Dc) [mm]
Duct Containing Single Cable Pipe Containing Three Cables
10 -
50 -
100 100
- 115
- 130
150 ref model 150 ref model
200 200
250 250
Pipe Length
The longest straight section modeled was selected to be 1000m as this approaches the longest
cable lengths that can be shipped on standard reels for the larger conductor sizes 138-345kV
cables. The shortest length modeled is 100m representing for example a short link in a power
plant.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-9
Table 7-5
Lengths of straight pipe modeled
Length of Straight Span [m]
100
200 ref model
400
1000
Bend Radius and Bend Angle
The values are given in Table 7-6. A 35D radius was taken as the reference, this value having
been present in the 138kV 1500kcm route (2.9m) and is in the range of 25-35D listed for ducts in
Chapter 4. A radius of 8D was used in the design of manhole for the 230kV 2500kcm cable
(1m). The 150D value is representative of a gentle route bend.
The bend angles are given in Table 7-7, the smallest being 5
o
, which simulates small horizontal
and vertical deviations in pipe routes, whilst also providing a position for pattern initiation.
Table 7-6
Bend Radii
Bend Radius
Ratio on Cable
Diameter D
Radius [m]
5D 0.42
10D 0.83
35D ref model 2.91
50D 4.15
100D 8.3
150D 12.45
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-10
Table 7-7
Bend Angles
Bend Angle
[Degrees]
5
30
60
90 ref model
Angle of Inclination
For the reference case the model is held horizontally. The angle was varied between 30
o
as
shown in Figure 7-1. Generally the slopes encountered within cities are less than 10
o
as the
routes follow the gradient of the road surface. Results for smaller angles in the range of 7.5
o

are given in Chapter 10. The critical angle of slip for the maximum slope of 30
o
occurs with a
coefficient of friction of 0.58, which is within the range studied.
Table 7-8
Slope Angle
Slope Angle [Degrees]
+30
+20
+10
0 ref model
-10
-20
-30
Coefficient of Friction
The reference coefficient of friction between the cables and between the cable and the pipe is
0.1, this being the typical value measured when pulling-in XLPE cables with polyethylene
jackets into well lubricated pipes and ducts. A value of 0.3 is recommended in reference
[ii]
. It is
reasonable to assume that the value will increases with time after the lubricant has dried out and
the surfaces of the cable and pipe have deteriorated. The extreme values of 0 and 1 are included
to provide information.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-11
In the sensitivity study for the pipe system, two studies were performed. One kept the cable to
cable constant at 0.1 and the other kept the pipe to cable constant at the same value.
Table 7-9
Coefficient of friction
Coefficient of Friction
Critical Angle of Slip
[Degrees]
0 0
0.1 ref model 5.7
0.3 16.7
0.5 26.6
0.7 35
1.0 45
Cable Designs
Cable designs was chosen to cover the full range of cables in the project. The following designs
were selected and are arranged in order of bending stiffness EI ( lowest EI at the top):
69kV, 1000kcm, foil laminate, polyethylene jacket.
138kV 1500kcm, foil laminate, polyethylene jacket. (Reference cable)
345kV 1000kcm, lead sheath, polyethylene jacket
230kV 3000kcm, lead sheath, polyethylene jacket
138kV 3000kcm, corrugated aluminum sheath, polyethylene jacket.
345kV 3000kcm, corrugated aluminum sheath, polyethylene jacket.
The 69kV and 345kV cables were selected to cover the lowest and highest system voltages in the
project. Conductor sizes varied from 1000kcm to 3000kcm, being the range expected to be
encountered in present and future use. Three metallic sheath systems were included, extruded
lead, extruded aluminum and copper tapes with aluminum foil laminate. The jacket material was
polyethylene.
The values for EI, EA and GJ were derived for each temperature and for each of the three
component beams in the model. The weights of each of the designs were calculated and these
were used for the weight sensitivity.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-12
Calculation of Cable Component Stiffnesses
The stiffnesses that require to be determined for each of the three cable components are:
EI: bending stiffness [Nm
2
]
EA: axial stiffness [MN]
GJ: torsional stiffness [Nm
2
]
The recommended approach is to measure values from cable samples, with similar methods to
those described in Chapter 3, using wherever possible long sample lengths to overcome end
effects.
Properties for the complete cable are first measured The jacket and sheath layers are then
removed to reveal the extruded core (semiconducting shields and XLPE insulation), which
encapsulate the conductor. The properties are remeasured at the core-only stage. Unfortunately
the properties of the XLPE insulation and conductor are complex and interactive.
The work in this project has confirmed that the conductor is the key component in determining a)
stiffness and b) thermomechanical performance of the cable. Unfortunately it is not possible to
expose the conductor and measure its properties for the reasons given below:
Conductors have traditionally been designed to have a flexible construction comprised of
individual wires to permit the cable to be wound onto the despatch reel and to be pulled-in
around bends. Various conductor constructions exist, the most common being the stranded
circular design present in the 138kV 1500kcm cable, this being used typically up to 2000kcm,
and stranded segmental milliken designs used for larger sizes, as present in the 230kV
2500kcm cable. Uncompacted round wire conductors are the most flexible, but have a low
occupancy of area, which increases the overall size and cost of he cable. It is usual to increases
occupancy by dieing down and compacting the wires and layers, resulting in the conductor
increasing in axial and bending stiffness. Thus different cable designs made by different
manufacturers will have different stiffnesses. The XLPE extrusion and crosslinking processes
cause the XLPE layer to contract and compress the conductor, which also increases the stiffness.
XLPE is a stiff material (high elastic modulus) at ambient temperature. Above 60
o
C XLPE a)
softens significantly and b) expands at an increasing rate, thereby relaxing its grip on the
conductor and permitting the conductor to become more flexible. At the cable operating
temperature of 90
o
C and emergency temperatures of 105-110
o
C, the stiffness of the XLPE layer
and in particular the conductor will be particularly low.
If the stiffness of the conductor is measured with the XLPE insulation removed, the conductor
wires will be free to flex, expand outwards and slide in a way that cannot occur within the cable.
Thus the conductor would exhibit an artificially reduced stiffness (low EI).
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-13
The method of separating the elastic stiffness of the XLPE insulation from the conductor is to
calculate the bending stiffness over a range of temperatures and to subtract this from the
experimentally measured stiffness for the combined insulation and conductor.
The measured and derived stiffnesses EI, EA and GJ described below are valid parameters. Care
should be taken when separating the geometric parts I, A and J from the elastic modulii E and J
as the values cease to have a traceable physical meaning. The approach taken below is to
calculate the values of I, A and J as though the particular cable component is a geometrical solid
with the same outer and inner diameters. The value of the accompanying modulus then becomes
the effective value representing both the material properties and the constructional geometry of
the layer (e.g. the stranded conductor, the composite XLPE insulation/semicon shield layer and
the composite sheath/jacket layer).
EI: Bending Stiffness
The values of EI for the cables in the sensitivity model were derived in a similar manner to those
in the route models. The EPRIsolutions test laboratory measurements gave values for two cables
at three temperatures: ambient, 60C, 90C and 105C. The samples were measured a) as
complete cable and b) in the core-only stage, comprising the conductor covered with the
triple extruded XLPE insulation and semiconducting shields.
However the model requires three sets of stiffness values corresponding to the conductor, the
insulation and the sheath/jacket system. Theses values are temperature dependent.
Jacket/Sheath EI
s

EI
s
was determined by subtracting the measured value for the core-only sample, (comprising the
conductor and XLPE insulation), from the value for the total cable:
EI
s
= EI
t
EI
cx
Equation 7-2
Where:
EI
s
: bending stiffness of sheath/jacket [Nm
2
]
EI
t
: total bending stiffness of complete cable as measured [m
2
]
EI
cx
: measured bending stiffness of conductor and XLPE insulation [Nm
2
]
Insulation EI
x

A value of EI
x
for the XLPE insulation only was obtained from the following equation:
( )
4 4
x x
d D
64
E E =

Equation 7-3
Where;
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-14
E
x
: effective modulus of elasticity for XLPE insulation and shield layer [Nm
-2
]
I
x
: second moment of area of XLPE insulation along its axis [m
4
]
D: outside diameter over the semiconducting insulation shield [m]
d: inside diameter of the semiconducting conductor shield [m]
The value of E
x
, the effective modulus of elasticity for the XLPE insulation, was derived from
measurements carried out by EPRI
[iii]
. The values for the modulus under compression and the
modulus under tension at each temperature are given in this report, Chapter 5 Modelling. The
temperature is taken at the midpoint of the insulation. It is necessary to calculate the temperature
drop from the conductor for the particular ampacity rating and conductor operating temperature.
Upon bending the cable, the inside of the insulation is under compression and the outside under
tension, an average value of the compressive and tensile modulii is required to be calculated.
Tables of values for the 138kV 1500kcm and 230kVkcm cables are given in Chapter 5.
Conductor EI
c

The conductor EI
c
is calculated in Equation 7-4 from the measured EIcx and calculated EIx from
Equation 7-3:
EI
c
= EI
cx
EI
x
Equation 7-4
For the sensitivity study model only, the value for the effective modulus of elasticity E
c
for the
stranded copper conductor was then determined from Equation 7-5.
4
c
c
D
64 E
E

= Equation 7-5
Where:
E
c
: effective modulus of elasticity of the stranded conductor [Nm
-2
]
EI
c
: bending stiffness of the conductor [Nm
2
]
D: diameter of the conductor [m]
This process was repeated at each temperature step.
Total Cable EI
Having derived the effective values of E for each of the three cable layers, then the value EI was
calculated from Equation 7-6 for each of the other four cables in the study.
Where D is the outer diameter of each layer and d is the inner diameter.
( )
4 4
d D
64
E E =

Equation 7-6
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-15
The total cable bending stiffness is given by Equation 7-7:
s x c t
E E E E + + = Equation 7-7
EA Axial Stiffness
Values of EA were part derived following the procedure described above for EI, from a)
measured values and b) calculated values using Equation 7-8.
( )
|
|
.
|

\
|
=
4
d D
E EA
2 2

Equation 7-8
Where:
EA: axial stiffness of cable component [MN]
E: effective elastic modulus in compression of each component [Nm
-2
]
D: outer diameter of a given layer [m]
D: inner diameter of a given layer [m]
The total axial stiffness for a particular cable design is the addition of each layer, Equation 7-9:
( )
s x c t
EA EA EA EA + + = Equation 7-9
GJ Torsional Stiffness
Values of GJ were part derived following the procedure described above for EI, from a)
measured values and b) calculated values using Equation 7-10 and Equation 7-11.
( )
4 4
d D
32
J =

Equation 7-10
( ) +
=
1 2
E
G Equation 7-11
Where:
GJ: torsional stiffness [Nm
2
]
J: polar second moment of area [m
4
]
D: outer diameter of the particular cable component [m]
D: inner diameter of the particular cable component [m]
: Poissons ratio for XLPE (0.3)
The total torsional stiffness for a particular model cable design is the addition of each layer,
Equation 7-12:
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-16
( )
s x c t
GJ GJ GJ GJ + + = Equation 7-12
Cable Parameters Used in the Sensitivity Study
The reference cable design comprises:
138kV system voltage,
1500kcm circular stranded copper conductor,
Triple extruded XLPE insulation and semiconducting shields
Copper tape ground conductor
Aluminum co-polymer coated foil water barrier, bonded to an extruded polyethylene jacket.
The stage diameters are:
Conductor diameter: 32.9mm
Insulation and shield diameter: 73.9mm
Sheath/jacket diameter: 83mm
Weights for the Range of Cables
Table 7-10 is the list of total cable weights that were varied in the sensitivity study. They are
referred to by their ratio to the weight of the 138kV reference cable of 11.9kgm
-1
. The 0.67
ratio represents the weight of a 69kV 1000kcm copper conductor foil sheathed cable and the 3.08
ratio a 345kV 3000kcm copper conductor lead sheathed cable. For comparison, the samples of
230kV 2500kcm copper conductor lead sheathed cable tested by EPRIsolutions weighs
34.3kgm
-1
(a ratio of 2.9).
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-17
Table 7-10
Total Cable Weights
Total Cable Weight
Ratio to Ref Cable Total Cable Weight
[kgm
-1
]
0.67 8
1.03 12.3
1.64 19.5
2.36 28
3.08 36.6
Coefficient of Thermal Expansion for the Range of Cables
The values that were selected for the model are given in Table 7-11. An axi-symmetric
cylindrical FEA model of an XLPE cable was constructed to investigate the effective coefficient
of thermal expansion of the cable. It was found that the copper conductor predominated over the
other coefficients of expansion within the cable. The value for the cable was only marginally
increased from 17 to 18 x 10
-6

o
C
-1
, despite the high coefficient of expansion of the thick layer of
XLPE insulation. The value subsequently measured on a sample of 138kV 1500kcm XLPE cable
was 18.5 x 10
-6

o
C
-1
. It is recommended that this value be taken when designing pipe and duct
systems.
Table 7-11
Values of coefficients of thermal expansion
Coefficient of Thermal Expansion
[
o
C
-1
x 10
-6
]
Comment
14 Lower limit for information
17 (ref case) Property of copper
20 From technical papers for XLPE cables
23 Property of aluminum
Stiffness Parameters for the Range of Cables
The elastic stiffnesses for each of the five cable designs studied in the model are given for EI in
Table 7-12, EA in Table 7-13 and GJ in Table 7-14. These parameters were based on cable
measurements taken at the beginning of the test program. It was later realized that the axial
stiffness values from the axial stiffness rig were too low. These should be replaced with values
from the thermally constrained thrust test and are given in Table 7-15 for the 138kV and 230kV
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-18
cable designs. It is recommended that the Table 7-13 values be used as the basis for the design of
pipe and duct systems.
The key value in interpreting the performance of the model cable is the total stiffness. Each
model cable is constructed with the values for each of the three layers. The values referred to in
the results of the sensitivity study are the total values of EI, EA and GJ for each of the 5 cable
constructions.
Table 7-12
Bending stiffness EI for the cable designs
EI of Cable Components [Nm
2
] Cable Design
Conductor Insulation Sheath Total Cable
20C
138/1500/Foil/PE (Ref Cable) 1699 167.5 816 2682.5
345/3000/CSA/PE 5429.5 858 9780 16069.7
138/3000/CSA/PE 5429.5 488 6610 12527
230/3000/Lead/PE 5429.5 679 4830 10935
345/1000/Lead/PE 604 374 2960 3939.2
69/1000/Foil/PE 604 104 573 1279.75

60C
138/1500/Foil/PE (Ref Cable) 623.6 36.99 528.16 1188.8
345/3000/CSA/PE 2499.7 185 8570.5 11254.2
138/3000/CSA/PE 2499.7 105 5778.3 8383.3
230/3000/Lead/PE 2499.7 147 2900 6733
345/1000/Lead/PE 278 80.8 2402 2765
69/1000/Foil/PE 278 22.4 353 653.2

90C
138/1500/Foil/PE (Ref Cable) 366.8 12.1 762.2 1141.1
345/3000/CSA/PE 1466.7 60.6 9530.5 10958.4
138/3000/CSA/PE 1466.7 34.4 6369.4 7870.5
230/3000/Lead/PE 1466.7 47.9 4620 6127.6
345/1000/Lead/PE 163 26.4 2800 2990.4
69/1000/Foil/PE 163 7.3 509 679.3

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-19
Table 7-13
Axial Stiffness EA for cable designs
EA of Cable Components [MN] Cable Design
Conductor Insulation Sheath Total Cable
20C
138/1500/Foil/PE (Ref cable) 17.5 0.223 0.5 18.2
345/3000/CSA/PE 34.9 0.388 4.38 39.7
138/3000/CSA/PE 34.9 0.273 3.84 39.1
230/3000/Lead/PE 34.9 0.336 2.30 37.6
345/1000/Lead/PE 11.7 0.336 2.30 14.4
69/1000/Foil/PE 11.7 0.136 0.45 12.2

60C
138/1500/Foil/PE (Ref cable) 18.7 0.055 1.1 19.9
345/3000/CSA/PE 37.4 0.0951 5.25 42.7
138/3000/CSA/PE 37.4 0.0669 4.61 42.1
230/3000/Lead/PE 37.4 0.0824 3.07 40.5
345/1000/Lead/PE 12.5 0.0697 2.65 15.2
69/1000/Foil/PE 12.5 0.0334 0.991 13.5

90C
138/1500/Foil/PE (Ref Cable) 19.1 0.0151 1.22 20.3
345/3000/CSA/PE 38.2 0.0262 5.42 43.6
138/3000/CSA/PE 38.2 0.0185 4.76 42.9
230/3000/Lead/PE 38.2 0.0227 3.23 41.4
345/1000/Lead/PE 12.7 0.0192 2.79 15.5
69/1000/Foil/PE 12.7 0.00921 1.10 13.8


EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-20
Table 7-14
Torsional stiffness GJ for cable designs
GJ of Cable Components [Nm
2
] Cable Design
Conductor Insulation Sheath Total Cable
20C
138/1500/Foil/PE (Ref Cable) 1133.1 113 424.9 1671
345/3000/CSA/PE 4499.6 566 6593.9 11660.0
138/3000/CSA/PE 4499.6 322 4449.2 9270.4
230/3000/Lead/PE 4499.6 448 3033.5 7980.9
345/1000/Lead/PE 504.25 247 1817.1 2568.6
69/1000/Foil/PE 504.25 68.4 303 875.5

60C
138/1500/Foil/PE (Ref Cable) 877.7 24.7 611.2 1513.5
345/3000/CSA/PE 3485.9 124 7326.9 10824.8
138/3000/CSA/PE 3485.9 70.2 4951.7 8507.8
230/3000/Lead/PE 3485.9 97.7 3481 7064.6
345/1000/Lead/PE 390.6 53.9 2152.9 2597.4
69/1000/Foil/PE 390.6 14.9 435.6 841.1

90C
138/1500/Foil/PE (Ref Cable) 674.64 8.06 592.0 1274.7
345/3000/CSA/PE 2679.4 40.4 7251.8 9971.6
138/3000/CSA/PE 2679.4 22.9 4899.6 7602
230/3000/Lead/PE 2679.4 31.9 3435 6146.3
345/1000/Lead/PE 300.28 17.6 2118.6 2436.5
69/1000/Foil/PE 300.28 4.87 422 727.15

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-21
Table 7-15
Axial stiffness EA for the complete 138kV and 230kV Cables
Axial Stiffness EA [N x 10
7
] Temperature [C]
138kV 1500kcm EA 230kV 2500kcm
15 6.76 11.3
20 6.26 10.5
30 5.35 8.92
40 4.47 7.45
50 3.72 6.20
60 3.00 5.00
70 2.34 3.90
80 1.80 3.00
90 1.47 2.45
100 1.29 2.15
110 1.20 2.00
120 1.13 1.88
125 1.13 1.88
Load Case for the Sensitivity Study
Process time step 0-1seconds:
The cable was first placed in a neutral position in the bottom of the pipe by being dropped from a
height of 1mm under gravitational force. The ends of the cable were then anchored such that they
would not exhibit curvature, in this way unwanted initiation of patterns would be avoided. The
ends were held parallel to the axis of the pipe and were free to translate across the end of the
pipe.
Process time step 1-6 seconds:
The temperature was uniformly increased from 15
o
C to 125
o
C. Output was requested from 44
frames with 2.5
o
C increments. The ambient of 15
o
C was taken as an average annual temperature
for the cable system. The temperature rise of 110
o
C was selected, because some installations in
North America could in principle be installed and jointed at 0
o
C and subsequently be heated to
an emergency loading temperature of 110
o
C.
Formation of Thermomechanical Patterns in the 5 Bend Model
In Chapter 6 the effect of a small kink in the cable was studied. The kink comprised a 20mm
offset at the center of two 0.3 m element lengths. A comparison is now made with the effect of a
small deviation in the pipe route comprising a 5
o
bend and the reference radius of 2.9m. The
resulting arc of the pipe bend being of short length at 0.25m.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-22
The model showed that an incipient deformation pattern was present at 37.5
o
C, Figure 7-5. It has
a closely identical shape to that formed by one cable kink, as shown in Figure 6-3 at 42.5
o
C. To
aid comparison the 11.8m distance from the bend to the edge of the Figure 7-5 frame is the same
as from the centre of the kink to the frame edge. (The node separation of 300mm and the lateral
scale of 10:1 are also the same).

Figure 7-5
Duct system with 5
o
bend, showing a pattern forming at 37.5
o
C
Lateral scale 10:1


Figure 7-6
Duct system with 5
o
bend, showing the pattern at the turn-over temperature of 50
o
C
Lateral scale, 10:1

The pattern at the turn-over temperature of 50
o
C is shown in Figure 7-6. In the plan view a
pattern of three waves can be seen with an overall length of 21.6m. The absorption by the pattern
of thermal strain limits the force to a peak of 6,350N at B and 8,650N at A. It is of note that the
pattern initiated by the kink at 50
o
C, Figure 6-4, has a closely similar shape and 17.6m length.
The patterns are both comprised of the same wave formations with a half formed helical loop
growing out of the sine waves. Thus the deformation patterns formed by the bend and kink are
identical.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-23
Table 7-16
Duct and Pipe Systems: Comparison of peak force and turn-over temperatures for pattern
initiation by either a cable kink, a 5
o
bend or a 90
o
bend
Duct System Pipe System Pattern
Initiator
Position in
Route
Peak Force
[N]
Turn-over
Temperature [
o
C]
Peak Force
[N]
Turn-over
Temperature [
o
C]
A 8,942 42.5 9,837 (1) 45 (1) Cable Kink
B 8,937 42.5 9,804 (1) 45 (1)
A 8,650 42.5 10,676 (2,3) 47.5 (2,3) 5
o
Bend
35D radius B 6,350 50 8,451 (1) 45 (1)
A 7,520 70 10,458 (3) 60 (3) 90
o
Bend
35D radius B 5,377 72.5 7,322 (3) 55 (3)
Legend: (1) is the cable core number and position in the pipe system that has the highest
force. Core 1 is at the apex of the trefoil group and cores number 2 and 3 proceed in a
clockwise direction viewing from A to B.
Table 7-16 for duct and pipe systems, compares the peak forces and turn-over temperatures for
patterns initiated by the 5
o
bend with those of the cable kink and he reference model 90
o
bend.
The table is arranged in descending order of the force at B.
In making the comparison it is of note that the routes are not identical. Although the routes are
each 200m long they behave in different ways. The route with the kink is straight , but the kink is
at the center of the span and effectively divides it into two routes of 100m. In chapter 6 it is
shown that the frictional force front had reached the end of the model at A 100m away and had
relieved the available thermal strain at the temperature of 42.5-45
o
C. In the 5
o
bend model the
force front will continue to travel a further 100m before it reaches A, releasing additional thermal
strain into the deformation pattern. This alters the relationship between the force in the pattern
region at B with the force at the far end of the span at A. The values at position B are more
appropriate in comparing the mechanism of pattern formation.
The following points explain the differences in pattern inception and turn-over temperatures
between the 5
o
bend and the cable kink models. The duct and pipe models can be seen to behave
in similar ways with the exception that the axial force at the turn-over temperature is higher,
reflecting some restriction in lateral space for cable curvature to form. The 90
o
bend model is
described in the next section.
In the 5
o
hockey stick model the minimum radius of cable curvature occurs at the pre-set
bend radius of 2.9m, this is half the 6.1m value occurring naturally in the kink pattern at
42.5
o
C. If the axial forces are the same, the bending moment in the 5
o
model will be twice as
high and will form a pattern at a proportionally lower temperature. Thus a small bend in a
pipe route is shown to be more efficient than a cable kink in initiating deformation patterns.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-24
The 5
o
bend can absorb more than 6.5mm of cable expansion, this is equivalent to a
temperature rise of 2
o
C. Thus the bend delays the turn-over temperature at which the force
reaches a sufficient value to initiate the incipient thermomechanical pattern.
The peak forces occur at a higher temperature than the temperature of formation of the
incipient pattern. The turn-over temperature occurs when the rate of growth of the wave
amplitude a) rapidly accelerates to absorb thermal strain and b) changes the pattern from a
state of axial stiffness to axial flexibility. In some respects the pattern can be said to have
elastically buckled. Overall: a) the absorption of the 5
o
bend delays the turn-over temperature
from 42.5
o
C to 50
o
C and b) the small radius causes the critical bending moment to be
reached at a lower level of axial force.
Formation of Thermomechanical Patterns in the Reference Duct Model
Figure 7-7 shows that the turn-over temperature at which the force peaks in the reference model
is 72.5

C. The bend, Figure 7-1, has an angle of 90


o
, a radius of 2.9m and an arc length of 4.6m.
The cable absorption capacity of a 90
o
bend with a 75mm radial clearance is 236mm, which is
equivalent to the thermal strain from a 67
o
C temperature rise. However the force peaked in the
model at a rise of 57.5
o
C, indicating that 86% of the thermal strain from the 206m cable length
had been released into the bend and pattern.
The following features are present in Figure 7-7:
Position 1: The cable expands comparatively easily into the bend up to 30
o
C requiring a
minimal axial thrust of 192N.
Position 2: At 50
o
C and 1,108N, the cable no longer moves readily into the bend. The rate of
rise of now reaches a maximum for this model at 221N
o
C
-1
. This is 65% of the maximum
thrust of 338 N
0
C
-1
for a rigid straight cable. (Calculated from EA the reference cable axial
stiffness value 19.9MN at 60
o
C, Table 7-13, and the coefficient of thermal expansion of 17 x
10
-6

0
C
-1
). It is shown that some of the thermal strain is being absorbed a) into the bend and
b) into an incipient deformation pattern formed at the entrance to the bend, B.
Position 3: At 65
o
C, 4,634N, a single pronounced wave has formed at B and has absorbed
strain, Figure 7-8.
Position 4: The turn-over temperature occurs at 72.5
o
C and limits the peak force to 5,377N.
The pattern has grown sufficiently in amplitude and number of waves to be able to absorb
locked-in strain from the adjacent cable at a higher rate than generated, Figure 7-9.
Position 5, the force falls to a minimum of 4,190N at 82.5
o
C.
Position 6, the operating temperature of the cable at 90
o
C at which the force has risen to
4,920N, although this is below the turn-over force of 5,377N. Figure 7-10 is the plan view of
the 43.5m pattern length. Figure 7-11 is a foreshortened view looking from B towards A,
showing that two helical loops and six sine waves are present.
Positions 7 (95
o
C; 5,250N) and 8 (115
o
C; 5210N), more helical loops form and limit the
force.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-25
Position 9, the end of the study at 125C. The force has risen to 5,440N, which is now above
the first turn-over peak of 5,377N at 72.5
o
C, position 4.

Figure 7-7
Reference duct system: force-temperature characteristic at position B




Figure 7-8
Reference duct system: incipient wave at 65
o
C
Plan view, lateral scale, 5:1
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-26

Figure 7-9
Reference duct system: pattern at turn-over temperature of 72.5
o
C
Plan view, lateral scale, 5:1

Figure 7-10
Reference duct system: pattern at operating temperature of 90
o
C
Plan view, lateral scale, 5:1
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-27

Figure 7-11
Reference duct system: pattern at 90
o
C showing helical loops and sine waves
Foreshortened projection view, lateral scale, 5:1
Comparison of 5
o
and 90
o
Bend Models
Table 7-16 for duct and pipe systems compares the peak force and turn-over temperatures at
positions A and B for the 5
o
and the 90
o
bend models. The models have a) identical cable and
duct/pipe geometry and b) identical 35D bend radii.
The increase in the bend angle has increased the turn-over temperature at position B from 50
o
C
to 72.5
o
C for the duct system and from 45
o
C to 55
o
C for the pipe system. This results from a
theoretical increase of 111mm in the capacity of the bend to absorb thermal strain and is
equivalent to a temperature rise of 33
o
C. However this is significantly higher than the rise that
occurred for the duct system of 22.5
o
C and the pipe system of 10
o
C. (Note: the peak forces and
turn-over temperatures are different in each of the three cores in a pipe system making the
comparison less specific than with the duct system).
The increase in bend angle has resulted in a 15% drop in the peak axial force from 6,350 to
5,377N for the duct system and in a 13% drop from 8,451 to 7,322N for the pipe system. It
would have been expected that had the maximum curvature of the cable in the deformation
patterns remained the same, then the peak axial force would not have changed. The explanation
is that the bending stiffness properties of the cable EI are temperature dependent. As given in
Table 7-12 for the reference cable, the value reduces by 55% from 2,683 Nm
2
at 20
o
C to 1,189
Nm
2
at 60
o
C. Thus in the temperature rise from 50
o
C to 72.5
o
C the stiffness of the cable can be
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-28
calculated to have reduced by 14%. This reduces both the bending radius and the axial force to
achieve the critical bending moment by 14% and is in agreement with the model output.
At position A the average peak force in the four models is 36% higher than in the pattern region
at B. This is explained by the difference in frictional constraint along the 200m duct/pipe span.
Formation of Thermomechanical Patterns in the Reference Pipe Model
The pipe model contains three cables. The mechanism of pattern formation at the entrance to the
bend is the same as that for the duct system.
A significant difference is that the length of the pattern region is longer in a pipe system, because
a) helical loops cannot form and b) the amplitude of the sine waves is restricted. At 90
o
C, the
pattern length in the pipe is 130m, which is three times longer than the pattern in the duct of
43.5m. Figure 7-12 is a plan view at 90
o
C of the 25m length of cable adjacent to the bend.
Within the pattern the average wave length is 4.4m.
Figure 7-13 is a foreshortened projection view looking towards B at the entrance to the bend.
The three cables in the bend can be seen to have been pressed into a vertical flat alignment by the
axial thrust of the cable in the 200m length of straight pipe. They experience significant sidewall
force at this position.


Figure 7-12
Reference pipe system: patterns at 90
o
C
Plan view, scale 1:1
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-29

Figure 7-13
Reference pipe system: view of patterns adjacent to bend at 125
o
C
Foreshortened projection view, scale: 1:1
Method of Analysis of the Sensitivity Study Results
Computational runs and analysis were undertaken on duct and pipe models with 12 categories of
parameters being varied. Load cycling was performed on 4 models with different angles of slope.
In total 131 models were constructed and analyzed.
The objective of the analysis is to rank the variables by the order of their effect on the limiting
parameters identified in Chapter 4. These being:
Peak axial force in the cable
Temperature at peak turn-over force
Numbers and types of thermomechanical waves present
Cable bending radius in the straight pipe section
Cable bending radius in the pre-formed pipe bend
Sheath strain in the 200m straight section
Maximum sidewall force in the pipe bend
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-30

Figure 7-14
Diagrammatic force-temperature plot used to characterize sensitivity study output
The force-temperature plot shown diagrammatically in Figure 7-14 is representative of the
behavior of all the models. Temperature rise is the key driver in producing thermal strain and the
resulting cable axial force is the prime mover in determining cable thermomechanical behavior.
To reduce the volume of data to be analysed, a) the force-temperature characteristic abstracted
from the model was limited to A, B and C, Figure 7-1, as these were the positions where a joint
manhole could be located and b) the force was abstracted from five points on the force-
temperature characteristic shown in Figure 7-14 at a, b, c, d, and e.
The sensitivity of thermomechanical behavior to each of the 12 parameters was obtained by:
Plotting axial force against the percentage variation in the parameter (typically 5 variations)
Interpolating the change in axial force resulting from a 50% variation in the parameter
Ranking the parameters in order of their effect on axial force for the 50% change
The sensitivity results given below are based on the peak force at the turn-over temperature. In
the majority of cases the turn-over temperature was below the operating temperature of 90
o
C.
One exception being when cables were installed in tight fitting ducts or pipes.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-31
Sensitivity Study Results
Duct System Sensitivity Results
Figure 7-19 is the bar chart of parameter ranking of maximum axial force at position A.
Figure 7-20 is the accompanying bar chart of turn-over temperature.
Figure 7-32 is the bar chart of parameter ranking of maximum axial force at position B.
Figure 7-33 is the accompanying bar chart of turn-over temperature.
Back-up Information:
Figure 7-21 to Figure 7-31 are graphs of axial force - % parameter variation at position A.
Figure 7-34 to Figure 7-44 are graphs of axial force - % parameter variation at position B:
Figure 7-45 to Figure 7-56 are bar charts of axial force, turn-over temperature, number of
waves, bending radius, sheath strain and sidewall force.
Pipe System Sensitivity Results
Figure 7-57 is the bar chart of parameter ranking of maximum axial force at position A for
Cable 1.
Figure 7-58 is the bar chart of parameter ranking of maximum axial force at position A for
Cable 2 or 3.
Figure 7-71 is the bar chart of parameter ranking of maximum axial force at position A for
Cable 3. Figure 7-72
Back-up Information:
Figure 7-59 to Figure 7-70 are graphs of force - % parameter variation at position A.
Figure 7-73 to Figure 7-84 are graphs of force - % parameter variation at position B.
Figure 7-85 to Figure 7-98 are bar charts of force, turn-over temperature, bending radius,
sheath strain and sidewall force.
Sensitivity Study Results for Cyclic Loading
Thirty load cycles comprising heating from 15
o
C to 90
o
C and then cooling back to 15
o
C were
applied to the reference model with slope angles of +7.5
o
, +3
o
, -3
o
and -7.5
o
.
The cable in the models with the +3
o
and -3
o
slopes changed position during the load cycling.
This is because the slope angles are below the critical angle of slip. As described in Chapter 2,
Basic Theory, sliding will be assisted or opposed depending on the direction of the
thermomechanical force. Thus the cable will ratchet up and down the slope during heating and
cooling, but with a different degree of movement. At the end of the first load cycle the cable will
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-32
neither have returned to the same position nor have regained the original neutral force
distribution:
Figure 7-99 is the variation of axial force in the conductor during each of the 30 load cycles at
the entrance to the bend, position B, for the +3
o
slope model. Position B is at the top of the slope.
The figure shows that:
The amplitudes of the peak force varies with a random sequence
The force at 90
o
C varies by 21% from a maximum value in load cycle 1 of 4,830N to a
minimum of 3,630N in load cycle 14
The maximum force is only reached a second time at load cycle 26
At the end of the first cycle the cable at B does not return to zero force, but stays slightly in
compression.
The force only returns to zero for a second time at the end load cycle 4.
Figure 7-100 compares the thermomechanical patterns at 90
o
C in load cycle 1 and load cycle 30:
The thermomechanical patterns have the same overall length
The pattern in load cycle 1 has formed two helical loops close to the bend, whereas the
pattern in load cycle 30 has formed one helical loop
The pattern difference is accompanied by a force reduction between load cycle 1 and 30
of 7%
The models with the -3
o
slope behaved in the same way with variations in the maximum force
and a difference in shape between the first and last load cycles. Compared to the +3
o
slope the
conductor axial force reached a 9% higher level of 5,250N, which is explained by position B
being at the bottom of the slope and experiencing the additional gravitational cable force.
Interestingly upon cooling, the cable returned to zero axial force, whereas it would have been
expected to have retained some compressive force.
The models with the +7.5
o
and -7.5
o
slopes showed no change in the force distribution or in the
pattern shapes during the 30 load cycles. The before and after force graphs and deformation
patterns were superimposed and were identical. This was attributed to: the slope angles being
above the critical angle of slip of 5.7
o
. Thus during heating and cooling the cable was always
able to a) slide, b) redistribute the thermal strain and force and c) return to its starting condition
at the end of each cycle.
Ranking of Sensitivity Study Results
Table 7-17 to Table 7-24 rank the parameters by a) order of peak axial force and b) by type of
effect. Load cycling is neither a change in geometry nor in a physical property and so was not
included in the parameter ranking.
The sign notation followed in the ranking tables is:
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-33
A positive force % is when an increase in axial force results from a 50% increase in the
parameter.
A negative force % is when a decrease in axial force results from a 50% increase in the
parameter.
The duct clearance and the bend angle are the only parameters that resulted in a decrease in
force.
The legend for the type of thermomechanical effect is:
Classification Symbol Variation in the parameter results in:
Group I Force
Temp
Increase in peak force
Decrease in turn-over temperature
Group II Force
Temp
Increases in peak force
Increase in turn-over temperature
- Force



Temp
Variation in peak force
Constant turn-over temperature
- Force
Temp


Constant peak force
Variation in turn-over temperature
Two of the parameters have inappropriate reference values:
The coefficient of friction reference value of =0.1 is varied between 0.05 and 0.15. This
results in the variation in axial force being artificially low and the importance of the
parameter passing unnoticed in the ranking list. In compensation a higher reference value has
been selected with =0.3 and has been varied between 0.15 and 0.45. This has raised the
ranking of friction to close to the top of the list.
The bend angle reference value of 90
o
was given a 50% negative variation down to 45
o
, as an
angle of greater than 90
o
was considered to be infrequent and misleading. Thus the effect of a
+50% increase is not available. It was observed that the relationship between force and angle
is linear , thus the total variation in force was doubled to represent a notional 50% variation
between 45
o
and 135
o
.
The right hand column entitled curve shape indicates whether the variation in force is linear, or
is sufficiently so to be interpolated for the purpose of designing a pipe or duct system.
Common Rankings in Pipe and Duct Systems
The clearance between the cable and duct/pipe has the largest overall effect on the axial
force. It ranks first in the pipe system at positions A and B and in the duct system at B. It
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-34
ranks second in the duct system at A. Pipe clearance is the parameter that provides the
geometry for the thermomechanical patterns to form and to absorb thermal strain from the
sections of straight cable.
The torsional stiffness GJ of the cable is low and ranks at the bottom of all the lists as having
no effect.
The bend radius has a small effect and ranks towards the bottom of the lists, more so in the
duct system than in the pipe system. This is in line with the theory that the bend radius has no
effect in the absorption of thermal strain. Cable absorption is the product of bend angle and
pipe-cable radial clearance. The reason for the influence of bend radius in the pipe system is
attributed to the smaller bending radii restricting the space available for the cables to change
from sinusoidal trefoil geometry to vertical flat formation at the entrance to the bend. A
comparatively long transition zone can be seen in the change of pattern shapes between the
sinusoidal waves and the vertical flat formation, Figure 7-12. Another possible reason is the
observation that a smaller bend radius forma an incipient deformation pattern at a reduced
level of axial force and temperature.
Some parameters rank at the opposite ends of the list, depending on the position in the route,
A or B.
A is the position most remote from the patterns and so is influenced strongly by
external influences. When the slope angle is positive, A is at the bottom of the slope
and experiences an increased gravitational compressive loading which is
superimposed upon the thermomechanical compressive loading. When the angle is
negative it experiences tensile loading and this reduces the thermomechanical
loading. Thus the slope angle ranks at the top of the list. It should be noted that the
variation in slope angles studied, of 30
o
, is predominantly in the region where the
cable will overcome frictional constraint and slide. Slopes in pipe routes are more
usually in the range of 5
o
, in which the slope angle has a small effect on the axial
force, but the cable movement is more complex, as described in the previous section
on load cycling.
B is the position of initiation of the thermomechanical patterns. The patterns are
highly efficient in absorbing strain, in reducing the force and keeping it low. This has
the effect of isolating B from external influences, such as the slope angle. Thus for B,
the slope angle ranks at the bottom of the list.
Some of the parameters can be categorized depending on their affect on peak force and turn-
over temperature. The decrease of the axial and bending stiffness of the XLPE cable with
increasing temperature is an important variable. This categorization is common to each of the
four cases, although the ranking in terms of force varies between route positions A and B.
Group I. The peak axial force at which the patterns form increases, whilst the turn-
over temperature decreases. Group I parameters primarily influence the cable
longitudinal behavior necessary to reach achieve the peak force and turn-over
temperature.

The coefficient of thermal expansion , the axial stiffness EA and the angle of the
pipe/duct bend were always in this group. The parameters EA and increase the rate
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-35
of rise of axial force with temperature and so reduce the turn-over temperature at
which the necessary peak force is reached. They are the primary drivers in producing
a) maximum axial force with temperature, EA F = and b) maximum thermal
strain with temperature, L=L.

Length L influences Group I in a more complex secondary interaction. The variation
of turn-over temperature with length is found to exhibit a non-linear U shaped
characteristic. When a short span is lengthened the force also increases, but turn-over
temperature initially falls. For longer span lengths both the force and the temperature
increase; cable length then behaves as a Group II parameter.

The angle of the pipe bend has the reverse effect. Larger bend angles accommodate
more thermal strain and initially reduce the rate of rise of axial force. This increases
the turn-over temperature at which cable patterns limit axial force to the peak value.
The cable bending stiffness EI is temperature dependent, falling to half its ambient
value at 60
o
C. This permits the axial force necessary to produce the critical bending
moment to fall as temperature increases.
Group II: The peak axial force and the turn-over temperature both increase or both
decrease. Group II parameters primarily influence the cable lateral behavior necessary
to absorb thermal strain in the patterns.

An increase in the bending stiffness EI increases the cable radius and reduces the
moment arm. A higher axial force is required to produce the critical bending moment
and bend the cable laterally to form the pattern. A higher temperature is needed to
generate the force.

Increases in either, or both, the cable weight and the coefficient of friction increase
the axial force necessary to move the cable laterally up the pipe walls to increase the
amplitude of the deformation pattern. At the critical ratio of amplitude to wavelength,
the wave becomes capable of absorbing strain at an accelerating rate and then
forming a train of sinusoidal or helical waves.

The pipe clearance produces a non-linear reduction in both the peak axial force and in
the turn-over temperature. Starting with a close fitting pipe, a small increase
disproportionally reduces the peak axial force. In a wider space the lateral cable
deformation can achieve the critical ratio of amplitude to wavelength at which
thermal strain is absorbed at an increasing rate. For wide pipe-cable spacings the
force and temperature fall at a decreasing rate until they are limited by the increases
in cable bending and axial stiffness that occurs at lower temperatures.
Outside the common aspects of ranking and grouping listed above, the relative positions of the 9-
10 main parameters change significantly according to a) the position in the route and b) whether
it is a pipe or duct system. This, together with the interactions identified, confirms that cable
thermomechanical behavior in pipes and ducts is a multi-variable problem and explains the past
inability to derive a simple mathematical equation for engineering design.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-36
Duct System: Ranking of Parameters at Position A
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
7-17 in descending order of total percentage change in force.

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-37
Table 7-18 ranks each parameter numerically and by the type of interaction between the turn-
over temperature and the peak axial force.
The reference model values for Table 7-17 are:
Peak axial force: 7,520N (20% of straight cable maximum)
Turn-over temperature: 70
o
C

Table 7-17
Duct at position A: Ranking list of parameters by peak force at turn-over temperature
-50% +50% Parameter Total
Change
[%]
% Force [N] % Force [N]
Curve Shape
Slope angle 106 -53 3,555 53 11,478 non-linear
Clearance 65 +47 11,054 -18 6,197 non-linear
(Friction changed to =0.3) (58) (-29) - (+29) - linear
Length 49 -26 5,574 23 9,257 linear
Weight 48 -24 5,709 24 9,290 linear
Bend angle 48 +24 9,308 - - linear
Friction 36 -11 6,705 25 9,432 linear
Expansion 26 -11 6,696 15 8,646 linear
EI 26 -15 6,393 11 8,343 non-linear
EA 11 -5 7,115 6 7,592 non-linear*
Bend radius 0 0 7,487 0 7,505 flat
GJ 0 0 7,520 0 7,523 flat
*can be approximated as linear over part of the range
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-38
Table 7-18
Duct System at A: Ranking by type of effect
Force
Temp
Force
Temp
Parameter Total Change
[%]
Group I Group II
Force


Temp
Force
Temp


Slope angle 106 X
Clearance 65 X*
(Friction =0.3) (58) X
Length 49 X*
Weight 48 X
Bend angle 48 X
Friction ref =0.1 36 X
Expansion 26 X
EI 26 X
EA 11 X
Bend radius 0 X
GJ 0 - - - -
*Non-linear U shaped temperature characteristic
Duct System: Ranking of Parameters at Position B
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
7-19 in descending order of total percentage change in force.
Table 7-20 ranks each parameter numerically and by the type of interaction between the turn-
over temperature and the peak axial force.
The reference model values for Table 7-19 are:
Peak axial force: 5,372N (14% of straight cable maximum)
Turn-over temperature: 72.5
o
C
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-39
Table 7-19
Duct at position B: Ranking list of parameters by peak force at turn-over temperature
-50% +50% Parameter Total Change
[%] [%] Force [N] [%] Force [N]
Curve Shape
Clearance 153 +131 12,411 -21 4,252 non-linear
Expansion 43 -22 4,209 +23 6,587 linear
EI 36 -23 4,119 +13 6,081 non-linear
Length 23 -19 4,364 +4 5,600 non-linear
Weight 22 -11 4,769 +11 5,979 non-linear*
(Friction changed to
=0.3)
(22) (-11) - (+11) - linear
EA 17 -21 4,230 -3 5,214 non-linear
Bend Angle 16 +8 5,798 - - linear
Friction 9 -4 5,172 +5 5,649 linear
Slope angle 7 -2 5,247 +5 5,624 non-linear*
Bend Radius 0 -3 5,208 -3 5,198 flat
GJ 0 0 5,363 0 5,379 flat
* can be approximated as linear over part of the range
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-40
Table 7-20
Duct System at B: Ranking by type of effect
Force
Temp
Force
Temp
Parameter Total Change
[%]
Group I Group II
Force


Temp
Force
Temp


Clearance 153 X*
Expansion 43 X
EI 36 X
Length 23 X*
Weight 22 X
(Friction =0.3) (22) X
EA 17 X
Bend angle 16 X
Friction ref =0.1 9 X
Slope angle 7 X
Bend radius 0 X
GJ 0 - - - -
*Non-linear U shaped temperature characteristic
Pipe System: Ranking of Parameters at Position A
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
7-21 in descending order of total percentage change in force.
Table 7-22 ranks each parameter numerically and by the type of interaction between the turn-
over temperature and the peak axial force.
The reference values for Table 7-21 are:
Peak axial force: 10,000N (27% of straight cable maximum)
(axial force in top cable 7,960N and bottom cables 10,600N and 10,000N)
Turn-over temperature: 57.5
o
C
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-41
Table 7-21
Pipe cable 3 at position A: Ranking list of parameters by peak force at turnover
temperature
-50% +50% Parameter Total
Change
[%]
[%] Force
[N]
[%] Force
[N]
Curve
Shape
Clearance 224 +176 27,574 -48 5,242 non-linear
Slope angle 92 -63 3,237 +29 12,858 non-linear*
(Friction pipe-cable,
changed to =0.3)
(58) (-31) - (+27) - non-linear
Expansion 46 -19 8,114 +27 12,730 linear
Weight 46 -25 7,468 +21 12,100 linear
Length 40 -13 8,732 +27 12,741 non-linear*
Friction, pipe-cable 32 -8 9,232 +24 12,375 non-linear
EA 31 -11 8,867 +20 12,070 non-linear
Bend radius 20 -14 8,600 +6 10,600 non-linear
Bend angle 14 -7 10,711 - - non-linear
nearly flat
Friction, cable-cable
(changed to =0.3)
11
(6%)
-6
(-3)
9,402
-
+5
(+3)
10,466
-
non-linear*
EI 10 -7 9,335 -3 9650 non-linear*
GJ 0 +4 10,434 +4 10,434 flat
* can be approximated as linear over part of the range

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-42
Table 7-22
Pipe System at A: Ranking by type of effect
Force
Temp
Force
Temp
Parameter Total Change
[%]
Group I Group II
Force


Temp
Force
Temp

Clearance 224 X*
Slope angle 92 X
(Friction pipe-cable, =0.3) (58) X
Expansion 46 X
Weight 46 X
Length 40 X*
Friction, pipe-cable ref =0.1 32 X
EA 31 X
Bend radius 20 X
Bend angle 14 X
(Friction cable-cable, =0.3) 11
(6%)
X
EI 10 X
GJ 0 - - - -
*Non-linear U shaped temperature characteristic
Pipe System: Ranking of Parameters at Position B
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
7-23 in descending order of total percentage change in force.
Table 7-24 ranks each parameter numerically and by the type of interaction between turn-over
temperature and peak axial force.
The reference model values for Table 7-23 are:
Peak axial force: 7,322N (20% of straight cable maximum)
(axial force in top cable 5,628N and bottom cables 7,035N and 7,322N)
Turn-over temperature: 55
o
C

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-43
Table 7-23
Pipe cable 3 at position B: Ranking list of parameters by peak force
-50% +50% Parameter Total
Change
[%]
[%] Force
[N]
[%] Force
[N]
Curve
Shape
Clearance 281 +253 25,860 -28 5,303 linear
Expansion 49 -22 5,750 +27 9,282 non-linear
EA 42 -21 5,789 +21 8,855 non-linear*
(Friction cable to cable
changed to =0.3)
(32) (-16) - (+16) - non-linear*
EI 25 -10 6,615 +15 8,448 non-linear*
Weight 24 -12 6,444 +12 8,179 non-linear*
Friction, cable-cable 14 -1 7,273 +13 8,238 non-linear*
Slope Angle 10 +3 7,570 -7 6,816 non-linear*
(Friction pipe to cable changed
to =0.3)
(20) (-10) - (+10) - linear
Friction, pipe-cable 5 -3 7,104 +2 7,496 linear
Length 5 -4 7,043 +1 7,422 non-linear
Radius 5 -6 6,911 -1 7,243 non-linear*
Bend Angle 4 -2 7,200 - - non-linear
nearly flat
GJ 1 +1 7,424 +2 7,442 flat
* can be approximated as linear over part of the range


EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-44
Table 7-24 Pipe System at B: Ranking by type of effect
Parameter Total Change
[%]
Force
Temp
Force
Temp
Force


Temp
Force
Temp

Group I Group II
Clearance 281 X*
Expansion 49 X
EA 42 X
(Friction cable-cable, =0.3) (32) X
EI 25 X
Weight 24 X
Friction cable-cable, =0.1 14 X
Slope angle 10 X
(Friction pipe-cable, =0.3) (20) X
Friction, pipe-cable ref =0.1 5 X
Length 5 X*
Bend radius 5 X
Bend angle 4 X
GJ 1 - - - -
*Non-linear U shaped temperature characteristic
Duct and Pipe System: Comparison of Thermomechanical Wave Patterns
Table 7-25 gives the extremes of the wave pattern dimensions in order of peak ascending axial
force. It is shown that:
As the peak axial force increases, a) the number of waves increases, b) the percentage of
helices to sinusoids increases, c) the length of the pattern region increases and d) the
wavelength decreases. The exceptions to the trend are the long route, the coefficient of
expansion (COF) and the bending stiffness.
The lowest axial force occurs with the largest amplitude duct clearance of 250mm. Sinusoid
patterns with a large ratio of amplitude to wavelength can easily form and so the patterns
absorb the thermal strain whilst exhibiting low longitudinal stiffness, thereby reducing the
force. The absorption of strain by the sinusoids is a lower energy minimum energy condition
than storing the energy in helices.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-45
The sine waves are restricted to small amplitude when the duct clearance is decreased from
the reference value of 150mm, through the practical installation clearance of 50mm to the
minimum clearance of 10mm. It is possible for sine waves to grow in amplitude and to pass
360
o
around the circumference, but he minimum energy condition is now for helices to form.
Helices can absorb more thermal strain. High axial forces result as there is now insufficient
storage capacity in the duct, even allowing for the capacity of the bend. The high force
causes the patterns to a) occupy a high percentage of the duct length; greater than 86%, b)
predominantly form helices and c) form the highest number of waves with the shortest
wavelength. The patterns in a tight fitting pipe form at high temperature. The patterns would
be difficult to observe in a tight fitting duct system, because the amplitude of the patterns is
small.
The routes with 100m, 200m and 1000m lengths contain cable with exactly the same
properties. Increasing the length has a marked effect on a) increasing the axial force, b)
increasing the length of the pattern, c) increasing the percentage of helices and d) reducing
the wavelength. The key parameter is the ratio of the storage capacity of the bend to the
length of the straight section. In the longer route there is a higher percentage of locked-in
thermal strain to be released into the pattern and so longer patterns with shorter wavelengths
are formed to absorb it. For a given coefficient of thermal expansion the length of thermally
constrained cable also increases.
The presence of a high coefficient of friction of 1, limits longitudinal and axial movement,
producing a) high force, b) the shortest pattern length of 35m, c) short wavelengths and d) a
high percentage of helices.
A high bending stiffness EI has a marked effect on producing a) a high axial force, b) the
longest wavelengths and c) a reduced percentage of helices. The ratio of amplitude to
wavelength is reduced, which reduces the ability to accommodate thermal strain, thereby
increasing the axial thrust.

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-46
Table 7-25
Duct System: Examples of thermomechanical wave patterns
Wave Details
Number at 125
o
C Wavelength
[m]
Model Comment Axial Peak
Force at
Turn-over
Temperature,
Position B
[N]
All Waves Helices,
N
o
& %
90
o
C 125
o
C
Length at
125
o
C
[m]
Largest
clearance,
250mm
smallest
number of
waves
3,650 9 0
0%
5.7 5.4 53
(27%)
Shortest route
100m
- 4390 9 0
0%
4.2 5.1 47.2
(47%)
Reference
route, 200m
- 5377 20 6
30%
4.6 4.8 91.5
(46%)
Longest
route, 1000m
- 5890 46 41
89%
3.6 3.6 165
(17%)
Highest COF,
=1
shortest O/A
length
9770 8 5
63%
3.9 3.9 35
(18%)
Highest
bending
stiffness, EI
longest
wavelength
15914 14 6
43%
8.7 7.6 108
(54%)
Practical duct
clearance
50mm
- 19917 58 52
90%
2.9 2.9 171
(86%)
Smallest duct
clearance,
10mm
shortest
wavelength
35897 96 96
100%
2.63 2.63 197
(99%)
Pipe System: Thermomechanical Wave Patterns
Table 7-26 gives examples of wave pattern dimensions ranked in order of ascending peak axial
force. In comparison with Table 7-25 for the duct system it is shown that:
The pipe system containing three cables produces thermomechanical waves with closely
similar dimensions to those in a duct system of the same dimensions containing one cable.
Thus the presence of two other cables in the pipe does not significantly restrict the formation
of patterns, or the reduction in axial force that results. The differences are listed below.
The pipe system cannot form helical loops. All of the patterns are cylindrical sinusoids,
which are not so efficient as absorbing thermal strain.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-47
The axial force levels are slightly higher in the pipe system, resulting in slightly longer
pattern lengths and slightly shorter wavelengths.
The axial force in the tightest fitting duct (10mm clearance) of 35,897N is significantly
higher than the 28,498N value in the tightest fitting pipe (a 5mm installation clearance on the
trefoil group). However this comparison is artificial. The true clearance to one individual
cable in the pipe is significantly larger at 100mm, thereby providing space for the formation
of thermomechanical patterns, which absorb strain and reduce force.
Table 7-26
Pipe System: Examples of thermomechanical wave patterns in cable 2 (bottom)
Wave Details
Number at 125
o
C Wavelength
[m]
Model Comment Axial Peak
Force at Turn-
over, Position
B
[N]
All
Waves
Helices,
N
o
& %
90
o
C 125
o
C
Length at
125
o
C
[m]
Largest
clearance,
250mm
smallest
number of
waves and
shortest O/A
length
2,870 13 0 5.3 5.3 62
(31%)
Shortest route
100m
- 3516 14 0 4.8 5.1 68
(68%)
Highest cable
COF, =1
5491 17 0 3.7 3.7 66
(33%)
Reference
route, 200m
- 5,509 28 0 4.2 4.7 117
(58%)
Longest
route, 1000m
- 8,543 61 0 3.5 3.7 196
(20%)
Practical pipe
clearance
115mm(group
clearance
20mm)
- 12,258 43 0 3.13 3.3 145.6
(73%)
Highest pipe
COF, =1
shortest O/A
length
16,884 16 0 2.4 4.9 8
(42%)
Highest
bending
stiffness, EI
longest
wavelength
18,129 20 0 8.5 7.4 146
(73%)
Smallest duct
clearance,
100mm,
(group
clearance
5mm)
shortest
wavelength
28,498 53 0 0 2.1 120
(60%)
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-48
Parameters that Limit the Life of Duct and Pipe Systems
Chapter 4 listed the values of the design parameters that can limit the service life of a
transmission class XLPE cable. In this section a comparison is made with the extreme
magnitudes of the limiting parameters, that were present in the models. The limiting values
were abstracted for the duct system from Figure 7-15 to Figure 7-56 and for the pipe system
from Figure 7-57 to Figure 7-98 and are given below:
Axial Force
Duct System
The highest axial force occurred with the smallest duct clearance of 10mm, because the cable
had insufficient lateral clearance to develop strain absorbing patterns. The second highest force
at 32,574N occurred with a coefficient of friction of 1.0 at position A in the reference model. The
high value of friction isolated end A from the benefits of the thermomechanical pattern at B and
so the cable was unable to release locked-in thermal strain. The benefits of increasing the
clearance to the practical installation clearance of 50mm can be seen in the large reduction in
force.
Pipe System
The same comments apply, however the axial forces are higher than in the duct system for the
clearance of 150mm used in the reference model, because there is slightly less free space and
because helical loops cannot form. In consequence the turn-over temperatures at which the peak
force occurred are lower. A turn-over temperature of 125
o
C is the final temperature step set in
the study and indicates that the force had not been limited and could have risen to a higher value
limited.
The values are listed for the core number in the trefoil formation that exhibited the maximum
force.
Limiting Values
The cable construction is not directly imperiled by the magnitude of compressive axial loads.
The highest compressive loads occur in the tightest fitting pipes in which the cable receives
lateral support from the duct or pipe.
The joints and to a lesser extent the terminations would require to be designed to withstand the
compressive axial forces, particularly in the positions along the route where an unbalanced load
can be expected to exist across the joint. The joints within an in-line casing in a pipe system
would require to be reinforced to prevent the cores from splaying out and buckling.

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-49
Table 7-27
Duct system: Examples of peak axial force levels
Peak Force [N] Turn-Over Temperature [
o
C] Model Comment
A B C A B C
Reference model Ref values 7,520 5,377 5,022 70 72.5 77.5
50mm clearance - 13,078 19,917 18774 75 125 125
10mm clearance maximum 33,042 35,897 33,608 125 125 125

Table 7-28
Pipe System: Examples of peak axial force levels
Peak Force [N] Turn-Over Temperature [
o
C] Model Comment
A B C A B C
Reference model Ref values 10,458
(3)
7,322
(3)
5,512
(3)
60
(3)
55
(3)
75
(3)
115mm clearance - 15,141
(2)
11,814
(2)
15,729
(3)
60
(2)
117.5
(2)
115
(3)
100mm clearance maximum 27,574
(1,2,3)
30973
(1)
34791
(3)
95
(1,2,3)
125
(1)
125
(3)
Legend: (3) refers to core 3 in the trefoil formation in which core 1 is at the top and 2 and
3 are at the bottom.
Sidewall Force in Bend
Duct System
The effect of duct clearance dominates both the axial thrust and the sidewall force in the bend,
producing a x 6.9 range of sidewall force. The use of the practical installation clearance of 50mm
is beneficial, but the force of 6,379Nm
-1
is still of significant magnitude. The bend radius is 2%
smaller than the preformed bend, resulting from the cable rising to the horizontal position at the
entrance and exit to the bend. The choice of a bend radius of larger than 35D would be of benefit
as the force is inversely proportional to the bend radius and a so will further decrease the
sidewall force.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-50
Pipe System
The pipe system requires an optimum pipe clearance to be selected. The large clearance of
250mm allows the cable in the preformed bend to be bent into a 77% tighter radius of 1.64m.
Limiting values
The limiting values for sidewall force recommended in Chapter 4 for the 138kV 1500kcm
reference cable are 19,740 Nm
-1
at 90
o
C and 6,580 Nm
-1
at 105
o
C.
In the duct system these limits are approached when the cable is installed with a tight clearance.
Consideration should be given to selecting, a) a clearance of 50mm or greater and b) a bend of
radius of 35D or greater.
In the pipe system the sidewall forces are generally higher than in the duct system. The sidewall
forces do approach the limiting values for small pipe clearances. Consideration should be given
to selecting, a) a clearance of 115mm or greater and b) a bend of radius of 35D or greater.
The minimum radius in a pre-formed duct bend for a 33-200kV class XLPE cable with a foil
laminate sheath is recommended in Chapter 4 to be 35D (2.86m). The cable in the duct system is
close to this value.
The cable bend radius in the pipe system is significantly less than 35D. Consideration should be
given to a) checking the suitability of the cable construction to withstand a 20D bend radius
under sidewall load or b) depending on the pipe clearance for a particular application, to increase
the minimum bend radius to 45D. Although the bend is in a semi-controlled location, the
particular cable construction would require to be checked for suitability.
Table 7-29
Duct system: Examples of maximum sidewall force in the bend
Bending Radius [m] Sidewall Force [Nm
-1
] Model Comment
90
o
C 125
o
C 90
o
C 125
o
C
250mm clearance minimum 2.86 2.86 1,117 1,274
Reference model Ref values 2.86 2.86 1,512 1,688
50mm clearance - 2.85 2.87 4,649 6,379
10mm clearance maximum 2.86 2.86 7,760 11,630
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-51
Table 7-30
Pipe system; Examples of maximum sidewall forces in the bend
Bending Radius [m] Sidewall Force [Nm
-1
] Model Comment
90
o
C 125
o
C 90
o
C 125
o
C
250mm clearance (3) minimum 1.65 1.64 2,501 2,934
Reference model (3) Ref values 1.95 1.89 3,528 4,243
115mm clearance (1) - 2.2 2.23 5,276 5,539
100mm clearance (3) maximum 2.48 2.48 8,776 12,371
Minimum Bending Radius in the Straight Section
The smallest bending radius is the most severe on the cable construction, increasing both the
sidewall pressure and the sheath strain.
Duct System
The smallest radii of 3.1m in the patterns is numerically acceptable as it is slightly larger than the
value recommended in Chapter 4 of 35D (2.86m) for the preformed bend.
It is interesting to note that the effect of increasing the route length from 100m to 1000m is to
bend the cable into a 2.3 times tighter radius. In the 1000m model the capacity of the bend to
absorb thermal strain becomes a small fraction of the strain generated by the longer route. This
demonstrates the benefit of having more preformed bends in the route a) to reduce the length of
the straight section, b) to reduce axial force, c) to inhibit the formation of short wavelength
patterns and d) to prevent the formation of small bending radii in the cable.
Pipe System
The smallest bending radii of 2.77m is tighter than in the duct system and is 5% less than the
recommended radius of 35D for a preformed bend. The cable construction requires to be checked
for suitability in repeated bending down to 30D
Limiting Value
Unlike the preformed bend in which the cable is pressed immobile against the duct wall, the
cable in the pattern will flex during load cycling, introducing the possibility of fatigue failure.
(Fatigue is dealt with in the following section on sheath strain).

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-52
Table 7-31
Duct System: Examples of minimum cable bending radii in the straight duct
Minimum Cable Bending Radius [m] Model Comment
90
o
C 125
o
C
1000m straight section minimum 3.1 3.1
100m straight section maximum 7.1 7.1

Table 7-32
Pipe System: Examples of minimum cable bending radii in the straight pipe
Minimum Cable Bending Radius [m] Model Comment
90
o
C 125
o
C
1000m straight section (1) minimum 2.77 2.77
100m straight section (2) maximum 6.47 6.47
Sheath Strain in the Straight Section
The highest sheath strains at 90
o
C are produced by the cables with the lowest bending stiffness
EI, the highest weight and the highest friction.
Sheath strain is inversely proportional to the bending radius of the cable. For a given bending
moment, the lowest EI produces the smallest cable radius. The cables in the models with the
smallest duct are prevented from forming small bending radii by the lack of space and so have
lower sheath strain than the reference cable. The cables with the highest weight and highest
friction both develop a) high frictional constraint to both longitudinal and circumferential
movements, b) high axial force, d) short wavelengths and c) small bending radii.
Duct System
The levels of sheath strain are of sufficient magnitude to require the careful design of the cable
and duct system to minimize the level of cyclic strain. Table 7-33 shows that the cables with the
smallest duct clearances have the lowest sheath strain, however it has previously been shown that
this will result in an increases in the sidewall force. Thus an optimum design in required to suit
each application.
Pipe System
The levels of cyclic sheath strain are high compared with a) the limiting values and b) the duct
system, (30-90% higher).
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-53
Limiting Values
The values of sheath strain are significantly higher than the fatigue limits given in Chapter 4 for
extruded metallic sheaths, i.e. lead of 0.024-0.048% and aluminum of 0.25%. However, the
values in Table 7-33 and
Table 7-34 are the absolute values in strain resulting from heating from an ambient temperature
of 15
o
C to 90
o
C, a rise of 75
o
C. Due to the high thermal time constants the cable is unlikely to
experience a daily load cycle temperature rise of more than 10-15
o
C (ie 20% of the 75
o
C rise).
As an approximation, 20% of the values in Table 7-33 and
Table 7-34 can be taken, giving prospective fatigue strains of 0.06% to 0.28%.
In a duct system to reduce the possibility of premature service fatigue failures, a lead sheathed
cable would a) have to be installed in a tight fitting duct and b) have the fatigue life calculated
carefully for the particular cable design, service life and circuit load pattern. A corrugated
extruded aluminum sheathed cable has the prospect of being suitable for each of the duct
applications, but its design and fatigue life should also be checked carefully for the particular
application.
In the majority of the pipe systems studied, the levels of cyclic sheath strain are too high for the
application of extruded lead sheaths. The levels in extruded corrugated aluminum sheaths are
satisfactory for the majority of the applications, although the pipe system should be designed to
minimize the cyclic strain. Experimental evidence on the fatigue performance of foil laminate
sheaths over a range of temperatures is required to judge their fatigue life, this is unlikely to be
as long as that of the corrugated sheath. It is particularly important to select a pipe clearance for
the particular cable size and application to minimize cyclic sheath strain.
Table 7-33
Duct System: Examples of maximum sheath strain in straight section
Sheath Strain [%] Model Comment
90
o
C 125
o
C
10mm clearance - 0.31 0.33
50mm clearance - 0.61 0.59
Reference Ref values 0.63 0.67
Highest COF, =1.0 - 0.61 0.86
Low EI - 0.85 0.83
High Weight, 2.36 x ref, maximum 0.68 0.97

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-54
Table 7-34
Pipe system: Examples of maximum sheath strain in straight section
Sheath Strain [%] Model Comment
90
o
C 125
o
C
100mm clearance (2) - 0.26 1.16
Highest COF pipe-cable, =1.0 - 0.78 (1) 0.82 (2)
115mm clearance (3) - 1.03 1.18
Reference (3) Ref values 1.08 1.19
Low EI (3) - 1.26 1.39
High Weight, 3.08 x ref, (3) maximum 1.28 1.42
Discussion
The key parameter in determining a) the magnitude of axial force that is reached and b) the
presence and type of thermomechanical patterns, is the clearance between the cable and pipe
or duct, expressed as an absolute measurement, D=Dp-Dc, [mm].
Thermal expansion, if unrelieved, produces high axial thrust in the cable. The maximum
value occurs in a tight fitting pipe or duct such that the cable cannot move laterally. The
maximum thrust for the 138kV 1500kcm reference cable at 125
o
C (a rise of 110
o
C) is
F=EA, giving a value of 37,200N (3.7t). Thus both EA the axial stiffness and the
coefficient of thermal expansion are important parameters, but they are not the only ones.
Duct and pipe systems can be designed to be very efficient absorbers of cable thermal
expansion, achieving significantly reduced axial thrust into bends and joints.
A pipe system inherently has a clearance that gives the advantage of reduced force. For
example the 138kV 1500kcm pipe system studied in this project has a clearance on one cable
of 123mm (82% of the 150mm reference case) and on the trefoil group of 28mm (compared
with 55mm in the reference case), this will reduces the thrust to 33% of the maximum value.
Cable in a duct system is usually fitted into the smallest practical pipe and in consequence
will exhibit a higher level of thrust than a pipe system. The 230kV 2500kcm cable route
studied in this project has a clearance of 37mm that reduces the thrust to 67% of the
maximum possible.
Duct Systems
Figure 7-15 is the axial force - % clearance characteristic for the reference duct system.
With the smallest clearance of 10mm (6.7% of 150mm), the resulting thrust developed at
position A, would be 35,500N i.e. 96% of the maximum possible. The side wall force in
bends and thrust on accessories would also be close to maximum values. The reason why the
force is not at the maximum is that a) the bend can absorb a little of the thermal expansion, ie
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-55
8mm, which is 2% of the total thermal strain of 385mm over the 206m length and b)
thermomechanical patterns of very small amplitude form and absorb another 2% strain.
An increase in clearance to a practical installation value of 50mm (33.3% of 150mm) would
very significantly reduce the thrust to 20,000N, 53% of the maximum. A further increase to
90mm clearance would reduce the force to 10,000N, 27% of the maximum. Thereafter
further gains are small. The reason why the 50mm clearance is so effective is a) the bend can
absorb 39mm of thermal strain, (10% of the cable strain), thus dropping the force in
proportion and b) the thermomechanical patterns are sufficiently well developed to absorb
37% of the thermal strain. However the presence of the thermomechanical patterns increases
sheath strain and introduce the risk of fatigue failure of the metallic foil or sheath. The
capabilities of helical loops to absorb thermal strain are significantly greater than those of the
sine wave patterns. Helical loops have the disadvantages of higher strain and cyclic
movement. Thus the selection of a) too wide a pipe clearance and b) too high an axial force
should be avoided. The optimum clearance is interactive with the properties of the particular
cable.
Pipe Systems
Figure 7-16 is the axial force % clearance characteristic for a pipe system. The pipe is an
interference fit with the three cables at 178mm diameter giving a minimum clearance of
95mm (63% of 150mm) to an individual cable.
Comparison of Duct and Pipe Systems
Figure 7-17 shows that the maximum force that can be developed with an installation
clearance of 5mm on the group of three cables in the pipe is significantly lower than that for
the duct system. The axial force for the duct system is 96% of the theoretical maximum
compared to 74% for the pipe system. The reason is that the thermomechanical behavior is
determined by the 100 mm clearance between the individual cable and the pipe. The
clearance gives a) a 79mm absorption of thermal strain in the bend, equivalent to 20% of the
total cable system thermal strain and b) a 6% absorption in the thermomechanical patterns.
Figure 7-18 gives the axial force-% clearance characteristics for each of the three cables in
the pipe system and compares them with that of a single cable in the duct system. Above a
clearance of 130mm, (87% of 150mm) the two systems have very similar force
characteristics. The reference pipe system developed regions of thermomechanical patterns
27% longer than in the duct system.

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-56

Figure 7-15
Duct system: Force - % clearance



Figure 7-16
Pipe system: Force - % clearance

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-57

Figure 7-17
Duct and pipe systems: Force - % clearance



Figure 7-18
Duct and pipe systems: Force - % clearance, all cables plotted

EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-58
Conclusions
The following points summarize the preceding sections. The magnitudes of the numerical effects
can be obtained from the appropriate section.
The mechanism of thermomechanical behavior of cables in pipes and ducts is now
understood. Thermomechanical patterns are initiated by the combination of a) cable
curvature at the entrance to the pre-formed bend in the route and b) the axial thrust arising
from the thermal expansion of the cable in the straight span. The patterns are comprised of
regular waves that grow along the straight span as the cable temperature is increased.
Patterns are prevented from growing around the bend by the presence of high sidewall force.
A small deviation in the route, for example a 5
o
bend angle initiates thermomechanical
patterns. The mechanism is exactly the same as that described in Chapter 6 for pattern
initiation by a small kink in the cable. The 5
o
bend was shown to be a more efficient initiator
as it formed the incipient pattern at a lower temperature. The preformed pipe bend enforces a
lower bending radius than can form naturally in the cable next to a kink. This increases the
bending moment applied to the adjacent straight cable and initiates pattern formation at a
reduced magnitude of axial force. Thereafter, the patterns at each temperature step grow in
identical patterns and shapes. This is an important finding as small pipe bends provide the
means of initiating thermomechanical patterns at a) predetermined positions along the route
and b) lower temperature than those that will form at the normal bends in the route geometry.
Larger bend angles are beneficial in providing the storage capacity to absorb thermal
expansion strain from the cable. They significantly delay the rate of rise of force with
temperature. The storage capacity is proportional to the product of pipe-cable radial
clearance and bend angle [radians].
All of the cases studied for duct and pipe systems follow the same characteristic force-
temperature curve. The relative shape of the curve depends upon a) the pipe-cable diameter
clearance, b) the storage capacity of the bend and c) proximity to the bend. Upon initiating
heating, the cable first expands into the bend. Having occupied the available space, the rate
of rise of force increases to the theoretical maximum of F=EA (where EA is the axial
elastic stiffness, is the coefficient of thermal expansion and is the temperature rise). As
soon as the force rises at the higher rate, the bending moment at the entrance to the pipe bend
initiates very small deformation patterns. Initially the patterns have small amplitude. They
exhibit high longitudinal stiffness and can only absorb a little strain.
At a certain bending moment, the rate of growth of the pattern amplitude accelerates;
a) absorbing thermal strain from the route at a higher rate than it is being produced
and b) exhibiting reduced longitudinal stiffness. This event occurs at the turn-over
temperature at which the axial force in the cable falls rapidly until the time when
residual axial thrust reaches equilibrium with the low axial stiffness of the cable. This
is a rapid event that can be called dynamic buckling. The maximum force reached is
named the peak force.
As the temperature continues to rise, a train of sinusoidal and/or helical waves grow
along the straight span of cable, absorbing its locked-in thermal strain. Thermal strain
is released from the cable for long distances in advance of the pattern (10-100m,
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-59
depending upon friction and force differential). In a horizontal route the distance is
determined by the frictional constraining force of the cable, F= WgL, (where is
the coefficient of friction between the cable and the pipe, Wg is the weight of the
cable per meter and L is the length). The frictional constraint isolates the remote end
of the cable from the low axial force within the thermomechanical pattern. Thus the
axial force at the remote end is significantly higher than the peak force. This is an
important design consideration as the axial force acting at a joint chamber will
depend upon its relative positioning either close to, or remote from a bend.
A sensitivity study has ranked the parameters in the way they affect the performance of a 69-
345kV range of transmission class XLPE cables with a) copper and aluminum conductor
sizes from 1000 to 3000kcm, and b) different combinations of sheath/jacket finishes. The
number of models examined in detail was reduced to 131 by the principle of normalizing the
model geometry on the clearance between the pipe and cable, shown by basic theory to be
the key variable. The method of normalization was validated by modelling a 69kV and a
345kV cable with different outer diameters, but shared pipe-cable clearances and physical
properties. The cables behaved identically.
The parameters of route geometry and cable physical properties were ranked upon their
effect on the magnitude of the axial force. 11 parameters were studied for the duct system
and 12 for the pipe system. The ranking was undertaken at two positions along the route.
Overall, the cable to pipe clearance was shown to be the most important parameter. At the
end of the straight span, remote from the bend, it was found that the angle of slope ranked of
equal importance. The parameter that had no effect was the cable torsional stiffness GJ. The
radius of the preformed bend had minimal effect in the duct system and a slightly stronger
effect in the pipe system.
The parameters were also ranked by their effect on either, increasing force and
reducing temperature (Group I), or by increasing both peak force and turn-over
temperature (Group II).
Group I parameters primarily influence the longitudinal behavior of the cable in
generating the peak axial force needed to form wave patterns. They are the prime
movers that determine the magnitudes of thermal strain and axial force generated in
the straight, undeformed cable. The Group I parameters are axial stiffness (EA) and
coefficient of thermal expansion (). The angle of the pipe/duct bend and the length
of the straight span are also included in this group, although they have more complex
interactions.
Group II parameters primarily influence the lateral formation of the cable patterns.
They are the bending stiffness EI, cable weight and the pipe-cable and cable-cable
coefficients of friction. The pipe clearance, slope angle and radius of the pipe bend
are included in this group, although they have more complex interactions.
Most importantly it was found that a) the order of the parameter rankings changed,
depending upon the particular position in the route, in both duct and pipe system, b)
the parameters interacted and c) the temperature dependence of the axial and bending
stiffnesses resulted in some key interactions. This confirms that the
thermomechanical behavior of duct and pipe systems is a multi-variable problem and
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-60
explains historically why it has not previously been possible to develop a simple
design equation by analytical mathematics.
Load cycling was found to cause longitudinal movement in both uphill and downhill routes
with slopes of less than the critical sliding angle, e.g. 5.7
o
for a coefficient of friction of 0.1.
The ratcheting movement changed the magnitude of the peak axial force at 90
o
C and the
minimum force at 15
o
C in a random way, such that there was no evidence of a trend to
equilibrium after 30 loading cycles. Load cycling had no effect on routes with angles of slope
above the critical angle of slip.
Pipe systems follow the same mechanism of pattern formation and behave in a closely
similar way to duct systems, exhibiting similar wavelengths. Most importantly it is the
clearance between the pipe and the individual cable and not the clearance to the trefoil group
that determines thermomechanical behavior. The three cables are able to form patterns in
synchronism and in so doing make use of almost all of the pipe to cable clearance.
Differences are that the duct system can form two wave patterns, these being cylindrical
sinusoids and helical loops. Pipe systems can only form cylindrical sinusoids due to the
restriction of three cables moving in synchronism. The axial force is found to be higher in
pipe systems and the overall length of thermomechanical patterns is longer.
Helical loops form preferentially in situations where the axial force and the related stored
strain are higher, for example: smaller duct clearances, longer lengths, higher bending
stiffness and higher coefficient of friction. In most duct routes a proportion of both waves
form, with some sine waves changing to single or multiple helical loops as the temperature is
increased. The helical loop stores more strain than a sine wave and therefore is in a lower
energy condition. The minimum energy theorem explains that the three transitions a) straight
cable to sine wave, b) straight cable to helical loop and c) sine wave to helical loop occur
when the total energy stored in the system can be reduced from the present level to a lower
state. The present energy state required for a helix to form is higher, but the resulting level
is lower. This theorem is classically applied to the problem of elastic buckling. In the cable
model energy is continuously being added to the system by the temperature increase. Graphs
of the increase of system energy of the model with temperature were analysed and showed
that, at the formation of a sine wave or a helical loop, the system energy continued to
increase, but at a reduced rate. Once the storage capacity of the particular wave or loop was
occupied, the energy increased at the higher rate until another wave transition occurred.
Cables in ducts are traditionally installed in comparatively close fitting pipes. This restricts
the clearance for lateral patterns to form and for bends to absorb thermal strain. Consequently
the peak force and turn-over temperature are likely to occur above the 90
o
C operating
temperature of the cable. Thus the axial force at 90
o
C will be the theoretical maximum
(F=EA ) for a rigid straight cable. The sidewall forces at bends and longitudinal forces at
joints and terminations will also be maximum values. On the beneficial side, cable lateral
movement is small and the risk of fatigue failure of the metallic sheath is reduced.
The highest force in a pipe system occurs when the exscribed circle around the triangular
group of three cables is equal to the pipe diameter. This limits the minimum clearance
possible between the pipe and the individual cable to 1.15D. The minimum clearance for the
reference 138kV cable is 93mm. Good practice is to allow a sensible clearance on top for
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-61
installation purposes. Thus there is more likelihood of experiencing high axial forces in a
tight duct system in which the clearance may be 30-50mm, than in a pipe system in which the
clearance starts at 93mm.
The temperature dependent properties of the XLPE insulation significantly reduce the cable
axial stiffness, EA, and the bending stiffness, EI, at higher temperatures. EA and EI play key
roles in thermomechanical behavior. At ambient, the temperature dependence of EA
increases the rate of rise of axial force and at higher temperatures reduces it. The maximum
theoretical axial force is also reduced. At higher temperatures the magnitude of EI falls and
increases the cable curvature, thereby permitting the critical bending moment to reduce to a
lower level. This usefully reduces the magnitudes of the turn-over temperature and peak
force.
The pattern wavelength is influenced by the pipe clearance and cable properties. Cable
wavelength reduces significantly under high axial force, becoming shorter under high thrust.
High thrust also reduces the bending radius. However the space for small radii is limited in
the tighter fitting pipes and ducts. Thus excessive levels of sheath strain actually occur in
pipes with larger clearances. For a particular cable geometry there is an optimum pipe
clearance that will reduce axial thrust without excessively increasing sheath strain.
The limiting values of the parameters that can shorten the life of an XLPE cable system were
derived in Chapter 4. They are compared with the values exhibited by the cable in the pipe
and duct models. The parameters are listed below in decreasing level of importance to the
reliability of the cable system:
Axial Thrust

Very high values of axial thrust occur in tight fitting pipes and ducts. The values fall
rapidly to low levels when larger clearances are selected. The cable construction is
capable of withstanding high levels of axial force as these occur in the situations
when the cable is afforded lateral support by the tight fitting pipe.

The accessories are the weakest part of the system in withstanding cable axial force.
Differential movement in an accessory risks short term electrical failure. It is essential
that a) the joint and b) the cleating system are designed to withstand the axial force. A
straight joint is required to withstand, without movement or disturbance, the
difference in force exerted by the adjacent pipe/duct spans. An anchor joint cannot
move, but is required to withstand the full cable thrust. The force is significantly
reduced by selecting an increased pipe clearance.
Sidewall Force

The level of sidewall force that can occur in the bends of tight fitting ducts and pipes
is close to the limit under which the axial force will either a) cause the insulation to
be penetrated, or b) move the conductor more than 5%-10% through the XLPE
insulation. Penetration of the insulation risks immediate electrical failure. This risk
can be reduced by selecting a) an increased pipe/duct clearance to reduce the thrust
and b) selecting an increased bend radius to reduce the sidewall load. The selection of
a sensible clearance in a duct system is particularly important. It has been practice to
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-62
select the duct diameter as the next increment in a standard range of diameters. The
inadvertent selection of too close a fit will result in a disproportionate increase in
axial and sidewall force. The selection of a cable construction capable of
withstanding sidewall force is equally important. To avoid local penetration and
disruption; the semiconducting insulation shield, the ground wires/tapes and the
metallic foil sheath should be protected with adequate cushioning layers.
Sheath Strain

The absolute levels of sheath strain are significant in pipe and duct systems, but are of
themselves not a limiting factor, providing that the minimum bending limits are not
exceeded. The important limiting factor is the level of cyclic fatigue strain at which
cracking occurs within the required life of the cable system. Cracking of the sheath
may result in local overheating of the cable by the interruption of sheath currents, this
can directly imperil the XLPE insulation by thermal degradation. Moisture may also
penetrate the sheath, leading to longer term water tree failure of the XLPE insulation.

The highest levels of sheath strain occur in a pipe system with maximum levels of
absolute strain up to 1.4%. Cyclic strain depends upon the particular daily loading
curve and the thermal capacity of the cable. For HV and EHV cables, with large
conductors and thick insulation, the daily fluctuation in temperature will be small.
Taking a 20% fluctuation in temperature to be representative, the average cyclic
strain will be reduced to 0.28%.

Except in the circumstance of a constant, non varying load, a lead alloy sheathed
cable would be beyond its fatigue limit of 0.05% and should not be selected for pipe
system applications. Extruded corrugated aluminum sheaths with a wall thickness of
greater than equal to 2mm have a suitable fatigue performance, with a limiting fatigue
strain of 0.5%, however they are unlikely to be acceptable for pipe installation
because of their increased diameter.

The limiting design value for fatigue strain of 0.15% has been proposed in Chapter 4,
Table 4.9, based on published test results for a development design of four layer
aluminum foil laminate 275kV cable. Foil laminate cables are likely to be suitable for
pipe systems with medium clearances. The importance is emphasized of a)
performing fatigue tests at 80
o
C to validate the limiting value and b) calculating the
optimum pipe clearance, using either the FEA or NS calculation, methods given in
this report.

Thin wall, corrugated welded aluminum, copper and stainless steel sheaths are
available for some designs of HV and EHV cables. These sheaths are likely to have a
test performance superior to foil laminate cables, especially at the typical jacket
temperature of 80
o
C. In the absence of published test results it is recommended that a
fatigue limit of less than 50% of that of the thick wall aluminum sheath be taken i.e.
0.25%. On this basis thin wall welded sheaths are likely to be suitable for use in
pipe installations. As with foil laminate sheaths it is important that fatigue tests be
performed to validate the design limits for fatigue strain.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-63

The sheath strain in a duct system has a lower absolute level, with a calculated
maximum value of 1.0%; thus the cyclic strain is likely to be less than 0.2%.

The comments made on sheath suitability for pipe systems are equally applicable to
duct systems. The use of a lead sheath should only be considered for the tightest
fitting pipe as this will reduce the level of cyclic strain, however other design
measures will be required to limit the consequent level of high axial thrust occurring
in bends and in manholes. Corrugated aluminum, extruded thick wall sheaths have the
best fatigue performance, followed next by the thin wall welded sheaths and then by
foil laminate sheaths. It is recommended that fatigue tests be performed to validate
the design limits for fatigue strain.
Minimum Bending Radius

In the preformed pipe bend the cable curvature is dictated by the specified radius,
such as 35D. In the larger clearance pipe systems the cable crosses the entrance to the
bend at an angle resulting in a reduction in cable radius, for example to 20D.

The cable curvature within the wave patterns can be less than the 35D radius
specified for the preformed bend, particularly with the larger pipe clearances. This is
an increased risk situation as the cable is unsupported and flexes during load cycling.

The suitability of the cable construction should be checked by cyclic fatigue testing at
the required bending radius as calculated in this report
The mechanism of thermomechanical behavior has been shown to be predictable with the
variation of axial force for each individual parameter following continuous functions, in
some cases a near linear function. However the sensitivity study confirmed that
thermomechanical behavior is a multi-variable problem. From the sensitivity study alone it is
not possible to conclude whether, for a specific cable design, a simple linear
summation/multiplication of parametric factors will be sufficiently accurate for engineering
design.
The finite element modelling technique has been demonstrated to yield consistent and
verifiable results over a wide range of parameters.
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-64
Sensitivity Study Bar Charts and Graphs

Figure 7-19
Duct position A: Bar chart ranking of turn-over force for % change in parameters

Figure 7-20
Duct position A: Bar chart ranking of turn-over temperature for % change in parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-65

Figure 7-21
Duct position A: Angle of route inclination

Figure 7-22
Duct position A: Duct to cable clearance
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-66

Figure 7-23
Duct position A: Length of straight span

Figure 7-24
Duct position A: Weight
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-67

Figure 7-25
Duct position A: Angle of duct bend

Figure 7-26
Duct position A: Coefficient of friction, cable to duct
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-68

Figure 7-27
Duct position A: Coefficient thermal expansion

Figure 7-28
Duct position A: Bending stiffness, EI
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-69

Figure 7-29
Duct position A: Axial stiffness, EA

Figure 7-30
Duct position A: Radius of duct bend
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-70

Figure 7-31
Duct position A: Torsional stiffness, GJ

Figure 7-32
Duct position B: Bar chart ranking of turn-over force for % change in parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-71

Figure 7-33
Duct position B: Bar chart ranking of turn-over temperature for % change in parameters

Figure 7-34
Duct position B: Pipe to cable clearance
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-72

Figure 7-35
Duct position B: Coefficient of thermal expansion

Figure 7-36
Duct position B: Bending stiffness, EI
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-73

Figure 7-37
Duct position B: Length of straight span

Figure 7-38
Duct position B: Weight
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-74

Figure 7-39
Duct position B: Axial stiffness, EA

Figure 7-40
Duct position B: Angle of duct bend
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-75

Figure 7-41
Duct position B: Coefficient of friction, cable to duct

Figure 7-42
Duct position B: Angle of route inclination
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-76

Figure 7-43
Duct position B: Radius of duct bend

Figure 7-44
Duct position B: Torsional stiffness, GJ
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-77

Figure 7-45
Duct positions A,B,C, Bar Chart part 1: Axial force variation with parameters

Figure 7-46
Duct positions A,B,C, Bar Chart part 2: Axial force variation with parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-78

Figure 7-47
Duct positions A,B,C, Bar Chart part 1: Turn-over temperature

Figure 7-48
Duct positions A,B,C, Bar Chart part 2: Turn-over temperature
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-79

Figure 7-49
Duct, Bar Chart part 1: Numbers of total and helical waves along straight route

Figure 7-50
Duct, Bar Chart part 2: Numbers of total and helical waves along straight route
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-80

Figure 7-51
Duct, Bar Chart part 1: Minimum cable bending radius in straight section

Figure 7-52
Duct, Bar Chart part 2: Minimum cable bending radius in straight section
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-81

Figure 7-53
Duct, Bar Chart part 1: Maximum sheath strain in straight section

Figure 7-54
Duct, Bar Chart part 2: Maximum sheath strain in straight section
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-82

Figure 7-55
Duct, Bar Chart part 1: Maximum sidewall force in bend

Figure 7-56
Duct, Bar Chart part 2: Maximum sidewall force in bend
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-83

Figure 7-57
Pipe position A cable 1: Bar chart ranking of turn-over force for % change in parameters

Figure 7-58
Pipe position A cable 2 or 3: Bar chart ranking of turn-over force for % change in
parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-84

Figure 7-59
Pipe position A: Clearance pipe to one cable

Figure 7-60
Pipe position A: Angle of route inclination
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-85

Figure 7-61
Pipe position A: Coefficient of thermal expansion

Figure 7-62
Pipe position A: Weight
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-86

Figure 7-63
Pipe position A: Length of straight span

Figure 7-64
Pipe position A: Axial stiffness, EA
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-87

Figure 7-65
Pipe position A: Radius of pipe bend

Figure 7-66
Pipe position A: Coefficient of friction, pipe to cable
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-88

Figure 7-67
Pipe position A: Angle of pipe bend

Figure 7-68
Pipe position A: Bending stiffness, EI
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-89

Figure 7-69
Pipe position A: Coefficient of friction, cable to cable

Figure 7-70
Pipe position A: Torsional stiffness, GJ
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-90

Figure 7-71
Pipe position B cable 3: Bar chart ranking of turn-over force for % change in parameters

Figure 7-72
Pipe position B cable 3: Bar chart ranking of turn-over temperature for % change in
parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-91

Figure 7-73
Pipe position B: Pipe clearance to one cable

Figure 7-74
Pipe position B: Coefficient of thermal expansion
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-92

Figure 7-75
Pipe position B: Axial stiffness, EA

Figure 7-76
Pipe position B: Bending stiffness, EI
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-93

Figure 7-77
Pipe position B: Weight

Figure 7-78
Pipe position B: Coefficient of friction, cable to cable
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-94

Figure 7-79
Pipe position B: Angle of route inclination

Figure 7-80
Pipe position B: Coefficient of friction, pipe to cable
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-95

Figure 7-81
Pipe position B: Length of straight span

Figure 7-82
Pipe position B: Radius of pipe bend
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-96

Figure 7-83
Pipe position B: Angle of pipe bend

Figure 7-84
Pipe position B: Torsional stiffness, GJ
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-97

Figure 7-85
Pipe, Bar Chart position A, cables 1,2,3, part 1: Axial force variation with parameters

Figure 7-86
Pipe, Bar Chart position A, cables 1,2,3, part 2: Axial force variation with parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-98

Figure 7-87
Pipe, Bar Chart position A, cables 1,2,3, part 1: Turn-over temperature variation with
parameters

Figure 7-88
Pipe, Bar Chart position, cables 1,2,3, A part 2: Turn-over temperature variation with
parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-99

Figure 7-89
Pipe, Bar Chart position B, cables 1,2,3, part 1: Axial force variation with parameters

Figure 7-90
Pipe, Bar Chart position B, cables 1,2,3, part 2: Axial force variation with parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-100

Figure 7-91
Pipe, Bar Chart position B, cables 1,2,3, part 1: Turn-over temperature variation with
parameters

Figure 7-92
Pipe, Bar Chart position B, cables 1,2,3, part 2: Turn-over temperature variation with
parameters
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-101

Figure 7-93
Pipe, Bar Chart, cables 1,2,3, part 1, Cable minimum bending radius in straight section

Figure 7-94
Pipe, Bar Chart, cables 1,2,3, part 2, Cable minimum bending radius in straight span
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-102

Figure 7-95
Pipe, Bar Chart, cables 1,2,3, part 1, Maximum sheath strain

Figure 7-96
Pipe, Bar Chart, cables 1,2,3, part 2, Maximum sheath strain
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-103

Figure 7-97
Pipe, Bar Chart, cables 1,2,3, part 1, Maximum sidewall force in pipe band

Figure 7-98
Pipe, Bar Chart, cables 1,2,3, part 2, Maximum sidewall force in pipe band
EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-104

Figure 7-99
Duct system: Variation in axial force at position B during load cycling 15-90
o
C

Figure 7-100
Duct system: Comparison of the 1st and 30
th
patterns at 90
o
C Foreshortened view, scale
1:1



EPRI Licensed Material

Thermomechanical Performance of Cable in a Route Containing Bends and Slopes: Sensitivity Study
7-105

i
Mechanical effects on extruded dielectric cables and joints installed in underground transmission systems in North
America, technical update, August 2002, EPRI, Palo Alto,CA:2002.1001848
ii
Underground Transmission Systems Reference Book, 1992 Edition, EPRI, Palo Alto, TR-101670, Research
Project 7909-01
iii
Research to determine the acceptable emergency operating temperatures for extruded dielectric cables, EPRI, Palo
Alto, CA, 1978, EL-938, Project 933-1
EPRI Licensed Material
8-1
8
PERFORMANCE OF SELECTED DESIGNS OF XLPE
CABLES IN THE SYSTEM VOLTAGE RANGE 69-345KV
Introduction
The objectives of this chapter are:
To model a range of 69-345kV duct and pipe cable sizes in the hockey stick route, Figure
7-1. Figure 8-1 illustrates the range of cable sizes studied.
To assess the effect on the results of modelling updated physical parameters analysed from
the cable test program, Chapter 3, for the 138kV 1500kcmil and 230kV 2500kcmil reference
cables
To calculate the equivalent axial forces for the above duct and pipe systems using the
sensitivity study data from Chapter 7. Based on a comparison with the modelling results, to
assess the accuracy of the method as a simplified engineering approach to the design of duct
and pipe systems.

Figure 8-1
Photographic comparison of 138kV 1500kcm and 230kV 2500kcm cables
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-2
Parameters
A new set of physical parameters was selected for the study of 69-345kV systems in this chapter
using up to date values from the cable tests at EPRIsolutions laboratory.
Cable Axial Stiffness EA
The EA values were analysed from the thermally constrained tests performed on the 138kV
1500kcmil reference cable, these measured the increase of axial compressive force with
temperature. The data set was designated EA
4
, this being the forth set used in the project. This
supersedes the reference set EA
ref
, which was based on the

results of the measurements of axial
stiffness under tension for the same cable.
Figure 8-2 compares EA
4
and

EA
ref
values for the 138kV 1500kcmil foil reference cable
extrapolated to extend over the 15
o
C to 125
o
C temperature range used in the study. EA
4
is 2.7 x
higher at 20

C and 1.7 x higher at 90


o
C. This results in the magnitude of axial thrust being
increased in the same proportion for a cable exhibiting the characteristics of a rigid bar i.e.
longitudinally constrained with anchored ends and laterally constrained within a tight fitting
duct.
Experimental measurements of axial thrust under thermally constrained conditions were not
available for other cable constructions and so it was necessary to calculate these by taking the
138kV 1500kcmil EA values pro-rata to the required conductor area and then adding the EA
values for each of the component layers.

Figure 8-2
138kV 1500kcmil cable EA
4
axial stiffness values compared to the previous EA
ref
reference
set
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-3
Coefficient of Thermal Expansion
The 17 x 10
-6
C
-1
reference value used in the Chapter 7 sensitivity study was increased to 18.5 x
10
-6
C
-1
, as measured by EPRIsolutions. The effect is to increases the cable axial force in
proportion i.e. by 9%.
Coefficient of Friction
The 0.1 reference value used in the Chapter 7 sensitivity study was increased to 0.3, this being
more representative of the general condition in service after the pulling-in lubricant has dispersed
and dried out. The effect is to increase the frictional constraint by 30%. In the longitudinal
direction, a given length of cable will be able to support a bigger difference in force between end
A and bend B. In the transverse direction this affects the minimum energy condition at which
large amplitude waves form.
Axial Force Coefficient EA
The axial force is directly proportional to EA for a given temperature rise. The rigid bar
values for each cable construction over the 15
o
C-125
o
C range are plotted in Figure 8-3. The
force-temperature characteristic for the 138kV 1500kcmil cable is closely similar to the
measured values. It is probable that the calculated force characteristics for the 2500 and
3000kcmil conductor sizes will be higher than can be generated. I is known that self-contained
fluid filled cables with 4,000kcmil conductors of the segmental milliken construction experience
relaxation under thermally constrained conditions, limiting the maximum achievable thrust to
60,000N. There was no experimental information available for XLPE cable conductor sizes
above 2,500kcmil and so no allowance could be made for relaxation. The thrust-temperature
characteristics in Figure 8-3 are the rigid bar reference values for the studies in this chapter.
Duct System Cable Models
The cables and ducts in Table 8-1 are representative of the range of cables, weights and
stiffnesses likely to be encountered in duct installations.

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-4

Figure 8-3
Force-temperature characteristics of 69kV-345kV cables in the rigid-bar condition

Table 8-1
Details of cables for the duct system studies
System Voltage 69 kV 138 kV 230 kV 345 kV
Properties (EA, , )
4
(EA, , )
4
(EA, , )
4
(EA, , )
4

Conductor [kcmil] 1500 1500 2500 3000
Insulation
thickness [mm]
13 20 27 30
Sheath type Corrugated
seamless aluminum
Copper tape ground
conductor and
aluminum foil laminate
Lead alloy Corrugated
seamless aluminum
Cable outer
diameter [mm]
96 83 125 133.4
Duct internal
diameter [mm]
136.5 233 162.7 213.2
Duct-cable
clearance [mm]
40.5 150 37.7 79.8
Cable weight
[kgm
-1
]
14 11.9 34.3 25.1
The calculation of the temperature dependent elastic stiffness parameters EA, EI and GJ
followed the method given in Chapter 7.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-5
The 69kV 1500kcmil cable dimensions are those of a cable sample held in reserve for testing at
EPRIsolutions. The duct internal diameter of 136.5mm (5.37ins) is that of the nominal 5ins
duct taken from a list of FRE standard conduit sizes (FRE: fiber reinforced epoxy). The
136.5mm duct size was chosen to give a practical clearance for pulling-in, this being 40.5mm.
The 345kV 3000kcmil design was selected as having a large conductor area and the
comparatively thick XLPE insulation of 30mm. An extruded corrugated seamless aluminum
(CSA) sheath was chosen to suit this large diameter cable. The CSA sheath has a good
combination of a) low weight for pulling-in and b) high fatigue resistance to withstand load
cycling strains expected in service life. A standard internal duct diameter of 213.2mm (8.39ins)
was selected, giving a generous clearance of 79.8mm for pulling-in.
The 230kV 2500kcmil cable dimensions are those of the lead sheathed cable tested at
EPRIsolutions. The duct internal diameter of 162.7mm (6.4ins) is the size that the service cable
is installed in, giving a clearance of 37.7mm. One reason for selecting this cable is that it had not
previously been modeled in the hockey-stick reference route. Another reason is that the
insulation thickness of 27mm, is suitable for use in the higher stressed XLPE insulations likely to
be employed in a 345kV cable. Thus the thermomechanical results are directly applicable to both
230kV and 345kV applications.
Figure 8-4 is a photograph of the 230kV 2500kcm cable in a 152mm internal diameter example
duct. The clearance of 27mm is 11mm less than the 37.7mm clearance in the service installation.
The conductor is of the five segment milliken type. The slight non-circularity of the lead sheath
at the 12o/c and 6o/c positions reveals the presence of two small tubes containing optical fibers.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-6

Figure 8-4
230kV 2500kcmil lead sheathed cable in a 152mm duct, with 25mm clearance
Duct Systems Results
The axial forces generated in the four models is summarized in Table 8-2 at a) the temperature at
which the thermomechanical pattern forms (turn-over temperature), at the 90
o
C operating
temperature and at the 125
o
C study limit temperature. The duct systems are listed in order of
ascending clearance.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-7
Table 8-2
Duct systems: axial forces at positions A and B
Turn-Over Force
[N]
Turn-Over
Temperature
[
o
C]
Force at 90
o
C
[N]
Force at 125
o
C
[N]
Cable and
(Properties)
Duct
Clearance
[mm]
A B A B A B A B
230kV,
2500kcmil,
lead
(EA,,)
4

37.7 68,568 48,499
#
70 60-
67.5#
68,705 62,055 46,177 76,671
69kV
1500kcmil,
CSA
(EA,,)
4

40.5 34,486 24,080
#
70

57.5# 30,442 37,537 20,631 46,162
345kV,
3000kcmil,
CSA
(EA,,)
4

79.8 56,655 48,499
#
65 60-
67.5#
44,750 54,745 36,370 72,615
138kV
1500kcmil,
foil (EA,,)
4

150 14,607 7,284 60 62.5 11,540 7,375 8,228 5,765
Legend: # indicates that thermomechanical wave patterns formed, causing an inflection
point in the force-temperature characteristic, but not producing a significant drop in force.
230kV 2500kcmil Lead Sheath Cable, (EA,, )
4

, Clearance 37.7mm
Figure 8-5 compares the axial force-temperature characteristics at positions A and B; values are
recorded in Table 8-2. A photograph of the cable is shown in Figure 8-4:
Position B, Bend
The cable had moves into the bend up to 27.5
o
C. The axial force then rises at rigid-bar rate. At
60
o
C wave patterns are seen to have been initiated. A point of inflection can be seen in the force
curve at 60-67.5
o
C, however the force does not fall. At 67.5
o
C there are 5 sine waves present. At
67.5
o
C the force is 48,499N. Including the absorption effect of the bend, the force is 74% of the
rigid-bar value. Excluding the absorption of the bend the force is 96% of the rigid-bar value.
At 90
o
C the patterns extend 90m along the duct. The patterns and bend have reduced the axial
force to 62,055N, 75% of the rigid bar force.
At 125
o
C the patterns have extended the complete 200m along the duct, Figure 8-6. The patterns
and bend have reduced the axial force to 76,671N, 79% of the rigid bar force.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-8
Position A, Span End
The force rises at the rigid-bar rate from 15
o
C to reach a turn-over value at 70
o
C of 68,568N
(99% of rigid-bar). The cable does not slip during the rise. Slip does occur at 70
o
C shortly after
the pattern has formed at B. The force is then limited to a maximum value of 69,852N at 75
o
C.
At 90
o
C thermal strain is released along the frictional force gradient into the approaching
patterns, reducing the force to 68,705 N, 83% of the rigid bar force.
The force characteristics at A and B cross over at 95
o
C, because the frictional force gradient
which travels in advance of the pattern had reached A , releasing thermal strain. Whereas at B,
the small amplitude waves were saturated and unable to absorb the thermal strain being
generated, thus permitting the force to rise.
At 125
o
C the pattern has reached A and reduced the force to 46,177N, 48% of the rigid bar
maximum.
Summary
The 37.7

mm clearance is the smallest studied in this chapter. Despite the comparatively close fit,
thermomechanical patterns were able to form and had a significant effect in reducing the force at
90
o
C to 75% at B and to 83% at A. The effect was more marked at 125
o
C, with reductions to
79% and 48% at A and B.
A new variant of thermomechanical behavior is seen, associated with the close fitting duct, in
which at 125
o
C the force at bend B significantly exceeded that at A the anchored end.

Figure 8-5
230kV 2500kcmil lead cable, (EA, , )
4
: Force-temperature graphs at A and B
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-9

Figure 8-6
230kV 2500kcmil lead cable, (EA, , )
4
: Patterns at 125

C
69kV 1500kcmil CSA, (EA,, )
4
Clearance 40.5mm
Figure 8-7 compares the axial force-temperature characteristics at positions A and B, values are
recorded in Table 8-2.
The cable produces thermomechanical patterns at B, the entrance to the bend, but does not
exhibit a true turn-over temperature in which the force drops to low level. The maximum thrust
of 46,162N occurs at 125
o
C at position B.
Position B, Bend Entrance
The cable can be seen to have moved into the bend between 15 and 25
o
C. The force then rises
linearly to 57.5
o
C at the rigid bar rate. Wave patterns form at 55
o
C and produce a point of
inflection in the force characteristic at 57.5
o
C.
At 90
o
C the patterns have extended to 119.5m in length. The thermomechanical patterns and
bend have reduced the force to 74% of the rigid bar value.
At 125
o
C the patterns have extended along the complete 200m length, Figure 8-8.
At 125
o
C the presence of patterns and bend have reduced the maximum force to 79% of the rigid
bar case.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-10
Position A, Span End
At position A the force rises at the rigid-bar rate with a small fluctuation occurring at 25
o
C. This
being slippage caused by cable moving into the bend at B and the force differential between A
and B exceeding the frictional constraint. Pronounced slip occurs at 60
o
C, following the
formation of thermomechanical patterns at B.
The force peaks at a turn-over of 70
o
C due to slip and falls to values of 30,442N (60% of rigid-
bar) at 90
o
C and 20,631N (35% of rigid-bar) at 125
o
C. The fall is due to thermal strain being
released by the frictional gradient in advance of the approaching wave patterns.
At 78
o
C the force characteristic at A falls below that at B. This is because the pattern growing
from A towards B is now long. The force at the tip of the pattern is low and the frictional force
differential has reached A, releasing thermal strain and reducing force. Whereas, the wave
pattern at B can no longer absorb strain due to the small duct clearance and so the force rises
with temperature.
Summary
Thermomechanical patterns formed at the entrance to the bend at B. A turn-over temperature
resulted, but the force did not fall, as the rate of generation of thermal strain was greater then the
absorption capability of the waves. The rate of increase of force was however limited. The
patterns rapidly extended along the complete 200m length and at 125
o
C had a major effect in
reducing the force at A to 35% of the rigid bar value.

Figure 8-7
69kV 1500kcmil CSA cable: Force-temperature characteristics at A and B
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-11

Figure 8-8
69kV 1500kcmil CSA cable: Thermomechanical wave pattern at 125
o
C
Foreshortened view, scale 1:1
345kV 3000kcmil CSA Sheath Cable, (EA,, )
4
Clearance 79.8mm
Figure 8-9 compares the axial force-temperature characteristics at positions A and B; values are
recorded in Table 8-2:
Position B, Bend Entrance
The cable moves into the bend between 15-32.5
o
C. The large duct clearance of 79.8mm has
permitted the bend to absorb a higher proportion of thermal strain. The force then rises at the
rigid-bar rate until patterns are seen to form at 60
o
C. A point of inflection occurs at 62.5
o
C with
a very small reduction in force. At 62.5
o
C the force is limited to 59% of the rigid-bar value,
taking the benefit of the bend into account. Removing the benefit of the bend, the force is
reduced to 97% of the rigid bar value.
At 90
o
C the patterns have extended to 98.2

m in length, Figure 8-10. Taking the combine effect
of the bend and patterns, the force is reduced to 56% of the rigid bar.
At 125
o
C the force is reduced to 64% of the rigid bar.
Position A, Span End
The force rises at the maximum theoretical rate. The cable slips at 25
o
C because of the force
differential of 14,003N that has built up at A, caused by the cable moving into the bend at B.
The cable slips again at 65
o
C, shortly after the pattern has formed at B. This limits the maximum
force to 56,655N, which is 68% of the rigid-bar force.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-12
The force characteristic of A crosses that at B at 75
o
C. This is because the pattern at B is
saturated and can no longer absorb the rate of increase of thermal strain. Whilst at A, the
approaching pattern is still absorbing strain.
The force falls above 65
o
C, because the frictional force differential in advance of the pattern has
reached A.
The force at 90
o
C of 56,655N is 45% of the rigid-bar.
The force at 125
o
C of 36,370N is 32% of the rigid bar.

Figure 8-9
345kV 3000kcmil CSA cable,(EA,, )
4
: Force-temperature graphs at A and B,
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-13

Figure 8-10
345kV 3000kcmil CSA cable,(EA,, )
4
: Thermomechanical pattern at 90
o
C,
Foreshortened view, scale 1:1
138kV 1500kcmil Foil Laminate, (EA,, )
4
Clearance 150mm
Figure 8-11 compares the axial force-temperature characteristics at positions A and B, values are
recorded in Table 8-2.
Position B, Bend
The large clearance significantly increases the absorption of the bend. This delays the rise of
force until a temperature of 50
o
C. The force then rises in a linear way from 50
o
C to 57.5
o
C.
Waves form at 57.5
o
C and a turn-over occurs at 62.5
o
C limiting the force to 7,284N. The
combined effect of the bend and patterns is to limit the force to 20% of the rigid bar force. The
force then falls to approximately 5,000N
Figure 8-12 shows a 48m long region of waves present at the 90
o
C operating temperature. The
pattern comprises three helical loops and a train of sine waves. The force at 90
o
C of 7,375N is
16% of the rigid-bar. The force at 125
o
C of 5,765N is 11% of rigid-bar.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-14
Position A, Span End
Position A is initially isolated as the force rises immediately at the rigid bar-rate. At 25
o
C the
force differential between A and B exceeds the frictional constraint and the cable slides 200m to
feed thermal strain into the incipient pattern at B. Slippage occurs at three temperatures up to
47.5
o
C. The force then rises linearly to reach a maximum of 14,607N at 60
o
C.
The force at 90
o
C of 11,540N is 24% of rigid-bar.
The force at 125
o
C of 8,228N is 15% of rigid-bar.
Summary
The 138kV cable and duct combination develops significantly lower thrusts than the other
designs. This results from the large duct clearance of 150mm. The clearance permits a) more
thermal strain to be stored in the bend and b) large amplitude wave patterns to form and act as a
constant force limiting device.

Figure 8-11
138kV 1500kcmil foil cable, EA
4
: Force-temperature graphs at A and B

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-15

Figure 8-12
138kV 1500kcmil foil cable, EA
4
: Pattern at 90
o
C
138kV 1500kcmil Foil Laminate: Effect of Varying EA, and
Table 8-3 compares the effect of changing the parameters from (EA,,)
ref
to (EA,,)
4
for the
same cable and duct geometry:
For a given temperature rise to 60
o
C the EA ratio is 2.22, which agrees fortuitously with
the 1.94 ratio of peak forces at the point of slip at A. Thus turn-over occurred at a 94%
higher level.
The key effect is at B where the higher rise of force reduces the turn-over temperature at
which patterns form by 7.5
o
C.
The peak turn-over force occurs at a 35% higher level. This shows that the formation of
large amplitude waves does not simply occur at a constant value of force/bending moment,
but is a multi-variable function.

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-16
Table 8-3
138kV duct system with 150mm clearance, effect of varying axial stiffness EA
Peak Force [N] Turn-Over Temperature [
o
C] 138kV cable
Parameters
A B A B
(EA,,)
ref
7,520 5,377 70 72.5
(EA,,)
4
14,607 7,284 62.5 65
Ratio: (EA,,)
4
/(EA,,)
ref

1.94 1.35 - -
Duct System Summary
The study is summarized in Table 8-4 in order of increasing clearance:
Beneficial thermomechanical effects are present in each of the four cable systems in the form
of absorption of thermal strain a) in the bend and b) in the wave patterns.
Clearance is the key parameter in reducing thermomechanical force, as predicted in the
Chapter 7 normalization and sensitivity studies.
Thermomechanical forces make a beneficial reduction in force at the 90
o
C operating
temperature of 20% even with the smallest clearance of 37.7mm
Clearances of greater than 79.8mm reduce the force by more than 50%.

Table 8-4
Axial forces in duct systems listed in order of duct clearance
Axial force as
Proportion of Rigid-Bar
[%]
90
o
C 125
o
C
Clearance
[mm]
System
Voltage
Conductor
Area [kcmil]
Axial
Stiffness EA
at 90
o
C [MN]
Rigid-Bar
Force at 90

o
C [N]
A B A B
37.7 230kV 2500 59.7 82,830 83 75 48 79
40.5 69kV 1500 36.3 50,370 60 74 35 79
79.8 345kV 3000 71 98,460 45 56 32 64
150 138kV 1500 34 47,175 24 16 15 11

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-17
Pipe System Cable Models
The cable and pipe diameters are listed in Table 8-5.

Table 8-5
Details of cables for the pipe system studies
System Voltage 69kV 138kV 138kV 138kV
Properties (EA, , )
4
(EA, , )
4
(EA, , )
4
(EA,,)
ref

Conductor [kcmil] 1000 1500 1500 1500
Insulation thickness [mm] 18 20 20 20
Sheath type Copper wire
ground
conductor and
aluminum foil
laminate
Copper tape
ground
conductor and
aluminum foil
laminate
Copper tape
ground
conductor and
aluminum foil
laminate
Copper tape,
ground
conductor and
aluminum foil
laminate
Cable outer diameter [mm] 74.4 83 83 83
Pipe internal diameter [mm] 206.4 233 206.4 206.4
Pipe cable clearance [mm] 132 150 123 123
Cable weight [kgm
-1
] 7.97 11.9 11.9 11.9
The 69kV cable was selected as having a) a small conductor size of 1000kcmil and b) a low
stress insulation giving an insulation thickness of 18mm, which is also suited to 110-138kV.
Thus, the 69kV design is also indicative of the thermomechanical performance of higher stressed
110-138kV cables. A ground conductor of copper wires and a sheath of aluminum foil laminate
were selected to give a combination of reduced diameter and weight. A standard internal pipe
diameter of 206.4mm (8.125ins) was selected, giving a clearance on one of the three cables of
132mm.
The two 138kV cables are identical and have the design of the reference cable in the Chapter 7
sensitivity study, but are now installed in the correct pipe size for the service application. This
being 206.4mm (8.125ins), giving a clearance of 123mm. The reference model clearance is
150mm. One of these models is given the (EA)
ref
parameter set and the other the (EA, , )
4
parameter set
.

Figure 8-13 is a photograph of the 138kV 1500kcmil cables in an example pipe with an internal
diameter of 200mm. The clearance to an individual cable is 117mm, this being within 6mm of
the 123mm clearance in the service installation.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-18

Figure 8-13
138kV 1500kcmil XLPE cables in a 200mm diameter pipe with a clearance of 117m
Pipe System Results
The thermomechanical performance of the three cables in the pipe is found to be consistent.
Cable 1, at the top of the trefoil pattern, has a lower frictional constraint than cables 2 and 3
underneath it. Thus cable 1 tends to slip longitudinally at a lower axial force.
The forces are given in Table 8-6 for each of the three cables at a) the turn-over temperature, b)
the 90
o
C operating temperature and c) the study limit of 125
o
C. The average force per cable is
given to assist comparison.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-19
Table 8-6
Pipe systems: peak axial forces at positions A and B
Pipe [mm] Turn-Over
Force [N]
Turn-Over
Temperature
[
o
C]
Force at 90
o
C
[N]
Force at 125
o
C
[N]
Cable and
Properties
ID D
Core
N
o

A B A B A B A B
1 7,129 5,344 47.5 50 3,453 5,830 2,524 7,845
2 14,951 5,550 47.5 50 10,043 6,556 7,485 7,479
3 15,781 8,587 47.5 45 10,907 7,327 8,160 8,877
69kV
1000kcmil
aluminum
foil
(EA,,)
4


206.4

132
Av 12,620 6,494 - - 8,134 6,571 6,056 8,067
1 14,040 11,662 45 50 4,921 9,366 4,222 12,521
2 26,050 12,724 47.5 45 14,864 11,698 11,214 14,140
3 26,049 15,833 47.5 45 14,867 13,107 11,218 14,259
138kV
1500kcmil
aluminum
foil,
(EA,,)
4


206.4


123
Av 22,046 13,406 - - 11,551 11,390 8,885 13,640
1 9,488 7,947 50 50 4,910 5,679 3,747 7,176
2 21,408 8,335 50 52.5 14,852 8,446 11,663 9,895
3 23,823 12,440 50 47.5 16,863 10,187 13,219 10,784
138kV
1500kcmil
aluminum
foil,
(EA,,)
4


233

150
Av 18,240 9,574 - - 12,208 8,104 9,543 9,285
3 13,900 10,676 60 55 9,584 10,075 7,946 11,333 138kV
1500kcmil,
foil,
(EA,,)
ref

206.4 123
% of
EA
4

53% 80% - - 64% 88% 71% 79%
69kV 1000kcmil Aluminum Foil Sheath Cable, (EA,, )
4
, clearance 132mm
Figure 8-14 compares the increase of axial force with temperature at positions A and B. Figure
8-15 shows the thermomechanical deformation patterns present at the 90
o
C operating
temperature at the entrance to the bend, position B:
Upon heating from 15
o
C to 30-37.5
o
C, the cable expands into the bend at B thus initially holding
the axial force at low magnitude.
In comparison , the force at the start of the route at A in cable 1 rises immediately from 15
o
C to
20
o
C, at which the force differential between A and B exceeds the frictional constraining force
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-20
along the 200m cable length. Cable 1 then slips from A to B, releasing a proportion of the
locked-in thermal strain into the bend. The forces in Cable 2 and 3 continue to rise.
The force at B continues to rise until thermomechanical patterns form at turn-over temperatures
of 42.5-45
o
C, limiting the axial forces to 5,344N-8,587N.
At A, the force continues to rise to 7,129-14,951N at the turn-over temperature of 47.5
o
C, at
which the force differential between A and B in cables 2 and 3 rises to approximately 9,401N.
The frictional restraining force along the 200m length is exceeded and the cable slips, releasing
thermal strain into the pattern at B.
At B, the thermomechanical patterns continue to grow along the cable length with temperature,
absorbing thermal strain and holding down the axial force. Figure 8-15 shows the
thermomechanical patterns adjacent to the entrance to the bend at the cable operating
temperature of 90
o
C. The overall length of the pattern is 71m compared to the 33m length in
view.

Figure 8-14
Pipe, 69kV 1000kcmil foil, clearance 132mm, (EA,, )
4
: Force temperature graph at A and B
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-21

Figure 8-15
Pipe, 69kV 1000kcmil foil, clearance 132mm, (EA,, )
4
: Thermomechanical pattern at 90

C
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,, )
4
, Clearance 123mm
Figure 8-16 shows the force/temperature characteristic at positions A and B for each of the three
cables installed in trefoil formation. Figure 8-17 shows the thermomechanical pattern at 90
o
C
adjacent to the bend, B.
At B the three cables expand into the bend as the temperature rises from 15 to 27.5
o
C.
At end A the three cables exhibit an identical rapidly increasing force characteristic from 15
to 20
o
C, the top cable then slips towards B. The bottom two cables do not slip because of the
higher frictional constraint, resulting from the additional weight of the top cable.
At B, above 27.5
o
C, the force continues to rise until thermomechanical patterns form at 45
o
C
at forces of 11,662-15,833N.
At A above 20
o
C, the force in cable 1 rises to the comparatively low maximum level of
14,040N, at a turn-over temperature of 45
o
C, at which the cable slips. Whereas the forces in
cables 2 and 3 rise to a maximum 26,050N at a turn-over temperature of 47.5
o
C.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-22

Figure 8-16
Pipe, 138kV 1500kcmil foil, clearance 123mm, (EA, , )
4
: Force temperature graph at A and B


Figure 8-17
Pipe, 138kV 1500kcmil foil, clearance 123mm, (EA, , )
4
: Thermomechanical pattern at
90C
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-23
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,, )
4
, Clearance 150mm
This case is included to quantify the effect of changing the parameters from the Chapter 7
sensitivity study reference set to the (EA,,)
4
set.

The cable to pipe clearance of 150mm is the
reference case value. The results are given in Table 8-6.
The force-temperature characteristics are very similar to those of the same cable in the smaller
pipe clearance of 123mm, shown in Figure 8-18 and Figure 8-19. The differences are mainly that
the forces in the larger pipe are lower in magnitude. The differences are:
Position B, Bend
The cables moves into the bend and fill it at a smaller temperature rise. Cables 2 and 3 move into
the bend and fill it by 32.5C, a rise of 17.5C and force of 2,581N, compared to the smaller pipe
with a temperature rise of 12.5C and force of 2,865N.
Similarly cable 3 reaches a turn-over temperature of 47.5C and force of 12,440N, whereas the
smaller pipe values are 45C and 15,833N.
The length of the thermomechanical pattern at 90C and 125C is shorter at 68m and 115mm,
compared to the smaller pipe lengths of 89m and 134m.
The force at 90C is lower at 10,187N compared to the smaller pipe of 13,107N.
Position A, Span End
At 20C cable 1 slips twice in the larger pipe and once in the smaller pipe.
Cables 2 and 3 rise to a turn-over slip force of 21,408N-23,823N at 50C compared to the
smaller pipe values of 26,050 at 47.5C.
The force at 90C is the reverse, being 16,863N compared to 14,867N in the smaller pipe. This is
because the thermomechanical pattern has grown to a longer length in the smaller pipe, such that
the frictional force differential reached end A earlier and has released more thermal strain.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-24

Figure 8-18
Pipe, 138kV 1500kcmil foil, 150mm clearance, (EA,, )
4
: Force temperature graph at A and B

Figure 8-19
Pipe, 138kV 1500kcmil foil, 150mm clearance, (EA,, )
4
: Pattern at 90C
138kV 1500kcmil Aluminum Foil Sheath Cable, (EA,, )
ref
, Clearance 123mm
The values of force are recorded in Table 8-6 for comparison with the results from the same
clearance pipe with the (EA,,)
4.
It can be seen that the effect is large as it increases the force at
the turn-over temperature at position A by x 2 and at position B by x 1.48. At 90
o
C the effect is
similar with an increase in force at A and B of x 1.55 and x 1.3.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-25
Pipe System Summary
The forces are summarized by comparison to the rigid bar-forces and are given as percentages in
Table 8-7. In each of the cases the force at the turn-over temperature was the highest in the
temperature range studied up to 125
o
C.
The forces are given as percentages of the rigid bar values a) at a nominal turn-over temperature
of 50
o
C and at the operating temperature of 90
o
C. The table is ranked in order of increasing
clearance between one cable and the pipe diameter. It was found that:
The results are in agreement with those for both the duct study in this chapter and with the
sensitivity study in Chapter 7.
Increasing the clearance reduced the forces. The 123-150mm range of clearances studied was
comparatively small, nevertheless the percentage of rigid-bar force at 90
o
C fell from 50 to
40%.
The reduction due to thermomechanical effects was similar, whether there was one or three
cables in the pipe. Table 8-4 shows that at 90
o
C for one 138kV cable in the duct the forces at
90
o
C at A and B are 24% and 16%, whereas three cables in the same size pipe produce 39%
and 20% of the rigid-bar force.

Table 8-7
Axial forces in pipe systems listed in order of increasing clearance
Axial Force
At Turn-Over
Force
At 90
o
C
Clearance
[mm]
System
Voltage
Conductor
size
[kcmil]
Value
A B A B
Turn-Over Force [N] 22,046 13,406 - - 123 138kV 1500
[%] of Rigid Bar at
T/O
80% 48% 47% 28%
Turn Over Force [N] 12,620 6,494 - - 132 69kV 1000
[%] of Rigid Bar at
T/O
68% 35% 40% 21%
Turn Over Force [N] 18,240 9,574 - - 150 138kV 1500
[%] of Rigid Bar at
T/O
66% 35% 39% 20%


EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-26
Normalized Span Calculation Method
This section investigates whether the peak axial force (and thence other key design parameters)
can be calculated with the level of accuracy necessary for engineering design purposes. The
calculation method is named the NS technique (Normalized Span). The NS technique is based on
a) comparing the ratios of the parameters for a specific pipe or duct system with those of the
138kV 1500kcmil reference model and b) for each parameter ratio, interpolating the change in
force using the sensitivity study data in Chapter 7.
Thermomechanical behavior is shown in Chapter 7 to be a predictable, multivariable problem.
Additionally, axial force is a discontinuous function of temperature, arising from the formation
of peak force levels at an intermediate turn-over temperatures. These transitions are caused by
a) formation of thermomechanical patterns and b) frictional slip towards the part of the route
containing either a bend or a thermomechanical pattern. More usefully, the Chapter 7 sensitivity
study showed that when one individual parameter was varied, the graph of peak force and the
percentage change in that parameter was a continuous function, in many cases being near linear.
The forces calculated by the NS technique are compared with the output from the 69kV-345kV
study of 4 duct and 4 pipe system models.
NS Calculation Method
Step 1: Calculate the magnitudes of each of the parameters for the particular cable system. The
parameters are listed in Table 8-8 and Table 8-9.
Step 2: Identify the changes in parameters and dimensions between the study cable duct or pipe
system and the reference 138kV 1500kcm duct or pipe system.
Step 3: Calculate the ratio of each parameter to that of the reference case given in Chapter 7. The
reference is a 138kV 1500kcmil aluminum foil cable with a 150mm pipe clearance in a flat,
hockey stick shaped span (200m straight section with a 30D x 90
o
bend). The parameter
symbols and figure numbers are given in Table 8-8 and Table 8-9 for both duct and pipe systems.

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-27
Table 8-8
Geometric parameter ratios for both duct and pipe systems
Duct system
Figure No
Pipe system
Figure No
Geometric Parameter Ratios:
Actual Case/Reference Case [%]
Position Position
Symbol Description
Force
Ratios
[per unit]
A B A B
G
D
Clearance between individual
cable and pipe diameters, (D
P
-
D
c)
F
D
Figure
7-22
Figure
7-34
Figure
7-59
Figure
7-73
G
L
Length of straight section of
span
F
L
Figure
7-23
Figure
7-37
Figure
7-63
Figure
7-81
G
R
Radius of pipe bend F
R
Figure
7-30
Figure
7-43
Figure
7-65
Figure
7-82
G

Angle of pipe bend F

Figure
7-25
Figure
7-40
Figure
7-67
Figure
7-83
G

Angle of slope F

Figure
7-21
Figure
7-42
Figure
7-60
Figure
7-79

Table 8-9
Cable parameter ratios and graphs for duct and pipe systems
Duct Systems
Figure No
Pipe Systems
Figure No
Cable Parameter Ratios:
Actual Case/Reference Case [%]
Position Position
Symbol Description
Force
Ratios
[per unit]
A B A B
C
W
Weight per meter

F
W
Figure
7-24
Figure
7-38
Figure
7-62
Figure
7-77
C
EA
Total axial stiffness EA F
EA
Figure
7-29
Figure
7-39
Figure
7-64
Figure
7-75
C
EI
Total bending stiffness EI F
EI
Figure
7-28
Figure
7-36
Figure
7-68
Figure
7-76
C
GJ
Torsional stiffness GJ F
GJ
Figure
7-31
Figure
7-44
Figure
7-70
Figure
7-84
C

Total coefficient of thermal
expansion (linear)
F

Figure
7-27
Figure
7-35
Figure
7-61
Figure
7-74
C
CP
Coefficient of friction, cable to
pipe
F
CP
Figure
7-26
Figure
7-41
Figure
7-66
Figure
7-80
C
CC
Coefficient of friction, cable to
cable
F
CC
- - Figure
7-69
Figure
7-78
Step 4: For each parameter percentage ratio read off the axial force from the appropriate figure
numbers given in Table 8-8 and Table 8-9. Calculate the force ratio, by dividing the force read
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-28
from the graph by the reference case force. (The ratio is calculated as a per unit value and not as
a percentage).
Step 5: Calculate the axial Force Factor by multiplying together the individual force ratios
identified in Table 8-8 and Table 8-9.
Step 6: Note the Reference Forces for the 138kV 1500kcmil reference duct or pipe system at
positions A (span end) and B (bend entrance). (These are the peak values that occurred at the
turn-over temperature when thermomechanical wave patterns formed). The Reference Forces are
obtained from the Chapter 7 sensitivity study and are:
Duct system at A: 7,520N
Duct system at B: 5,373N
Pipe system at A: 10,458N (cable 3, bottom right of trefoil group)
Pipe system at B: 7,322N (cable 3, bottom right of trefoil group)
Step 7: Calculate the peak force by multiplying the Reference Force in Step 6 by the Force
Factor in Step 5, for both positions A and B. Note that the resultant peak force may occur at
either the formation of a thermomechanical pattern (turn-over temperature), or in the case of a
close fitting duct at the study limiting temperature of 125
o
C.
Example NS Calculation for a 69kV Duct System 1500kcmil CSA Sheathed Cable
The outline dimensions of the 69kV 1500kcmil CSA sheathed cable and duct have the outline
dimensions recorded in Table 8-1. The duct clearance is 40.5mm.
Calculation of Axial Force at Position A in the 200m Hockey Stick Route
The force at the turn-over temperature at which the thermomechanical patterns form is calculated
for the 69kV 1500kcmil cable in the following steps:
Step 1: Calculate the magnitudes of each of the 11 geometric and cable parameters for the
duct or pipe system to be studied as given in Table 8-8 and Table 8-9.
Step 2: Identify the changes in parameters and dimensions between the 69kV and 138kV
cables. In this 69kV duct example, 6 of the parameters change, they are:
G
D
Clearance
C
W
Weight
EA Axial stiffness
EI Bending stiffness
C

Coefficient of friction
C
CP
Coefficient of thermal expansion
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-29
Step 3: The parameter ratios are calculated as the percentage ratio of the 69kV parameter to
the reference 138kV parameter. The calculated percentages are listed in row 1 in Table 8-10.
For example, the 69kV duct clearance is 40.5mm. The reference 138kV clearance is 150mm.
The G
D
ratio is:
% 27 100 x
150
5 . 40
= |
.
|

\
|
.
Step 4: The force is read off the appropriate Chapter 7 sensitivity study graphs as listed in
Table 8-8 and Table 8-9 for position A. This was divided by the 138kV 1500cmil reference
case force for each of the parameter ratios. The resultant force ratios are listed in row 2 of
Figure 8-10.
Example 1; the graph of force-duct clearance is copied in this chapter as Figure 8-20, for a ratio
of G
D
=27%, a force ratio of F
D
=2.03 is calculated.
Example 2; the graph of force-bending stiffness is copied as Figure 8-21, from which, for a
ratio of C
EI
=212%, a force ratio of F
EI
=1.26 is calculated.
Step 5: The Force Factor is then obtained by multiplying the individual force ratios together
in the second row of Table 8-10, = 6.02.
Step 6: The 138kV reference peak turn-over force at A of 7,520N is obtained from Table
8-3, (or from the sensitivity study graphs).
Step 7: The turn-over force for the 69kV cable at A is obtained by multiplying the position A
reference force by the Force Factor: 7,520N x 6.02 = 45,270N.
Comparison of NS Answer with Model Analysis
The calculated peak force at position A of 45,270N is 31% higher than the peak force in the
model analysis of 34,486N, which occurred at a turn-over temperature of 70
o
C, at which
frictional slip occurred. (The rigid-bar force for this cable at 90
o
C is 50,370N.)
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-30
Table 8-10
Peak axial force factors for the 69kv 1500kcmil duct system at position A
Ratios Diameter
Clearance
Weight Axial
Stiffness
EA
Bending
Stiffness
EI
Coeff of
Friction
Coeff of
Thermal
Expansion
Force
Factor
Param
r

[%]
G
D=
27% C
W
=118% C
EA
=229% C
EI
=212% C

=109% C
CP
= 300% -
Force
Ratio
[per unit]
F
D
=2.03 F
W
=1.092 F
EA
=1.145 F
EI
=1.26 F

=1.043 F
CP
=1.804 6.02
Position B in 200m Hockey Stick Route
The same procedure is repeated for position B except that a) a different set of reference graphs
specifically for position B is referenced, see Table 8-8 and Table 8-9 and b) the 138kV reference
turn-over force has to be selected for position B from Table 8-3. The reference force is 5,373N.
The resultant force ratios for position B are given in Table 8-11. The calculated Force Factor is
8.14. (It should be noted that both individual force ratios and the force factor are different from
those at position A).
The turn-over force at position B is 5,373N x 8.14 = 43,736N. This is 5% less than the model
analysis peak force of 46,162N.
Table 8-11
Peak axial force factors for the 69kV 1500kcmil duct system at position B
Ratios Diameter
Clearance
Weight Axial
Stiffness
EA
Bending
Stiffness
EI
Coeff of
Friction
Coeff
Thermal
Expansion
Total
Force
Factor
Param
r

[%]
G
D=
27% C
W
=118% C
EA
=229% C
EI
=212% C

=109% C
CP
= 300% -
Force
Ratio
[per unit]
F
D
=4.17 F
W
=1.06 F
EA
=1.12 F
EI
=1.34 F

=1.18 F
CP
=1.04 8.14


EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-31

Figure 8-20
Duct: Force- % clearance characteristic at A

Figure 8-21
Duct: Force- % bending stiffness EI characteristic at A

EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-32
Force Factors for the Duct Models
The calculated Force Factors for the four duct system models are given in Table 8-12. It was
found that the resulting magnitudes of the Force Factors are large. A significant proportion is the
Force Ratio for the duct clearance alone.
The peak force levels for the reference 138kV 1500kcmil hockey stick duct system are:
Position A: 7,520N
Position B: 5,373N
Table 8-12
Force factors for duct system calculations
Duct Clearance Force Factor for
Position A
Force Factor for
Position B
Cable
[mm] Percentage of
150mm Ref
Clearance Total Clearance Total
230kV 2500kcmil
lead sheath
37.7 25% 1.93 14.2 4.38 18.80
69kV, 1500kcmil,
CSA sheath
40.5

27% 2.03 6.02 4.17 8.15
345kV 3000kcmil
CSA sheath
79.8 53% 1.43 12.28 2.13 12.97
138kV,
1500kcmil,
aluminum foil
sheath
150 100% 1.13 2.12 1.09 1.35
Force Factors for the Pipe Models
The calculated Force Factors for the three pipe models are given in Table 8-13. The magnitudes
of the Force Factors are significantly smaller than the duct models. The Force Ratio for the duct
clearance is now a minor proportion of the factor.
The peak force levels for the reference 138kV 1500kcmil hockey stick pipe system are:
Position A: 10,458N
Position B: 7,322N
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-33
Table 8-13
Force factors for pipe system calculations
Pipe Clearance Force Factor for
Position A
Force Factor for
Position B
Cable
[mm] Percentage of
150mm Ref
Clearance Total Clearance Total
138kV 1500kcmil
aluminum foil
sheath
123 82% 1.19 2.80 1.3 2.35
69kV, 1000kcmil,
aluminum foil
sheath
132

88% 1.02 1.65 1.16 1.33
138kV,
1500kcmil,
aluminum foil
sheath
150 100% 1 2.35 1 1.81
Comparison of NS Calculation Method Results with FEA Model Analysis
The accuracy of the calculated method is compared with the modeled method for pipe systems in
Table 8-14 and duct systems in Table 8-15.
Accuracy of NS Calculation for the Pipe Systems
For pipe systems the range of variation of the NS calculated peak force is acceptably low (-
7% to +13%). Expressing the difference as a percentage of the rigid bar-force at 90
o
C
reduces this to -2% to +7%.
Some of the sensitivity study models need to be re-run with a wider range of axial stiffness.
The present graphs have force-EA characteristics which allow for maximum EA values up
to 215% of reference value. The effect of adopting the EA results from the EPRIsolutions
thermally constrained test creates a need to widen the specification to allow for a 500%
increase. It is recommended that the EA sensitivity models be re-run.
The accuracy of the calculation for pipe systems is surprisingly high, considering that the
point of reference was one of the two lower cables, cable 3, in the trefoil group. The model
output shows that a) the lower cables generate higher forces and b) there is invariably some
variation between the two bottom cables. The small variances between the NS calculation
method and the FEA model analysis arise from:
The small differences between the model pipe clearances and the sensitivity study
clearance. This resulted in the total range of multiplication Force Factors being small
i.e. x 1.65 to x 2.8.
The clearance between one cable and the pipe is inherently large in pipe systems; in
the study being 123-150mm. This provides space for large amplitude
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-34
thermomechanical waves to form. These waves very efficiently reduce the
thermomechanical force to typically 50-55% of the rigid-bar force at 90
o
C. This
reduces the differences in force due to variations in cable and geometric parameters.
The recommended methods of calculation are:
For accurate solutions with minimum uncertainty, it is recommended that the FEA
thermomechanical modelling technique described in Chapter 5 be used. This method
has been successfully proven in the modelling of combinations of route geometry up
to 16km in length, containing multiple spans. Expertise is required in preparing the
model, running the analysis and in analysing the results.
The preparation of the stiffness parameters for the particular cable design is tedious
and it is recommended that this be transferred to CD ROM based software.
Fore outline designs of pipe systems, it is recommended that the NS method
described in this chapter is sufficiently accurate. The NS method uses the output data
from the FEA sensitivity study analysis in Chapter 7. Although the method is simple,
the work is tedious, because it uses a large number of look-up graphs. The example
given in this chapter was to calculate the limiting parameter of peak axial force.
Exactly the same procedure is followed to calculate the other limiting parameters,
these being sheath strain, minimum bending radius and sidewall pressure. For
practical engineering use it is recommended that the NS method be prepared in the
form of CD ROM based software.

Table 8-14
Pipe systems: difference in peak axial force between calculated and model methods
Pipe
Clearance
Ratios of Calculated Peak Force Fc to
Modeled Force Fm
Position A Position B

Cable

Pipe
Diameter
[mm]
[mm] % of
ref
case
Max Rigid
Bar Force
90
o
C [N]
Fc/Fm
[N/N]
Fc/Fm
[%]
Fc/Fm
[N/N]
Fc/Fm
[%]
138kV
1500kcmil
206.4 123 82% 47,175

29,313/
26,050
113%
(7%)
17,189/
15,833
109%
(3%)
69kV
1000kcmil
aluminum foil
206.4 132 88% 31,635 17,234/
15,781
109%
(5%)
7,999/
8,587
93%
(2%)
138kV
1500kcmil
233 150 100% 47,175

24,566/
23,823
103%
(2%)
13,222/
12,440
106%
(2%)
Legend: (xyz %) in the Fc/Fm column is the variance (Fc-Fm) expressed as a % of the rigid
bar force at 90
o
C.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-35
Accuracy of Calculation for the Duct System
For duct systems, the range of variation of the calculated peak force is high (-5% to +63%).
The variance is reduced slightly to -5% to +44%, by expressing the difference between
calculated and model force as a percentage of the rigid-bar force at 90
o
C.
The maximum levels of force from the analysis of the 69-345kV duct system models, with
small duct clearances, are in the range 73-92% of the rigid-bar force. Thus the variance from
the simple calculation method gave a force that was either at, or above, the rigid-bar
theoretical limit. System design engineers would previously have designed duct circuits with
a) comparatively small clearance and b) maximum rigid-bar force. For large conductor cables
this would have resulted in an axial forces of excessively high magnitude. On modern HV
and EHV XLPE cables with large conductor sizes, stringent design measures would have to
be taken to reinforce cables and joints in manholes, to the extent that raises doubt about
practicability. The work in this project has shown that a small increase in the cable to duct
clearance significantly reduces the axial force at manholes to levels that permit practical
solutions. To take advantage of this knowledge, it is essential to be able to select the
optimum clearance and calculate the reduced level of force with confidence. This is achieved
with the FEA modelling method.
The NS calculation method has the prospect of being suitable for the outline design of
duct systems, but it is necessary for the accuracy to be improved. This can be
achieved by making the data sets from the sensitivity study specific to duct
applications.
The large variance in force found in duct system calculations arose from the selection in the
Chapter 7 sensitivity study of one set of model dimensions to study both duct and pipe
installations. The objective was to compare and understand the then unknown
thermomechanical mechanisms with either one or three cables present in the same pipe. The
use in this chapter of the same data to calculate forces in duct systems with small clearances
has identified the following areas of uncertainty:
There were insufficient models analysed with small clearances, thus the shape of the
Figure 8-20 characteristic for clearances of less than 50% is uncertain i.e. clearances
of less than 75mm. Three of the four cable clearances studied fall within this zone of
uncertainty.
The total multiplying Force Factors are too large and introduce unacceptable
uncertainties. The sensitivity study reference clearance of 150mm is too large
compared with available duct diameters. The Force Ratio to calculate the effect of
clearance in one case being as large as x 4.3. The resulting Force Factors of x 6 to x
18 are too high to achieve accuracy.
At the entrance to the bend, position B, it was found that wave patterns always
formed and that a discernable turn-over temperature was present. However, only in
one duct case, with the reference clearance of 150mm, were the patterns able to act as
constant force device, which either leveling-off, or reducing the force, as present in
each of the pipe systems studied. In the particular Chapter 7 sensitivity study cases in
which the force continued to rise, the maximum force occurred at the maximum
applied temperature of 125
o
C and so this temperature was entered into the set of data.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-36
The same approach was taken in analysing the 69-345kV models in this chapter. This
introduces two different physical phenomenon as the criteria for peak force. The
analysis of the sensitivity study in Chapter 7 has already shown that there are multi-
variable interactions present. Thus the particular behavior of ducts with small
clearances, in which two criteria are used, increases calculation uncertainty.
In the evaluation of the NS calculation method, the axis of axial EA graph obtained
from Chapter 7 was 215% x EA
ref
. However the EA values from the latest
EPRIsolutions measurements were 450% x EA
ref
. This required the trend line to be
extrapolated into a region of uncertainty, which was clearly unjustifiable. This is
particularly true for values of greater than 200% EA
ref
, because the trend line moves
from a linear to a polynomial function.
For the design of duct systems the recommended methods of calculation are:
For accurate solutions with minimum uncertainty, the FEA implicit
thermomechanical modelling technique be used as described in Chapter 5.
For outline designs to be undertaken quickly, the simple linear interpolation method
described in this chapter has merits, but requires improved data from further FEA
modelling. The reference case needs to be positioned with each parameter in the
center of the range of variation and not offset, as in the present data set.
In particular it is recommended that the sensitivity models for variation of a) clearance and b)
axial stiffness EA, be re-run to update the data and improve the accuracy of the calculation
method.
Table 8-15
Duct system: difference in peak axial force between calculation and model methods
Duct Clearance Ratios of Calculated Force to Modeled
Force
Position A Position B

Cable

[mm] % of ref
case
Max Rigid Bar
Force 90
o
C
[N]
Fc/Fm
[N/N]
Fc/Fm
[%]
Fc/Fm
[N/N]
Fc/Fm
[%]
230kV 2500kcmil
lead sheath
37.7 25% 82,833 106,619*/
69,852
153%
(44%)
100,995*/
76,671
132%
(29%)
69kV, 1500kcmil,
CSA sheath
40.5

27% 50,366 45,270/
34,486
131%
(21%)
43,736/
46,162
95%
(5%)
345kV 3000kcmil
CSA sheath
79.8 53% 98,457 92,346/
56,655
163%
(36%)
69,711/
72,615
96%
(3%)
138kV, 1500kcmil,
aluminum foil sheath
150 100% 47,175 16,003/
14,607
110%
(3%)
7,239/
7,284
99%
(0.1%)
Legend: (xyz %) in the Fc/Fm column is the variance (Fc-Fm) as a percentage of the rigid
bar force at 90
o
C.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-37
Conclusions
The work in this chapter has quantified the performance of selected 69kV-345kV cable sizes
in pipe and duct systems. The thermomechanical behavior was confirmed to be in agreement
with that predicted by the Chapter 7 sensitivity study.
Thermomechanical wave patterns were formed in each of the cases studied in both pipe and
duct systems, at temperatures that were within the chosen 15-90
o
C XLPE cable operating
range.
Thermomechanical wave patterns were confirmed to have the beneficial effect of reducing
axial force. The magnitude of reduction is primarily determined by the clearance within the
duct, or pipe, to one cable. This knowledge permits major design advantages to be gained by
reducing a) cable side wall force in bends and b) forces and movements in joint manholes.
An optimum clearance exists beyond which fatigue strain in the metallic sheath/foil may
become unacceptable.
In the duct systems studied, the force was reduced to 83%* ( t) for a 230kV/345kV
2500kcmil cable design with a close fitting clearance of 38mm. The force was reduced to
56% ( t) for a 345kV 3000kcmil cable by selecting a wider clearance of 80mm. (*The
reduction in force is compared to the force developed by a fully constrained cable at 90
o
C in
the rigid-bar condition).
In the 69kV to 138kV pipe systems studied the force was reduced to 32%-36% ( t) for cable
designs in the range of 69-138kV and 1000-1500kcmil conductor size with clearances of
123-150mm. The large force reductions result from the large clearances inherent in pipe
system design.
Improvements were successfully made to the set of parameters to better represent 69-345kV
cable performances. Increases were made to the magnitudes of axial stiffness (+104% at
60
o
C), coefficient of friction (+200%) and coefficient of thermal expansion (+9%). The result
was to increase the force by +90% at the end of a straight span and by 40% at the entrance to
a bend. The coefficient of friction between the cable and pipe/duct wall is shown to be a key
parameter.
Two methods of calculation have been successfully developed to calculate thermomechanical
behavior. The first method is the backbone of this project and analyses the output of a
detailed FEA (Finite Element Analysis) model of the route geometry. The second is
evaluated in this chapter and is called the NS technique (Normalized Span). It is based upon
the interpolation of look-up data from the sensitivity study, based on comparison with the
normalized geometry of a reference span. The recommended applications for each of the
methods is given below:
Detailed Design
The FEA modelling technique is recommended. Accuracy and minimum uncertainty is
required in the detailed design of new systems and modifications to existing systems as these
are required to exhibit reliability over a long service life. The FEA method is described in
Chapter 5. The FEA method has been specifically adapted to model pipe and duct routes. It is
now an efficient, powerful and reliable technique, which has been proven in the modelling of
routes with multiple spans. The advantage is that the technique closely models the real route.
EPRI Licensed Material

Performance of Selected Designs of XLPE Cables in the System Voltage Range 69-345kV
8-38
Disadvantages are that the computing is non-trivial and powerful software and workstation
computers are required, together with the expertise to prepare and analyse the large volume
of data. The FEA technique is fully operational for use.
Outline Design
The NS technique described in this chapter is recommended. Speed is required in the
preparation of feasibility studies when a) planning new route geometries and b) selecting
outline cable and pipe/duct dimensions to meet ampacities. The advantage is that the
technique can be transferred to run on a standard PC with data being entered using a simple
dialogue box. The disadvantage is that knowledgeable decisions have to be taken on the sub-
division of complex span geometry into manageable sections for comparison with the
reference route. The RS technique is suitable for application to pipe systems now. The NS
technique requires an updated parameter set to achieve the accuracy required for application
to duct systems.

EPRI Licensed Material
9-1
9
EFFECTS OF VIBRATIONS
Introduction
The longitudinal movement of different types of cable associated with traffic vibration has been
experienced by many utilities, but in general has not been satisfactorily explained. It has been
experienced that cable installations with clearances inside pipes, ducts and extruded aluminum
sheaths are more susceptible to movement.
This chapter commences with a review of existing work on traffic vibration with the objective of
abstracting the vibration data necessary for the modelling work. The method of modelling
vibration from the journey of a truck driving near to a cable route is then described.
The work in this chapter shows that vibration precipitates movement in a cable containing a non
uniform distribution of axial strain. The necessary distribution arises from the gravitational and
thermomechanical forces present in normal service operation. The mechanism is believed to be
new and is described at the start of the chapter to assist the reader in the interpretation of the
detailed results. Finally, the detailed modeling results are given of duct and pipe systems with
different angles of inclination to the horizontal. The effect of vibration due to repeated truck
journeys is also considered.
Published Work on Traffic Vibration
Previous Investigational Work on Cable Problems
Longitudinal movement of cable core within metallic sheaths, pipes and ducts associated with
vibration has been experienced in many countries, with records dating from the 1920s. A paper
i

from an Australian CIGRE symposium describes the electrical failure of a straight joint in a
330kV 1200mm
2
, fluid filled self contained cable circuit in 1992 after 13 years in service and the
subsequent radiography and replacement of five further joints. Regular radiography revealed a
progressive core movement of 96mm over a five year period. The cause was attributed to traffic
induced movement of the paper insulated core within the extruded aluminum sheath. The typical
criteria associated with cable creepage were said to be present in the vicinity of the 330kV spans:
Poor subsoil composition, muddy soil and silt
Water saturated subsoil, below the water table
Depressions and cracking in the road surface
EPRI Licensed Material

Effects of Vibrations
9-2
The cable was directly under a long length of highway over and cable creepage was in the
direction of traffic flow
Vehicle movements and weights had increased since installation
The paper
[i]
is of interest as it reviews contributory causes together with a good historical
bibliography. The 1997 paper was influenced by an experiment performed in 1937 in which a
cable was laid in ducted road crossings in marshy clay. One cable was installed in a duct
underneath an ordinary road surface and a second cable was installed underneath a specially laid
smooth road surface. The ordinary road surface produced both depression and vibration of the
road resulting in a cable movement of 110mm during 3 months. The special surface produced
depression and no vibration, resulting in 100mm movement. It was reported that the predominant
effect was depression of the surface and that the most important variables were a) the strength of
the road surface, for example concrete spreads the load and b) the elasticity of the subsoil which
transmits the depressive displacement through the ground to the duct.
Mechanisms of failure were postulated for the 1992 failure, although not confirmed. These were
based on the possible elastic deformation of the road and backfill beneath vehicle wheels
resulting in cyclic longitudinal movement of the cable sheath. The cable core within the elastic
aluminum sheath was postulated to have been moved progressively in one direction only by
either a surfing, conveyor belt, or hybrid mechanism. These mechanisms are not directly
applicable to pipe and duct systems because of the high longitudinal rigidity of the pipe wall and
of the solid backfills that encases them.
However the generation and magnitude of vibration by passing trucks on an uneven road surface
and the application of the vibration to the pipe or duct, is of interest to the FEA modeling work.
The possibility to be investigated is that vertical vibration could momentarily lift the cable from
the floor of the pipe or duct. The longitudinal frictional constraints that hold the cable from
moving under gravitational or thermomechanical forces would also momentarily be overcome.
The consequences, if any, can then be analysed and quantified.
Previous Work to Quantify the Magnitude of Traffic Induced Vibrations
A literature search resulted in contact with GR Watts, an expert at the Transport and Road
Research Laboratory (TRRL), Crowthorne UK and author of Traffic Induced Vibrations in
Buildings
ii
. This is a formal report issued by the Department of Transport, a UK Government
body. The report records experimental measurements that quantify a) the magnitude of ground
vibrations generated by trucks passing over measured experimental irregularities in a road
surface at the TRRL research track and b) the vibration transmission characteristics of different
types of ground. A comprehensive list of 46 references is included. This report provided the
input data for the applications of vibration to duct and pipe cable systems.

EPRI Licensed Material

Effects of Vibrations
9-3

Figure 9-1
Road surface test profiles
Figure 9-1 shows the shape of four experimental profiles on the test track. Earlier work had
measured the vibrations due to the movement of concrete slabs when vehicles passed over the
joints between slabs, however it was considered that the results would not be suitable for
experimental replication. Eight trucks termed HGVs (heavy goods vehicles), with different
types of suspension, loaded and unloaded, were driven at a range of speeds over the
irregularities. These irregularities represented the range of defects that might arise from poorly
repaired road surfaces and were chosen to generate repeatable vibrations.
The HGV speeds were up to 80kmh
-1
(50mph). The vertical velocity of the road surface, ground
or building foundation was defined as the peak particle velocity. In the first experiment,
velocity was measured at lateral distances of 2m and 6m from the wheel at the track surface by
triaxial geophone displays. The minimum level of vibration perception by a person is 0.3 mms
-1
.
Linear correlations were obtained between the vertical ground velocity measured at 6m distance
and the following parameters:
Vehicle speed: a maximum ground velocity of 1.4 mms
-1
was produced at 80kmh
-1
(50mph),
with a surface bump of 25mm height and 600mm length.
Vehicle weight, a maximum ground velocity of 0.8mms
-1
is produced at 80kmh
-1
, for a two
axle vehicles loaded up to 15t passing over a surface bump of 25mm height and 600mm
length at 48kmh
-1
(30mph). Note that modern European HGVs with 4 to 5 axles are now
loaded up to a 38t limit.
Height of bump in road: a maximum velocity of 1.7 mms
-1
was produced with a height of
50mm at a speed of 58kmh
-1
(36mph).
EPRI Licensed Material

Effects of Vibrations
9-4
Peak accelerations were measured at public dwellings exhibiting perceptible vibration; one
dwelling being in London and four in Swindon. The peak accelerations of building foundations
ranged from 42-130mms
-2
with dominant frequencies of typically 12Hz, but in the range of 12-
60Hz. A peak sinusoidal acceleration of 130mms
-2
at the accompanying measured frequency of
12.7Hz is equivalent to a peak vertical velocity of 1.63mms
-1
and a peak vertical displacement of
0.02mm.
The forcing frequency of 10-12Hz results from the wheel hop mode of vibration of the vehicle
suspension. It was noted that different types of suspensions have different characteristics and in
consequence the vibration from an unloaded vehicle at speed may be higher than that from a
loaded vehicle.
Propagation tests on the transfer of vibrations through different types of ground were made at 13
sites over a range of 3-50m lateral distances by applying a falling weight deflectometer impact
test to the road surface. It was shown that the modulii of the ground transfer functions between
the 6m measurement position and a 1m distance was typically 1:3. It was shown that the
modulus of the vibration transfer function for damp peat soil is 10 times greater than London
clay, 30 times higher than sand and 100 times higher than boulder clay and rock.
Equation 9-1 is given
[ii]
to predict the vibration at the foundation of a building:
x
v
6
r
48
p . t . v . a . 028 . 0
G
(

= Equation 9-1
Where:
G
v
: peak particle velocity at the building foundation [mms
-1
]
a: maximum height or depth of road surface defect [mm]
v: maximum speed of HGV truck [kmh
-1
]
r: distance of building foundation from road defect [m]
t: ground scaling factor; 0.1 for chalk, 0.94 for sand/gravel, 3.84 for peat,
7.07 for alluvium
p: wheel factor, if defect is in path of one wheel = 0.75, in two = 1.0
x: power factor for soil type, average values being in range of -0.74 to -1.19
Vibration measurements were made where HGV trucks passed over known road irregularities
and these were compared with predictions from Equation 9-1. The highest measured vibration
velocity was plotted as 3.5 mms
-1
. The correlation of the plotted comparison appeared to be
reasonable for measured velocities up to 1.5 mms
-1
.
The researchers
[ii]
then designed an experiment to apply vibration to the foundation of a house at
2m distance from a geophysical vibrator. They selected a frequency of 13Hz and a vibration
acceleration of 2.5-3.0 mms
-1
at foundation level adjacent to the vibrator.
EPRI Licensed Material

Effects of Vibrations
9-5

Figure 9-2
Waveform of ground vertical velocity representing traffic vibration
The resulting vertical velocity measured at the building foundation is shown in Figure 9-2. The
peak velocity is given as 2.6 mms
-1
, the principal frequency is shown to be approximately 13Hz
and the duration of the wave train is 0.8s. The reason for the duration is not given. It is not
recorded whether this is representative of the period of oscillation of the truck suspension for one
wheel alone, or for all of the wheel axle sets in succession passing over the bump. The distance
traveled by a truck at 45kmh
-1
(28mph) in 0.8 s is 10m, which corresponds to the typical length
of a truck. The resulting vibration recorded in the structure of the house was eight times the
minimum magnitude of a persons perception, with the highest levels being recorded in
suspended floors and ceilings, these were said to be very noticeable such that they would have
been unacceptable to most occupants.
The peak velocity of 2.6 mms
-1
at 13Hz yields a peak displacement of 0.032mm, Equation 9-7,
which is in agreement with the 0.03mm figure given in the CIGRE AP21 paper
[i]
.

Overall, the selection of the wave pattern and magnitude of vibration was supported within the
report
[ii]
.
The vibration waveform in Figure 9-2 was therefore chosen to be applied to the modeling of duct
and pipe cable systems.
Applying Vibration to Duct and Pipe System Models
Model Geometry
The span geometry selected was that of the reference model used in the sensitivity study, Figure
7-1 and Figure 7-2. The span starts at position A and has a 200m straight section leading to
position B at the start of a 90
o
bend with a radius of 2.91m (35D, where D is the cable diameter).
A one meter straight section connects the exit of the bend to the end of the span at C. The ends of
the cable at A and B are anchored.
EPRI Licensed Material

Effects of Vibrations
9-6
As shown, the angle of slope is referenced to end A. A slope is defined to be uphill when B and
C are above A; the angle of slope is defined as positive. A downhill slope is when B and C are
below A; the angle of slope is then negative.
The cable and pipe dimensions are those of the reference case from the sensitivity study in
Chapter 7 and are the same for both the pipe and duct systems. The details of the cable design
are: 138kV system voltage, with a 1500kcmilil circular copper conductor, XLPE insulation,
copper tape ground conductor, aluminum foil water barrier and polyethylene jacket. The cable
outer diameter is 83mm and the weight is 11.9kgm
-1
. The pipe internal diameter is 233mm
giving a clearance to the single cable of 150mm (reference case value).
Modelling Traffic Vibration
Figure 9-3 shows diagrammatically the action of the model when a truck drives along a road
from A to B, above the 200m length of pipe containing three cables. In this diagram the slope is
negative (downhill).

Figure 9-3
Pipe system: model of traffic vibration caused by a moving truck
The speed of the truck is set at 48kmh
-1
(30mph), this being equivalent to 13.3ms
-1
(43.7fts
-1
).
The vibration waveform of vertical velocity Figure 9-2 was converted to an applied vertical pipe
displacement by the following relationships ( Equation 92 to Equation 9-7) :
u = U sin t Equation 9-2
s = -S cos t = -(U/) cos t Equation 9-3
EPRI Licensed Material

Effects of Vibrations
9-7
S = U/ Equation 9-4
a = A cost = (U) cost Equation 9-5
A = U Equation 9-6
= 2f Equation 9-7
Where:
u: velocity with time [ms
-1
]
U: peak vertical velocity [ms
-1
]
s: vertical distance with time [m]
S: amplitude of peak vertical distance [m]
a: acceleration with time [ms
-2
]
A: amplitude of peak acceleration [ms
-2
]
t: time [s]
: angular frequency of vibrations [radians s
-1
]
f: frequency [Hz]
Defects on the road surface are taken to be distributed along the length. The impact of a wheel on
a defect causes the suspension to vibrate and impart a vibration to the road surface at 13Hz for
0.8s, spreading out downwards through the ground to force an amplitude of vibration on the pipe
of 0.032mm. At the end of the 0.8s period of vibration the truck will have traveled 10.7m at the
speed of 12.5ms
-1
. For modelling it is taken that:
the truck is 10m long
the wheel encounters

a succession of defects spaced at 10m
the truck causes the 10m section of pipe beneath it to vibrate
the 10m section of vibrating pipe moves along the route at the same speed as the truck
The pipes are modeled as rigid inelastic surfaces. The 10m section of vibrating pipe is modeled
as a separate piece of pipe superimposed on the main pipe such that the inner surfaces are exactly
coincident. The cable surfaces detect the geometric presence of both pipes.
When the sinusoidal waveform is applied to the 10m pipe it accelerates the cable upwards to a
minimum amplitude of 0.032mm. The pipe then descends, following exactly the specified wave
pattern. Momentum permits the cable to continue to rise until the gravitational retardation force
returns it to the original pipe surface in accord with Newtons laws of motion.
Table 9-1 records the results of a model sensitivity study, it shows that the cable continues to rise
off the pipe surface an additional 0.048mm to a height of 0.08mm. A range of vibration
amplitudes was applied. It was observed that the ratio of resultant cable displacement to pipe
displacement varied, this being attributed to the differences between the duration of the applied
cycle and the cable time of flight; these times would not be expected to be in synchronism.
EPRI Licensed Material

Effects of Vibrations
9-8
Table 9-1
Vertical displacement of model cable resulting from vibration of the duct
Duct
Vibration Peak Velocity
[ms
-1
]
Vibration Peak
Amplitude[mm]
Cable
Vertical Displacement
[mm]
0.0026 0.032 0.08
0.0082 0.1 0.2
0.0409 0.5 0.7
0.0817 1.0 1.8
0.1226 1.5 4.0
The 10m pipe descends on the negative half cycle to below its starting position, however the
cable is prevented from following it downwards by the static presence of the main pipe. Thus the
model cable experiences vibration from only the positive half cycles and can only be accelerated
through half the peak to peak height. In this respect the model cable is treated more gently than
in the real world. For future modelling applications the full vibration waveform can be applied
by the application of a modified sine wave together with a bias equal to the amplitude.
The 10m section of vibrating pipe was specified to travel along the 200m straight span from A to
B. The 10m cylindrical pipe section cannot traverse the 4.56m length of cable in the pipe bend
from B to C, because of the necessity to have curved pipe geometry. This had no consequence on
the outcome as:
the bend is only 2% of the span length
the effects of vibration occur long before the vibrating pipe section reaches the bend
pre-existing thermomechanical sidewall forces in the bend restrict vertical vibrational
movement by having lifted the cable to the 3-6
o
/c axis and pressing it to the outer surface
Load Cases
The model was constructed to the required angle of slope, the angles being varied between +7.5
o

and -7.5
o
.
The single cable was dropped onto the floor of the duct to fall in a neutral position. In the case of
the pipe, the three cables were dropped in trefoil formation.
The 138kV cable was then heated from a 15
o
C ambient to the cable operating temperature of
90
o
C to permit the cable to form the thermomechanical patterns studied in Chapter 7.
EPRI Licensed Material

Effects of Vibrations
9-9
The 10m section of vibrating pipe was then moved along the 200m straight span at the speed of
48km.h
-1
, in a time of 15s.
Output was requested in 40 steps, these each being equivalent to a) time steps of 0.375s and b)
incremental distances traveled by the truck of 5m. (As recorded in Chapter 5, software
computations are made in much smaller time steps).
At time zero the 10m long truck was parked outside the span. Thus in the first 5m time step it
had partially overlapped the span and applied vibrations to the first 5m. In the second time step
the truck had entered the span over its full 10m length. Thereafter vibrations were continually
applied to the cables by the smoothly moving 10m pipe section, with progress being reported by
output snapshots at 5m increments.
Avalanche Cable Movement
Avalanche Mechanism
The analysis of the model outputs revealed a new mechanism of cable movement, which was
found to occur suddenly and to affect a long length. The new mechanism is given the name of
avalanche movement, because of the similarity of a) release of locked-in strain by a small
vibration and b) cascade movement. It is of note that the use of the explicit FEA modelling
technique was key in discovering this dynamic mechanism.
Prior to an avalanche, the weight of snow is held in equilibrium by shear strength at the interface
with an underlying snow surface. The snow is prospectively impelled to move downhill by
gravity. In the case of the cable, the locked-in axial thermal strain is held in equilibrium by
friction with the pipe surface. The cable is prospectively impelled to move longitudinally from a
region of high axial force to low axial force. The region of low axial force is formed by release
of axial strain due to cable movement a) longitudinally into a bend, b) laterally into
thermomechanical deformation patterns and c) longitudinally and laterally into a joint or
termination casing. In the mechanical modelling sense, shear and frictional constraints are
closely similar, both having a longitudinal strength and a defined limiting value.
Sudden movement in the cable is initiated over a comparatively long distance (50-180m). In the
case of pipe and duct systems, the magnitude of cable movement is small and is limited by the
comparatively small absorption capacity within thermomechanical patterns and bends, both of
which are dictated by the cable to pipe clearance. The equivalent limit to movement in a snow
avalanche would be the very close proximity of an avalanche fence.
Response of Pipe and Duct Systems to Vibration
The behavior of the pipe system is explained diagrammatically in Figure 9-4, Figure 9-5 and
Figure 9-6 using the spring analogy of thermomechanical pattern formation from Chapter 2.
EPRI Licensed Material

Effects of Vibrations
9-10

Figure 9-4
Spring analogy of duct system at 90
o
C: before application of truck vibration
Equilibrium Before the Application of Vibration
The diagram in Figure 9-4 represents the 200m straight section of pipe in the hockey stick
shaped span, between end A and the start of the bend at B. The cable is shown as already having
been heated to a 90
o
C operating temperature. During initial heating, there is no scope for
movement in the straight pipe at A and so the axial force increases immediately with temperature
rise, producing the highest theoretical force F
1
for the particular cable construction. At B the
cable is permitted to expand and moves into the bend between B and C (not shown), when the
absorption capacity of the bend is full the axial force rises at the theoretical maximum rate. Thus
at any particular temperature, a difference in axial force is present along the 200m length
between A and B.
As the temperature is increased, the cable continues to behave as a straight rigid bar and
develops high locked-in thermal axial force along the span, this being illustrated in Figure 9-4 by
the stiff steel spring with tight coils at A, generating force F
1
.
At the entrance to the bend B the combination of axial thrust and cable curvature form incipient
wave patterns comprising a few half loops of low amplitude. At a certain critical bending
moment the patterns rapidly grow transversely to form a) large amplitude cylindrical sine waves
in the case of a pipe system and b) combinations of sine waves and helical loops in the case of a
duct system. The mechanism of pattern formation is explained by the minimum energy theorem.
The wave patterns in Z-B behave as a steel spring with low stiffness; being able to absorb high
strain at low axial force F
3
. In Figure 9-4 the spring Z-B is shown with widely spaced coils.
EPRI Licensed Material

Effects of Vibrations
9-11
However, unlike a real spring, Z-B behaves as a constant force device. The spring Z-B absorbs
the strain released from the adjacent rigid cable by increasing its length. In this respect Z-B is the
opposite of a steel spring, which under compressive force reduces its length in the storage of
more strain.
If the coefficient of friction between the cables and pipe is zero, then it will not be possible for a
force differential to exist along A-Z. All of the stored thermal strain in the rigid cable will be
released into the pattern at A until the force falls everywhere along the span to near F
3
. The
spring Z-B will grow in length to absorb the strain, Figure 9-6 (i.e. the thermomechanical
patterns grow along the pipe).
In reality the cable experiences friction with the pipe wall. Friction is illustrated in Figure 9-4 by
the notched surface, which holds each coil of the spring. The frictional constraint along the
cable length between X and Z is able to support a force differential of (F
1
-F
3
). For a horizontal
route, the gradient of the force differential
|
|
.
|

\
|
Z X
1 2
L
F F
is (W), where is the coefficient of
friction and W is the weight per unit length of cable (for a single cable). This gradient is always
constant for given weight and coefficient of friction. If the gradient is exceeded, cable slip will
occur and the released thermal strain will reduce the axial compressive force. In Figure 9-4 if the
force is increased each coil of the spring will gain sufficient energy to jump a defined number of
notches; thus releasing strain, reducing force and opening the distance between coils.
The cable between A and Z is still straight and rigid. It is represented by one spring of high
stiffness. The first part of the spring A-X is mechanically isolated from the rest of the span by the
frictional constraint and so it experiences maximum compressive force F
1
, illustrated by the coils
being tightly compressed. In the region of frictional constraint X-Z the axial force linearly
decreases and is illustrated by the coils being tightly compressed at X and the spacings becoming
wider with distance until Z is reached at force, F
3
.
In the particular case shown in Figure 9-4, there are two plateaus of constant force separated by
the frictional gradient. The plateau A-X is at maximum theoretical rigid bar force. The plateau at
Z-B is at a low force dictated by the low longitudinal stiffness of the thermomechanical patterns.
If the temperature is increased further, the increase in locked-in thermal strain along the whole
cable length between A and Z will attempt to increase the axial force F
1
in proportion. However
the temperature will be reached when there is insufficient length of cable between A and Z to
support any increase in the force differential. The gradient of the force differential cannot be
increased above (W) and so the cable will slip and feed thermal strain into the pattern at Z-B,
which will grow in length to absorb it. This is illustrated by the spring coils jumping notches and
feeding strain to Z. The thermomechanical models studied in this work indicate that in some
cases avalanche slip between A and B may already have occurred during normal cable heating,
as evinced by the presence of intermediate force plateaus.
At the constant temperature condition shown in Figure 9-4 a state of equilibrium exists between
the springs in the two cable plateau regions A-X and Z-B and in the transition zone of frictional
constraint separating them.
EPRI Licensed Material

Effects of Vibrations
9-12

Figure 9-5
Spring analogy of duct system: transition diagram at initial application of vibrations
Figure 9-5 illustrates the transition events occurring between the equilibrium diagrams Figure
9-4 and Figure 9-6. The events are triggered by the truck driving over part of the friction
gradient. The truck is shown overlapping the gradient length between X-Y. The vertical vibration
of the pipe momentarily lifts the cable off the surface, thus releasing the frictional constraint to
longitudinal movement and unbalancing equilibrium. This is the point when the avalanche is
initiated and the strain in cable X-Z suddenly slips into the thermomechanical pattern Z-B.
The locked in strain in A-X is redistributed over the longer length affected by vibration between
A-Y and reduces the plateau force from F
1
to F
4
. However the new gradient
|
|
.
|

\
|
Z Y
3 4
L
F F
is too
steep and exceeds the gradient of frictional constraint (W), thus the cable suddenly slips
towards Z to restore equilibrium. This is illustrated by the coil springs gaining sufficient force to
jump the notches.
EPRI Licensed Material

Effects of Vibrations
9-13

Figure 9-6
Spring analogy of duct system: after disturbance by vibrations
Equilibrium After the Application of Vibration
Figure 9-6 shows the cable in equilibrium after the avalanche has stopped. The release of stored
thermal strain by the avalanche into Z-B has reduced the force at Y and restored the friction
gradient between new positions Y
11
and Z
11
to (W). At the other end of the span the length of
the pattern between B-Z- has grown to Z
11
to absorb the released strain. Overall the following
changes have occurred:
The plateau axial force at A has dropped from F
1
to F
4

Locked in thermal strain has cascaded into the pattern at Z-B
The pattern has grown in length from B-Z to B-Z
11

The distribution of axial force has become more uniform
Angle of Slope
The slope of the route affects both the length of the thermomechanical pattern and the force
distribution before and after the application of vibration. The analogy with a snow avalanche is
not as helpful in these cases as snow only experiences a downhill gravitational force, whereas the
analogous prime mover for cable is the locked-in thermal strain and associated force distribution.
Thermomechanical force can impel the cable strain to move uphill.
It is more helpful to imagine the spring diagrams, Figure 9-4, Figure 9-5 and Figure 9-6, as
model springs mounted on a board that can be inclined to the horizontal.
EPRI Licensed Material

Effects of Vibrations
9-14
In each case the amount of thermal strain generated is the same for a given temperature rise. The
slope affects the proportion of locked-in strain to released strain.
Uphill Slopes
In Figure 9-4 the definition of an uphill slope is when the bend at B is above A. The angle to the
horizontal is defined as positive.
Before the cable is heated, the gravitational component of cable weight Wsin acting
downhill places the cable at B in tension and at A in compression
During heating and before a pattern forms at B, the cable behaves as a rigid rod in
compression. The thermal locked-in strain and resulting stress are distributed uniformly. The
gravitational profile of stress is superimposed in the linear thermal distribution. As the
compressive thermal strain increases, a) the tensile force at the top of the route B reduces and
becomes compressive and b) the magnitude of the compressive force at A increases further.
The same combination of compressive force and cable curvature is required to initiate the
pattern formation at B in each of the slope angles. However, a higher temperature will be
required to achieve the required magnitude in an uphill slope. This same temperature will
produce higher levels of compressive force and locked-in strain at A.
When the pattern forms, the force at B will fall and the strain will attempt to move from A
uphill into it. Equilibrium is reached when the frictional constraint equals (F
1
-F
3
) in Figure
9-4. However the slope angle effectively increases the constraint to movement. Compared to
a flat route, the coefficient of friction is increased from Wcos to W(cos+sin). Thus the
force differential that can be supported is increased. This steepens the gradient of frictional
force thereby a) reducing the amount of strain that can be released into B and b) retaining a
higher level of locked-in strain at A. The resulting thermomechanical pattern at B will be of
short length.
When vibration is applied to one part of the route containing the frictional force gradient the
cable is lifted off the surface and the local frictional constraint falls to Wsin. The local
gradient becomes shallower, thus releasing strain to flow uphill to B. The movement of the
lorry from A to B progressively releases stored strain from the steep frictional gradient until
equilibrium is reached.
The release of the higher level of stored strain permits the pattern to grow longer in length
towards A. The cable pattern has low internal longitudinal stiffness and so the free strain
within the transverse waves can move downhill to its lowest extremity, a) increasing the
amplitude of the downhill sine waves and b) draining strain out of the uphill patterns.
The result is that, of all the cases considered, vibrations permit the thermomechanical
patterns to grow to their longest lengths in the steeper uphill route angles.
Downhill Slopes
Before the cable is heated, the gravitational component of cable weight (Wsin) places the
cable at A in tension and at B in compression.
EPRI Licensed Material

Effects of Vibrations
9-15
During heating and before a pattern forms at B, the cable behaves as a rigid rod in
compression. As the compressive thermal strain increases a) the tensile force at the top of the
route A reduces and becomes compressive and b) the magnitude of the compressive force at
B increases.
The same combination of compressive force and cable curvature is required to initiate the
pattern formation at B. However a reduced temperature will achieve the required magnitude.
The low temperature will reduce the levels of force and locked-in strain at A.
When the pattern forms, the force at B will fall and the released strain will move from A
easily downhill into it. Equilibrium being reached when the frictional constraint equals (F
1
-
F
3
) in Figure 9-4. The slope angle reduces the frictional constraint to downhill movement.
Compared to a flat route the effective coefficient of friction is reduced from Wcos to
W(cos- sin). It can be seen that when sin> cos the resistance to movement becomes
zero and the cable will slide. For a coefficient of friction of =0.1 this occurs at 5.7
o
. Thus on
downhill slopes the thermomechanical force differential that can be supported is reduced.
This a) increases the proportion of strain that can be released into B and b) reduces the
locked-in strain and compressive force at A. Thus the thermomechanical pattern at B at the
bottom of downhill spans will be of longer length.
When vibration is applied to a part of the frictional force gradient, the cable is lifted off the
surface and the local constraint W(cos-sin) to downhill thermomechanical movement is
eliminated and instead becomes an assisting downhill force Wsin . The local gradient
becomes shallower, thus releasing strain to flow downhill into B. As there was little stored
strain before vibration and, as residual strain after vibration is easily released, the movement
of the truck from A to B releases stored strain almost instantly.
The release of the low level of stored strain permits the pattern at B to grow slightly in length
towards A.
The result is that, in a downhill route, there will be a lower level of thermomechanical strain
available to be released by vibration and thus the pattern increases by the shortest length of
the cases studied.
The above explanation is a simplification of model cable behavior. Behavior varies with a) the
cable temperature at which vibration is applied and b) the particular combination of cable
stiffness, cable weight, pipe clearance, coefficient of friction and span geometry.
In the case studied of the 138kV cable with pipe and duct clearances of 150mm, it was found that
a significant amount of thermal strain had already been released into the pattern. This is because
the 90
o
C temperature is significantly above the turn-over temperature at which
thermomechanical patterns form and limit the magnitude of the peak axial force. In consequence
the patterns were already of significant length and the longitudinal stiffness of the pattern region
had risen, such that the force was slightly higher than in the adjacent cable. The distribution of
axial force both before and after the vibration was found to be more complex. Nevertheless the
application of vibration to the start of the route had a significant and sudden effect on increasing
the length of the thermomechanical patterns and on altering the shape of the force distribution.
EPRI Licensed Material

Effects of Vibrations
9-16
These effects became dramatic for the steeper angles of uphill slope, as described in the
following section.
Results from Model Spans Subject to Vibration
The avalanche effect produced a similar degree of sudden movement in both duct and pipe
systems. These models all had the same 150mm clearance between the outer diameter of one
cable and the inner diameter of the duct or pipe. The presence of three cables in the pipe system
increased the frictional constraint on the bottom two cables, but still allowed the avalanche effect
to occur.
In the following sections the behavior of duct systems is considered first, followed by pipe
systems.
Duct and Pipe Systems: Summary of Effect of Vibrations on Different Slope
Angles
Table 9-2 and Table 9-3 show that traffic vibrations had a very significant effect on the change in
length of existing thermomechanical patterns at 90
o
C over the range of slopes likely to be
encountered in duct and pipe routes. The distances are given at which the truck had driven into
the route when the effect ceased. Not shown is that the sudden movement commenced
immediately the truck entered the route.
It can be seen that the percentage growth of thermomechanical patterns increases when the slope
angle in increased from -7.5
o
through 0
o
to +7.5
o
.
The avalanche effect is explained in detail below for each of the slope angles, starting with the
horizontal route span for a duct system.
Table 9-2
Effect of traffic vibrations on duct routes with different slope angles
Length of Thermomechanical Pattern Slope Angle
[degrees]
At 90
o
C
[m]
After Traffic
Vibrations [m]
Change in
Length [%]
Effect Ceased When Truck
Traveled Beyond
[m]
+7.5 (uphill) 38 181 376% 75
+3.0 42 90 114% -
0 (flat) 43 83 93% 20
-3.0 45 79 76% -
-7.5 (downhill) 47 70 49% 15

EPRI Licensed Material

Effects of Vibrations
9-17
Table 9-3
Effect of traffic vibrations on pipe routes with different slope angles
Length of Thermomechanical Pattern Slope Angle
[degrees]
At 90
o
C
[m]
After Traffic
Vibrations [m]
Change in
Length [%]
Effect Ceased When Truck
Traveled Beyond
[m]
+7.5 (uphill) 57 159 181% 100
0 (flat) 63 103 63% -
-7.5 (downhill) 69 80 16% 10
Duct System: Horizontal Span
Figure 9-7 shows that the truck only needed to travel a distance of 20m to initiate the avalanche
and cause the thermomechanical pattern to double in length at the opposite end of the route. No
further change occurred when the truck continued along the 200m route. In this case, the shape
of the wave patterns was unaltered, with the two helical loops remaining near to B. The doubling
of length occurred in the region containing sinusoidal waves.

Figure 9-7
Duct system (horizontal): thermomechanical patterns before and after 20m truck
movement
Figure 9-8 shows that at 90
o
C, before the truck arrived a) the force differential along the cable
had already reached end A, such that no plateau existed and b) the force in the pattern at B had
already risen above the adjacent cable. The avalanche caused by the vibration is shown to have
a) dropped the force at A by 33%, b) reduced the minimum force mid-route and c) increased the
force in the pattern. Overall the momentary reduction in friction by traffic vibration released
thermal strain and produced a more uniform force distribution with lower magnitude.
EPRI Licensed Material

Effects of Vibrations
9-18

Figure 9-8
Duct system (horizontal): axial force before and after 200m truck movement
Figure 9-9 shows the length of thermal strain released by the cable a) during the period of
heating from 15-90
o
C and b) with the distance traveled by the truck. If the cable had been in a
tight fitting duct in a straight span with no bends there would have been a zero release of thermal
strain. The resulting maximum locked-in strain of 261mm would have produced a force at 90
o
C
of 25,880N, significantly higher than the 6,653N maximum in Figure 9-8
Upon heating, thermal strain is released a) by the cable moving into the bend and then b) when
patterns form and absorb strain. The total strain released is 207mm, 79% of the maximum
possible of 261mm. The passage of the truck over the short distance of 20m releases another
16mm, making 85% in total. It is noteworthy that the release of such a small length of strain had
a large effect on changing the pattern length and axial force distributions, as shown in Figure 9-7
and Figure 9-8.
EPRI Licensed Material

Effects of Vibrations
9-19

Figure 9-9
Duct system (horizontal): release of cable free length during a) heating to 90
o
C and b)
200m truck movement
Duct System: Uphill Spans
Uphill Span +7.5
o

Before the Truck Applies Traffic Vibration
Figure 9-10 and Figure 9-11 show a plan view of the distribution of thermomechanical patterns
along the 200m span with an uphill slope of 7.5
o
. Figure 9-10 compares 16 positions of the truck
(one before and 15 after the truck has moved in 5m increments). The top pattern is at 90
o
C,
before the truck has entered the span (the truck can be seen parked on the left adjacent to A). A
thermomechanical pattern has already formed.
Figure 9-11 shows the details of 4 selected patterns (one before and the others after 40m, 55m
and 75m of movement).
Figure 9-12 is the distribution of axial force along the span before and after the passage of the
truck.
EPRI Licensed Material

Effects of Vibrations
9-20

Figure 9-10
Duct system (+7.5
o
uphill route): growth of thermomechanical patterns as truck drives
along span

Figure 9-11
Duct system (+7.5
o
uphill route): details of thermomechanical patterns before and after
truck moves 40m, 55m and 75m
At 90
o
C the initial length of the thermomechanical pattern is 38.4m, this being slightly shorter
than the 43m length in the flat route. However the flat route has two pronounced helical loops
whereas the 7.5
o
route has two loops that are only part formed. The difference in pattern length
EPRI Licensed Material

Effects of Vibrations
9-21
and helical loop shape are significant, showing that there is significantly less strain stored. This
is because the cable is restricted from moving uphill into the thermomechanical pattern at B. The
upward thermal expansion is opposed by the downward component of cable weight, this doubles
the effective coefficient of friction from the specified value of 0.1 to 0.23 (in comparison with
the flat route).

Figure 9-12
Duct system (+7.5
o
uphill route): axial force before and after 200m truck movement
Figure 9-12 shows the force distribution at 90
o
C. It can be seen that the forces are higher, in
consequence there will be a higher proportion of locked-in strain compared to the flat route
shown in Figure 9-8:
the force at the bottom of the hill at A is 34% higher at 8,884N
the force at the top of the hill is similar at 4,221 N, being only 7% lower than 4,534N.
the friction gradient of 4,663N is the subtraction (8,884N-4,221N) and extends over a 178m
length compared to the flat case of 2,119N (6,653N-4,534N) over 180m. The force gradient
is thus 2.2 x steeper than in the flat route, which agrees well with the theoretical ratio of
frictional gradient of 2.3.
When the Truck Applies Traffic Vibration
Figure 9-10 and Figure 9-11 shows that strain is released into the thermomechanical patterns
causing them to grow when the truck overlaps the route by only 5m. The growth continues
linearly in proportion to the distance traveled by the truck until it has moved 75m, after which
the patterns no longer grow.
EPRI Licensed Material

Effects of Vibrations
9-22
In total, the release of strain by the vibrations has permitted the pattern length to increase from
38m to 181m, extending almost over the complete route length of 200m. In consequence the
axial force has fallen from a maximum of 8,884N to a near uniform, flat distribution at 3,700N.
It can be seen in Figure 9-11 that the thermomechanical patterns have significantly changed their
shape during the progress of the truck. Before the truck enters the span, there are large amplitude
sine waves and partly formed helical loops at B. After it has reached 75m the pattern at B has
lost the helical loops and the sine waves have become smaller in amplitude. At the downhill end
however the newly formed sine waves are larger in amplitude. The pattern shape reflects the
distribution of strain released by the vibrations. Before the truck drove over the span, the strain
distribution comprised two components, gravitational and thermomechanical. In both cases the
strain is higher at the bottom of the hill, as shown by the force distribution in Figure 9-12. The
distribution of the sine wave amplitudes reflects the original strain distribution.
Overall, a steep uphill slope appears to be a unique case in which there is a higher level of
locked-in strain available to be released. The initial release is still a sudden event, which is
followed by a progressive and linear release of strain at the same speed as the advancing truck.
The traffic vibration very efficiently removed the locked in strain and produced the most uniform
distribution of thermomechanical patterns within the study.
Uphill Span +3
o

The downhill component of the cable weight is smaller than the 7.5
o
degree case such that the
effective coefficient of friction is reduced from to 0.23 to 0.15. The effect of the vibration is very
similar to that of the flat slope. The length of the pattern is increased from 42m to 90m. Figure
9-13 shows that the distribution of axial force along the span is of a closely similar shape and
magnitude to that in Figure 9-8.
EPRI Licensed Material

Effects of Vibrations
9-23

Figure 9-13
Duct system (+3.0
o
uphill route): axial force before and after 200m truck movement
Duct System: Downhill Span
The critical angle of slip occurs when the tangent of the slop angle is equal to the coefficient of
friction. In the cases studied the coefficient of friction is 0.1 and thus the critical sliding angle
under gravity alone is 5.7
o
.
The results of two cases are given below, one below 5.7
o
and one above.
Downhill Span -7.5
o

The start of the span at A is higher than the bend at B. The -7.5
o
slope angle is greater than the
critical sliding angle for the coefficient of friction of 0.1 requiring that the cable be anchored
during installation and jointing to prevent movement. After jointing is completed and the
temporary constraint has been removed, the cable would be free to slide downhill and attempt to
fill the storage capacity of the bend at B.
Before the Truck Applies Traffic Vibration
During heating the cable initially behaves as a rigid rod and generates the maximum theoretical
axial force which is superimposed on the gravitational component of cable weight.
EPRI Licensed Material

Effects of Vibrations
9-24
Inspection of the nearly flat force distribution in Figure 9-14 and comparison with the reducing
thermomechanical force distribution in Figure 9-8 confirms that the difference is the
gravitational profile. The gravitational profile comprises a tensile force at A (positive force),
rising to a compressive force (negative) at B. It is indicated that thermal locked-in strain is
available to be released.

Figure 9-14
Duct system (-7.5
o
downhill route): axial force before and after 200m truck movement
Figure 9-14 shows that compared to the flat route in Figure 9-8, the force is decreased at A from
6,653N to 3,764N and at the bend B, is increased from 4,534N to 5,531N. Thus a) the
thermomechanical pattern will form at B at a lower temperature and b) thermal strain will be
released into the pattern because the cable experiences low frictional constraint. Friction still has
an effect as it reduces the downhill component of force from 0.13W for a frictionless surface by
77% to 0.03W for a coefficient of friction of 0.1 (W: weight of cable per meter).
EPRI Licensed Material

Effects of Vibrations
9-25

Figure 9-15
Duct system (-7.5
o
downhill route): details of thermomechanical patterns before and after
truck has moved 20m
At 90
o
C the 47.3m thermomechanical pattern, Figure 9-15, is the longest pattern in the +7.5
o
to -
7.5
o
range (the pattern in the flat route is 43m). Significantly two fully formed helical loops are
present and these will have absorbed more strain than the cylindrical sine waves, thereby limiting
the length of the sine wave patterns.
When the Truck Applies Traffic Vibration
After the truck has entered the route, Figure 9-15, and traveled only 15m, the vibrations have
released the available thermal strain and no longer have an effect. The release of strain has a)
grown the pattern length to 70m and b) produced a third helical loop. Figure 9-14 shows that the
release of strain has reduced the axial force distribution at A by 44% to 2,105N and at B by 15%
to 4,703N. The force distribution is now seen to follow the gravitational downhill profile,
indicating that the thermomechanical strain is more uniformly distributed and is of reduced
magnitude.
Downhill Span -3
o

The slope is below the critical angle of slip of 5.7
o
. The downhill component of cable weight
reduces the effective coefficient of friction from 0.1 to 0.05 compared to the flat route.
The possibility exists that during installation and jointing, traffic vibration could lift a sufficient
proportion of the cable off the pipe floor and initiate sliding. Calculation shows that 132m of the
EPRI Licensed Material

Effects of Vibrations
9-26
200m length would have to be lifted before cable slid downhill. Although this may seem to be
unlikely, the possibility exists that in heavy traffic a) 13 x 10m long trucks could simultaneously
produce vibrations on a rough road surface and b) some types of ground may be able to transmit
vibrations to more than 10m of pipe, thereby requiring fewer trucks. It is good practice to anchor
the cable during installation and jointing to eliminate this possibility.
The length of the pattern at 90
o
C is 45m and is increased to 79m when the truck enters the span.
Figure 9-16 shows that the distributions of axial force before and after the truck vibrations are
closer to those for the -7.5
o
slope than for the flat route. The vibrations reduce the force at A by
33% to 3,577N and at B by 17% to 4,703N.

Figure 9-16
Duct system (-3.0
o
downhill route): axial force before and after 200m truck movement
Duct System: Repeat Truck Journeys
Three repeated truck journeys were made from A to B along a downhill slope with an angle of -
1
o
. Figure 9-17 shows that in the first journey the pattern length increased from 96.3m to 123.4m.
The second and third journeys had no further effect on either the pattern length or shape.
EPRI Licensed Material

Effects of Vibrations
9-27

Figure 9-17
Duct system (-1.0
o
downhill route): thermomechanical patterns before and after 1, 2 and 3
truck journeys along span
Pipe System: Horizontal Span
Before the Truck Applies Traffic Vibration
Figure 9-18 in 3D view at 90
o
C shows that there is a 62.9m long thermomechanical pattern
present in each of the three cores adjacent to the bend at B. The patterns are comprised of
cylindrical sinusoidal waves. The presence of three cables prevents the formation of helical
loops, this being one of the reasons why the pattern is of longer length than in the duct, Table 9-2
and Table 9-3. The cables have been pushed to the outside of the bend in vertical alignment. The
sidewall forces prevent the formation of patterns.
EPRI Licensed Material

Effects of Vibrations
9-28

Figure 9-18
Pipe system (horizontal): thermomechanical patterns at 90
o
C before truck movement
Figure 9-19, Figure 9-20 and Figure 9-21 show the distributions of axial force in each of the
three cables. The distributions and magnitudes of force are similar in the bottom two cables in
the trefoil formation. These distributions have a similar shape to the single cable in the duct,
Figure 9-8, with the axial force being high at end A and falling to the thermomechanical pattern
at B. The magnitudes are higher in the pipe than in the duct system reflecting a) the increased
frictional constraint and b) the greater longitudinal stiffness of the sine patterns. The axial force
in the top cable (number 1) has a nearly flat distribution of lower magnitude.
EPRI Licensed Material

Effects of Vibrations
9-29

Figure 9-19
Pipe system (horizontal): axial force in cable 1 (top) before and after 200m truck movement

Figure 9-20
Pipe system (horizontal): axial force in cable 2 before and after 200m truck movement

EPRI Licensed Material

Effects of Vibrations
9-30

Figure 9-21 Pipe system (horizontal): axial force in cable 3 before and after 200m truck
movement
When the Truck Applies Traffic Vibration
Figure 9-22 shows that the vibrations from the truck released thermal strain into the patterns at B
and permitted the length to grow suddenly to 103.2m.
Figure 9-19, Figure 9-20 and Figure 9-21 show that the vibrations have released thermal strain in
all three cores as the magnitudes of force have fallen significantly. The distribution of force is
similar in the bottom two cables (number 2 and 3) and is also of similar shape to the case of the
single cable in the duct system.

EPRI Licensed Material

Effects of Vibrations
9-31

Figure 9-22
Pipe system (horizontal): thermomechanical patterns at 90
o
C after truck movement along
200m span
Pipe System: +7.5
o
Uphill Span
Figure 9-23 shows the line plot of the three cores over the complete 200m route length. The
lateral scale has been increased by 10:1 to magnify the deflection of the thermomechanical
patterns in the three cables. The pattern is seen to be 56.6m long and, as with the +7.5
o
duct
system, is shorter than the pattern in the horizontal route.
The arrival of the truck in the span at A initiates an immediate growth in the pattern, which
ceases when the truck has traveled 100m (comparing to 75m in the +7.5
o
duct system). The
181% increase in pattern length to 159m is large, this compares with 376% in the +7.5
o
duct
system.
EPRI Licensed Material

Effects of Vibrations
9-32

Figure 9-23
Pipe system (+7.5
o
uphill route): details of thermomechanical patterns before and after
truck has moved 100m
Pipe System: -7.5
o
Downhill Span
Figure 9-24 shows that the thermomechanical pattern at 90
o
C is longer at 68.7m than that in the
flat route. The sudden release of thermal strain ceases after the truck has traveled only 10m
(compared to 15m in the -7.5
o
duct system). The resulting increase in pattern length to 80m is
small at 16%.
EPRI Licensed Material

Effects of Vibrations
9-33

Figure 9-24 Pipe system (-7.5
o
downhill route): details of thermomechanical patterns
before and after truck has moved 10m
Conclusions
The finite element explicit modelling technique has been successfully adapted to model cable
spans in its full high speed dynamic mode. Traffic vibrations of 0.032mm amplitude and 13Hz
frequency, for a duration of 0.375s, were applied to replicate the suspension hop when a truck
wheel encounters a defect in the road surface. The truck was modeled to travel above the pipe
span at a speed of 48kmh
-1
(30mph) with defects being present at 10m spacings along the road
surface (for example resulting from potholes or from movement and joints between concrete
slabs. The data on traffic vibrations was abstracted from measurements made by the UK
Transport and Road Research Laboratory from real roadside situations and from a road test track.
The vibrations were measured at the foundation level of buildings. This was the closest location
that could be found to a buried pipe or duct system beneath, or parallel, to a road.
The following conclusions were reached:
The application of traffic vibrations to both pipe and duct systems heated to the 90
o
C XLPE
cable operating temperature resulted in the sudden release of thermal strain and growth of
thermomechanical patterns. The discovery and investigation of this mechanism is believed to
be new and has been named the avalanche effect.
The avalanche effect depends on the presence of a non-uniform distribution of thermal strain.
The distribution results from the presence of a) high magnitude locked-in thermal strain and
axial force within a straight section of pipe at one end of the span, b) low level axial force
within a thermomechanical pattern at the opposite end of the span adjacent to a bend and c) a
EPRI Licensed Material

Effects of Vibrations
9-34
gradient of frictional constraint separating the regions of high and low force. The traffic
vibration lifts a short length of cable off the duct or pipe surface, momentarily reducing the
frictional constraint and unbalancing equilibrium. This results in the sudden release of strain,
which cascades along the cable into the thermomechanical pattern, which grows in length to
accommodate it. The mechanism is closely similar to a snow avalanche except that the
distance that can be traveled by the released cable strain is very much shorter, being
restricted to the small absorption capacity of a) deformation patterns, b) bends and c) space
within cable joint and termination casings.
In the reference case studied the thermomechanical patterns had formed at low temperature
and by 90
o
C had already released and absorbed the largest proportion of locked-in thermal
strain. However the release of the small percentages of residual strain was still dramatic in
suddenly lengthening the thermomechanical patterns and reducing the magnitude of axial
force.
The avalanche effect occurred at each of the span angles studied. The effect increased when
the angles were increased progressively from the downhill angle of -7.5
o
, through horizontal,
to the uphill angle of +7.5
o
. (Uphill is defined when the end of the span containing the bend
is above the end with the straight section). At +7.5
o
, the pattern grew in length from 38m to
181m. For the flat route the pattern grew from 43m to 83m. The difference is explained by
there being a greater proportion of residual thermal strain available to be released in the
uphill route. In the downhill span a higher proportion of the thermal strain had already been
released into the thermomechanical pattern with the assistance of gravity.
The pattern growth commenced suddenly when the truck producing the traffic vibration first
entered the route. The avalanche effect commenced within the smallest time frame of 0.375s,
equivalent to a 5m truck movement. The release of thermal strain ceased according to the
route angle, the maximum coinciding with a truck journey of 75-100m for the uphill duct and
pipe systems and the minimum being 10-15m for the downhill systems.
Repeated truck journeys had no effect in releasing additional thermal strain.
The release of strain and the degree of movement is likely to be more severe close to the
cable temperature at which the maximum locked-in thermal strain is stored. This is after the
formation of incipient thermomechanical half waves and immediately before the formation of
large amplitude waves.
To trigger the avalanche the truck does not have to travel above or parallel to the cable
span, but could simply cross it at a road junction. This is based on the effect occurring after
the truck has moved 5m (or less). However, the truck does have to pass over part of the route
containing the axial force differential. In the cases studied with a low coefficient of friction
of 0.1 this occupied typically 50- 180m. With higher coefficients of friction the distance is
reduced.
Study of the axial force distributions have revealed force plateaus which may indicate the
avalanche effect can also occur during normal circuit heating. This would not have been
detected by in previous modelling studies because the time steps were too large to capture
dynamic events.
EPRI Licensed Material

Effects of Vibrations
9-35
The question arises whether the avalanche effect is harmful to the cable system? The most
vulnerable position is at joints and terminations. The avalanche has two effects a) there will
be an impulsive change in force, positive or negative, depending where the joint is located
and c) additional thermal strain will be released to move into the joint casing or manhole. If
the insulation of the joint is sufficiently disturbed then electrical failure may be initiated.
Design measures are to reinforce the accessory and cable within manholes to a) prevent
longitudinal and lateral movement and b) withstand the forces generated.

Having reinforced the accessory design, the traffic vibration can be seen to be beneficial as it
produces a lower and more uniform distribution of axial force.

An area of risk is that movement of cable will be initiated by traffic vibration during
installation and jointing. This has not been studied by modeling. The analysis of the thermal
work shows that the risk of avalanche is increased by a) steeper slopes, b) length of road
containing surface defects and c) presence of force differentials, for example from residual
pulling-in strain and changes in ambient temperature between pulling-in and jointing. It is
good practice to anchor the cable after pulling-in and during jointing.




i
Toirkens E and Bucea G, Core Creepage in 330kV Oil Filled Cables, CIGRE Australian National Committee and
Panel 21, Third Cable Technology Seminar, 1997, Sydney
ii
Watts GR, Traffic Induced Vibrations in Buildings, Research Report 246,Transport and Road Research
Laboratory, 1990, ISSN 0266-5247


EPRI Licensed Material
10-1
10
EFFECT OF SHORT CIRCUITS
Introduction
This chapter studies the effects of short circuits on the thermomechanical behavior of cables in
pipes and ducts. Short circuit currents generate electrodynamic forces between the cables. In pipe
and duct systems the cables are free to move in both the lateral and longitudinal directions during
a short circuit.
Cable circuits are part of the transmission network and are designed to be able to safely carry the
rated short current in the event of a system through-fault. One design level of short circuit
current is usually specified for all of the cable circuits in a particular transmission system. In
practice the current level is determined by the circuit impedance such that a cable outlet from a
power plant will experience the highest current, whilst a distant cable will experience the lowest.
As a guide, the rated short circuit level is typically 30 x the continuous rated current. A cable
remote from the power plant will experience one third to one half of this value. The interaction
between the current in the reference conductor and the magnetic field from the two adjacent
phase conductors produces a force proportional to the square of the current. Thus the maximum
electrodynamic force during a short circuit will be up to 900 x higher than in normal service
operation. The time duration of the short circuit is a specified design parameter, usually related
to the clearance time of the circuit protection, usually specified within the range of 0.1s and 60s.
The force is a vector quantity in both direction and in time phase and so each of the three cables
experiences a force pulsating at either the fundamental, or a harmonic of, the 60Hz frequency.
In the study of short circuit forces producing mechanical forces at the rate of 120 impulses per
second, the capabilities of the explicit FEA software are used in full dynamic mode to solve
this high speed problem. The expressions for the time varying electromagnetic forces were
derived and were programmed into the model for each of the three cables, so that each cable
could interact with each other.
Pipe and Duct Spans Models
The reference model from the sensitivity study in Chapter 7 was selected, it comprises a 200m
hockey stick shaped span, Figure 10-1. The model contains either three single ducts or three
cables in one pipe. The reference cable is a 138kV 1500kcmilil cable with a diameter of 83mm.
The cable to pipe clearance is 150mm. The internal diameter of the pipe or duct is 233mm. The
parameters of the cable are the reference set listed in Chapter 7, with a cable to cable and cable
to pipe coefficient of friction of 0.1. The bend in the model has a 35Dc bending radius, which
EPRI Licensed Material

Effect of Short Circuits
10-2
for a cable diameter of 83mm gives a radius of 2.9m. A 1m straight section was added at the end
of the bend for the purpose of modelling.
The short circuit force is inversely proportional to the distance between the conductors and so it
is necessary to model the thermomechanical and short circuit performance of three parallel ducts
simultaneously. During the short circuit each cable is free to move. The vectors of force will also
change in direction, thus the behavior of the three cables in the model is interactive. This is also
true for the interaction between the three cables in the pipe system.
The following models were constructed and solved:
Three ducts in flat formation with a centerline spacing of 300mm.
Three ducts in trefoil formation with a centerline spacing of 300mm.
A pipe system containing three cables with an at-rest centerline spacing of 83mm

Figure 10-1
Diagram of model route
Not to scale
EPRI Licensed Material

Effect of Short Circuits
10-3
Short Circuit Force
Basic Theory of Short Circuit Forces
The expression for the force experienced by the conductor is given in Equation 10-1, for the
diagram shown in Figure 10-2.

Figure 10-2
Short circuit force diagram

BA A A
H F =

Equation 10-1
Where:
F
A
: force per unit length experienced by the reference conductor A [Nm
-1
]
I
A
: current in the reference conductor A [A]
H
BA
: magnetic field strength at the reference conductor A produced
by a separate source at B [Am
-1
]
: magnetic permeability of surrounding medium [Hm
-1
]
The permeability of XLPE insulation is that of free space: = 4 10
-7
[Hm
-1
]
The strength of the magnetic field around a current carrying conductor is given by Equation
10-2:
EPRI Licensed Material

Effect of Short Circuits
10-4
BA
B
BA
r 2
H

=

Equation 10-2
Where:
H
BA
: field strength produced by I
B
at r
B
[Am
-1
]
I
B
: current at B [A]
r
BA
: radial distance between B and A [m]
The case chosen for study was that of a symmetrical three phase fault. For example when a cable
circuit feeds a three phase fault on an overhead line. The three phase fault was chosen in
preference to a single phase fault to ground, in which the zero phase impedance of the cable to
ground-return circuit depends upon the particular local conditions. In the case of a single phase
cable fault the current will predominantly return in the sheaths and metallic pipe, depending
upon the type of sheath bonding, with a small proportion in the ground return. The resulting
cable forces and movements will be smaller in a single phase fault. The reasons for studying the
three phase fault are that the force is higher and that the impedances are a) determined from
known cable geometry, b) independent of the surroundings and c) readily calculable.
In a balanced three phase fault the currents are time displaced by 120
o
(2/3). The subscripts A,
B and C refer to the cable phase designation shown in Figure 10-2.
t sin i
A A
=

Equation 10-3

|
.
|

\
|
+ =
3
2
t sin i
B B



Equation 10-4

|
.
|

\
|
+ =
3
4
t sin i
C C


Equation 10-5
Where:
i
A
: instantaneous magnitude of alternating current in conductor A [A]
I
A
: amplitude of peak alternating current in conductor A [A]
: = 2f: angular frequency of alternating current [radians.s
-1
]
f: supply frequency [Hz]
Application of Short Circuit Force to the Model Cables
The model of the cable span used the radial thrust connector utility of the ABAQUS software
to connect each pair of nodes between parallel cables. The radial thrust connectors were
programmed to apply the electrodynamic force to each of the nodes in the model, dependent
upon their relative positions, they imposed no constraint on the radial or longitudinal movement.
EPRI Licensed Material

Effect of Short Circuits
10-5
The electrodynamic forces are vectors having two components, spatial and temporal
displacement.
Temporal Displacement of Current
For parallel currents flowing in the same direction, the direction of force of
BA
F acting on A is
towards B. If the current in B flows in the opposite direction (anti-phase) then the direction of the
force is to repel each of the conductors. The phase displacement between the currents is
important. In the balanced three phase case any two currents are 120
o
displaced, ie nearly in anti-
phase (180
o
), thus the force will be predominantly repulsive.
Software script was written to apply force to each of the three pairs of cable from Equation 10-6,
Equation 10-7 and Equation 10-8. The equations are for a repulsive force along the axis
connecting the two cables.
The currents are balanced and thus I=I
A
, I
B
, I
C.
Each force is time displaced. This explains the phenomenon seen in high speed photography of
short circuit testing, in which each cable is kneaded in turn by the rotating magnetic field.
( ) |
.
|

\
|
+

=
3
2
t sin t sin
r 2
F
BA
2
BA

Equation 10-6
( ) |
.
|

\
|
+

=
3
4
t sin t sin
r 2
F
CA
2
CA

Equation 10-7
|
.
|

\
|
+ |
.
|

\
|
+

=
3
4
t sin
3
2
t sin
r 2
F
CB
2
CB

Equation 10-8
Spatial Displacement of the Conductor Positions
The cables will not be in straight alignment, because even in flat formation they are free to move
within the ducts, thus vector spatial addition must be applied according to the geometries shown
in Figure 10-2. Note that in the figure the forces F
A
, F
B
and F
C
are in the directions of the
resultant forces, following the vector additions of for example
CA BA
F F + . The direction of force
is continuously recalculated by the ABAQUS software according to the equations of motion.
Verification of Modeled Short Circuit Forces
The spatial and temporal vector equations were solved separately to check the output of the
ABAQUS solver, Equation 10-9:
CA BA A
F F F + = Equation 10-9
EPRI Licensed Material

Effect of Short Circuits
10-6
Cables in trefoil formation, a) three cables in a pipe and b) three ducts.
Equation 10-10
The value of the prefix constant in Equation 1010 of
|
|
.
|

\
|

4
3
is 1.73 x 10
-7
, this is in agreement
with reference
[i]
. Of particular interest is that the force is outwards in direction and that the
( )
2
t sin term is that of a unidirectional impulsive force, with 120 pulses per second. The
average force may be obtained from the integration of the expression giving a factor of 21%, this
reduces the prefix constant to 0.36 x 10
-7
.
Ducts in flat formation.
Equation 10-11
The force on the centre cable Equation 10-11 is initially in the direction 2 to 1 and is alternating
in direction between 1 and 2 at a frequency of 120Hz. As a transmission class cable has
significant weight and inertia, the centre cable will tend to remain stationary.
The outer cables experience a force with a more complex waveform, the peak value of which is
given by the additional factor of 0.93, thus reducing the prefix constant
[i]
to that in Equation
10-12.
Equation 10-12
The trigonometric expression for the outer cables can be approximated by a ( )
2
t sin term,
showing that, as in the trefoil case, they experience unidirectional repulsive forces at 120 pulses
per second.
Load Cases
The following load cases were applied:
Position cables in neutral starting position by dropping them 1mm onto the pipe floor under
gravity.
Heat from 15
o
C to 90
o
C with 10 computational steps of 7.5
o
C.
( )
2
AB
2
A
t sin
r 4
3
F


|
|
.
|

\
|
|
|
.
|


\
|
=
( ) t 2 sin
r 4
3
F
AB
2
B



|
|
.
|

\
|
|
|
.
|

\
|
=
|
.
|


\
|
=

AB
2
7
A
r
10 x 61 . 1 F

EPRI Licensed Material

Effect of Short Circuits
10-7
Apply the short circuit forces to each pair of cables for a duration of 0.1 s (6 supply
frequency cycles), with 10 computational steps of 0.01s.
Remove the short circuit forces and allow a period of 2s for the cable to regain equilibrium,
with 10 computational steps of 0.2s.
Electricity Utilities provided examples of their short circuit levels:
Utility X: 50kA design with 40kA achievable.
Utility Y: 40kA design with 23kA achievable.
During the development of the technique, the short circuit forces were initially simplified to be
constant unidirectional forces for the 0.1s duration. This is a reasonable simplification as the
cables are heavy and will not respond instantly to rapid fluctuations in force. In a following
development step sinusoidally varying forces were applied. Finally the expressions for the time
varying and phase displaced forces were programmed, Equation 10-6, Equation 10-7 and
Equation 10-8. The forces were always applied as vector quantities in line with the axis between
each pair of cables.
The recommended future stage of development is to make the forces a variables of the distance
between the cable centerlines. The use of the at-rest spacings in the reported work was
considered to be a reasonable simplification as a) in the pipe system movement from the atrest
spacing of 83mm was restricted to less than plus or minus 75mm and b) in the duct systems each
cable could move a maximum distance of plus or minus 75mm in relation to the duct centerline
spacings of 300mm.
The following peak currents and forces were applied:
70.7kAp (50kA rms), 0.1s, time varying peak force of 1,933Nm-
1
between cable pairs.
28.3kAp (20kA rms), 0.1s, time varying peak force of 309Nm
-1
between cable pairs
Installation Formations and Slope Angles:
Ducts in trefoil
Ducts in flat
Three cables in pipe
Slope angles at 50kA: +7.5
o
, +3
o
, 0
o
, -3
o
, -7.5
o
.
Slope angles at 20kA: +3
o
, 0
o
, -3
o
.
Description of Thermomechanical and Electrodynamic Effects
In the following sections the most complete descriptions and explanations of the interactions
between the electrodynamic and thermomechanical effects is given for the 50kA short circuit
case applied to the horizontal models. These phenomena were found to be common to each of
the configurations, to the different angles of slope and to the 20kA short circuit level. However
EPRI Licensed Material

Effect of Short Circuits
10-8
there are important differences that add to the overall understanding. For the sake of brevity the
numbers of figures and tables of results have been reduced and limited to those that illustrate the
key differences.
Ducts in Flat Formation: 50kA Flat Route
The application of the short circuit is shown to be a violent event, which generates major
movements and changes in both axial compressive force and lateral sidewall pressure. The effect
on the region containing the naturally formed thermomechanical pattern is dramatic. The lifting
of the cable off the pipe floor permits the patterns to grow along the complete 200m length of the
cable in 1.5 seconds after the end of the short circuit. The cable experiences a significant shock,
during and after the short circuit.
Table 10-1 records the increase in the lengths of the thermomechanical patterns that occur when
the cables are heated from 15
o
C to 90
o
C.
To permit small output time steps to be requested during and after the short circuit, the
temperature output steps were increased to 7.5
o
C. Although the temperature and force values are
accurate, care is required in interpreting the exact temperature and force at which patterns form
during heating. Details of the events are given in chapter 6, using 2.5
o
C steps.
Table 10-1
Thermomechanical pattern lengths-Ducts in flat formation 50kA case
Temperature [C] Load Case Average Pattern Length [m]
67.5 Heating 8.4
75 18
82.5 29.1
90 Before short circuit 43.5
90 During short circuit 47.4
90 2s after short circuit 200
Figure 10-3 is a plan view of the three ducts showing the thermomechanical patterns that were
present at 90
o
C in the 45m length closest to the bend, before the application of the short circuit.
The lateral scale has been expanded in the ratio of 4:1 to permit the patterns to be seen.
EPRI Licensed Material

Effect of Short Circuits
10-9

Figure 10-3
Ducts in flat formation. Thermomechanical patterns at 90
o
C in 45m length next to bend
Plan view, scale 4:1
Figure 10-4 is a three dimensional visualization of the three cables before the application of the
short circuit. To permit the thermomechanical patterns to be seen a) only the last 150m of the
200m straight span is shown and b) the straight section is foreshortened by taking a downward
projection view looking from B to A; the foreshortening ratio being 37:1. This has had the effect
of giving the bend radius of 2.9m the appearance of a right angled bend. The 3D view shows that
the patterns in each cable have the same lengths, but with slightly different shapes. This is
attributed to the different bending radii of the three ducts in flat formation as they pass around
the bend. Helical loops are present in the 3D view which appear in the plan view to be longer
wavelength sine waves. Where the cables leave the bend they form one sine wave, followed by
two helical loops and then five further sine waves, extending over a distance of 43.5m.
The cables in the bend remain straight. They are not free to form patterns because they are
pressed outwards by high axial thrust from the 200m straight section. In the bend where the
cables approach the 1m straight section at end C, they descend to the bottom of the duct and then
lift again at C. This movement results from the type of end boundary which represents the
mirrored span geometry of a continuing span.

EPRI Licensed Material

Effect of Short Circuits
10-10

Figure 10-4
Ducts in flat formation before the 50kA short circuit, position B
Plan projection view, 1:1 scale
Figure 10-5 is the view during the 50kA short circuit along the last 150m length of ducts towards
point B at the entrance to the bend. Figure 10-6 is the plan view at the same moment. The views
are after 0.1s of short circuit current have been applied and immediately before the short circuit
is removed.
The cables in the straight duct are no longer in horizontal alignment and have been moved by
the electrodynamic forces as far apart as possible. The two outer cables are being forced
outwards and upwards and the centre cable is being forced downwards. It is possible to see
small gaps between the centre cable and the roof and sides of the pipe showing that the cable
pattern has been compressed to a smaller diameter.
The thermomechanical wave patterns have been significantly changed by the short circuit
forces. The outer cables, which experience the highest forces have each lost the two helical
loops, such that it is now possible to have an unobstructed view along the 150m length into
the bends. The center cable, which experiences a smaller resultant force, has retained two
loop next to the bend.
The length of the pattern region has been increased by 10%, this is attributed to the lateral
compression of the patterns.
The cable circular cross sections shown in the foreground are the positions of the 86m
lengths of straight cable in front of the regions containing thermomechanical patterns.
EPRI Licensed Material

Effect of Short Circuits
10-11
The cables can be seen in the background as they pass around the bend (from left to right).
In the cross sectional view the inner cable is slightly higher, whereas the center cable is
slightly lower. A complex situation exists in the bend. The cables experience outwards
thermomechanical thrusts in the horizontal plane. The electrodynamic force adds to the
thermomechanical force on the outer cable and subtract from the inner cable, however it also
increases the axial thrust from the straight section into the bend

Figure 10-5
Ducts in flat formation during the 50kA short circuit, position B
Section view towards B, 1:1 scale
The lengths of the patterns in each of the three cables, before and during the short circuit is
recorded in Table 10-2.

Table 10-2
Length of thermomechanical patterns at 90
o
C before, during and after the short circuit
Ducts in Trefoil Ducts in Flat Formation Pipe Containing Trefoil
Cable Number 1 2 3 1 2 3 1 2 3
At 90
o
C 44.1 44.1 44.1 43.2 43.5 43.8 129.6 129.6 130.2
Short Circuit 42.3 44.1 51.9 45.9 47.4 47.4 141.9 141.6 148.2
2 Sec After 200 200 200 200 200 200 200 200 200

EPRI Licensed Material

Effect of Short Circuits
10-12

Figure 10-6
Ducts in flat formation during the 50kA short circuit, position B
Plan projection view, 1:1 scale
Figure 10-7 is the plan view 2 seconds after the short circuit has ended.
The short circuit permitted the thermomechanical patterns to grow within a 1.5s period from
a 47m length to occupy the complete 200m span.
During the short circuit the outer cables were lifted high off the duct floor. The
thermomechanical patterns in all three cables were laterally compressed and longitudinally
lengthened. At the end of the short circuit the compressive strain was freed, allowing the
cables to spring back to the floor of the duct under elastic and gravitational forces. During
this period the cables were also freed from the longitudinal constraint of friction along their
complete length. This permitted the locked-in thermal compressive strain in the straight
section to be released into the thermomechanical patterns at position B, which in less than
two seconds, grew to occupy the complete 200m length
The center cable experienced a smaller disturbance. The plan view shows that the patterns
doubled in length, although the patterns also spread along the complete 200m span length.
EPRI Licensed Material

Effect of Short Circuits
10-13

Figure 10-7
Ducts in flat formation after the 50kA short circuit, position B
Plan projection view, 1:1 scale
Figure 10-8 compares the distribution of axial force in the conductor, before, during and after the
short circuit. The conductor was selected because ABAQUS provides a route plot along the
length of each of the cable components, but not of the complete cable. The conductor
experiences more than 90% of the complete cable force and so is a god indicator of
thermomechanical effects.
Before the short circuit the force is highest at end A at 6,240N and falls linearly to a minimum of
4,200N at the thermomechanical pattern 175m away. The difference in force is supported by the
frictional constraint.
At the end of the short circuit period, the force has doubled in the 125m length closest to the
bend entrance at B, reaching a peak of 10,829N. At end A it has fallen to 4,290N. At the end of
the 2s aftermath period the force has fallen to give a longitudinal distribution that is a low and
nearly constant value in the range of 2,730-3,420N. This demonstrates that a high percentage of
the thermal strain has been released into the uniformly distributed patterns.
EPRI Licensed Material

Effect of Short Circuits
10-14

Figure 10-8
Ducts flat formation. Distribution of axial force in the conductor of cable 3 span before,
during and after the 50kA short circuit
Figure 10-9 and Figure 10-10 are the ABAQUS history plots which provide the axial force
characteristic for the complete cable in duct 3 at end A and at the entrance to the bend at B. The
discontinuous shape of the graphs is due to the large output step of 7.5
o
C. Table 10-3 and Table
10-4 record the magnitudes of the peak forces and the value of the turn-over temperatures at
which the sine wave patterns form. The turn-over temperatures occur in the range of 67.5-75
o
C
after which the force falls when thermal strain is absorbed. The force has started to rise again
before 90
o
C.
The effect during the short circuit on axial force is minimal at A, where the cable is straight, but
is very large at B, where the force is increased by 2.5 times. The thermomechanical wave
patterns at B have been compressed laterally against the pipe wall, where they have been
changed in shape and increased in length by 10%. The electrodynamic compression and the
pattern curvature have produced a high component of compressive force in the longitudinal
direction.
After the removal of the short circuit, the axial forces fall back to low levels within 1.5 seconds
at both A and B.
EPRI Licensed Material

Effect of Short Circuits
10-15

Figure 10-9
Ducts in flat formation. Axial force in cable 3 at position A before, during and after 50kA
short circuit

Figure 10-10
Ducts in flat formation. Axial force in cable 3 at position B before, during and after 50kA
short circuit
EPRI Licensed Material

Effect of Short Circuits
10-16
Table 10-3
Peak axial forces in cable 3 at end of span, position A
Model Cables per
Pipe
Formation Peak
Force
[N]
Turn-over
Temperature
[
o
C]
Force at
90
o
C [N]
Peak Force During
Short Circuit [N]
Duct 1 Flat 7,608 67.5 6,644 6,644
Duct 1 Trefoil 7,471 67.5 6,745 7,735
Pipe 3 Single 10,276 67.5 8,610 16,382

Table 10-4
Peak axial force in cable 3 at B, entrance to bend
Model Cables per
Pipe
Formation Peak
Force [N]
Turn-over
Temperature
[
o
C]
Force at
90
o
C [N]
Peak Force During
Short Circuit [N]
Duct 1 Flat 5012 75 4916 14,333
Duct 1 Trefoil 4996 75 4887 16,452
Pipe 3 Single 5821 75 6084 10,297
The sidewall forces (SWF) are recorded in Table 10-5 and Table 10-6 at the region either side of
the entrance to the bend. Almost as soon as the temperature starts to rise the thermomechanical
thrust pushes the cable into the outside of the bend and lifts it up the wall to the 3o/c-9o/c axis.
As explained in Chapter 3 the cable experience an SWF equal to the axial force divided by the
bend radius. Friction significantly increases this force at the entrance to the bend and decreases it
at the exit.
Table 10-5 and Table 10-6 give the values of the SWF at 90
o
C. The measuring positions 1 to 4
given in the table are:
Position 1, the first peak SWF in the wave pattern, when approaching from A to B.
Position 2, the second peak SWF
Position 3, the entrance to the bend, point B.
Position 4, the highest SWF in the bend in the direction B to C.
The SWF increases in magnitude when moving along the pattern towards the entrance to the
bend at B. This is because the waves become larger in amplitude and the bend radius decreases.
In this region there are two peaks of SWF arising from particular variations in the pattern shapes;
these occur at Positions 1 and 2. In moving from the pattern past B and into the bend the SWF
continued to rise; its value at the bend entrance being recorded for information as Position 3. The
EPRI Licensed Material

Effect of Short Circuits
10-17
SWF then reached the highest value in the model at typically 3m into the bend; the peak value
and its distance being at Position 4.
The SWF is higher a) in the 2.9m bend radius than in the wave patterns in the adjacent straight
span and b) in cable 1, as this duct has a smaller bend radius.
The effect of the short circuit force is to increase the SWF, by a factor of 1.4-3.5. The short
circuit force in each cable is time varying and phase displaced from the other cables. At the
moment the measurements were taken from at the end of the 0.1s period, cable 1 experienced the
maximum peak impulsive SWF of 2,923N m
-1
superimposed on the steady state SWF of
1,708N m
-1
, giving a total of 4,631N m
-1
. (The theoretical check value of impulsive force from
Equation 1010 is 2,900N m
-1
).
The resultant peak SWF of 4,613Nm
-1
is smaller than the 19,740 Nm
-1
limit derived in Chapter 4
for the XLPE insulation at 90
o
C, but is closer to the 6,580 Nm
-1
limit at 105
o
C. However the
force is applied for less than 0.1s and the possibility of XLPE creep in this short time in minimal.
Designs of cables with wire shields and foil sheaths are likely to be the most at risk of local
damage due to impulsive force. The adequacy of existing cable qualification test specifications
should be reviewed. Some specifications have a) short circuit tests performed on straight samples
and b) impact tests simulating the impact of a heavy wedge. Similarly the adequacy of the
designs of duct bends should be reviewed, particularly if the external surface of the bend is not
fully supported.
After the removal of the short circuit, the SWF falls to typically 70% of the pre-short circuit
value. This is attributed to the release of the thermal strain into the patterns and the reduction of
axial thrust.

Table 10-5
Sidewall force cable 1 (inside cable)-duct flat
Peak Sidewall Force in Thermomechanical Pattern
In Straight Duct Bend
Position 4
Load
Case

Pattern
Length
[m]
Position 1
[Nm
-1
]
Position 2
[Nm
-1
]
Entrance,
Position 3
[Nm
-1
] Force
[Nm
-1
]
Distance
[m]
90
o
C 46 523 729 468 1,708 2.6
50kA SC 54 1049 1054 1643 4,631 1.7
2s after
SC
200 269 272 323 1,217 2.0
EPRI Licensed Material

Effect of Short Circuits
10-18
Table 10-6
Sidewall force cable 2 (centre cable)-duct flat
Peak Sidewall Force in Thermomechanical Pattern
In Straight Duct Bend
Position 4
Load
Case

Pattern
Length
[m]
Position 1
[Nm
-1
]
Position 2
[Nm
-1
]
Position 3
Entrance
[Nm
-1
] Force
[Nm
-1
]
Distance
[m]
90
o
C 46 437 737 470 1,602 2.5
50kA SC 50 910 1,422 1,596 2,394 0.8
2s after
SC
94 160 589 589 1,209 2

Table 10-7
Sidewall force cable 3 (outside cable)-duct flat
Peak Sidewall Force in Thermomechanical Pattern
In Straight Duct Bend
Position 4
Load
Case

Pattern
Length
[m]
Position 1
[Nm
-1
]
Position 2
[Nm
-1
]
Position 3
Entrance
[Nm
-1
] Force
[Nm
-1
]
Distance [m]
90
o
C 45 496 701 465 1,458 3.1
50kA SC 49 1,214 1,968 1,594 2,992 0.9
2 s after
SC
200 212 474 288 1,001 3.9
Ducts in Trefoil Formation-50 kA Flat Route
Figure 10-11 is a plan view of the cable patterns in the three ducts at 90
o
C before the application
of the short circuit. Figure 10-12 is the cross sectional position of the cables at the end of the 0.1s
50kA short circuit. In the top duct the complete cable is seen to have been lifted and thrust
against the roof of the duct. This is readily explained as the weight of the cable is 117N (11.9kg),
compared to a short circuit force of 2,900N, giving a net upward force of 2,783N. The laws of
motion confirm that the cable will be accelerated upwards and will impact the roof within the
0.1s period.
The two bottom cables have been forced downwards and outwards. The cables in the foreground
are straight and so have retained their triangular formation with apex uppermost. If the cross
EPRI Licensed Material

Effect of Short Circuits
10-19
sectional view is taken within the region of wave patterns the orientation of the triangle rotates
through 20
o
.
The forces are sufficiently high that two helical loops present in each of the three cables have
been flattened against the side of the duct into sine waves, giving an unobstructed view to the
bend, 150m away.

Figure 10-11
Ducts in trefoil formation. Thermomechanical patterns at 90
o
C before the short circuit in
45m next to bend
Plan view, scale 4:1

Figure 10-12
Ducts in trefoil formation during the 50kA short circuit
End sectional view towards B, 1:1 scale
Figure 10-13 is a foreshortened plan view during the short circuit. In the straight section, the top
cable patterns can be seen in plan view because it is flattened against the duct roof. The patterns
EPRI Licensed Material

Effect of Short Circuits
10-20
of the bottom cables can only partially be seen because they are viewed sideways on. Of interest
are the larger patterns close to the exit to the bend, although the view is heavily foreshortened,
they are worthy of future examination as there may be some difference in the forces due to the
different bend radii. (The view appears to show that duct N
o
1 is not at the apex of the trefoil
group in the bend, however this is not the case and is the result of the foreshortened view).

Figure 10-13
Ducts in trefoil formation during the 50kA short circuit
Plan projection view, 1:1 scale
Figure 10-14 is the position of the cables two seconds after the end of the short circuit. Frictional
constraint has been momentarily overcome to permit the release of a significant proportion of the
locked-in thermal strain from the straight cable length into the patterns. The patterns have
extended along the complete 200m length.
EPRI Licensed Material

Effect of Short Circuits
10-21

Figure 10-14
Ducts in trefoil formation after the 50kA short circuit
Plan projection view, 1:1 scale

Figure 10-15
Ducts in trefoil formation. Distribution of axial force along span in cable 3 before, during
and after 50kA short circuit
EPRI Licensed Material

Effect of Short Circuits
10-22
Figure 10-15 is the distribution of axial force in cable 3 along the cable length from end A to the
entrance of the bend B and then to end C. The pre-short circuit force varies uniformly between
5250N at A to 4,000N within the pattern at B. The application of the short circuit results in a 5 x
increase to 24,800N at the mid span. This is a dynamic time varying force that happened to be
sampled at the end of the 0.1s period. It is attributed to a) the longitudinal thrust caused by the
lateral compression of the wave patterns and b) the resulting change in the length of frictionally
constrained cable. After the short circuit the axial force has fallen to a near uniform level of
3,000-4,000N.

Figure 10-16
Ducts in trefoil formation. Axial force in cable 3 at position A before, during and after 50kA
short circuit
Figure 10-16 and Figure 10-17 show the axial force characteristic for cable 3 before, after and
during the short circuit at positions A and B. Table 10-3 and Table 10-4 record the magnitudes of
the forces. The force characteristics are similar to those for ducts in flat formation. A greater
increases in force occurs during the short circuit (x 1.15 at A and x 3.36 at B), however this is
dependent upon the precise moment of sampling as the short circuit force is time varying at 120
impulses per second.
EPRI Licensed Material

Effect of Short Circuits
10-23

Figure 10-17
Ducts in trefoil formation. Axial force in cable 3 at position B before, during and after 50kA
short circuit
Three Cables in Pipe - 50kA Flat Route
Figure 10-18 is a plan view of the 45m length closest to the bend, before the short circuit. The
top cable has a sine wave formation with a significantly larger amplitude and longer wavelength
at B, the entrance to the bend. The two cables at the bottom of the trefoil group are parallel and
snake laterally. Table 10-2 records that the 130m average length of the three cable patterns in the
pipe is x 4 longer than those of the two duct formations of 43-44m. Unlike the duct systems the
pipe system is unable to produce helical loops due to interference between the cables, in
consequence the patterns have smaller amplitudes and longer length.

Figure 10-18
Three cables in a pipe system. Thermomechanical patterns at 90
o
C in the 45m next to bend
Plan view: 4:1 scale
EPRI Licensed Material

Effect of Short Circuits
10-24
Figure 10-19 and Figure 10-20 show the cable positions during the short circuit after 0.1
seconds. The three straight lengths of cable in the foreground can be seen to have been forced
outwards in the shape of a isosceles triangle. This is also true for the region containing cable
patterns except the sine patterns intermesh and the triangular shape oscillates. The outward force
is sufficiently high and the geometry sufficiently precise that at the center of the patterns a clear
cylindrical space of 41mm diameter is formed for the complete length in view of 150m.

Figure 10-19
Pipe system during 50kA short circuit at position B
Section end view on B, 1:1 scale

Figure 10-20
Pipe system during 50kA short circuit at position B
Plan projection view. 1:1 scale
EPRI Licensed Material

Effect of Short Circuits
10-25
Figure 10-21 is a foreshortened view of the patterns two seconds after the short circuit has ended.
At B, the exit to the bend in the direction towards A, the top cable has slipped sideways. This
cable then reoccupies the top position for a further 60m, before again falling to the side. The
patterns occupy the complete 200m length of the pipe.

Figure 10-21
Pipe system after 50kA short circuit at position B
Plan projection view, 1:1 view
Figure 10-22 shows the distribution of axial force in the conductor of cable 3 along the span
before, during and after the short circuit. The pattern lengths and axial forces in the cable at
positions A and B are recorded in Table 10-2, Table 10-3 and Table 10-4. The axial forces at
90
o
C before the short circuit are on average 24-29% higher than those in the two duct systems.
The reason is the same as that given for the longer pattern length in that the ability of the patterns
to absorb thermal strain is restricted by the presence of three cables.
EPRI Licensed Material

Effect of Short Circuits
10-26

Figure 10-22
Three cables in pipe. Distribution of axial force in conductor along the cable 3 span before,
during and after 50kA short circuit
Figure 10-23 and Figure 10-24 give the force characteristics with temperature and time for cable
3 at position A and B. At position A the axial force was increased by x 1.9 to 16,400N. At
position B the axial force was increased by x 1.7 to 10,300N. The net outward electrodynamic
force during the short circuit is significantly higher for the three cables in the pipe than in the
duct, because the initial cable spacing is reduced from 300mm to 83mm, increasing the force by
x 3.6.
Figure 10-24 shows at B in the pattern region the axial force exhibits oscillating values in the
range of 0-10,000N during the 0.1s period. Smaller variations are also present during the short
circuit applied to the duct systems. The radial, circumferential and longitudinal movements of
the three cables are complex. During this period a) the cable is experiencing outwards force at
120 impulses per second and b) the forces in each cable are time displaced from each other.
EPRI Licensed Material

Effect of Short Circuits
10-27

Figure 10-23
Ducts in trefoil formation. Axial force in cable 3 at position A before, during and after the
50kA short circuit

Figure 10-24
Ducts in trefoil formation. Axial force in cable 3 at position B before, during and after 50kA
short circuit
EPRI Licensed Material

Effect of Short Circuits
10-28
Figure 10-22 for the conductor axial force shows that the axial force after the short circuit falls to
a near constant value in the range 4,400-5100N, this being higher than in the duct systems of
3,500-4,000N (trefoil) and 3,000-3,750N (flat spaced).
The overall effect of the short circuit in the pipe system is to uniformly release a significant
proportion of the locked-in strain from the straight cable into the patterns and thereby reduce the
axial thrust. However the events are more violent than in the duct systems and result in a greater
disturbance to the regularity of the cable patterns. It is recommended for future work that a) this
disturbance be studied and b) the effect of the side wall forces in the bend be quantified as these
will be greater than given for the duct system.
Effect of Slope Angles 50kA Short Circuit
Models of a trefoil duct system and a pipe system were constructed at five angles of slope, +7.5,
+3, 0, -3 and -7.5 degrees. They were subjected to the 50kA short circuit for 0.1s.
The pattern lengths for the 50kA short circuit case are recorded in Table 10-8, Table 10-9, Table
10-10, Table 10-11 and Table 10-12.

Table 10-8
Pattern lengths at +7.5 degree slope for a 50kA short circuit
Pattern lengths [m]
Ducts in Trefoil Three Cables in Pipe System
Cable Number 1 2 3 1 2 3
90
o
C 40.8 42 41.1 153.3 163.2 163.5
During S/C 43.2 53.5 48 200 200 200
After S/C and 2s 200 200 200 200 200 200

EPRI Licensed Material

Effect of Short Circuits
10-29
Table 10-9
Pattern lengths at +3 degrees slope for a 50kA short circuit
Pattern lengths [m]
Ducts in Trefoil Three Cables in Pipe System
Cable Number 1 2 3 1 2 3
90
o
C 42.3 42.6 42.3 153.9 156 154.2
During S/C 44.4 42.9 53.7 200 200 200
After S/C and 2s 200 200 200 200 200 200

Table 10-10
Pattern lengths at 0 degree slope for a 50kA short circuit
Pattern lengths [m]
Ducts in Trefoil Three Cables in Pipe System
Cable Number 1 2 3 1 2 3
90
o
C 44.1 44.1 44.1 129.6 129.6 130.2
During S/C 42.3 44.1 51.9 141.9 141.6 148.2
After S/C and 2s 200 200 200 200 200 200

Table 10-11
Pattern lengths -3 degree slope for a 50kA short circuit
Pattern lengths [m]
Ducts in Trefoil Three Cables in Pipe System
Cable Number 1 2 3 1 2 3
90
o
C 46.2 47.1 45.3 126.3 126.3 126
During S/C 48.3 45 52.8 200 200 200
After S/C and 2s 200 200 200 200 200 200
EPRI Licensed Material

Effect of Short Circuits
10-30
Table 10-12
Pattern lengths at -7.5 degree slope for a 50kA short circuit
Pattern lengths [m]
Ducts in Trefoil Three Cables in Pipe System
Cable Number 1 2 3 1 2 3
90
o
C 50.4 47.1 47.7 95.7 97.2 103.5
During S/C 44.4 47.4 52.2 182 95.7 123.3
After S/C and 2s 103.8 107.4 179.7 200 200 200
The study showed that the angle of slope did not affect the cable patterns during the application
of the 20kA and 50kA short circuit levels, this was attributed to the electrodynamic force
clamping the cable to the walls of the duct and pipe. However the angles slope does have some
effect on the thermomechanical patterns before and after the short circuit.
Trefoil Duct System
At 90
o
C before the short circuit, the angle of slope influences the length of the
thermomechanical pattern. The span with +7.5degree slope has pattern length of 42m
compared to 49m for the -7.5 degree slope. The 7.5 degree slopes are above the critical angle
of slip of 5.7 degrees, which permits more thermal strain to feed downhill from A into the
pattern at B and less uphill from A to B. There was no significant difference between the
pattern lengths with slopes in the range of +3 to -3 degrees as these were less than the critical
angle.
After the short circuit had been removed there were some differences in the appearance of
the patterns. In the +7.5 degree model the pattern had formed along the complete length and
had straightened out the cable for the 50m closest to the bend at B. This was attributed to the
slope being above the critical angle of slip. The short circuit had lifted the cable and
permitted a) the release of the thermal strain to lengthen the sinusoidal patterns along the
complete 200m span and b) the weight of the cable at the bottom of the span at A to
straighten the pattern at the top end. There was no difference with slopes between +3 and -3
degrees.
Pipe System
At 90
o
C before the short circuit, the angle of slope influenced the length of the
thermomechanical pattern. The thermomechanical patterns were longer when they were at
the bottom of the downhill slopes, 47.7m (-7.5
o
) and shorter when they were at the top of
uphill slopes, 41m (+7.5
o
).
After the short circuit had been removed, the downhill slope of -7.5
o
grew a slightly shorter
pattern length of 180m. The patterns for the other slope angles extended the complete 200m
span length. The explanation for the reduced pattern length was that the slope was above the
critical angle of slip so that the thermal strain was relieved into the patterns at the bottom of
EPRI Licensed Material

Effect of Short Circuits
10-31
the slope at B, where the highest compressive forces existed, leaving the top part of the span
at A under tension.
The models with slope angles in the range of +3 to -3 degrees did not influence the performance
of the cable during the short circuit.
The behavior of a -3 degree slope is described in the following section.
Flat Formation - Reduced Short Circuit Level of 20kA
The 20kA short circuit level is more likely to be encountered in service than the typical 50kA
design level. Error! Reference source not found. shows that the force is proportional to I
2
, thus
the resultant force is reduced to 16%. This is still sufficient to overcome the frictional constraint,
permitting patterns to grow and axial force to fall. In the following examples the route was
inclined to -3% below the horizontal, Figure 10-1, with the bend (positions B and C) downwards.
The effect of the short circuit was not influenced by the slope.

Figure 10-25
Ducts in flat formation, -3
o
slope, during 20kA short circuit at position B
Section end view, 1:1 scale
The following comparisons are made between the 20kA and 50kA cases:
The cross sectional positions of the cables during the 20kA short circuit, Figure 10-25,
compared to the 50kA, Figure 10-5, show that the outer cables have been lifted to 40mm at
20kA and 137mm at 50kA.
EPRI Licensed Material

Effect of Short Circuits
10-32

Figure 10-26
Ducts in flat formation, -3
o
slope, thermomechanical patterns during 20kA short circuit at
position B
Foreshortened plan view, 1:1 scale
The plan foreshortened views in Figure 10-6 and in Figure 10-26 show that the 20kA force
acting on the two outer cables was not able to flatten the two helical loop patterns adjacent to
the bend, whereas the 50 kA force did. This is confirmed at 20kA in the Figure 10-26 cross
sectional view. It is no longer possible to see along the duct due to the helical turns at the far
end.
EPRI Licensed Material

Effect of Short Circuits
10-33

Figure 10-27
Ducts in flat formation, -3
o
slope. Distribution of force along the cable 3 span, before, during
and after 20kA short circuit
A comparison of the distribution of forces in the conductor of cable 3, Figure 10-8 at 50kA and
Figure 10-27 at 20kA, show that:
Before the short circuit the average force is slightly higher in the flat route at 5,200N
compared to 4,800N in the -3
o
route.
During the short circuit the average axial force falls from 9,000N at 50kA to 2,600N at 20kA.
The level of force within the pattern region is increased at 50kA and reduced at 20kA.
After the short circuit, the average axial force has fallen to a similar level of 3,300N at 20kA
and 3,100N at 50kA.
The effect of the -3
o
slope is seen after the short circuit. The force at the top of the slope at A
has fallen to 2,600N, whereas that at the bottom of the slope has increased to 4,000N.
Although the electrodynamic force at 20kA is much reduced, it is still sufficient to relieve the
frictional constraint and release thermal and gravitational strain.
EPRI Licensed Material

Effect of Short Circuits
10-34

Figure 10-28
Ducts in flat formation, -3
o
slope. Axial force in cable 3 at position A during, before and after
20kA short circuit

Figure 10-29
Ducts in flat formation, -3
o
slope. axial force cable 3 at position B during, before and after
20kA short circuit.
EPRI Licensed Material

Effect of Short Circuits
10-35
The force characteristics at 20kA for cable 3 at A and B are given in Figure 10-28 and Figure
10-29 and are compared with those at 50kA, Figure 10-9 and Figure 10-10:
During the short circuit the effect on axial force is minimal in both cases at position A, where
the cable is straight.
Immediately after the short circuit on the -3
o
route at 20kA there is a significant fall in force
at A, compared to a small reduction in the flat route at 50kA. The 20kA force overcame
frictional constraint and permitted the cable to move downhill under gravitational force,
reducing the force at A. Thus the application of a 20kA short circuit to a new unloaded cable
is capable of removing the restriction of the critical angle of slide; permitting the axial strain
to be redistributed. If there is space at the bottom of the span, such as the bend at B, or a joint
manhole, then the cable will slide into the space.
The effect of the short circuit is more marked within the thermomechanical pattern at
position B. During the 50kA short circuit a significant increases in axial force is shown in
Figure 10-10, although the force falls afterwards. At 20kA the force is reduced, Figure 10-29,
afterwards the axial force increases, because the cable from the top of the slope slides
downhill and applies additional compressive load.
Ducts in Trefoil Formation-Reduced Short Circuit Level of 20kA

Figure 10-30
Ducts in trefoil; during 20kA short circuit at position B
Section end view towards B, 1:1 scale
EPRI Licensed Material

Effect of Short Circuits
10-36

Figure 10-31
Ducts in trefoil formation, -3
o
slope, thermomechanical pattern during 20 kA short circuit
at position B
Plan projection view, 1:1 scale
Figure 10-30 is the cross sectional view and Figure 10-31 is the foreshortened plan view of the
cables during the 20kA short circuit. These views are informative when compared to Figure
10-12 and Figure 10-13 for the 50kA case:
At 50kA the cables were pressed hard against the walls of the ducts and the helical loops had
been flattened, giving a clear view along the 150m length of duct. At 20kA the force is
insufficient to reshape the 2-3 helical turns in Figure 10-13. However a small radial gap
exists in the ducts at the top of the helical turns in the two lower cables, confirming that they
are experiencing compressive forces.
At the moment of sampling 0.1s into the 20kA short circuit the top cable can be seen to have
lifted the cable close to the centre of the duct, a distance of 75mm. This can be checked by an
approximate calculation of the force from Equation 1010, assuming that the cables stay at
300mm centers, and then applying the equations of motion. The resultant upward
electrodynamic force can be calculated to be 464Nm
-1
, at 120 pulses per second. Subtracting
the cable weight of 117Nm
-1
gives a net upward peak force of 347 Nm
-1
. In the 0.1s time
period a constant force of 347 Nm
-1
would lift the cable 146mm. Integrating the waveform
gives an average force of 73Nm
-1
, which would lift the cable 30mm. These heights are
sufficiently close to the 75mm dynamic computation of the software to confirm the
interpretation.
EPRI Licensed Material

Effect of Short Circuits
10-37
Upon removal of the 20kA short circuit the patterns spread along the 200m span length as in
the 50kA case.
Three Cables in the Pipe System - Reduced Short Circuit Level of 20kA
Figure 10-32 and Figure 10-33 show the cable positions in the pipe during the 20kA short circuit
compared to Figure 10-19 and Figure 10-20 for the 50kA case.
During the 20kA short circuit the electrodynamic forces have repelled the three cables
outwards to form the same shape of isosceles triangle as seen at 50kA. The electrodynamic
forces are 1.9 x higher in the pipe compared to the duct case, because of the closer proximity
of the three cables.
During the short circuit a difference can be seen within the bend in the cross sectional views.
At 20kA, Figure 10-32, a gap is present at the top of the pipe showing that the cable has not
been lifted to the top of the pipe. At 50kA, Figure 10-19, a) the top cable has been pushed
upwards very close to the top of the pipe and b) a gap can be seen between cables 1 and 2.
After the 20kA short circuit effect is the same as that at 50kA in permitting the cables to
overcome the frictional constraint and release the thermal and gravitational strain. The
thermomechanical patterns grow to occupy the full 200m span.

Figure 10-32
Three cables in pipe, -3
o
slope during 20kA short circuit
Section end view towards B, 1:1 scale
EPRI Licensed Material

Effect of Short Circuits
10-38

Figure 10-33
Three cables in pipe, -3
o
slope. Thermomechanical patterns at 90
o
C in 150m next to bend
Plan projection view, 1:1 scale
Conclusion
A method exists to design duct and pipe systems to withstand the combined effects of short
circuit electrodynamic forces and thermomechanical forces.
The application of a 50kA three phase short circuit is shown to be a violent event, which applies
high axial and sidewall force to the cable, causing rapid changes in cable position.
The effect on regions of pipe and duct containing naturally formed thermomechanical patterns is
dramatic and sudden. The cable system experiences a significant shock, during and after the
short circuit:
During the short circuit the cable is accelerated off the pipe floor within the 0.1s period. In
the majority of 50kA cases studied, the cable is pressed hard against the pipe wall and any
helical thermomechanical loops present are flattened into sine waves. The compression
results in a 10% increase in the length of pre-existing thermomechanical patterns.
After the short circuit, the release of the elastic and potential energy stored in the cables
cause them to spring back and seek a new position. They momentarily overcome the
longitudinal frictional constraint between the cable and the pipe wall, permitting the sudden
release of the locked-in thermal and gravitational strain from the full length of cable into the
EPRI Licensed Material

Effect of Short Circuits
10-39
thermomechanical patterns. The existing patterns grow rapidly to occupy the complete pipe
span (200m) within 1.5 seconds.
During the short circuit a reduction in the current level from 50kA to 20kA was seen to reduce a)
the violence of the forces (a theoretical reduction of x 6.3) and b) the degree of cable
displacement. After the short circuit, the same sudden release of locked-in thermal strain and
growth of patterns occurred. It is indicated that this would also occur with a) high level single
phase short circuits and b) lower level three phase short circuit currents.
The effect is greater in the straight sections of pipe containing a region of thermomechanical
patterns than in other regions containing straight cable. The axial force is momentarily increased
by typically 2.5 times and the sidewall force by 3.5 times.
The short circuit permits the release of locked in thermal strain irrespective of the angle of slope
within the 7.5
o
range studied in both uphill and downhill directions. Short circuits release both
locked-in thermal and gravitational strain. A short circuit is capable of releasing gravitational
strain and permitting movement in a new and unloaded circuit. During the release of locked-in
strain, cable movement can be distributed over long distances.
Short circuit violence is greater in a pipe system than in a duct system, because the distances
between the phase conductors is smaller, resulting in significantly higher electrodynamic forces
(x 3.6 for the cases studied). However the effects are greater in the duct system. Electrodynamic
force significantly alters the shape of cable thermomechanical patterns. In the duct system helical
loops are flattened into sine waves and the pattern region is increased in length by 10%. In pipe
systems only thermomechanical sine waves can be formed, although these are also compressed
against the wall resulting in an increase in pattern length of 9-14%.
Cables experience a high sidewall load in route bends during the short circuit. The sidewall force
in the pipe system being the highest. The possibility of XLPE creep under the action of the side
wall forces in the bend in the 0.1s short circuit duration is minimal. Designs of cables with a)
shield wires and b) sheaths of foil laminate are the most at risk of local damage due to impulsive
force.
The question arises whether the passage of short circuit current is beneficial or detrimental to the
cable system. Overall, if the system can survive the violence of the short circuit, the effects are
beneficial:
The after effects are beneficial as the release of locked-in thermal strain produces uniformly
distributed thermomechanical patterns and reduced axial and sidewall loads.
The normal effects of load cycling over a long period tend to result in a more uniform
distribution of thermomechanical patterns and lower level forces. Thus the consequence of
the short circuit is to hasten the process.
The violence during and immediately following the short circuit is clearly detrimental as the
constructions of cable and duct/pipe experiences momentary high axial and sidewall
impulsive loads superimposed on the thermomechanical loads
EPRI Licensed Material

Effect of Short Circuits
10-40
In particular the cable is likely to impose high longitudinal loads on the joints and the cleats
in manholes indirectly through the presence of thermomechanical patterns in the adjacent
pipe/duct, in addition to the directly applied electrodynamic force.
The construction of the cable can be selected to withstand the application of short circuit and
sidewall force during development and qualification testing, although the adequacy of
existing test specifications should be reviewed.
In principle, pipes and ducts and, most importantly their surrounding supports, can be
designed to contain the thermomechanical and electrodynamic forces. Although it is doubtful
that adequate design practices and test specifications exist at present.
The joint and manhole can be designed to withstand the forces. However the possibility of
design variations and inadequacies are greater. Short circuit effects are more likely to result
in electrical failure of the joint and severe local mechanical damage to the cable in the
manhole. This subject is worthy of further work.


i
Heinhold L, Power Cables and their Applications, Siemens Aktiengesellschaft, Berlin and Munchen, 1970
EPRI Licensed Material
11-1
11
MECHANICAL PERFORMANCE OF A ONE-PIECE
STRAIGHT JOINT
Introduction
Chapters 6, 7 and 8 have studied the thermomechanical performance of the cable in pipe and
duct systems. Chapter 4 considered the factors that may limit their service life. As part of the
project, information on the forces experienced by joints was analysed from the performance of
joints within spans of a) an in-service 138kV 1500kcmil pipe route and b) a 230kV 2500kcmil
duct route. The forces that are exerted by the cable within the manhole are relevant to all designs
of joints in pipe systems, both anchor and one-piece. One-piece and anchor joint designs and
their critical factors are described in greater detail in Chapter 5 of the 2002 EPRI report

reviewing the technology of XLPE EHV cable systems
[i]
.
This chapter applies the information to detailed models of a one-piece joint and analyzes the
effects that cable thermomechanical forces have on joint performance.
The joint in the 230kV duct route is an anchor design in which the cable conductor is fixed to the
manhole chamber via a rigid epoxy resin insulating body, thereby preventing longitudinal
movement, but prospectively experiencing maximum differential thrust. Lateral movement is
prevented by the presence of cleats on either side of the joint. Additionally these joints are
protected from the full cable thrust by the constraining effects of a) cleats and b) offset cable
bends. The performance of these joints is therefore predictable and there is no point in modelling
them.
The joints in the 138kV pipe system are installed within an in-line casing and as such do not
have inherent design measures to protect them from lateral and longitudinal movements. The
138kV joint design was therefore selected as the subject of this chapter.
Although one particular design of 138kV 1500kcmil joint is studied, the limiting factors
identified are relevant to the design family of one-piece joints. It is understood that the 138kV
design has exhibited a satisfactory electrical performance in operational systems despite the
reported presence of thermomechanical disturbance within the casing. It is the objective of this
chapter to identify key factors for use when a) designing constraining systems in manholes and
b) selecting joints. The one-piece design family includes a) rigid and plug-in conductor
connectors, b) EPR or silicone insulating and semiconducting compounds and c) different
assembly methods, such as pulling the molding directly onto the cable, or indirectly pre-
stretching it onto a parking tube.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-2
The chapter first reviews the performance of the joints within the casing to quantify a) the axial
cable forces and b) the radius to which the joints are bent. The construction of a detailed FEA
joint model is described, followed by the analysis of joint bending and cable pull-through
performance. The performance of the joint is compared to benchmark limiting factors that may
affect service life. Finally outline recommendations are given on development and pre-
qualification tests to demonstrate thermomechanical performance.
Joint Performance in a Full Sized Route
Figure 11-1 shows the 763m pipe system model selected for the joint performance study,
comprising two spans, with the joint manhole at position X. Span A-X is 238m long and is
substantially straight with a downhill slope angle of -2.3
o
. Span X-E is 525m long, having a
pronounced S bend starting close to the manhole. X-E has a slight downhill slope of 0.33
o
. The
distance from manhole X to the start of the bend at B, is X-B = 13.5m. The first bend B-C has an
angle of 54.3
o
and radius 45.7m. The second bend C-D has an angle of 58.6
o
and radius of
18.3m.

Figure 11-1
138kV pipe system route profile showing joint position
From previous studies it would be expected that the two bends would separate the route into two
halves A-C and C-E, by the locking effect of the sidewall forces. The first half of the route model
A-C is similar to the hockey stick shaped reference model A-B-C used in Chapters 7 and 8 with
the inclusion of a joint position close to the entrance of the bend at B.
The behavior of span A-C can be analyzed in isolation. Upon heating, the cable in the straight
section A-B would attempt to thrust into the bend B-C. The bend has a storage capacity of
0.5(Dp-Dc) = 58.3mm, where Dp is the 206.4mm pipe internal diameter, Dc is the 83mm cable
outer diameter and is the 0.95 bend angle in radians. If the joints behaved like the cable within
the pipe, they would transmit the 58mm movement through the casing into the bend. The cable
thrust at B would then initially be of low magnitude, whilst at higher temperatures it would
initiate thermomechanical patterns, again limiting the thrust. However the joints cannot behave
like the cable as a) the clearance within the casing is greater, b) the cables have to curve out to
the increased centerline spacing of the joints, c) the adjacent cables sag under gravity and d) the
joints are heavier and experience higher frictional constraint.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-3
The cables within the casing experience a bending moment formed by the cable curvature and
cable compressive thrust. The bending moment increases the cable curvature, such that the joints
are rotated transversely across the casing until limited by contact with the casing. There is ample
thermal strain stored in A-X to generate thrust and deflect the joints. The thermal strain
generated in A-X is 307mm for a temperature rise of 15 to 90
o
C and coefficient of thermal
expansion of 18.5x 10
-6

o
C
-1
.
The presence of the downhill slope of -2.3
o
from A to X reduces the effective coefficient of
friction from 0.3 to 0.26, facilitating movement of cable into the casing.

Figure 11-2
Thermomechanical deformation inside 138kV joint casing at 90
o
C: elevation view

Figure 11-3
Thermomechanical deformation inside 138kV joint casing at 90
o
C: plan view
The deformation within the casing at 90
o
C in the first load cycle is shown in side view in Figure
11-2 and in plan view in Figure 11-3. The three cables and joints are constrained to move
together by a) a central aluminum spider that holds the joints parallel and b) tape wrappings
binding the cables and joints together. The resulting deflection in Figure 11-2 closely replicates
that seen in service.
The most severe bending occurs where the cable enters the joint molding; the minimum bend
radii for joint 1(top of trefoil group) being 1.8-4m, for joint 2 being 2.6-2.7m and for joint 3
being 2.2-3.5m; the average being 2.8m. The reason for the ranges of bending radii being given
is that the model of this particular joint was embedded within the model of the full sized pipe
spans. This enforced discrete changes in element sizes and properties at the transition positions
between the cable and the end of the one-piece molding, this being the position where the
bending was the most concentrated. A separate, more detailed model of the joint was also
produced, which had smooth transitions and a smooth distribution of bending radii. The detailed
model had short lengths of cable and pipe attached, to which was applied the force-extension
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-4
characteristic of a typical span length of cable. Overall it was considered that the joint embedded
within the full sized spans gave the closest simulation of service performance.
The longitudinal movements of the cables at the casing are given in Table 11-1:
25mm of cable movement was required to form the deflection pattern. The ratio of cable
absorption to cable length in the casing is 1:100, giving an average angular deflection of 8
o
.
the length of cable moving into the casing from section A-X was 37mm in the first load cycle
and 36mm in the third.
the length of cable moving out of the joint to bend B in the first load cycle was 12mm and in
the third 23mm.
the joints moved slightly further than the cable indicating that they started to move
lengthwise before the patterns fully developed.

Table 11-1
Cable strain stored in the casing and movement through casing
Cable Strain Stored in
Casing [mm]
Cable Movement Into Casing
in Direction A-X [mm]
Cable Movement Out of
Casing in Direction X-B [mm]
Cable
Number
1
st
Load
Cycle
3
rd
Load
Cycle
1
st
Load
Cycle
3
rd
Load
Cycle
1
st
Load
Cycle
3
rd
Load
Cycle
Cable 1 28.1 10.7 51.7 35.4 23.6 24.7
Cable 2 23.3 8.5 29.6 35.6 6.2 27.1
Cable 3 22.8 18.7 29.95 35.8 7.2 17.1
Average 24.7 12.6 37.1 35.6 12.3 23.0
The cable axial forces thrusting into the joint casing during three load cycle of 15-90
o
C are
shown in Figure 11-4 and Figure 11-5 for cable 3. Cable 3 produced the highest forces of 8,750N
(side A-X) and 8,650N (side X-B). The joints and cable were unable to withstand these forces
and deflected from their original straight alignment.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-5

Figure 11-4
Axial force from cable 3 on A-X side of joint

Figure 11-5
Axial force from cable 3 on X-B side of joint
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-6
Figure 11-6 shows the differential forces acting across joint 3, the highest force being 1,595N.
The differential force is balanced by the reaction force where the joints and cables contact the
casing and joint cradle. The direction of differential force alternates from left to right twice per
cycle during heating and cooling. The differential force is generated by imbalance in the pipe
sections A-X and X-C on either side of the joint, which produce different magnitude forces at
each temperature step.

Figure 11-6
Differential force across joint 3, (negative movement is from joint to bend)
Figure 11-7 shows the differential force acting across all three joints, treated as one structural
body. The peak force is 1,729N, alternating in direction four times each cycle.
The differential force across joint three of 1,595N was checked by comparison with the
maximum of 3,500N from the Chapter 8 route. The route maximum was obtained from the
hockey stick model of the same 138kV cable and pipe in the Chapter 8 study. The force-
temperature characteristic at the entrance to the bend B was subtracted from that at A, the
anchored end of the span. The difference is plotted in Figure 11-8, yielding a maximum value of
14,000N, from which was subtracted the frictional constraining force along the span of 10,500N.
The frictional constraining force was calculated for the Chapter 8 horizontal route span of 200m,
cable weight of 3 x 11.9kgm
-1
shared between two cables and a coefficient of friction of 0.3.
The maximum compressive force in the joint of 8,750N is less than in the Chapter 8 route, at end
A of 26,000N at a turn-over temperature of 47.5
o
C and at end B of 16,000N at a turn-over
temperature of 45
o
C. It is probable that the curvature of the cable within the casing will always
initiate thermomechanical patterns in the adjacent pipe if the joint is located in a straight span,
thus the force at the entrance to the bend at B of 16,000N is a more realistic comparison.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-7

Figure 11-7
Differential force across all three joints, (negative movement is from joint to bend)

Figure 11-8
Maximum differential force, (F
A
-F
B
), where F
A
is the force at the end of the straight section
and F
B
is the force at the entrance to the bend in the reference route
In summary, the study of the 138kV 1500kcmil span showed that the individual joint and
adjacent cable is required to withstand nominal values of:
compressive axial thrust of 9,000N, with a possible maximum of 16,000N
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-8
differential longitudinal force of 2,000N, with a possible maximum of 4,000N.
minimum bending radii 2-3m, with a spot minimum of 1.8m.
longitudinal alternating movements of 25mm, with a theoretical maximum of 58mm.
138kV One-Piece Joint Design
Descriptions are first given of the joint design and assembly sequence to provide an
understanding in support of the selection of the limiting parameters. The benchmark parameters
will later be compared to the output from the joint models.
Description of Joint
The outline drawing of the 138kV joint is shown in Figure 11-9. Only length dimensions
referring to key positions are given. For the detailed modelling, the dimensions of the joint were
abstracted from a) the manufacturers technical information and b) from a cut section of a
molding. It is not expected that the dimensions will be precise as dimensions are known to
change due to the internal mechanical stresses which arise from a) the molding process and b)
the on-site assembly process. For example a) when the molding is stretched onto the cable, the
radial compressive stresses are increased and in consequence the diameter increases and the
length reduces and b) when the molding is cut open the internal manufacturing stresses are
relieved and the dimensions change. The effect of modelling the stretching of the molding onto
the cable, with an interface coefficient of friction =1, was to shorten the length overall by
26mm and between the base of the LV stress control profiles by 12mm.
The study one-piece joint is molded using ethylene propylene rubber, (EPR). The elastomer is
usually modified to permit peroxide crosslinking (EPDM). The molding is electrically self-
contained, comprising:
an integral HV (high voltage) electrode of semiconducting rubber to shield a) the cylindrical
aluminum heat sink covering the conductor connection, b) the two cut ends of the cable
XLPE insulation and c) the interface between the HV electrode itself and the cable interface.
two integral LV (low voltage ) stress cones of semi-conducting rubber to a) shield the
termination of the cable insulation shield and b) to limit the magnitude of the longitudinal
component of electrical stress along the cable-joint interface.
a bonded semiconducting layer over the center of the molding to provide an electrically
smooth interface and shield the EPR insulation.
high quality EPR insulation of sufficient thickness to withstand the voltages under normal
and transient operating conditions.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-9

Figure 11-9
Outline dimensions of 138kV one-piece joint showing key positions
The most vulnerable part of the joint is the electrically stressed interface between the one-piece
molding and the XLPE cable insulation. The electrical stress at the insulation shield in a 138kV
cable is sufficiently high to cause any air gaps (termed voids) trapped in the interface to break
down and spark (termed partial discharging). XLPE insulation has poor resistance to partial
discharge and at HV and EHV cable stresses eventual failure is inevitable. The report of the
CIGRE Joint Task Force 21/15 in 2002
ii
gives a qualitative account of their studies on the
parameters that effect interfaces in accessories. The joint designer endeavors to eliminate the
presence of air gaps by:
Stretching the molding onto the cable core such that it permanently exerts sufficient
interfacial pressure to ensure intimate contact and thereby excluding space for voids. The
magnitude of the interfacial pressure is expected to fall with time due to a) the inherent
characteristic of the elastomeric compounds of relaxation under tensile stress, c) incomplete
cross linking during manufacture (perfect crosslinking is not achievable and cross-linking the
thick radial sections is a difficult manufacturing process). It is good practice to add
performance margins into the design by a) selecting an EPR compound with low relaxation
properties and b) increasing the radial stretch onto the cable.
Manufacturing the joint molding with a very smooth and flat bore particularly at the triple
point interfaces at a) the base of the LV stress control profile where the EPR insulation,
semicon rubber and XLPE insulation meet and b) the interface between the HV electrode,
EPR insulation and XLPE insulation. These are shown in Figure 11-10.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-10
Developing reliable field processes to assist the jointer in removing the cable insulation
shield and in polishing the surface of the cable XLPE insulation, such that local depressions
and scratches capable of forming air voids are removed.
Developing a repeatable field process to taper the ends of the cable LV semiconducting
insulation shields to the required angle and shape such that a void is not present at the base of
the taper. A void can be formed by a) the jointer undercutting the taper into the cable
insulation, b) forming an incorrect convex shape instead of straight or concave shape and c)
the LV stress cone molding being unable to conform to the taper shape such that it bridges
the taper and forms an air void. The electrical termination of the cable LV shield will either
occur at the tip of the profile, or at the tip of a hand painted semiconducting lacquer applied
onto the XLPE insulation (as present in the study joint).
The reference positions used in this study are given in Figure 11-10:
a: interface of HV electrode with the EPR insulation in the bore of the molding
b: base of LV stress control profile
c: base of taper on cable insulation shield (the painted electrical termination of the LV shield is
12mm from this position)
d: top of taper on the cable insulation shield (this does not have a key technical significance, but
does have an identifiable peak in the interfacial pressure graph)

Figure 11-10
Reference positions in 138kV one-piece joint
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-11
Description of Assembly Sequence
The typical assembly sequence of a one piece-joint is:
To reveal the extruded core by removing a) the jacket and metallic sheath/foil, b) the
underlying ground wires/metallic tapes and c) the cushioning and water blocking tapes.
When the molding is stretched onto the cable, the radial compressive stresses are increased
and in consequence the diameter increases and the length reduces. To straighten the extruded
core by removing the residual curvature from the manufacturing process. This usually
involves heating the cable for a period of time and allowing it to cool whilst constrained, for
example within a tube. This also encourages the early relaxation of some of the locked-in
shrink-back at the insulation cuts.
To cut back the cable insulation and reveal the conductor.
To strip the semiconducting LV shield from the extruded core and to taper the termination of
the cable shield.
To smooth, polish and clean the exposed XLPE insulation.
To apply lubricant to the XLPE insulation (a silicone insulating fluid and in some cases a
silicone grease) and then to pull the one-piece molding onto the long, parking side of the
joint. This process forces the bore of the molding to be expanded by typically 25%.
Significant mechanical effort is required, thus an assembly tool is provided to generate the
pulling force. In some joint designs the molding is pulled onto a disposable tube and this is
parked loosely over the long side of the joint.
The conductor connection is then made usually with a compression ferrule or a welding
technique. The connection may be of the rigid type, or of the plug-in type, in which case it
is good practice to have means of locking the connection together to prevent separation under
the action of thermomechanical tensile forces that occur is service. The connection is then
covered with a split cylindrical metallic shield (called a heat sink in the study joint). The
shield will later support the molding and make electrical contact with the semicon HV
electrode. In some joint designs the shield is fixed to the ends of the XLPE insulation to
prevent shrink-back.
The molding is then pulled back towards the short end of the joint using an assembly tool.
It is finally positioned over the center of the HV shield.
Finally the LV electrical connections are made across the joint and between the cables and
molding. Protective coverings are applied overall
It is important that the dimensions of the molding incorporate sufficient cumulative length
margins to allow for:
Manufacturing tolerance of the molding components
Contraction of the mold body during the expansion and stretching that occurs during pulling-
over
Inaccuracy in centering the molding
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-12
Asymmetry between the positions of the LV stress control profiles due to frictional drag
Jointer measuring tolerance in the forming of the semiconducting cable shield tapers and
painted shield termination
XLPE insulation shrink back at the shield termination positions (likely to be larger in cables
with smooth or taped conductor surfaces)
The design margin to be incorporated into the molding and jointing dimensions may require to
be 10mm or greater.
The particular one-piece molding in the study is stated in the manufacturers information to have:
A nominal internal bore diameter of 57.15mm suitable to fit the 138kV 1500kcmil cable with
a nominal insulation diameter of 72.6mm. The resulting stretch of the molding during
assembly is 27%. This was taken as the reference case in the study.
A range taking ability to fit insulation diameters between 65.15mm and 74.88mm, giving a
minimum stretch of 14% and maximum stretch of 31% These values were selected as the
limits for the sensitivity study.
Limiting Parameters
When the joint enters service it is important for the satisfactory electrical performance that a) the
magnitude of the interfacial pressure is maintained and b) the positions of the triple interfaces
remain undisturbed.
Limiting Parameter: Interfacial Pressure
The choice of minimum interfacial stress should not be dissociated from the following key
factors:
cable stress at the insulation shield,
longitudinal component of stress at the base of the LV stress control profile
type of elastomer (EPR or silicone)
elasticity and hardness of elastomer
molding and machining methods employed to achieve a smooth bore
XLPE surface preparation method
type of lubricant
method of electrically terminating the cable insulation shield
The original technology came from the design of elastomeric stress cones used in cable
terminations, which had low stretch and low interfacial pressure, permitting them to be pushed
and stretched over the prepared XLPE surface by hand. The vulnerable points were a) the
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-13
electrical termination of the cable insulation shield, which was usually painted onto the XLPE
surface to minimize the change in radial geometry and b) the base of the stress control profile
which was easier to mold in a small stress cone. The stress cone also had the advantage that it
were immersed in an insulating fluid (or SF
6
gas) such that any voids present would be less
likely to dry out and discharge.
In early prefabricated joints the proven stress cone designs were simply incorporated into the
ends of the larger moldings
[iii]
. However the larger size of the molding made more difficult a)
molding a smooth bore and b) terminating the cable shield robustly. Key importance was placed
on the selection of a sufficiently high interfacial pressure to obtain and maintain an intimate fit.
Work in Japan in the mid to late 1980s showed that designers had selected a minimum
interfacial pressure of 0.15 MNm
-2
(1.5bar) for use in prefabricated composite joints in which the
interfacial pressure is achieved by placing an elastomeric stress cone in compression with an
annular bank of springs. At that time the Japanese cable makers had selected EPR in preference
to silicone rubber, as being more robust. These joints were designed to suit the lower stressed
XLPE cables in the Japanese voltage range of 66-154kV. A 1991 Jicable paper
[iv]
shows that the
pressure distribution in a UK design of 132kV prefabricated joint utilized 0.15 MNm
-2
at a) the
base of the stress cone and b) termination of the cable shield. A CIGRE 1988 paper
[v]
shows that
the pressure in an Italian design of 150-275kV one-piece EPR molding with a thin tail is 1-
0.08MNm
-2
at the stress cone profile and cable shield termination. A 1993 IEE conference
paper
[vi]
shows that the pressure at the shield termination in a French one-piece joint is
0.20MNm
-2
at 20
o
C

, falling to 0.15 MNm
-2
at 90
o
C.
Interfacial pressures were increased
[i][vii]
to greater than 0.4-0.5 MNm
-2
for prefabricated
composite joints to match the next generation of cable designs which employed the higher
electrical design stresses necessary for application in the voltage range 200-500kV. (The higher
EHV cable design stresses have since been employed to reduce cable diameters at HV system
voltages under 230kV and down to 110kV
[i]
). Supporting work
[viii]
showed that electrical
breakdown strength reduces at pressures under 0.3MNm
-2
, with a more pronounced reduction
under 0.25 MNm
-2.
A 1998 CIGRE paper describing the development of silicone rubber one-
piece joints for 400kV application
[ix]
shows that an interfacial pressure at the shield termination
of greater than 0.3MNm
-2
, is achieved by thickening the stress cone tail.
Recent development work in Japan is described in a Jicable 2003 paper
[x]
; tests on a 66kV
showed that satisfactory test performance could be obtained at the low pressure of 0.4MNm
-2
,
however this was attributed to the use of a low modulus (soft) silicone rubber to achieve an
intimate fit.
From the above literature review the benchmark minimum interfacial stress to evaluate the
thermomechanical performance of the EPR molding, in the particular 138kV 1500kcmil cable
size and stress, is selected to be 0.1MNm
-2
at the cable-joint interface and triple-point interfaces.
This does not mean that molding designs with lower pressures will not work electrically, but that
they will be less tolerant of variations in dimensions and quality that may occur in manufacture
and field assembly.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-14
The benchmark limiting parameters under the application of bending and differential axial loads
are:
the interfacial pressure shall not fall substantially at a triple point interface
the minimum interfacial pressure shall not be less than 0.1MNm
-2

Limiting Parameter: Longitudinal Movement
In the previous section it was described that a) sufficient length tolerances have to be included in
the design to ensure that the triple point interfaces are correctly overlapped to maintain shielding
and b) the cable-joint interface must not be disturbed in service. For example a) the termination
of the cable insulation shield must never protrude beyond the base of the LV stress control
profile and b) the base of the HV electrode must never expose the connector shield/XLPE
insulation cuts.
The limiting parameter is:
there shall be no longitudinal slip of the interface under cable axial load
Detailed Model of One-Piece Joint
The software selected for the modelling of the structural analysis was ABAQUS Standard, which
uses the FEA implicit method. Two similar models were constructed, a bending model of one
half of the joint and a cable pull-through model of the complete joint. Quadratic elements were
selected to give smooth interfacial pressure plots.
The cable was modeled in the core stage, comprising the conductor covered with XLPE
insulation, but with the outer jacket, foil and tape layers removed. The conductor and XLPE axial
stiffness, EA, and bending stiffness, EI, are those of the reference set for the 138kV 1500kcmil
cable at 20
o
C, listed in the Chapter 7 thermomechanical sensitivity study. The 90
o
C properties
were included as one case in the joint bending sensitivity study.
The one-piece molding was modeled using the elastic properties of typical EPR molding rubber
used in one-piece joints. The properties were measured in tension and compression at 20
o
C. The
two rubbers in the joint were characterised with a hyper-elastic material model, using the Van
der Waals form of the strain energy potential. A comparison of the effective elastic modulii at
30% strain shows that the insulating EPR is 25% more elastic than the semiconducting
compound:
insulating EPR: 2.66MNm
-2

Semiconducting EPR: 3.32MNm
-2

A sensitivity study was performed on the effect of varying EPR material elasticity. Models were
analysed using moldings comprising a) all-insulating EPR and b) all-semi-conducting EPR, for
both the insulating and semiconducting components.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-15
A sensitivity study was performed on varying the coefficient of friction at the interface. It is
experienced that a) immediately after pulling the molding onto the cable it can be repositioned
comparatively easily, thus the coefficient of friction is initially low, b) with time on long term
test the silicone fluid is absorbed and the friction becomes high such that movement is difficult
and c) traces of silicone grease persist longer than fluid in the interface. The thermomechanical
studies in Chapters 6, 7 and 8 have shown that the maximum axial forces occur early in the life
of the joint, usually during the first heating cycle, when some traces of the lubricant can be
expected to be present. Thus the values of coefficient of friction were selected to be 0.1. 0.2,
0.3, 0.4 and 0.5. The reference value of being 0.5.
The reference value of the one-piece joint stretch was taken to be 27%, the minimum 14% and
the maximum 31%.
The first load case in both of the models was to stretch the molding onto the cable at a coefficient
of friction of =0 and then to apply the required value.
Bending Model
The model Figure 11-11 has an axis of symmetry at the center of the molding. The total length of
cable is one meter. The molding half length is 336mm with 664mm of cable protruding. The
joint body was constrained at its central axis of symmetry and was supported horizontally on a
rigid plate to represent the support of the aluminum spider and adjacent two moldings.
Figure 11-13 shows the bending moment M applied at the cable end to replicate the bending
moments generated in the thermomechanical study. The average minimum bending radius
exhibited by the joint in the 138kV route being 2.8m, with an individual minimum of 1.8m.
Bending moments were calculated to achieve nominal bending radii of 20m, 10m, 5m and 2m,
assuming that the model comprised a one meter length of cable. This resulted in the achieved
bending radii being slightly smaller than the nominal value in the cable and larger in the main
body of the molding, as occurred in the 138kV route.

Figure 11-11
FEA model of 138kV joint stretched onto the cable for the bending study
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-16
Cable Pull-Through Model
The model is shown in Figure 11-12. A longitudinal tensile force F was applied at the right hand
end of the cable to represent the differential thermomechanical force. A fixed annular plate was
positioned at the right hand end of the molding to provide uniformly distributed constraint
against movement. The plate simulates the constraining presence of the joint casing and support
cradle within the manhole. (A similar plate is used during installation to pull the molding onto
the cable).

Figure 11-12
FEA model of 138kV joint for the cable pull-through study
Bending Behavior of One-Piece Joint
Results: Bending Radii
Figure 11-13 shows the joint bent to a nominal 2m radius. Figure 11-14 shows the distribution of
radii along the centerline of the model for the different load cases when the joint and cable is
bent through a nominal 20 to 2m radius. The centre of the molding over the conductor
connection and cylindrical shield is rigid and so has an infinitely large bend radius.
The reference positions used in this study are given in Figure 11-10:
a: interface of HV electrode with the EPR insulation in the bore of the molding
b: base of LV stress control profile
c: base of taper on cable insulation shield (the painted electrical termination of the LV shield is
12mm from this position)
d: top of taper on the cable insulation shield (this does not have a key technical significance, but
does have an identifiable peak in the interfacial pressure graph)
In the 20m radius case the cable is almost uniformly bent to 20m, but it is found that the radius
falls to a minimum value of 18.8m at 233mm within the end of the molding between points b and
c. This position is a) in the region where the LV stress cone tail has a marked reduction in
diameter and b) 54mm beyond the end of the rigid plate underneath the joint. This local
concentration of bending is present in each of the load cases with the bending radius being
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-17
reduced by a) 3% compared to the cable and b) 6-9% compared to the nominal bend value. A
similar concentration of bending at the tail was seen in an in-service 138kV installation as an
angular deflection. In the model output the deflection of the stress cone tail/cable to the main
body was measured as 8
o
.

Figure 11-13
FEA model of 138kV joint bent to 2m radius

Figure 11-14
Distribution of radii within 138kV joint for different applied bending radii
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-18
Results: Interfacial Pressure
Straight Cable and Joint
The reference case is an unbent joint with a coefficient of friction of 0.5 and a stretch of 27%.
The distribution of interfacial pressure is shown in a) Figure 11-15, the colour contour plot of the
inside of the molding and b) Figure 11-16, the graph of the distribution along the 12o/c, 9 o/c and
6 o/c axis.
The legend of the colour contour plot is that the darkest colour (darkest blue) is the zone of zero
pressure, this being present at the convex tip of the tail, where there is no surface contact.
Medium blue is 0.09-0.18MNm
-2
, this zone encompasses the base of the cable shield profile
close to the shield termination. The two lightest blue zones are 0.18-0.35 MNm
-2
, this being the
general background pressure under the tail of the stress cone. The next two zones (lightest and
darker green) are 0.35-0.53 MNm
-2
, this being the general pressure from the base of the stress
cone to the center of the molding. The graph in Figure 11-16 gives the values more precisely:
The pressure is highest underneath the thickest radial section of EPR in the main cylindrical
section.
The highest pressure in the centre body of 0.5MNm
-2
occurs a) underneath the HV electrode
and b) LV stress control profile, this is because the semicon EPR is 25% less elastic (and
harder) than the insulating EPR.
The pressure falls to typically 0.25 MNm
-2
between points b and c, where the outer diameter
reduces to that of the stress cone tail, however large fluctuations are present.
The interfacial pressure falls to a minimum of 0.14MNm
-2
at the base of the cable shield
profile. This is within 12mm of the tip of the painted shield termination, which is the highest
electrically stressed part of the joint. The pressure falls because the EPR molding act as a
structural bridge over the taper, with high pressures at the top of the profile and further along
the XLPE surface and with low pressure underneath.
The pressure distribution is identical along the 12o/c, 9 o/c and 6 o/c axis, resulting from the
straight alignment of the molding and cable.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-19

Figure 11-15
Interfacial pressure contours within 138kV joint: reference case straight

Figure 11-16
Interfacial pressure distribution within 138kV joint: reference case straight
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-20
Cable and Joint Bending
The colour contour plot of the molding bent to a 2m radius is shown in Figure 11-17, the plot
uses the same colour legend as Figure 11-15 for the straight case. The interfacial pressure
distribution is now circumferentially non-uniform. In particular, the 0.09-0.18 MNm
-2
low
pressure zone in the tail has increased in width along the top of the bore. A low pressure zone of
0.27-0.35 MNm
-2
has appeared in the main body at the 3o/c axis. At the same position a zone of
high pressure, 0.71-1.1 MNm
-2
, has appeared at the bottom of the bore above the supporting
plate. The external shape of the mold has been distorted by bending.

Figure 11-17
Interfacial pressure contours within 138kV joint: bent to 2m radius
Bending Sensitivity Study
Figure 11-18 shows that the interfacial pressure distributions are no longer equal at the 12o/c, 9
o/c and 6 o/c axis. The minimum pressure at the base of the cable shield taper has fallen at the
12o/c axis by 6% to 0.13 MNm
-2
. The longitudinal positions of the pressure peaks at the base and
top of the shield profile are no longer coincident due to a) the tensile stretch of the EPR on the
outside of the bend at the 12o/c position and b) the compressive reduction at the 6o/c position on
the inside of the bend.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-21

Figure 11-18
Interfacial pressure distribution within 138kV joint: bent to 2m radius
Two points of reference were taken for the bending sensitivity study. The first position is the
base of the cable shield profile, where the minimum pressure within the molding occurred,
Figure 11-19. The second position is the triple point at the end of the HV electrode, Figure
11-20.
The minimum interfacial pressure at the base of the taper is now 0.05MNm
-2
and occurred in the
2m bend case with the lowest molding stretch of 14%. This is a pressure reduction of 61%
compared to the reference case and is due primarily to the low stretch and not the bending.
Figure 11-21 is the color contour plot of the minimum stretch case showing that the pressures
have fallen throughout the molding. The zones of minimum pressure have grown significantly in
width under the tail, particularly at the 12o/c and 3o/c positions.
The highest interfacial pressure at the base of the taper of 0.18 MNm
-2
was produced by
increasing the stretch to 31%. This is an increase of 34% compared to the reference case.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-22

Figure 11-19
Variation of interfacial pressure at cable LV shield termination with different bending radii
under different conditions

Figure 11-20
Variation of Interfacial pressure at HV electrode to EPR insulation interface with different
bending radii under different conditions
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-23

Figure 11-21
Interfacial pressure contours in 138kV joint bent to 2m radius: minimum stretch sensitivity
case
Table 11-2 ranks the sensitivity study parameters in order of largest effect at the base of the
taper. It is shown that:
thermomechanical bending down to 2m radius has minimal effect (6% reduction) on the
minimum interfacial pressure
the reference case pressure of 0.13 MNm
-2
is above the benchmark value 0.1MNm
-2
, (11%
under it).
the stretch of the molding has the largest single effect; in particular low values of stretch
significantly reduce interfacial pressure, with the minimum pressure of 0.07 MNm
-2
being
below the benchmark of 0.1MNm
-2
.
the elasticity of the EPR selected for the molding has the second largest effect; high elasticity
(soft) EPR reduces interfacial pressure.
comparison with the interfacial pressure in the thick section, Figure 11-20, shows that the
radial thickness of the molding and in particular of the harder semicon electrode rubber has a
significant effect on increasing the pressure. Reduced the thickness of the molding, reduces
the interfacial pressure, especially underneath the tail of the molding.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-24
Table 11-2
Ranking of bending parameters on interfacial pressure at taper base
Interfacial Pressure Ranking Parameter
[MNm
-2
] [%]
Reference 27% stretch, straight molding 0.133 0
Interfacial Pressure Reduction
1 Minimum stretch, 14% 0.072 -46
2 All insulating EPR molding 0.106 -20
3 Bending radius, 2m 0.125 -6
Interfacial Pressure Increase
1 Maximum stretch, 31% 0.185 +39
2 Cable elasticitys at 90
o
C 0.153 +15
3 All semicon EPR molding 0.140 +5
Results: Cable Pull-Through Withstand Strength
The cable was pulled though the joint at a constant smooth rate with measurement of the cable
reaction force being made at 3.4mm intervals of cable movement. The coefficient of friction
was varied between 0.1 and 0.5. The cable movement was absolute, but the molding movement
comprised an elastic deformation and interface slip. The objective was to establish the force at
which the molding ceased to hold the cable by friction, allowing the interface to slip and the
cable to rapidly pull-through.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-25

Figure 11-22
Variation of pull-through force with cable movement and interfacial slip; the data points
are distributed uniformly at 3.44mm intervals. RH and LH refer to right and left hand ends
of molding
Interface Slip
A single, unique slip event did not occur at the interface. Figure 11-22 shows that the force rose
non-linearly with cable movement up to a value of 23,428N for a cable movement of 89mm, at
which point the cable continued to be pulled-through at a near constant force. This initially was
thought to be the point of slip, but close inspection of the EPR and XLPE elements on either side
of the interface showed that some slip had already occurred at lower force. The interface was
inspected for slip at a) the triple point at the base of the stress control profile on the right hand
side of the molding, this being the end in the direction the cable was being pulled, b) the mid-
point of the molding and c) the equivalent stress control profile on the left hand side of the joint.
It was found that:
The interface at the right hand end slipped almost immediately; the interface withstood
2,105N at a total cable movement of 3.4mm, but had slipped within the next 3.4mm. The
withstand point is shown on the graph.
The centre of the joint withstood a force of 17,404N at a total cable movement of 41.3mm
before the interface slipped. At this time, the right hand interface had already slipped by
(41.3-3.4)mm=37.9mm.
The left hand end withstood a force of 23,428N at a total cable movement of 89.4mm before
the interface slipped. At this time the right hand had already slipped 86mm and the centre
had slipped 48mm.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-26
When the region of slip reached the extreme left hand end of the molding, the cable pulled
out at the near constant force of 23,428N.
The force-movement characteristic is shown to be comprised of distributed slip, with the
point of slip progressing from right to left as the force builds.

Table 11-3
Cable movement and force for slip at different coefficients of friction
Coefficients of Friction
0.1 0.3 0.5
Position on
Interface
Cable
Movement
[mm]
Withstand
Force [N]
Cable
Movement
[mm]
Withstand
Force [N]
Cable
Movement
[mm]
Withstand
Force [N]
Right hand
of stress
cone base
3.4 1,839 3.4 2,098 3.4 2,105
Center of
molding
10.3 4,050 27.5 11,478 41.3 17,404
Left hand of
stress cone
base
24.1 6,255 65.4 16,651 89.4 23,428

Table 11-3 shows that for values of between 0.1 and 0.5:
The interface at the right hand stress control profile withstood a movement of 3.4mm, but
slipped before the next 3.4mm occurred
The slip withstand force at the right hand stress cone did not increase pro-rata with the 5 fold
increases in friction. Friction increased the pull-through force by 14%.
The pull-through force for slip of the complete interface was increased by x 3.7, which is
more comparable to the x 5 increase in friction.
The reason for the progressive slip was that the EPR molding deformed elastically under the
frictional grip, shortening its length and increasing its diameter. Thus particular points on the
molding were able to move towards the right in synchronism with the cable without slip. Figure
11-23 shows the distortion in the molding for =0.5
Void Formation at the Interface
During initial cable movement it was observed that the tail of the LV stress cone had distorted
and lifted off the cable, creating a void, as shown in Figure 11-23.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-27

Figure 11-23
Distortion of one-piece molding during cable pull-through with cable coefficient of friction
of 0.5
The interfacial pressure was plotted to investigate the point of void formation. Figure 11-24 is
the reference distribution prior to applying pull-through force. Figure 11-25 is the distribution at
the point the pressure falls to zero. This occurred at a force of 9,678N after a cable movement of
13.8mm, comprising 3.4mm elastic movement and 10.4mm interface slip. The location is 64mm
from the base of the cable LV shield profile, which is where LV electrode is required to shield
the exposed cable insulation. A void could be seen to have formed at the interface by the next
output step at a force of 11,237N and movement of 17.2mm.

Figure 11-24
Interfacial pressure in one-piece molding with coefficient of friction of 0.5, prior to
applying pull-through force
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-28

Figure 11-25
Interfacial pressure in one-piece molding with coefficient of friction of 0.5; formation of
void during cable pull-through
Table 11-4 shows that:
Interfacial pressure falls to near zero for a cable movement of 10.3mm and reaches zero at
the movement of 13.8mm.
The region of zero pressure grows to a width of 11mm for a movement of 41.2mm.

Table 11-4
Void formation under LV stress cone in molding during cable pull-through
Pull-Through
Force [mm]
Total Cable
Movement [mm]
Interface Slip
[mm]
Interfacial
Pressure [MNm
-2
]
Length of Zero
pressure [mm]
5,287 6.9 3.4 0.136 -
7,809 10.3 6.9 0.049 -
9,678 13.8 10.3 0 0
17,404 41.2 37.8 0 11
23,428 89.4 86.0 0 16
The formation of a void under the tail of a stress cone during insertion has previously been
reported in Japanese papers and in a UK paper
[iv]
in 1991. It is associated with the transition in
the molding from a region of hoop compressive stress in the body of the molding to a region of
hoop tensile stress in the tail.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-29
The pull-through study has shown that the limiting condition occurs when:
Interface slip occurs at the end of the joint above a withstand force of 2,000N with a
maximum allowable slip, of less than 3mm.
The interfacial pressure falls to near zero and an electrically stressed void forms at a force of
8,000N at a movement of 10.3mm.
For safety, the molding should not be subjected to any differential forces. Although it appears
to have a withstand capability of around 2,000N this would have to be demonstrated by long
term electrical tests under axial load to ensure a safe and repeatable performance.
Outline Test Schedule
The study has quantified the forces and movements that can occur in a pipe system and has
compared these to benchmark limiting parameters.
Chapters 7 and 8 have shown that large forces and movements can occur in duct and pipe
systems. Depending primarily on a) the conductor size and b) cable to pipe clearance these can
be significantly in excess of those considered in this chapter.
It is recommended that one-piece joints be protected from high compressive loads, differential
axial loads and from uncontrolled transverse and longitudinal movements. Designs of constraint
to protect joints within manholes are considered in the next chapter.
It is recommended that, if doubts exist on the effectiveness of the constraining system, then the
joints and manhole constraining system together, be subject to development and qualification
tests based on the following outline test schedule:
Criteria for absence of slip:
Whilst under the application of axial and bending loads, a)there shall be no differential slip at
any position in the joint-cable interface and b) the entry of the cable at the ends of the
molding shall remain straight and in-line with the axis of the molding. Axial loads shall be
calculated for the particular application and applied to the test rig. Differential slip at the
interface can be detected by such means as a) radial witness holes drilled through the
molding and cable after assembly or b) lead markers pre-located in the cable and molding
prior to assembly, for subsequent radiographic examination.
Criteria for absence of voids:
Whilst under the application of axial and bending loads, there shall be no voids present at any
position in the electrically stressed interface. The presence of voids at the interface can be
detected by a) ultrasonic examination, b) visual examination using a Plexiglas mandrel to
replicate the cable geometry or c) an electrical partial discharge test. It is noted that voids
filled with silicone lubricant can be detected ultrasonically, but not by partial discharge test.
Criteria to demonstrate minimum interfacial pressure:
The supplier shall declare, for the particular molding and cable size a) the minimum design
interfacial pressure and provide the distribution along the interface and b) the maximum
permissible cable insulation shield electrical stress and provide the interface stress
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-30
distribution. The joint in a configuration representative of the particular service installation
shall be subjected to mechanical tests involving load cycling to demonstrate that the
interfacial pressure does not fall below the declared design value. The presence of low
interfacial pressure in the mechanical tests can be detected by the pre-insertion of pressure
transducers into the cable surface.
Criteria to demonstrate satisfactory electrical withstand at the declared interfacial pressure:
The joint in a configuration representative of the particular service installation shall be
subjected to electrical qualification tests involving load cycling to demonstrate electrical
withstand and absence of partial discharge. This test is not required if the above mechanical
tests demonstrate that the design of manhole constraint fully protects the joint against axial
loads and bending.
Conclusions
The modelling of a manhole position has shown that the joints and cable within the casing
experience a pronounced deflection pattern similar to that seen in an operational 138kV
1500kcmil pipe system.
The movements and forces are explained by the location of the joint position being at the end
of a straight span adjacent to a bend. Thermal expansion of the cable in the straight span
creates high axial thrust and pushes cable though the casing into the bend. The compressive
loads cause the cables and joints to deflect laterally, such that they contact the casing and
support a differential longitudinal load. The sensitivity studies of cable thermomechanical
behavior in Chapters 7 and 8, indicate that larger loads and movements may be generated.
The loads and movements for the 138kV 1500kcmil joint are:
Compressive axial thrust of 9,000N with a possible maximum of 16,000N.
Differential longitudinal force of 2,000N with a possible maximum of 4,000N.
Minimum bending radii 2-3m with a spot minimum of 1.8m.
Longitudinal alternating movements of 25mm with a theoretical maximum of 58mm.
Detailed models were constructed of the 138kV one-piece molding. The joint behavior was
analyzed under the application of bending and cable pull-through forces. The output was
compared to benchmark limiting parameters:
The minimum pressure at the electrically stressed joint to cable interface should not
fall below 0.1MNm
-2
, at triple point interfaces: a) tip of the cable LV insulation
shield, b) base of the joint LV stress control profile and c) tip of the joint HV
electrode.
Differential longitudinal movement (slip) should not occur along any part of the
interface.
Bending to a 2m radius was shown to have minimal effect on reducing the interfacial
pressure i.e. a 6% reduction. This explains why the 1500kcmil joints in the operational
138kV circuit have been able withstand the lateral deflection. However a) the rate of pressure
reduction increases with smaller radii and b) bending was concentrated at the tail of the
molding, where the highest electrical stress exists.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-31
The sensitivity study on bending performance showed that:
The minimum interfacial pressure in the unbent molding is close to one of the
vulnerable triple point interfaces. At 1.33MNm
-2
the interfacial pressure is greater
than the 1MNm
-2
benchmark. This position is at the base of the cable shield profile
and is within 12mm of the electrical termination of the LV shield, the highest stressed
region in the joint.
The percentage stretch of the molding onto the cable diameter has the greatest effect
on interfacial pressure. The 27% stretch used in the model is in the upper part of the
manufacturers range. A 14% stretch reduced the interfacial pressure by 46% to
0.07MNm
-2
.
The elasticity of the EPR molding compound has the second largest effect on
interfacial pressure; a 25% increase in elasticity reduces interfacial pressure by 20%
to 0.11MNm
-2
.
The radial thickness of the molding and in particular of the semicon electrodes has a
significant effect; reduced thickness reduces interfacial pressure.
The joint is shown to have a withstand capability of 2,000N under longitudinal axial load.
Above this load part of the electrically stressed interface will start to slip. At a load of
8,000N the tail of the stress cone lifts and a void is formed over the unshielded cable
insulation. As the force was calculated to be 2,000N in the 138kV pipe system, the moldings
appear to be close to slipping. Possible explanations of why this has not become apparent are:
The use in the model of an annular constraining plate with a small bore may have
reduced the force at which the stress cone slipped compared to the service
configuration.
The differential reaction load supported by the casing is shared by both the joint and
the adjacent cable in the deflection pattern.
The axial component of force may have been reduced by the deflection geometry.
Friction at the interface may be above =0.5 and, if stiction is present, above =1.
The axial load may have relaxed with time.
The cable bending stiffnesses were modeled as being elastic. In reality the properties
also have components of creep and friction, which would cause the cable deflections
to remain in place during the cooling cycles, reducing the load variations and perhaps
increasing the restraining effects.
Interface slip of less than the values of 10.3-13.8 mm required to produce a void may
have occurred without initiating partial discharge.
The generic family of one-piece straight joints is not designed to withstand uncontrolled
bending and differential axial forces. However the study has shown that the particular 138kV
1500kcmil design does have certain performance capabilities. It is recommended that these
calculated values should not be relied upon without performing mechanical and electrical
tests to confirm a repeatable and safe performance.
EPRI Licensed Material

Mechanical Performance of a One-Piece Straight Joint
11-32
It is recommended in pipe and duct systems that one-piece joints be protected from high
compressive loads, differential axial loads and from uncontrolled transverse and longitudinal
movements. Designs of constraint to protect joints within manholes are considered in the
next chapter.
It is recommended that, to prove the efficiency of a particular design of constraint,
mechanical development and qualification tests be performed on a simulated manhole
assembly. The analytical work in this chapter has identified a) the limiting parameters and b)
the modes of distress that can be applied as withstand criteria. This requirement applies to
both conventional straight joint and anchor joint designs.


i
Cable System Technology Review of XLPE EHV Cables, 220kV to 500kV, EPRI, Palo Alto, CA: 2002. 1001846
ii
Interfaces in Accessories for Extruded HV and EHV Cables, CIGRE Technical Brochure 210, prepared by Joint
Task Force 21/15, August 2002
iii
138kV Splice for Extruded Dielectric Cables, EPRI, Palo Alto, Report EL354, 1977
iv
Attwood JR, Gregory B, Svoma R, A Range of Transition Joints for 33kV to 132kV Polymeric Cable, Jicable1991
Paper A.5.3
v
Parmigiani B, Premolded Accessories for High Voltage Extrudud Insulation Cables, Paper 21-08, CIGRE 1988
vi
Claeys F, Premolded Joints for Voltages from 72 to 245kV, IEE Third International Conference on Power Cables
and Accessories, 1993, London
vii
Okada M et al, Development of Prefabricated Joint for 500kV XLPE Cable, Jicable 99, Paper A.5.2
viii
Imai N, Andoh K, Development of Pre-fabricate Joints for 275kV XLPE Cables, Jicable 1991 Paper A.5.4
ix
Van der WijK GP, Pultrum E, Geene HTF, Development and Qualification of a New 400kV XLPE Cable System
with Integrated Sensors for Diagnostics, Paper 21-103, CIGRE 1998
x
Kobayashi et al, Development of Factory-Expanded Cold-Shrinkable Joint fo EHV WLPE Cables, Jicable 2003,
Paper A.5.1


EPRI Licensed Material
12-1
12
DESIGN OF CONSTRAINTS IN MANHOLES
Introduction
The thermomechanical modeling work on full sized cable spans in both pipe and duct systems
has shown that the most vulnerable position to axial force is the manhole containing the straight
joints.
It has also been shown that any curvature formed at the manhole entrance will initiate
thermomechanical wave patterns in the adjacent pipe or duct. There is thus a two way interaction
between the manhole and the pipe/duct span.
The studies of the effect of traffic vibrations and short circuit forces have shown that these
precipitate the sudden release of locked-in thermal strain into thermomechanical wave patterns.
The wave patterns are suddenly formed over a major part of the span. The momentum of this
longitudinal movement is likely to impose an impulsive shock to the manhole and its contents.
This is in addition to the direct application of short circuit electromechanical forces between the
cables and joints within the manhole. Thus the manhole constraining system requires a factor of
safety above that necessary to withstand the day to day thermomechanical forces in normal
service.
In this chapter the performance of existing designs of manhole constraints is modeled and
analysed and compared to the forces that can be generated by the cable in the adjacent spans.
Design improvements are identified and generic designs of constraint recommended for both
pipe and duct manhole systems.
Pipe System Constraints
Two basic types of constraints exist for a pipe system. The traditional method is to install the
joints in a larger diameter casing in line with the pipe. This permits the pipe and casing to be
welded together a) to make the system pressure/water tight and b) to provide a continuous
conductor for the passage of short circuit current. The alternative method is to allow the cables to
emerge from the pipe and to mount them for jointing via offsets vertically on the wall of the
manhole.
A variant of the design of cleated offset is the semi-flexible design that a) part cleats the cable
and b) part allows the joints to swing from chains fixed to the roof of the manhole. This was
evaluated at the beginning of the project, but as it did not meet the requirements to constrain the
EPRI Licensed Material

Design of Constraints in Manholes
12-2
levels of force and movement identified in the thermomechanical model, this method was not
pursued further.
Pipe System: In-Line Joint Constraints
Chapter 11 studied the performance of straight joints incorporating one-piece moldings in a
138kV pipe system. It was shown that the joints and cable within the in-line casing exhibited
significant deflection (Figure 11-2 and Figure 11-3). The cable and joints deflected laterally and
longitudinally and rotated within the casing. The cable and joints started to deflect at low
temperature i.e. within a 15 C temperature rise, this being before the cable in the pipe route
exhibited wave patterns at pipe bends.
The conclusion was that the joints and cables within manholes needed to have an improved
design of constraint to withstand the force and movement of the cable. Initial design methods
were first evaluated using detailed models of the joints within the casing.
In-Line Joint Model
Figure 12-1 shows cleats positioned on the cable at the ends of the one-piece moldings. These
cleats were initially given perfect characteristics i.e. of infinite holding strength. It was found
that although the joints were held in straight alignment, the cables exhibited excessive lateral
deflection.

Figure 12-1
Cable deflection with cleats at ends of joint moldings
Figure 12-2 shows two trefoil cleats positioned within the small sleeve adjacent to the pipe
entries. These perfect cleats fully protected the cable and joints within the casing.
The next stage was to achieve the ideal performance in reality.
EPRI Licensed Material

Design of Constraints in Manholes
12-3

Figure 12-2
Cable deflection with trefoil cleats at ends of casing
In-Line Joint: Improved Constraint Design
The objectives for the improved constraint were that the cable and joints should be constrained
better than within the adjacent pipe span by a) reducing lateral clearance and b) preventing
longitudinal movement. The following requirements were identified.
The cables and joints are to be centered along the axis of the casing to preserve symmetry
and minimize bending moments.
The cables are to be prevented from bowing outwards under compressive force by limiting
the span length between cleats.
The joints are to be held in parallel alignment with each other and with the casing.
The cables and joints are to be prevented from rotational movement to minimize bending
moments.
The cables are to be cleated to withstand longitudinal differential force and thence to prevent
longitudinal movement.
The joints are to be allowed to slide lengthwise to prevent differential force appearing
between the molding and cable core.
Figure 12-3 is the general assembly diagram showing the six positions where the cleats and
spiders are located.

Figure 12-3
General assembly showing positions of cleats and spiders
EPRI Licensed Material

Design of Constraints in Manholes
12-4

Figure 12-4
Trefoil cleat in smallest sleeve
Figure 12-4 shows the trefoil cleat which is located at the end of the small sleeve. The cleat
comprises an aluminum frame which is bolted through the steel sleeve. The bolt positions are
covered with external screw caps which are provided with O rings to permit the pressure
retaining properties of the system to be maintained. The cables are held in clamp cleats 100mm
long similar to the type which were tested at EPRIsolutions. The cleats are rubber lined to protect
the cable. The cleats are inclined to the longitudinal axis to permit the cables to splay outwards.
This is achieved by the base of the cleat being tapered at its interface with the aluminum frame
(not shown in this view).
EPRI Licensed Material

Design of Constraints in Manholes
12-5

Figure 12-5
Centralizing spider in medium sleeve
Figure 12-5 shows the spider which centralizes and binds together the cables within the casing,
thereby preventing the cables from bowing and increasing the rigidity of the structure. The spider
comprises an aluminum frame, into which the cables are located, together with three adjustable
legs. The cables are bound tightly to the frame by a synthetic fabric strap. Rubber liners are
provided to take up the tolerance between the cable and frame.
EPRI Licensed Material

Design of Constraints in Manholes
12-6

Figure 12-6
Joint constraint in large sleeve, style A
Figure 12-6 shows the three one-piece joint moldings. The assembly is an extension of the
present arrangement. It employs a central aluminum spider which the modeling showed to be
effective in keeping the joints parallel. The two support cradles have been modified in dimension
to center the three joints with the axis of the casing. Two design variants exist:
Style A, as shown in Figure 12-6 is the recommended method. The three joint moldings are
clamped by the fabric belts to the upper support cradle, which is allowed to slide freely
lengthwise on the lower support cradle. A vertical slot and peg ensure that the upper cradle can
move longitudinally but not rotate. This arrangement removes differential load between the
molding and cable core, thus minimizing the risk of disturbance to the electrically stressed
interface. This arrangement can be used within any manhole position along the route, providing
that the two trifurcating cleats at the ends of the casing can withstand the differential force.
Style B (not shown) is the method for use if a future joint design becomes available with a semi-
anchor facility that is capable of withstanding a guaranteed level of differential force. A fixing
bolt through the casing at the 6o/c position holds the two support cradles to prevent their
longitudinal movement. Two fabric binding belts clamp the joints to each other and to the casing
via the support cradle and fixing bolt. This arrangement increases the longitudinal withstand
strength of the joint and cables, but imposes a differential longitudinal force between the
molding and the cable core. The study in Chapter 11 showed that the differential force has to be
limited to less than 2000N to prevent the electrically stressed interface from slipping. The
thermomechanical modeling study has recommended that the differential force be calculated at
each manhole along the route.
EPRI Licensed Material

Design of Constraints in Manholes
12-7
Pipe Systems: Offset Joints
The positioning of joints on the wall of a manhole is the standard method for use with ducts. The
contents of this section are equally applicable to the design of duct manhole constraints and will
be referred to later.
The differences are:
Upon leaving the close fitting pipe the three cables have first to be cleated in line with the
pipe to withstand longitudinal force.
The cables then have to be spread out both horizontally and vertically using rigidly cleated
offsets.
Pipe systems have been shown in the previous chapters to generate a lower longitudinal axial
force than duct systems. Thus although the geometric arrangement of the cleats is more difficult,
their holding duty is reduced.
Offset Joint Model
Figure 12-7 is a perspective view of the FEA model looking downwards into the manhole (not
shown). In this particular diagram the joints are supported by guide cleats which do not clamp
the joint. They allow longitudinal movement, but prevent lateral movement.
Symmetrical forces were applied to the joint, thrusting into the manhole from each end.
The cleat design is based on that tested at EPRIsolutions Figure 12-8. Note that helical coil
springs have since been added to permit the radial thermal expansion of the cable to be taken up.
It is good practice to allow a minimum of 2mm of expansion movement for an XLPE cable
[i, ii]
.
This prevents the cable suffering permanent compression set when hot, followed by loss of cleat
holding strength when cold. The cleat was given the withstand characteristics of that measured
on the 138kV cable, as recorded in Chapter 3. The characteristic is that of the 100mm long, two
bolt cleat, tightened to 68Nm (50lbf ft). A typical cleat characteristic is shown in Figure 12-9,
with values being listed for 138kV and 230kV cables at different tightening torques in Table 12-
1. The cleat withstands the longitudinal load without slipping up to force P1. Up to force P1, a
longitudinal elastic stretch of the rubber liner of up to 4mm occurs. Between forces P1 and P2
the rubber liner has reached the limit of its extension and a maximum 1mm of slip occurs. At
force P2 the cleat allows the cable to slip through, but retains a constant frictional constraint.
EPRI Licensed Material

Design of Constraints in Manholes
12-8

Figure 12-7
Perspective view of model of offset joints in pipe system

Figure 12-8
100mm long clamp cleat, as tested at EPRI solutions, with springs added

EPRI Licensed Material

Design of Constraints in Manholes
12-9

Figure 12-9
Modeled withstand characteristic of clamp cleat from test results

Table 12-1
Sizes of clamp cleat used in the modeling work
Force [N] Cleat designation Cleat description
Elastic limit Slip
1300/2167 2 bolt, 100mm long,
125mm dia cable
1300 2167
3000/5000 2 bolt, 200mm long,
125mm dia cable
3000 5000
6500/10833 2 bolt, 100mm long,
83mm dia cable
6500 10833
15000/25000 8 bolt, 530mm long,
125mm dia cable
15000 25000
Figure 12-10 is an elevation view of the model of an offset manhole showing the cable deflection
in the vertical plane that occurred at 16,000N. The deflection has been magnified by a factor of
5. The 16,000N force was identified in Chapter 11 as being the maximum likely to be developed
with 138kV 1500kcmil cable in the hockey stick shaped model. The force derived from the
EPRI Licensed Material

Design of Constraints in Manholes
12-10
model of a full sized in-service span was 9,000N. It can be seen that the 800mm span length
between the trefoil cleat and first single clamp cleat is too long as the cables have bowed
inwards. A similar deflection can be seen in the 830mm span between the two cleats adjacent to
the joints. The cable between the center two cleats spaced at 630mm has a reduced deflection.

Figure 12-10
Deflection of cables between clamp cleats, elevation view
Figure 12-11 is a plan view of the same model as that depicted in Figure 12-10. Figure 12-11
shows the cable deformation in the horizontal plane. The deflections have been magnified by a
factor of 5. It can be seen that, as well as bowing in the vertical plane, deflections are present in
the horizontal plane. The spans with the greatest deflections are those with the longest distances
between cleats. These are the spans adjacent to the mouth of the duct and adjacent to the joint.
The modeling work showed that the distance between the cleats needed to be reduced.

Figure 12-11
Deflection of cables between clamp cleats, plan view
EPRI Licensed Material

Design of Constraints in Manholes
12-11
Offset Joint: Sensitivity Study on Offset Geometry and Cleat Strength
Figure 12-12 is the generic model used in the sensitivity study. The reference case model has an
offset angle of 60, a bending radius of 15D and cleats spaced at 300mm where D is the cable
diameter. Chapter 4 identified 15D as the recommended controlled bending radius for an XLPE
cable with an aluminum foil sheath. A bending radius of 12D is recommended for an XLPE
cable with an extruded lead or aluminum sheath. Table 12-2 lists the five sensitivity study cases.
These models were run for each of the four cleat characteristics in Table 12-1. Models were also
analysed in which a) the cleat spacing was increased to 600mm, b) a high strength cleat was
positioned at the mouth of the duct and c) a second high strength cleat was added adjacent to the
joint.

Figure 12-12
Model of generic clamped offset used in sensitivity study

Table 12-2
Geometry of offsets modeled in the sensitivity study
Offset design Radius R [m] Angle
[degrees]
No of cleats Offset Length
- L [mm]
Offset S
[mm]
1 5D 60 9 690 360
2 15D 60 15 2060 1050
3 15D 30 11 1160 280
4 15D 90 19 2550 2550
5 30D 60 23 4130 2100


EPRI Licensed Material

Design of Constraints in Manholes
12-12
Figure 12-13 shows the results from offset model No.4, employing 19 cleats. It shows that the
highest strength cleat design significantly increased the force differential that can be withstood
across the offset.
The cleat with the 1300/2167N withstand is the 100mm long prototype design tested on the
230kV 2500kcmil 125mm cable. In the 19 cleat offset the cleats had a withstand capability of
30,000N. It is shown that the first 9 cleats exceeded the limiting force P2 and slipped. The
remaining 10 cleats held.
The 3000/5000N cleat type is an intermediate strength 100mm long design. This has increased
the differential withstand to 40,000N. It is shown that the first 3 cleats slipped and the remaining
16 cleats held.
The 6500/10,833N cleat is the 100mm long design tested on the 138kV 1500kcmil 83mm cable.
This has increased the withstand force to 45,000N. All 19 cleats held in this model.
The 15000/25000N cleat is the 530mm long design used as an expansion cleat with the 230kV
2500kcmil 125mm cable. Note that this cleat is impractically long to be installed in the offset at
300mm spacing but was included to provide a limiting value for the sensitivity study. It is shown
that in this model the withstand strength of the offset was marginally increased to nearly
50,000N. All 19 cleats held in this model. Thus the number of cleats could have been reduced.
The shape of the force characteristics for cleat types 6500/10833 and 15000/25000 were
analysed and found to be of the form:
F/Fin = exp (-kn) Equation 12-1
Where:
F: Force at cleat n [N]
Fin: Cable input force [N]
n: Cleat number (or distance along offset, if cleats are evenly spaced)
k: A constant dependent on the cleat characteristic
EPRI Licensed Material

Design of Constraints in Manholes
12-13

Figure 12-13
Force distribution along offset, 19 cleats, 90
0
angle, 15D bend radius
Figure 12-14 compares the force characteristics for the lowest strength cleat type 1300/2167N in
the 60
0
angle and 15D bend radius model. The offset with 8 cleats spaced at 600mm supported a
force differential of 13,000N. Increasing the number of cleats to 15 by reducing the spacing to
300mm increased the differential to 18,000N.
The replacement of the first cleat at the input end with the 15000/25000N design of high strength
cleat increased the withstand significantly to 20,700N. The addition of a second high strength
cleat at the output end had a negligible effect.

EPRI Licensed Material

Design of Constraints in Manholes
12-14

Figure 12-14
Effect of a) increasing the number of cleats and b) addition of stronger clamp cleat(s), 60
0
angle and 15D bend radius
The bar charts in Figure 12-15, Figure 12-16 and Figure 12-17 summarize the performance of
each of the offset geometries in the sensitivity study. In each case an input cable force of
50,000N was applied:
The number of cleats is the most important variable, this being highest in the models with the
largest bending radii and largest offset angles.
The largest force differential (lowest output force) was achieved with the highest strength
cleat.
The shortest length model with a 5D bending radius and 60
0
angle containing 9 cleats was the
only model in which the three lowest strength cleats, types 1300/2167, 3000/5000 and
6500/10833 all slipped. In the other models types 1300/2167 and 3000/5000 slipped and
types 6500/10833 and 15000/25000 held.

EPRI Licensed Material

Design of Constraints in Manholes
12-15

Figure 12-15
Summary of sensitivity study: Percentage of cleats holding without slip

Figure 12-16
Summary of sensitivity study: Number of cleats holding without slip
EPRI Licensed Material

Design of Constraints in Manholes
12-16

Figure 12-17
Summary of sensitivity study: Variation of output force with offset geometry and cleat
characteristic
Offset Joint: Design of Offset for 138kV 1500kcmil Pipe System
A plan view of the manhole is shown in Figure 12-18. The manhole width and length are the
same dimensions as employed with the in-line design of joint. The joints have been moved to
within 400mm of the wall to provide room for jointing. Two separate circuits are shown entering
the manhole, their centerline distance has been reduced to 850mm to provide a greater offset.
The offset design selected has a 15D bending radius to suit the aluminum foil water barrier and a
30
0
angle to suit the manhole geometry. Ten cleats were selected, five on each side of the joint,
with a nominal spacing of 450mm.
EPRI Licensed Material

Design of Constraints in Manholes
12-17

Figure 12-18
Pipe System: Improved design of offset joints in vertical alignment on walls of manhole,
two pipe circuits are shown
Plan view
The cleat selected was the 6500/10833N, 100mm long design that was tested at EPRIsolutions,
with the addition of temperature compensating springs (Figure 12-8). The longitudinal force
withstand characteristic between the inlet and the exit of the manhole was calculated from Figure
12-19.
A manhole outlet force of 6500N was taken, this is representative of a) the holding strength of
one cleat or b) the magnitude of the opposing force from the cable span on that side of the
manhole. The force differential that can be supported is 9400N. This is greater than the 2000N
differential calculated across the in-line joints in a true 138kV manhole. The compressive force
present in the particular manhole was 9,000N, which would normally be nearly balanced by an
opposing force from the opposite cable. In the unlikely event that this load was completely
unbalanced, it could also be withstood by the cleats without slipping. The offset design was
judged to be satisfactory.
As the design of cleated cable offset withstood the force differential; it is unnecessary for the
straight joint to be of the anchor type. However to prevent any possible differential force or
movement occurring between the molding and the cable core the joint itself is not clamped. The
joint is allowed to move longitudinally within guide cleats. The allowance for movement is
8mm, which is the elastic movement of one cleat with a design margin of 2.
EPRI Licensed Material

Design of Constraints in Manholes
12-18

Figure 12-19
Input/output force characteristic across ten cleats, 30
0
15D model, 138kV cable,
6500/10833N cleat design
Duct System Constraints
An important difference between duct and pipe systems is that the clearance between one cable
and the duct is very much smaller e.g. 40mm compared to 130mm. Although thermomechanical
patterns form within the operating temperature range, the waves are unable to efficiently absorb
the thermal strain and so the axial force will be much higher. For example 80% of the rigid bar
force compared to 25% in a pipe system. Large cables with conductors of 2000kcmil/5000kcmil
are invariably installed in ducts. These cables develop axial thrusts approximately in proportion
to their conductor area. The theoretical rigid bar maximum for a 3000kcmil 345kV cable being
100,000N. Although conductor relaxation may reduce this, the resultant thrust into a manhole is
still of very high magnitude.
Two basic methods were selected to constrain the cable in duct systems.
The expansion cleat method is a semi-flexible system, which allows some of the thermal
strain to enter the manhole under the control of a spring-loaded expansion cleat, thereby
reducing the magnitude of the thrust.
The clamp cleat method is a rigid system which prevents all movement out of the duct from
entering into the manhole.
EPRI Licensed Material

Design of Constraints in Manholes
12-19
In addition to constraining the cable externally it is necessary to check that the internal friction
between the extruded core and the inner surfaces of the sheath is capable of withstanding
differential force and preventing movement inside the joint.
A key method of reducing the cable axial force is identified in this project. The clearance
between the cable and duct is deliberately increased to permit thermomechanical waves to grow
to larger amplitude and absorb a higher proportion of the locked-in thermal strain. Clearances of
50-80mm can usefully reduce the axial thrust to 50% of the rigid bar level. Care has to be taken
to ensure that the cyclic fatigue strain for the metallic sheath or foil is not exceeded. The method
used to calculate force and strain is given in Chapter 5 for the FEA modeling technique and
Chapter 8 for the NS techniques.
Duct System: Expansion Cleat System
Figure 12-20 is the plan view of a 230kV 2500kcmil manhole design employing the expansion
cleat system. The expansion cleat is located on the left hand side of the manhole. In this
particular case the manhole is at the bottom of an incline. The spring cleat allows the cable to
expand a calculated distance into the manhole, thereby reducing the axial thrust.
The cable movement is required to be absorbed within an s-shaped offset, the geometry of which
has to be designed to limit the cyclic strain within the fatigue life of the metallic sheath. The
shape of the offset itself provides a small resisting force which, in addition to the expansion cleat
springs, opposes the cable thrust. The cable has to be free to flex and thus cannot be cleated. The
expansion cleat is sometimes called a ratchet cleat, however this is a misnomer as the cleat
movement has to be cyclic such that the cable is always returned to its starting position upon
cooling.
The cable at the exit to the manhole which may be into a horizontal or downhill span may be
constrained by a second spring-loaded expansion cleat or as in the drawing with a cable stocking.
The joint in Figure 12-20 is of the anchor type in which the cable conductor is rigidly anchored
via an epoxy resin insulator to the manhole. Thus the thermomechanical performance of the two
cable spans on either side of the joint are completely segregated. The anchor joint is required to
withstand the unbalanced differential load.
EPRI Licensed Material

Design of Constraints in Manholes
12-20

Figure 12-20
230kV duct manhole with expansion cleat, cable stocking and anchor joint constraints
Figure 12-21 is a diagram of the expansion cleat which operates in the following manner. A
rubber lined cable cleat is clamped around the cable using spring loaded clamp bolts. The rubber
liner and sprung loaded bolts allow for cable radial thermal expansion. The length of the cleat is
chosen to give a pre-determined clamp holding strength. The cleat is attached to a plate able to
move along the longitudinal axis of the cable. This plate is connected to a fixed anchor position,
for example the wall of the manhole, via a number of helical compression springs of length L.
When the cable moves into the manhole due to thermal expansion, the springs are compressed by
a length dL. The spring force has to match the cable thrust. The characteristics of the springs and
their number combine to give a set force/distance relationship. The holding strength must not
exceeded before the maximum movement of the springs is reached.
Model of Thermal Expansion Cleat Constraining System
The manhole design shown in Figure 12-20 was modeled together with the adjacent cable span
and with a nominal cleat holding strength of 15,000N and an equivalent expansion spring
movement of 263mm. Figure 12-22 shows a plan view of the deflection of the cable at 90
0
C at
the expansion cleat end of the manhole. The cable has moved a) into the manhole by 151mm and
b) outwards into the S-bend by 198mm.
Figure 12-23 shows that the cable movement at the opposite end of the manhole is also inwards
with a cable movement of a) 150mm longitudinally and b) 216mm laterally into the S-bend.

EPRI Licensed Material

Design of Constraints in Manholes
12-21

Figure 12-21
Diagram showing operation of expansion cleat

Figure 12-22
Plan view of deflection of cable at expansion cleat offset, on left-hand end of manhole
EPRI Licensed Material

Design of Constraints in Manholes
12-22

Figure 12-23
Plan view of deflection of cable at stocking offset, on right-hand end of manhole
In the design of the expansion cleat system it is necessary to calculate the equilibrium diagram
for both a) the cable thrust and b) the expansion cleat reaction force characteristics. An example
is shown in Figure 12-24. Note that the cable in this model was given an early and lower value of
axial stiffness EA and in consequence the force extension curve has forces at 50% of the
expected value.
In the diagram the force extension characteristic of the cable is given in the line ac. If the cable
is fully constrained, the movement is zero and the force is greater than 30KN a. When the cable
is allowed to expand into the manhole the force reduces until zero force and maximum
movement (280mm) occur at position c. The constraint to movement is a) the expansion cleat
with characteristic def and b) the cable stiffness in the S-bend. Equilibrium occurs when the
cable force and constraint force characteristics cross at position b. In this model, equilibrium
occurred at a thrust of 14kN and a movement into the manhole of 170mm. It is important to
check that the curve does not reach point e, at which the force exceeds the cleat holding
strength of 15,000N and the cable slides through the cleat.
In the two examples shown in Figure 12-22 and Figure 12-24 the cable moved 151mm and
170mm into the manhole respectively, with the movement being absorbed by the deflection of
the cable within the S-bend. This movement is cyclic and imposes a significant strain in the
metallic sheath. It is important to calculate the level of strain to ensure that this does not exceed
the limiting values given in Chapter 4.
EPRI Licensed Material

Design of Constraints in Manholes
12-23

Figure 12-24
Expansion cleat equilibrium diagram

Figure 12-25
Chart of force at entry to the manholes at 90
o
C in the first and fifth load cycles
EPRI Licensed Material

Design of Constraints in Manholes
12-24
Figure 12-25 shows the forces at the entry to both sides of the manholes in the model. The
maximum force is 6,900N, which shows how effective the expansion cleat design is in reducing
the thermomechanical force by allowing expansion. The force is reduced to 10% of the 69,000N
maximum value given in Chapter 8 for the hockey stick route. The expansion cleat system has
allowed 300mm of the locked-in thermal strain of 540mm to be relieved. The balance of the
cable force is held within the duct by longitudinal friction constraint.
It is indicated that the reaction force at the joint is less than 5,000N. Thus if a clamp cleat is
placed adjacent to the joint, then it becomes possible to select a straight joint in place of an
anchor joint.
Figure 12-25 indicates that there is minimal change in force due to the action of five loading
cycles. A change occurs at the cable stocking end of JB2, when the force falls from 3,700N to
1,000N, indicating that the cable has moved downhill.
Duct System: Clamp Cleat Constraint
The design shown in Figure 12-26 is closely similar to that described earlier for the pipe system
offset constraint.
The manhole has the same dimensions as for the expansion cleat design. Two large clamp cleats
are mounted at the inlet and exit to the manhole. A number of smaller clamp cleats are located to
rigidly constrain the cable around a 12D, 60
0
offset. In this example 14 small clamp cleats are
employed together with two large clamp cleats. If the clamp cleat assembly design is capable of
fully withstanding the differential thrust across the manhole then it may not be necessary to
select an anchor joint. In this case a conventional straight joint may be employed, laterally
constrained by guide cleats. The guide cleats allow the joint to move longitudinally and thereby
prevent differential movement occurring between the joint and cable insulation. As with the pipe
system constraint, the allowance for movement is 8mm, comprising a maximum elastic
movement of 4mm for the cleat liner, with a design margin of 2.
EPRI Licensed Material

Design of Constraints in Manholes
12-25

Figure 12-26
Manhole layout showing clamp cleats, 60
0
, 12D radius offset and joint guide cleat
The differential withstand capability is calculated in the same way as in the pipe system. For
example an opposing force is taken at the outlet of the manhole equal to either the opposing
cable thrust or the cleat withstand force. Figure 12-27 shows the force diagram.

Figure 12-27
Diagram showing input, output and differential forces across clamped manhole
Figure 12-28 shows that for an opposing force F
out
of 2700N the differential that can be
withstood across 14 cleats of the 6,500/10,833N 100mm type is 10,538N. These cleats are
200mm long, twice the length of the 100mm cleat tested at EPRI solutions on the 230kV cable.
The peak differential force between locations A and B in the hockey stick model of the 230kV
2500kcmil cable used in Chapter 8 occurred at 35
0
C after a temperature rise of 20
0
C. The peak
EPRI Licensed Material

Design of Constraints in Manholes
12-26
differential force was 20,000N. This is twice the withstand capability of the clamped offset
model in Figure 12-26.
Three design options are possible to balance the forces:
To increase the clearance between the duct and the cable, thereby reducing the force
generated. Chapter 8 shows that increasing the clearance from 38mm to 80mm, will reduce
the peak thrust from 80% to 55% of the rigid-bar value.
To select an anchor joint to withstand the differential force, this being 10,000N. Anchor joint
designs should be capable of withstanding forces up to 60,000N in both compression and
tension.
To install a standard straight joint with high strength clamp cleats on either side, each capable
of withstanding 50% of the force. These cleats would supplement the existing clamp cleats
along the offset.

Figure 12-28
Input/output force characteristic across 14 cleats in the 60
0
, 15D clamped offset model,
with cleat withstand strengths of 6500N
Internal Constraint Within the Cable
It is possible for the extruded cable core to slide differentially within the cable sheath, even
though the outer surface of the cable is held by clamp cleats. Clamp cleats are not designed to
crush the jacket and metallic layers into the extruded core. This would damage both the outer
layers, the shields and the insulation.
EPRI Licensed Material

Design of Constraints in Manholes
12-27
Tests were performed at EPRIsolutions on the 138kV 1500kcmil and on the 230kV 2500kcmil
cables to measure the pull through strength of the core with the outer surface cleated in the shape
of a ninety-degree bend. A back tension was applied to ensure that the core contacted the sheath
so that the frictional constraint could be measured.
The equation for the frictional constraint is:
Fin/Fout = exp ( ) Equation 12-2
Where
Fin: applied tensile force [N]
Fout: back tension [N]
: coefficient of friction
: bend angle [radians]
The exponent was calculated to yield the effective values of friction/stiction, these being = 1.6
for the 138kV 1500kcmil cable and = 2.1 for the 230kV 2500kcmil cable.
The strength of the frictional constraint within the cable inside the 60
0
offset designs was
calculated for both the 138kV and 230kV pipe and duct systems. The offset has a total bend
angle of 240
0
. Taking a factor of safety of 2 on the value of friction gives a design value of 0.6.
For an outlet cable compressive force of 1000N the inlet force is calculated to be 28,500N giving
a differential withstand force of 27,500N thus indicating that differential movement between the
extruded core and the sheath will not occur in the offset designs.
Conclusions
Methods of constraining joints in both pipe and duct systems have been analysed and
compared to the thermomechanical forces and movements that can be generated from the
adjacent cable spans.
Improved methods of constraint have been identified and modeled and their performances
analysed to provide design data.
Two generic designs of constraint with improved withstand performances have been
produced for a 138kV pipe system. One comprising an in-line design and the other an offset
design.
Two design methods have been analysed for duct systems, one is a semi-flexible system
employing an expansion cleat and the other is a fully constrained system employing clamp
cleats and a cable offset.
The facility to reduce the cable axial thrust by increasing the clearance between the cable and
the duct/pipe was identified and quantified in Chapters 7 and 8. This is a key design aid in
permitting the constraining systems to safely protect the cables and joints within a manhole
whilst providing the flexibility necessary to design manholes with practical dimensions.
EPRI Licensed Material

Design of Constraints in Manholes
12-28

i
Development of 275kV XLPE cable with aluminium laminated tape and radial moisture barrier. Umeda,
Watanabe, Sakaguchi, Ohimo. Jicable 2003, Paper A.1.5

ii
Installation of a 220kV XLPE cable in a deep shaft for a hydroelectric power plant. Marini, Bisleri, Zaccone.
Jicable 2003, Paper A.4.6.
EPRI Licensed Material
13-1
13
APPLICATION GUIDE FOR THE DESIGN OF DUCT-
MANHOLE SYSTEMS AND PIPE SYSTEMS
CONTAINING UNDERGROUND HV XLPE CABLES
This guide will provide users with tools to design duct and pipe systems to withstand the
thermomechanical forces and movements that occur in service operation during the design life of
the cable circuit, by:
Selection of the optimum pipe clearance to:
Reduce the axial thrust a) at manholes and b) at bends
Avoid limiting the fatigue life of the metallic sheath or foil laminate.
Selection of manhole positions to protect the joints against differential force and longitudinal
movement
Selection of the type of cable best suited to duct and pipe installations
Selection of the type of joint to suit particular locations:
Straight joint
Anchor joint
Design of manhole constraint system to protect the joint and cable and to withstand:
Cable axial compressive force.
Cable differential force.
The guide describes the steps to be taken in the thermomechanical design of:
Duct systems containing one cable
Pipe systems containing three cables
Cable types:
System voltages 69kV to 345kV
Extruded XLPE insulation
System definition:
The system comprises pipes, ducts, cables and joints
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-2
The guide does not cover:
Constructional designs of duct blocks, manholes, or supports for cable terminations
Methods of cable installation
Methods of jointing
The guide is structured in a manner to provide the user with important information. The guide
proceeds in the following steps:
Explanation of thermomechanical behavior
Obtain cable and pipe geometries
Obtain limiting parameters for the cable
Calculate pipe/duct clearance
Calculate cable stiffness parameters
Calculate axial cable force
Obtain limiting parameters for the joints and cleats
Design manhole constraining system
Calculate withstand strength of manhole constraint
Compare calculated cable thrusts and manhole withstand strength
Calculate and check cable limiting parameters
Position manholes in outline route
Model route in detail and check cable and joint limiting parameters
Route Geometry and Thermomechanical Effects
Duct and pipe routes are installed in down-town, suburban and rural areas with routes dictated by
the road system. In consequence cable routes comprise a multiplicity of horizontal and vertical
bends.
An example of two typical pipe spans of a 138kV XLPE pipe system that were
thermomechanically analysed by three dimensional computer imaging is given in Figure 13-1.
The first span is 242m long and second is 520m long. The plan view is undistorted in length and
width (drawn to a 1:1 scale). The vertical depth scale is increased by 9:1.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-3

Figure 13-1
Two spans in a 138kV pipe system showing horizontal and vertical bends
The number and location of the bends is shown. It can be seen that a) the horizontal bend angles
are much larger than the vertical ones and b) there are twice as many changes in vertical
elevation as in the horizontal direction. The spans contains:
22 bends in total
11 bends in both horizontal and vertical direction
11 bends in the vertical direction only
The longest straight section is 48m, but is on an incline, it being 6% of the total 762m length.
In pipe and duct systems bends and inclines are the norm and long straight sections are rare.

Figure 13-2
Distribution of wave patterns in two spans of a 138kV pipe system at 90
o
C
Plan view, width to length scale of 10:1
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-4
The thermomechanical analysis of the two spans and of other spans in the same route confirm
that the majority of the length contains wave patterns at the cable operating temperature of 90
o
C,
Figure 13-2.
The three cables form regular intertwining waves. The amplitude of the waves is restricted to the
132mm clearance between one cable and the inside diameter of the pipe and would be too small
to be seen in a 1:1 plan view. To aid analysis, the images of two of the cables have been removed
from Figure 13-2 and the lateral scale has been increased by 10:1.
Figure 13-2 shows that the patterns cover nearly 90% of the span length and are linked with the
positions of the bends. The reasons are explained later in the guide.
Overall, the normal features of a pipe or duct system are:
Bends and inclines dominate the span geometry
Thermomechanical wave patterns dominate the length at medium to high operating
temperatures
Thermomechanical wave patterns and the accompany forces and movements are unique to pipe
and duct systems and, when design advantages are taken, are beneficial to their reliable
operation.
For a given XLPE cable construction there are three key design choices to be made:
The optimum thermomechanical clearance between the cable and pipe/duct as this controls
the magnitude of cable axial force
The location of the manholes, as this reduces the levels of differential force, movement and
axial thrust acting on the manhole
The design of the constraint system within the manhole, as this protects the cable and joints
The measurement of cable thermomechanical clearance is shown in Figure 13-3 for a 230kV
2,500kcmil duct system and in Figure 13-4 for a 138kV 1500kcmil pipe system.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-5

Figure 13-3
Thermomechanical clearance of 25mm in an example 230kV XLPE 2500kcmil duct system

Figure 13-4
Example of thermomechanical clearance of 117mm in an example 138kV 1500kcmil pipe
system
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-6
The possible reduction in axial force that can be obtained by selecting larger clearances is given
in Figure 13-5for a 138kV 1500kcmil XLPE cable in a duct. The curves for positions A and B
are the end of the straight pipe section and the entrance to the bend in the 200m normalized
span shown in Figure 13-7.
Based on previous engineering practice, clearances down to 30mm (20% of the reference 150mm
clearance) have been selected for duct systems; it is shown that a high axial force of 75% of
maximum value results.
With the assistance of the design method in this Guide, it is possible to select a clearance of
60mm or greater (40% clearance on the scale), to obtain a reduction in the axial force to 30% of
its maximum value.

Figure 13-5
Effect of varying duct to cable clearance for the 138kV reference cable
Note: The recommended value of axial stiffness shown in Figure 13-8 increases the rigid
bar force above that shown, by 28% to 47,200N at 90
o
C
The equivalent graph for the 138kV XLPE cable in a pipe system is shown in Figure 13-6. It is
not possible to install these 83mm diameter cables with a clearance of less than 95mm (63.6% of
150mm), as they would be an interference fit in the pipe. The tightest fit occurs when the
diameter of the exscribed circle around the trefoil group of three cables is 1.15D (where D is the
cable diameter). For installation reasons it is usual to install cables with a clearance of 85% or
higher (130mm). Figure 13-6 shows that this results in a significant reduction in axial force to
less than 30% of the maximum value.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-7

Figure 13-6
Effect of varying pipe to cable clearance for the 138kV reference cable
Note: The recommended value of axial stiffness shown in Figure 13-8 increases the rigid
bar force above that shown, by 28% to 47,200N at 90
o
C
Geometry and installation practices give the pipe system the inherent advantage of low axial
force compared to a duct system for the same cable size.
One disadvantage of the pipe system is that the level of fatigue strain in the metallic sheath, or
foil laminate, is higher.
The following is a step by step guide to pipe and duct system design, to take account of
gravitational and thermomechanical effects.
Route Geometry Modelling by FEA (Finite Element Analysis) Technique
Analysing the thermomechanical behavior of a route comprising the geometry illustrated in
Figure 13-1 is complex. The method used is the explicit FEA dynamic modelling technique.
The FEA method is described in detail in Chapter 5 of the EPRI project report.
The FEA technique is the recommended calculation method for the detailed design of new duct
and pipe routes. The technique is proven and has been successfully used to analyse in detail the
performance of multiple spans of pipe and duct installations.
The diagrams and explanations given in this design guide were obtained by use of the FEA
technique.
The FEA technique provided the sets of data essential to the Normalized Span calculation
method, described below.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-8
Route Analysis by the Normalized Span Technique
Thermomechanical performance is best understood by comparison with the analysis of a
simplified reference span containing one straight section and one bend, shown diagrammatically
in Figure 13-7.

Figure 13-7
Reference span
Not to scale
The reference span is normalized on:
Span geometry. The reference span is horizontal, comprising a 200m straight section and a
35D radius 90
o
pipe bend (D is the cable diameter),.
Cable to pipe/duct clearance. The reference is 150mm and is common to both duct and pipe
systems.
Cable properties. The reference cable construction is a 138kV, 1500kcmil stranded circular
copper conductor, triple extruded XLPE insulation and shields, double copper tape shield*,
aluminum foil laminate moisture barrier and polyethylene anti-corrosion jacket, as shown in
Figure 13-4. *This shield design is now superseded by a wire shield design.
There are 12 parameters necessary to be derived to represent the geometry of the span and the
characteristics of the cable. The method of the calculation of the cable stiffness characteristics
(axial stiffness, bending stiffness and torsional stiffness) is given in Chapter 7 of the EPRI
project report.
In Chapter 7 the performance of the normalized model is analyzed with each parameter varied to
cover the range of geometries, cable types and cable characteristics likely to be encountered in
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-9
North American applications. This process was undertaken separately for duct systems
containing one cable and pipe systems containing three cables.
The method to calculate the maximum axial force is as follows, an example calculation is given
in Chapter 8 of the EPRI project report:
The 12 parameters listed in Table 13-1 and Table 13-2 are calculated for the particular span
geometry and cable dimensions
The 12 parameter ratios are calculated by dividing each parameter by the equivalent
reference values given in Table 13-1 and Table 13-2. The ratio is expressed as a percentage.
The graphical data from the Sensitivity Study, recorded in Chapter 7, is accessed and the
value of force is read off. This is divided by the 100% reference value given on the graph to
obtain a Force Ratio. The Force Ratio is expressed as per unit value. Up to 12 Force
Ratios are required.
The Force Factor, F*, is calculated for positions A and B, as given in Equation 13-1 and
Equation 13-2, by multiplying together each of the Force Ratios listed in Table 13-1 and
Table 13-2.
The reference axial force F*
R
is obtained from the graphical data in Chapter 7. The values
are copied below in Table 13-3.
The magnitude of the forces, F
A
and F
B,
are calculated by multiplying together the Force
Factor F* and the reference axial force F*
R
.
F*
A
=( G
D.
x........x G

)
A
x (C
W
. x..x. C
CC
)
A
F
A
= F*
A
x F*
RA
Equation 13-1
F*
B
=( G
D.
x........x G

)
B
x (C
W
. x..x. C
CC
)
B
F
B
= F*
B
x F*
RB
Equation 13-2
The same method is used to calculate each of the following limiting parameters, from the
appropriate reference graphs given in Chapter 7.
F*
A
Axial force in the cable
S*
A
Sheath strain (absolute) in the cables metallic sheath
F*
A
Sidewall cable force in the thermomechanical patterns and in the preformed bends
R*
A
Bending radius of the cable along span (ie in the pattern)
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-10
Table 13-1
Geometric parameter ratios to calculate axial force for both duct and pipe systems
Force Factors
[per unit]
Geometric Parameter Ratios:
Actual Case/Reference Case
[%]
Duct
Positions
Pipe
Positions
Symbol Description
Reference
Span Value
Symbol
A B A B
G
D
Clearance between individual cable
and pipe diameters, (D
P
-D
c)
150mm F
D

G
L
Length of straight section of span 200m F
L

G
R
Radius of pipe bend 2.905m (35D) F
R

G

Angle of pipe bend 90
o
F


G

Angle of slope 0
o
F

Table 13-2
Cable parameter ratios to calculate axial force for duct and pipe systems
Force Ratios
[per unit]
Cable Parameter Ratios:
Actual Case/Reference Case [%]
Duct
Positions
Pipe
Positions
Symbol Description
Reference
Cable Value
Symbol
A B A B
C
W
Weight per meter

11.9kgm
-1
F
W

C
EA
Total axial stiffness EA (1) 19.25MN F
EA
(3) (3) (3) (3)
C
EI
Total bending stiffness EI (1) 1,910Nm
2
F
EI

C
GJ
Torsional stiffness GJ (1) (2) 1,470Nm
2
F
GJ

C

Total coefficient of thermal
expansion (linear)
17 x 10
-6 o
C
-1
F

(4) (4) (4) (4)
C
CP
Coefficient of friction, cable to
pipe
0.1 F
CP
(5) (5) (5) (5)
C
CC
Coefficient of friction, cable to
cable
0.1 F
CC
- -
Table 13-2 Notes:
(1): Detailed temperature dependent properties are required for FEA modelling. For the
Normalized Span calculation method take the mean of the values at 20
o
C and 90
o
C
(2): Variations of magnitude had negligible effect on the force
(3): Calculations are to be based on latest test values for 138kV cable,
for NS method use: EA = 41.7MN
(4): Recommended value: 18.5 x 10
-6

o
C
-1

(5): Recommended Value: 0.3
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-11
Table 13-3
Reference axial forces
Cable System Position in Normalized Span F*
R
Reference Force [N]
A, Span End 7,520 Duct
B, Bend Entrance 5,373
A, Span End 10,458 Pipe
(Cable 3, bottom right) B, bend Entrance 7,322
Gravitational Sliding Effects During Installation
The risk of cable sliding should be assessed when selecting the route.
Cable will not slide downhill either during or after installation, unless the angle of inclination to
the horizontal is greater than tan
-1
, where is the effective coefficient of friction between a) the
cable and the pipe and b) between cables.
Experience with XLPE cables has shown that the use of smooth polyethylene jackets in
combination with installation lubricants can reduce the from 0.3 to the low value of 0.1. This
increases the risk of slip, as the critical gradient of slip is reduced to 1 in 10, equivalent to an
angle of less than 5.7
o
.
The angle of slip is further reduced by a) increases in ambient temperature and b) traffic induced
vibration.
An increase in ambient temperature generates a thermomechanical thrust even in an unjointed
cable. (The explanation of generation of thrust in service operation is given in the following
sections). The general effect of thermomechanical force on a slope is that the cable will a) move
downhill at less than the angle of slip and b) uphill at greater than the angle of slip.
Vibration momentarily lifts the affected length of cable off the duct floor. This removes the
frictional constraint preventing movement, whilst adding the downhill component of weight to
the adjacent cable. For example, if sufficient length of cable is lifted, such that it applies a force
equal to 5% of the weight of the unaffected cable, then the angle of slip is reduced from 5.7
o
to
2.9
o
, a gradient of 1 in 20. Such gradients are likely to be present in pipe and duct routes. The
combination of vibration and thermomechanical effects is studied in detail in Chapter 9 of the
EPRI project report.
For safety, it is recommended that a temporary restraining system such as cleats be applied to
prevent cable movement after pulling-in, during jointing and when fixing cable cleats in
manholes.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-12
Thermal Expansion and Movement
When current flows, the conductor is heated and the cable temperature rises. The cable will
attempt to increase its length by thermal expansion. The increase in length L for a friction free
pipe is:
L L = Equation 13-3
Where:
L: length of pipe span [m]
L: free expansion length of cable [m]
: coefficient of thermal expansion of cable [
o
C
-1
]
: temperature rise [
o
C]
The typical coefficient of expansion for a 138kV 1,500kcmil XLPE cable is =18.5x10
-6

o
C
-1
.
This value is recommended for XLPE cables in the system voltage range of 69-345kV, with
stranded copper conductors up to 3,000kcmil.
When cable in a typical span length of 500m is heated from an ambient temperature of 15
o
C to
its operating temperature of 90
o
C, in a friction free pipe, the free expansion L is 694mm.
The consequences of thermal expansion to the cable system are:
Cable unconstrained at ends with zero pipe friction. Maximum movement occurs at ends with
zero axial force resulting.
Cable unconstrained at ends with pipe friction present. Limited movement occurs at ends,
with zero force resulting at ends. At the center of the span locked-in strain and axial force is
present.
Cable constrained longitudinally and laterally. Maximum locked-in strain and maximum
axial force are present along the length.
Cable constrained longitudinally and free to move laterally. Some locked-in strain results,
wave patterns form, internal movement occurs into the wave patterns and axial force is
reduced.
Thermomechanical Axial Force: Rigid Bar Behavior
The cable cannot expand lengthwise because it is anchored by the cable terminations at the ends
of the route.
The length of expansion L is said to be locked-in. The locked-in expansion is called thermal
strain and is expressed as the ratio
L
L
, thus 694mm of locked-in expansion over a 500m length
is a strain of 0.0014 (1.4mm per meter).
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-13
The locked-in thermal strain produces an internal thermomechanical force, which a) places the
cable in axial compression and b) applies a thrust F to the end fixings. If the free expansion
length L is pushed back level with the pipe, the compressive force required to do this is equal to
the thermomechanical force, F.
EA F = Equation 13-4
Where:
F: thermomechanical force exerted by a fully constrained rigid-bar cable [N]
EA: axial stiffness of the particular cable construction [N]
For an XLPE cable the axial stiffness EA is:
Primarily determined by the conductor axial stiffness and its cross-sectional area
Temperature dependent, being high at ambient temperature and low at operating temperature
For a 138kV 1500kcmil cable the values of EA are:
15
o
C: 50.5 x 10
6
N (50.5MN)
90
o
C: 34 x 10
6
N (34MN)
Equation 13-4 and Figure 13-8 show that the force is dependent on temperature. The force
calculated by this equation is called the rigid-bar force for the cable in its fully constrained
case.
At 90
o
C the thermomechanical force exerted by the 138kV 1500kcmil cable internally along its
length and at each cable end is 47,200N (4.7tonnes).
Magnitudes of rigid bar force for other cable sizes are given in Figure 13-8. In practice the
thermomechanical force is prevented from achieving the full rigid bar value. The key reason,
forming the foundation of this guide, is the ability of pipe and duct systems to reduce the
magnitude of thermomechanical force (and other limiting parameters) by the absorption of
thermal strain.
In this respect, pipe and duct circuits are unique. When designed with this knowledge, they have
a distinct advantage over other types of cable system.

EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-14

Figure 13-8
Force-temperature characteristic for XLPE cables constrained as rigid bars
Key Design Parameter: Clearance Between Cable and Pipe/Duct
The radial clearance between one individual cable and the inner surface of the pipe/duct directly
determines the length of expansion strain that can be absorbed and thence the amount the axial
force can be reduced. For convenience the clearance is expressed as the difference D between
the outer diameter of the cable and the inner diameter of the pipe/duct as shown in Figure 13-3
and Figure 13-4

c p
D D D = Equation 13-5
Where:
D: cable to pipe clearance [m]
Dp: inner diameter of pipe/duct [m]
Dc: outer diameter of a single cable [m]
The clearance determines the amount of cable strain that can be absorbed by a) pipe/duct bends
and b) thermomechanical wave patterns within the pipe/duct.
The reduction in axial force obtainable by increasing the clearance for duct systems is given in
Figure 13-5 and for pipe systems in Figure 13-6.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-15
Absorption of Thermal Strain and Locking Within Bends
Absorption of Strain in Bends
When the temperature first starts to rise, the cable initially easily moves outwards into the bend.
This is because the arc length around the outside of the bend is longer than the axial length. The
difference in length b is proportional to the angle of the bend. If the radial clearance is 75mm
and the bend angle is 90
o
, the bend will store 59mm of additional cable length. Thus a 90
o
bend
will store 6 times more strain than a 15
o
bend. This being 8% of the 694mm length of locked-in
strain in a 500m span, or 100% of the strain in a 40m length.
If 10 bends are present in the 500m span, each with an average angle of 30
0
, the cumulative
absorption would be 196mm, 28% of the total locked-in strain. The axial force would be reduced
by the same percentage in a friction free pipe.
Note that the cable absorption is independent of the bend radius of the pipe/duct.
360
. D .
b

=

Equation 13-6
Where:
b: cable absorption length of bend [m]
: angle of bend [degrees]
Locking Effect of Bends
The presence of friction between the cable and pipe causes the cable to become locked within
some bends, by a combination of the high sidewall reaction force and the frictional constraint.
Equation 13-7 gives the relationship.
180
exp F F
out in

= Equation 13-7
Where:
F
in
: Axial cable thrust into bend [N]
F
out
: Opposing axial thrust inwards from the other end [N]
: Coefficient of friction [per unit]
: Angle of bend [degrees]
The effect of the bend angle on the magnitude of the inwards axial force that can be withstood
without slipping is given in Table 13-4 for a coefficient of friction of 0.3 and a nominal opposing
force of 1,000N.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-16
Table 13-4
Locking effect of bend angle in withstanding differences in axial force
, Angle of bend
[degrees]
F
in,
Cable Axial Force
[N]
(F
in
- F
out
) [N] Difference in Locked-in
Thermal Strain [%]
5 1,026 26 3
15 1,082 82 8
30 1,170 170 17
45 1,266 266 27
90 1,602 602 60
The heated cable expands and moves into each bend. The movement is restrained by friction in
the span:
When the adjacent straight sections are longer than the frictional constraining length, then the
opposing cables will move equally into the bend, the forces are identical and so no movement
can occur. Force will rise on each side of the bend and wave patterns will form on each side.
When the span lengths are unequal and less than the frictional constraint length, the thermal
strain in one span is apportioned between the bends at each end, according to the comparative
bend angles. The levels of strain in the sections on either side of any bend in a real route are
likely to be different.
Thermal expansion will first fill the storage space in small angle bends. The axial rigid-bar
force will then rise. When the locked-in thermal strain is imbalanced between two spans by
more than 10%, small bend angles up to 15
o
will allow the cable to slide round (the
difference in force that can be held is less than 8%)
Medium angle bends of larger than 30
o
can separate spans with imbalances in strain of 17%,
this being a significant percentage of the locked-in expansion strain. For imbalances in strain
of less than 17% the bend will lock the cable from movement. It is then possible for
separate wave patterns to be formed on either side of the bend.
Large angle bends of 90
o
and above can lock cables with large imbalances in strain (and
force) of up to 60%. Note, that for an S-bend, the angles are additive, so that an equivalent
locking effect to 90
o
is obtained with two 45
o
bends close together.
Thermomechanical Patterns: Absorption of Strain, Turn-over Temperature
and Turn-Over Force
The heated cable within the pipe or duct span is a strut under compressive load, typically 500m
long. However, it is not possible for the cable and pipe, as supplied and installed, to be perfectly
straight and cylindrical. The double span route in Figure 13-1 of 760m length is divided by 22
bends with the longest straight pipe section being 48m. Deviations in the route geometry and in
cable shape generate the bending moments that cause the cable to deflect laterally.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-17
The formation of the lateral wave pattern occurs at moderate temperature rises in the range of
25
o
C-55
o
C. For an ambient temperature of 15
o
C, these are equivalent to operating temperatures
of 40
o
C-70
o
C. Below the temperature of wave formation the cable behaves as a rigid rod
generating high mid range forces, which can be read off Figure 13-8. Above that temperature the
force will either a) fall (turn-over), or b) continue to rise at a reduced rate.
Most cable circuits will experience the maximum axial thrust at the turn-over temperature early
in their service life, even if they are not operated at their full rated load current.
All of the 12 design parameters representing the cable properties and pipe/duct geometry play a
role in the mechanism of what happens before, during and after wave formation.
The magnitude of the maximum axial force can be significantly reduced by a) the careful design
of the pipe/duct geometry and b) by selection of the cable construction.
The following topics are covered:
13.9.1 Wave shapes
13.9.2 Inception and longitudinal growth of waves
13.9.3 Limiting effects of pipe friction on strain absorption
13.9.4 Turn-Over force due to transition to large amplitude waves
13.9.5 Turn-Over force due to frictional slip
Wave Shapes
A single cable in a duct can form both cylindrical sinusoidal waves and helical loops.
Cylindrical sinusoids (shortened to sine) and helices are present in the pattern in Figure 13-9.
This is a duct system at 90
o
C containing 138kV 1500kcmil cable with a large clearance D of
150mm. To enhance the wave shapes the view is deliberately foreshortened by the low viewing
angle perspective. Two helical waves are present at the entrance to the bend.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-18

Figure 13-9
Helical and sine wave patterns at 90
o
C in a 138kV duct with 150mm clearance
Foreshortened view, scale 1:1
The wave patterns in a pipe system at 90
o
C are shown in plan view in Figure 13-10 and at 125
o
C
in foreshortened view in Figure 13-11. Pipe systems can only form sine waves, because the three
cables compete for space. This imposes a regularity and uniformity that permits sine waves, but
excludes helical loops.

Figure 13-10
Waves at 90
o
C in a 138kV 1500kcmil pipe system with 150mm clearance
Plan view, scale 1:1

EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-19

Figure 13-11
Waves at 125
o
C in a 138kV 1500kcmil pipe system with 150mm clearance
Foreshortened view, scale 1:1
The wave lengths of the patterns are reasonably uniform along a pipe or duct. At the formation of
the incipient wave, the wavelength length is characteristic of the particular cable and pipe
parameters, however the wavelength shortens as the temperature rises and the axial thrust
increases.
When thermomechanical waves are viewed in an undistorted plan view Figure 13-10, for
example in a full size experiment, it would be a) almost impossible to see the waves and b) to
distinguish helical from large amplitude sinusoidal patterns. The aspect ratio of apparent
amplitude ( 0.5D) to wavelength is very small, being typically only 1-2%. The wavelengths of
both sine and helical waves are in the range of 4-7 meters.
An advantage of computer imaging in the post processing of FEA analysis for design work, is
that it is possible to distinguish wave patterns by:
Rotating the model of the route to give a foreshortened view, but keeping the 1:1 scale, as
shown in Figure 13-9 and Figure 13-11
Increasing the lateral scale of the image, for example by 10:1 to magnify the amplitude, as
shown in Figure 13-12 and Figure 13-16.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-20
Removing the images of the two adjacent cables in a pipe system and increasing the lateral
scale as shown in Figure 13-19.
Cylindrical Sinusoids
The most common wave shape is a cylindrical sinusoid. The amplitude of the wave is the
alternating circular distance it slides around the circumference.
When the wave alternates through an angular amplitude of 90
o
(e.g. between the 3o/c and 9o/c
axis) the amplitude is 0.25D, where D is the pipe to cable clearance (D
p
-D
c
). From a plan
view of the pipe (e.g. from a plan view of the computer image) the amplitude appears to be
smaller at 0.5D.
The wave can obtain a greater amplitude by simply sliding further around than the pipe than 3o/c
to 9o/c axis. Movements through 180
o
are common, i.e. an amplitude of 0.5D.
Some sine waves alternate through one complete revolution ie 360
o
with an amplitude of
D. The plan view of the wave is then confusing as the sine wave appears to be two helical
loops, each with half the wave length. This is easily verified by drawing a sine wave on a
transparent plastic sheet and rolling it up to form a cylinder.
The presence of a number of helical waves is a normal feature of all duct systems. The minimum
energy condition that results in the transition to large amplitude sine waves also applies to
whether a sine or a helical wave is formed. The factors that influence the relative percentage of
helical waves to sine waves are given in the Sensitivity Study in Chapter 7 of the EPRI project
report, in Figures 7-49 and 7-50. If the duct is tight fitting or the cable has a low weight, then the
waves may be 100% helical. The effect of increasing the cable temperature is to produce a higher
percentage of helical waves.
Helical waves usually form groups rotating in opposite directions such that the cable system
remains torsionally neutral. However, the torsional stiffness of the cable is the least sensitive
parameter to the magnitude of axial force, compared to the other 11 parameters.
Helical waves absorb significantly more thermal strain than sine waves and have a greater effect
in reducing the axial force.
A disadvantage of isolated helices of the type shown in Figure 13-9, is that cyclic fatigue strain
in the metallic sheath is also concentrated. In design, it is important to check that the limiting
value of strain is within acceptable limits throughout the route and particularly in isolated
helices. In routes with a high percentage of helices the strain distribution is uniform.
The four important mechanisms associated with wave formation are:
Wave inception. The pattern comprises a small number of half waves of small transverse
amplitude. The waves grow progressively in length as the temperature rises.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-21
The effect of pipe friction, in limiting the release of locked-in strain.
Wave transition. A dynamic event in which the sinusoidal waves suddenly increase in
amplitude and the force falls, (turns-over). In duct systems helical waves may also form.
The larger waves continue to grow in length as temperature rises.
Force reduction (turn-over), caused by frictional slip at the remote end of the span.
Inception and Longitudinal Growth of Waves
Figure 13-12 shows the shape of the half wave just after its inception. The lateral scale is
amplified by 10:1 as the wave is too small to be seen in 1:1 scale. The wave has been initiated by
the curvature within a very small 5
o
pipe bend, following a rise in temperature of 22.5
o
C. This is
a model of a cable in a duct with a clearance of 150mm. At the right hand side, at B, the cable
thrust can be seen to have pushed the cable to the outside of the bend creating a cranked shape.
The maximum distance it can move is half the clearance, ie 75mm. This displacement creates the
bending moment that forms the first wave.

Figure 13-12
Incipient half wave formed in a 138kV cable by a 5
o
bend at 37.5
0
C
lateral: vertical scale 10:1
The formation of the waves is explained diagrammatically in Figure 13-13:
The lateral displacement of the cable entering the pipe bend produces the bending moment
M
1
.
M
1
is applied to the adjacent cable and bends it to a radius of R
1
, forming a half wave.
The axial force F continues to increase, as a result of the temperature rise.
The force F and radius R
1
, apply a bending moment M
2
to the adjacent cable and this
produces another half wave of radius R
2
.
In turn, R
2
and F produce bending moment M
3
; and so on, progressively along the pipe
length in step with the rising temperature.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-22

Figure 13-13
Application of bending moments and growth of waves from pipe bend
A guide to the calculation of wave onset follows:
The cable generates its own bending moment M
1
at the entrance to the preformed bend in the
pipe. For the purpose of description; the initiating bending moment can be considered to
comprise a) the axial thermomechanical force F and b) a lever arm L
a
.
M=F.L
a
Equation 13-8
Where:
M: bending moment applied to cable [Nm]
F: thermomechanical axial force [N]
L
a
: moment arm, a function of the cable-pipe clearance D and bend geometry [m]
The lever arm L
a
is a convenient fictitious length, which is generated as a) the cable moves
dynamically outwards into the bend, b) is lifted up the pipe wall and c) experiences sidewall
force. It depends upon the pipe clearance, the bend geometry, the coefficient of friction, the cable
weight and the cable bending stiffness. As such it cannot be derived by simple calculation.
The method of determining the temperature, axial force and bending moment at which the cable
starts to bend is to analyse a dynamic model of the span geometry using the explicit FEA (finite
element analysis) technique, as described in Chapter 5 of the EPRI project report.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-23
The consequence of the bending moment is that it bends the adjacent cable into a radius of
curvature R, as given by the fundamental relationship:
M
EI
R = Equation 13-9
Where:
R: radius to which cable is bent [m]
EI: bending stiffness of cable [Nm
2
]
M: bending moment applied to adjacent cable [Nm]
EI is the bending stiffness. Cable designers select this to be low, so that the cable is sufficiently
flexible to be wound onto a shipping reel.
Equation 139shows that a flexible cable, ie with a low value of EI, will be bent into a tighter
bending radius. Within sensible limits this is beneficial as a) the wave will form at lower
temperature and b) it will absorb more thermal strain.
Unlike paper cables, XLPE cables have the distinct advantage of being mechanically robust. The
extruded insulation is a homogeneous cylinder, which bends in a predictably uniform and elastic
manner.
The bending stiffness EI, as with axial stiffness EA, is temperature dependent. Bending stiffness
decreases at higher temperatures as shown in Table 13-5.
The bending stiffness is influenced by the three main components of the cable design:
The conductor stiffness increases with its cross-sectional area and gains its temperature
dependence from the XLPE insulation
The XLPE insulation holds the conductor in compression particularly at lower temperatures
and gives the conductor its temperature dependence. However its own bending stiffness is
low.
The outer cable layers form a tube comprising the ground conductor, metallic sheath and
extruded polymeric jacket. A geometric reason for the importance of these outermost layers
is that the bending stiffness of a tube is proportional to the 4
th
power of diameter. The
stiffness is increased by a) extruded or welded metallic sheaths and b) HV and EHV cables,
having increased insulation thickness and larger diameters.
Examples of the bending stiffness of two different cable constructions are given in Table 13-5.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-24
Table 13-5
Values of bending stiffness EI
System
Voltage
Conductor
Area [kcmil]
Cable
Diameter
[mm]
Ground
Conductor and
Sheath
Jacket Temperature Bending
Stiffness EI
[Nm
-2
]
20
o
C 2,700 138kV 1,500 83 Copper tapes
and aluminum
foil
Extruded
PE
90
o
C 1,100
20
o
C 7,700 230kV 2,500 125 Extruded lead
alloy
Extruded
PE
90
o
C 5,600
Limiting Effect of Pipe Friction on Strain Absorption
The incipient half waves grow by absorbing locked in thermal strain from the adjacent rigid
cable.
In the theoretical case of the pipe being friction free, all of the locked-in strain would suddenly
be released from the whole span length and the wave pattern would grow immediately along the
pipe.
Real ducts and pipes possess distributed friction between the duct and the cable (and also
between the cables in a pipe system). Friction limits movement and sets up a force differential
along the route.
Figure 13-14 shows the force distribution at wave formation. In this example the wave is
initiated by a small bend at the center of the 200m span and one wave is formed on either side of
it. Note that the convention used in FEA post processing is for compressive force to be negative,
such that the curve is drawn below the axis of the graph. Thus Figure 13-14 shows F
1
as a
8,500N compressive force falling by 1,300N to F
2
at a level of

7,200N.
Before wave formation occurs, the force F
1
is constant along the span, at the rigid-bar value.

After wave formation, Figure 13-14 shows:
The force has dropped locally from F
1
to F
2
by
f
, Equation 13-10.
The force is reduced in the adjacent cable over a length L
f
= 76m, due to the frictionally
constrained cable ( = 0.1), as given in Equation 13-11.
g . W . L . F F
f 2 1 f
= = Equation 13-10
Re-arranging Equation 13-10;
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-25
( )
g . W ,
F F
L
2 1
f

= Equation 13-11
Where:

f
: frictional constraint force [N]
F
1
: Rigid-bar force before formation of incipient wave [N]
F
2
: Local force after formation of incipient wave [N]
: Coefficient of friction between cable and pipe
L
f
: Length of frictional constraint [m]
W: Weight of cable [Kgm
-1
]
g: Acceleration due to gravity, 9.81 [ms
-2
]

Figure 13-14
Frictional constraint force and length, after formation of incipient wave
Transition to Large Amplitude Waves
The sine waves in the thermomechanical patterns have the property that they can absorb an
increasing proportion of locked-in strain in a non-linear way by increasing their transverse
amplitude, as shown in Figure 13-15.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-26
To increase the wave amplitude, additional energy is required to bend the cable and to lift it up
and around the pipe. A critical energy state is reached when it is possible to reduce the total
energy stored in the cable system by the transverse growth of large amplitude waves.

Figure 13-15
Nonlinear absorption of strain by increase in sine wave amplitude
The transition to large waves is a point of dynamic instability, which occurs quickly. A large
proportion of the locked-in thermal strain within the elastically compressed cable is released by
the sudden growth of large wave amplitude waves.
Figure 13-16 is the same example of the 138kV duct at the formation of large amplitude waves
as shown in Figure 13-12 at the formation of incipient waves, following a further 12.5
o
C increase
in temperature. The amplitude of the wave has increased by 280%.
The transition to a larger wave amplitude has some similarities to the buckling event that
occurs in civil engineering, when an unsupported strut fails in compression. Following the
accelerated lateral movement under constant load, a strut fails immediately in an inelastic and
irreversible manner.
The cable is prevented from doing this by a) the constraining effect of the curved duct walls in
limiting transverse movement and b) the reduction in force following the release of locked-in
strain. The constraining effect of the duct environment delays the onset of the dynamic transition
by a further 12.5
o
C rise in temperature, such that two events occur. The second transition is also
one of elastic buckling, being reversible when the temperature is reduced.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-27

Figure 13-16
Large amplitude sinusoidal wave patterns at a turn-over temperature of 50
0
C
lateral: vertical scale 10:1
With the large cable-pipe clearances inherent to pipe systems, the sudden absorption of locked-in
strain reduces the axial force. The axial force characteristic turns-over and falls in magnitude at
the critical turn-over temperature, as shown in the right hand graph in Figure 13-17, for the
138kV 1500kcmil pipe clearance of 123mm.
The turn-over temperature is 45
o
C and the turn-over forces are 11,700N, 12,700N and
15,800N for each of the three cables. The reduction in force is such that the turn-over force of
15,800N is the maximum axial force that the cable experience at B up to 90
o
C. This is 33% of
the rigid-bar level that would otherwise have been experienced by the cable at 90
o
C; a very
beneficial reduction to the design and operation of the system.

Figure 13-17
Force-temperature characteristic, 138kV 1500kcmil cables with pipe clearance of 123mm.
Span end A on left. Pipe entry B on right
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-28
With the small clearances that have traditionally been selected in duct installations, a turn-over
force and turn-over temperature still occur. The waves are restricted in amplitude by the small
clearance in the pipe and so can only absorb a smaller proportion of locked-in strain. In
consequence the force does not fall completely, but rises with temperature at a reduced rate, as
shown in the right hand graph in Figure 13-18. The effect is still beneficial as at 90
o
C the axial
force is reduced to 80% of the rigid-bar force. By increasing the clearance, significantly larger
reductions are achievable.

Figure 13-18
Force-temperature characteristic 230kV 2500kcmil cable with duct clearance of 38mm.
Span end A on left. Pipe entry B on right.
The downside of thermomechanical patterns is that the bend radius within the wave locally
stretches the metallic sheath or foil by a small amount, called bending strain. As the cable heats
and cools the wave amplitudes expand and contract cyclically. This risks a reduction in the
fatigue life of the metal. Fortunately the pipe or duct is within a controlled geometric
environment, which limits and distributes the size and movement of the cable within the wave
patterns.
Turn-Over Force Due to Frictional Slip at the Span End
At any moment the magnitude of axial force is always different along the span, between end A
and the entrance to the bend at B, because:
The release of locked-in strain into the bend reduces the force at B.
The release of locked-in strain into the waves formed at B reduces the force
The frictional constraining force F separates the rigid-bar force at A, from the reduced force
at B.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-29
If the difference in axial force between A and B exceeds the maximum frictional constraint of the
200m length of cable, then the locked-in thermal strain will slip from end A towards, and into,
the bend at B. This can be seen to have occurred on two occasions in Figure 13-17, left hand
side, because:
At 20
o
C in cable 1 (top cable). The cable moves into the bend at B, reduces the force locally
and thereby increases the force differential along the span.
At 45-47.5
o
C in cables 1, 2 and 3. The formation of large amplitude waves at B, reduces the
force locally and thereby increases the force differential along the span.
The second occasion at 45-47.5
o
C, is followed by a progressive drop in force at A. This is also
defined as a turn-over temperature and turn-over force event. At A, the turn-over results
from frictional slip, whereas at B it results from wave formation. Wave formation is the common
controlling event. Once waves have formed at B it is inevitable that, following a sufficient
temperature rise, cable slip will eventually occur at A.
As the temperature increases the force falls progressively at A. This is because the wave pattern
from B is advancing rapidly towards A and is releasing the locked-in strain.
Note that up to 80
o
C the force at end A was significantly higher than at B. Above 80
o
C a reversal
occurs in which the force at A falls below that at B. This is because the waves at B cannot grow
further in amplitude and so the rate of generation of locked-in strain with temperature exceeds
the rate of absorption.
Frictional slip is also present in the 230kV 2500kcmil duct installation, with the smaller
clearance of 38mm, as shown in Figure 13-18, left hand graph:
Slip does not occur at low temperature, because the small clearance has reduced the
absorption of the bend at B. Thus the force rises almost as quickly at B as A, holding the
force differential to low magnitude.
Slip finally occurs at 70
o
C. This is because large wave patterns formed at 67.5
0
C at B,
reducing the local force and increasing the force differential.
Up to 105
o
C the force at A is higher than at B, at 105
o
C a reversal occurs and the force at A
falls below B.
Thermomechanical Patterns in the Full sized Span
The double span of 138kV 1500kcmil cable in Figure 13-1 has 22 bends counted separately in
the horizontal and vertical planes, with 11 of these being co-incident, thus leaving 11 separate
bend positions, all in the vertical plane.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-30

Figure 13-19
Thermomechanical pattern distributions in the double span at 15
o
C, 30
o
C, 45
o
C, 60
o
C, 75

o
C and 90
o
C
The positions of the wave patterns in the route for cable number 3 are shown in Figure 13-19.
(Cable number 3 is at the bottom right hand side of the trefoil group). There are 6 lines shown,
the first is at the starting temperature of 15
o
C and the others are at 15
o
C steps as recorded in
Table 13-6.
Table 13-6
Details of thermomechanical patterns in Figure 13-19
Line
Number
Temperature
[
o
C]
Temperature
Rise [
o
C]
Number of
Pattern
Sites
1 15 0 0
2 30* 15 1
3 45** 30 8
4 60 45 8
5 75** 60 9
6 90 75 8
* Wave turn-over temperature at joint manhole
** Wave turn-over temperature at pipe bends
This chart is divided into two overlapping diagrams to enlarge the patterns for greater clarity, in
Figure 13-20 and Figure 13-21, showing the patterns at:
45
o
C; the turn-over temperature at the bend in the normalized span in Figure 13-17, right
hand graph.
90
o
C; the cable operating temperature.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-31

Figure 13-20
A close-up view of patterns in left hand side of route at 45
0
C and 90
o
C

Figure 13-21
A close-up view of patterns in right hand side of route at 45
0
C and at 90
o
C
The analysis shows:
The first pattern was formed in the manhole at a temperature rise of 15
o
C. In the following
15
o
C step, the pattern had combined with that at the entrance to bend 7. Figure 13-22 shows
that at 90
o
C, the wave patterns have spread into the adjacent pipes.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-32
In the next15
o
C rise, patterns had formed a) at the entrances and exits to the 3 large angle
bends, with the cables in the bends having locked, these being [7,8,9], [10,11] and [20] and
b) around and on both sides of five bends with small bend angles.
One further pattern was formed at a rise of 60
o
C at a small-medium angle at bend 3.
At 90
o
C the wave pattern regions had grown to over 85% of the route length.

Figure 13-22
Deformation patterns inside 138kV joint casing have spread to form wave patterns in
adjacent pipes. Perspective view at 90
o
C
The major thermomechanical events can be predicted and are given below. Without the use of
the FEA modelling technique it is not possible to predict the events with certainty, nor is it
possible to predict minor events:
The complete cable length initially behaves as a rigid-bar
The cable moves into the nearest bends
The cable fills each bend in proportion to the bend angle
The cable locks itself into large angle bends by sidewall force, thus isolating the behavior of
each section.
The cable is restrained in small angle bends, but moves through them.
The axial force rises and, at the entrance to each bend, initiates incipient waves. If one
section has a larger effective pipe diameter it will initiate waves at a lower temperature. For
example, a) in an in-line joint casing, or b) in a duct with offset joints and wide cleat
spacings.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-33
The transition from small to large wave amplitudes occurs at most, but not all bends, at a
similar temperature. This results from a) the bends with large angles having locked more
effectively and b) some bends which are separated by only a short span length require a
higher temperature to generate the critical thermal force.
The wave patterns grow in length. Adjacent patterns merge to occupy the majority of the
span length.
Positioning of Manholes along the Route
To minimize risk of damage to joints and cable within manholes:
The axial force should be as low as possible
There should be no differential force across the manhole
There should be no cable movement through the manhole
The above requirements are met by positioning the manholes with a symmetrical
thermomechanical thrust on either side as shown in Figure 13-23 as marked as positions X, Y
and Z :
At the centre of two equal straight sections. Position X is the first choice, being at the center
of an almost straight and horizontal section, between the horizontal bend 20 and vertical bend
14. The bends in between 15 and 20 are minor. Position Z is the second choice, being
between horizontal bends 3 and 4, albeit on an incline.
If bends are close by to position the manhole equidistant between two bends of equal bend
angle and radius. Position Y is the third choice, being in the center of a large S bend between
9 and 10. It would be essential to confirm that the position is safe using FEA modelling i.e.
that the bends lock simultaneously and prevent differential force and movement.
At the top bump in the vertical profile. For example the position at 15 is remote from large
horizontal bends and is at a crest, albeit the sections on each side have an asymmetric
elevation.
As a fall-back choice to position the manhole at the bottom valley in the vertical profile.
However the combination of the gravitational loading and thermomechanical forces must be
less than the compressive withstand of the manhole constraining system. Such a position
should not be selected if the slopes are too steep. Position 12 is at the bottom of a valley but
is unsuitable, being at the entrance to the large horizontal bend.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-34

Figure 13-23
Preferred locations of manholes
Positions to be avoided are:
Next to a large angle bend, as the cable will attempt to move from the straight span through
the manhole into the bend. If the cable cannot move, a significant differential force will act
on the cables and joints in the manhole.
In the centre of a very long straight span with lengths of cable on either side longer than the
frictional constraint length. A manhole at this position will experience maximum rigid bar-
force. If this position is unavoidable then a) the manhole should be designed to withstand the
full force and b) consideration should be given to deliberately initiating wave patterns
symmetrically on either side of the manhole, for example by inserting small angle pipe
bends. (Note, the latter is an experimental technique and it would be essential that a full FEA
modelling study be undertaken at the design stage).
At a change of inclination angle to the vertical, particularly if one of the slopes is steep.
In a straight span and up a steep incline, with a large angle bend at the bottom of the slope.
The effective angle of slip will be reduced, such that friction will only support a small force
differential. Slip and cable movement will occur a significantly longer cable length, thereby
imposing differential loads across the manhole.
At the bottom of a steep straight incline with a bend at the top. The slope angle increases the
frictional constraint such that the cable cannot feed its thermal strain uphill into the wave
pattern at the bend. Thus the manhole experiences both gravitational load and a higher rigid-
bar thermomechanical force. If this position is unavoidable, consideration should be given to
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-35
introducing a small bend on the uphill side of the manhole to initiate wave formation. (Note
the latter is an experimental technique and it is essential that a full FEA modelling study be
undertaken at the design stage).
Calculation of Thermomechanical Effects
To design the pipe or duct system the following calculations are required:
Maximum axial force distribution at certain key positions along the span/route
Maximum sheath strain (absolute and cyclic)
Maximum sidewall force in bends
Minimum cable bending radius
Maximum differential forces at manhole positions
Maximum differential load on cleats
Maximum differential load on joints
In the detailed design of a new circuit it is recommended that the calculations be extended to:
The change of forces occurring during load cycling (a minimum of ten cycles is
recommended)
Emergency loading temperatures e.g. to 105
o
C
Low winter ambient temperatures e.g. depending on the location, to -30
o
C.
Calculation: FEA Method
The FEA explicit modelling method (finite element analysis) is recommended for the detailed
design of a new pipe or duct system.. The method is described in detail in Chapter 5 of the EPRI
project report. Example applications are:
The selection of the optimum pipe-cable clearance
The detailed design of the route geometry
The positioning of the manholes
The choice between straight and anchor joints at key positions
The identification of thermomechanical pinch points
The deliberate introduction of wave inducing pipe bends to minimize loads and protect
manholes.
Modelling cyclic loading conditions.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-36
Calculation: NS Method
The NS method (normalized span) is recommended for:
Outline design of duct and pipe systems.
Feasibility studies
Preparation of requests for tender.
Selection of the optimum pipe/duct clearance for thermomechanical and ampacity
calculations.
Preparatory selection of optimum pipe clearance before performing a detailed FEA
modelling study of the route.
The NS method is described in detail in Chapter 8 of the EPRI project report. The necessary sets
of reference data are given in Chapter 7 of the EPRI project report.
The method requires that the span geometry be divided and the subsections be compared to that
of the reference span in Figure 13-7 and the reference set of parameters given in Chapter 7 of the
EPRI project report. The method provides:
Maximum axial force at positions A and B
Maximum sheath strain (absolute only) at A and B
Maximum sidewall force in bends at B
Minimum cable bending radius at B within the wave pattern and in the bend
Maximum differential forces at manhole positions by subtraction of forces at A and B
If any doubt exists about particular positions on the route with a) extreme geometrical features,
such as steep slopes, sharp bends, long slopes, long section lengths or b) difficult manhole
positions, then the FEA calculation method should be selected.
Preparation of Input Data
The initial stages of obtaining geometric parameters for the cable, joint and pipe are common to
the explicit FEA method (Section 2) and to the NS calculation method (Section 3).
Cable and Pipe Data
Lists of the parameters required are given in Section 3, Table 13-1 and Table 13-2.
In particular the values of axial stiffness EA, the bending stiffness EI and, to a lesser extent the
torsional stiffness GJ, are required for the particular cable construction over a 15
o
C-110
o
C range
of temperatures.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-37
Chapter 7 of the EPRI project report gives methods of calculation of the cable stiffness values
based on the experimental measurements from two cable constructions (a 138kV 1500kcmil
aluminum foil sheath cable and a 230kV 2500kcmil lead alloy sheath cable).
If doubt exists on the applicability of the calculated cable stiffness values to a particular cable
design, or, if the thermomechanical performance of the cable or joints is calculated to be
marginal, then it is strongly advised that the cable stiffness parameters be confirmed by
experimental measurement. The test techniques are described in Chapter 3 of the EPRI project
report.
Route Data
The full [x,y,z] co-ordinates of the proposed pipe or duct route are required for both the FEA and
NS techniques. The co-ordinates from digital route surveying and mapping are suitable.
Alternatively co-ordinates can be read off existing route and span drawings in either scanned or
hard copy formats.
Ranking of Key Design Parameters
As an aid in the selection of cable designs, route geometries and duct/pipe clearances, the 12
parameters are ranked according to the magnitude of their effect for duct and pipe systems at A
and B.
The rankings are given relative to the normalized span in Figure 13-7 and to the reference 138kV
1500kcmil cable and 150mm clearance.
The 12 reference parameters are given in Table 13-1 and Table 13-2.
The magnitudes of the forces given in Table 13-7, Table 13-8, Table 13-9 and Table 13-10
should not be used directly for the design of a particular duct or pipe system. It is essential that
the calculation procedures given in Section 3 are followed. The magnitudes of the forces given in
the tables are lower than will occur in practical duct and pipe systems arising from a) the large
reference clearance of 150mm, b) the use of low values of axial stiffness EA, coefficient of
friction and coefficient of expansion . Updated values of EA, and are given in Table
13-2.
A commentary on the ranking study is given in Chapter 7 of the EPRI project report.
Duct System Ranking of Parameters at End of Span, A
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
13-7 in descending order of total percentage change in force.
Reference peak axial force: 7,520N
Turn-over temperature: 70
o
C
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-38
Table 13-7
Duct at position A: Ranking list of parameters by peak force at turn-over temperature
Duct System: Ranking of Parameters at Entrance to Bend, B
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
13-8 in descending order of total percentage change in force.
The reference model values are:
Peak axial force: 5,372N
Turn-over temperature: 72.5
o
C
-50% +50% Parameter Total Change
[%] % Force [N] % Force [N]
Slope angle 106 -53 3,555 53 11,478
Clearance 65 +47 11,054 -18 6,197
(Friction changed to =0.3) (58) (-29) - (+29) -
Length 49 -26 5,574 23 9,257
Weight 48 -24 5,709 24 9,290
Bend angle 48 +24 9,308 - -
Friction 36 -11 6,705 25 9,432
Expansion 26 -11 6,696 15 8,646
EI 26 -15 6,393 11 8,343
EA 11 -5 7,115 6 7,592
Bend radius 0 0 7,487 0 7,505
GJ 0 0 7,520 0 7,523
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-39
Table 13-8
Duct at position B: Ranking list of parameters by peak force at turn-over temperature
-50% +50% Parameter Total Change
[%] [%] Force [N] [%] Force [N]
Clearance 153 +131 12,411 -21 4,252
Expansion 43 -22 4,209 +23 6,587
EI 36 -23 4,119 +13 6,081
Length 23 -19 4,364 +4 5,600
Weight 22 -11 4,769 +11 5,979
(Friction changed to
=0.3)
(22) (-11) - (+11) -
EA 17 -21 4,230 -3 5,214
Bend Angle 16 +8 5,798 - -
Friction 9 -4 5,172 +5 5,649
Slope angle 7 -2 5,247 +5 5,624
Bend Radius 0 -3 5,208 -3 5,198
GJ 0 0 5,363 0 5,379
Pipe System: Ranking of Parameters at Span End, A
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
13-9 in descending order of total percentage change in force.
The reference values are:
Peak axial force for table reference: 10,000N
(axial force in top cable 7,960N and bottom cables 10,600N and 10,000N).
(Note that the reference force for the NS calculation method is 10,458N)
Turn-over temperature: 57.5
o
C

EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-40
Table 13-9
Pipe cable 3 at position A: Ranking list of parameters by peak force at turnover
temperature
-50% +50% Parameter Total Change
[%] [%] Force [N] [%] Force [N]
Clearance 224 +176 27,574 -48 5,242
Slope angle 92 -63 3,237 +29 12,858
(Friction pipe-cable,
changed to =0.3)
(58) (-31) - (+27) -
Expansion 46 -19 8,114 +27 12,730
Weight 46 -25 7,468 +21 12,100
Length 40 -13 8,732 +27 12,741
Friction, pipe-cable 32 -8 9,232 +24 12,375
EA 31 -11 8,867 +20 12,070
Bend radius 20 -14 8,600 +6 10,600
Bend angle 14 -7 10,711 - -
Friction, cable-cable
(changed to =0.3)
11
(6%)
-6
(-3)
9,402
-
+5
(+3)
10,466
-
EI 10 -7 9,335 -3 9650
GJ 0 +4 10,434 +4 10,434
Pipe System: Ranking of Parameters at Bend Entrance, B
The effect on the axial force of varying each parameter by 50% is ranked numerically in Table
13-10 in descending order of total percentage change in force.
The reference model values are:
Peak axial force: 7,322N
(axial force in top cable 5,628N and bottom cables 7,035N and 7,322N)
Turn-over temperature: 55
o
C

EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-41
Table 13-10
Pipe cable 3 at position B: Ranking list of parameters by peak force
-50% +50% Parameter Total Change
[%] [%] Force [N] [%] Force [N]
Clearance 281 +253 25,860 -28 5,303
Expansion 49 -22 5,750 +27 9,282
EA 42 -21 5,789 +21 8,855
(Friction cable to cable
changed to =0.3)
(32) (-16) - (+16) -
EI 25 -10 6,615 +15 8,448
Weight 24 -12 6,444 +12 8,179
Friction, cable-cable 14 -1 7,273 +13 8,238
Slope Angle 10 +3 7,570 -7 6,816
(Friction pipe to cable changed
to =0.3)
(20) (-10) - (+10) -
Friction, pipe-cable 5 -3 7,104 +2 7,496
Length 5 -4 7,043 +1 7,422
Radius 5 -6 6,911 -1 7,243
Bend Angle 4 -2 7,200 - -
GJ 1 +1 7,424 +2 7,442
Design Parameters for Cables in Pipe and Duct Systems
The design values are given for the following parameters, laid out in the order that they will be
required in the design of the system.
It is noted that the values are for thermomechanical design and not for installation. In some
circumstances it will be necessary to increase the radii of preformed bends to reduce sidewall
force and tension during pulling-in.
Choice of cable type
Bending radius of preformed bend
Bending radius in manholes
Sidewall pressure on XLPE insulation and PE jacket
Cable axial force
Cable radius of curvature
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-42
Sheath fatigue strain
Choice of Cable Type
Extruded insulation XLPE cables are in general more robust than fluid filled paper cables. They
are particularly suited to duct and pipe installations:
The bending performance of the insulation is predictable because a) the construction is
homogeneous and b) the bending mechanism is controlled by elastic properties and not by
sliding of individual paper tapes.
The duty on the metallic sheath is reduced as there is no requirement to contain pressure.
The duty on cables in a duct or pipe system differs from that in a fully constrained, laid direct
application:
They are required to withstand longitudinal and lateral movement and cyclic bending.
Fatigue strength is therefore a key design parameter pipe and duct cables.
The axial forces are lower, because of the freedom to move into bends and to form
thermomechanical patterns.
The sidewall forces will be lower for a given bend radius, because the axial forces are low.
Certain constructions of XLPE cable have vulnerabilities, which need to be noted when selecting
cable to suit pipe and duct systems:
The XLPE insulation softens significantly above 60
o
C. The sidewall force in a bend results in
the conductor attempting to push its way outwards through the insulation. Reduction of
insulation thickness can be minimized by a) reducing the axial force, b) increasing the radius
of the preformed bends and c) reducing the operating temperature. The design limits are
given in the following section.
The XLPE insulation has a high coefficient of thermal expansion and requires room to
expand radially to avoid damaging itself, the insulation shield or the outer components. The
cable should have an allowance for internal radial thermal expansion, this usually being
controlled by compressible semiconducting cushioning tapes laid over the extruded core.

The coefficient of thermal expansion of XLPE increases to a value of 120 x 10
-6

o
C at 90-
105
o
C, this being 6 times higher than the typical values for the metals in the construction.
An expansion allowance of 0.5-1mm will be required on the core diameter of an HV/EHV
cable. The compressible cushioning tapes will require to be twice this thickness.
XLPE cables are available with foil laminate sheaths and ground wire conductors as a lower
cost alternative to extruded lead or aluminum sheaths. Foil laminate constructions are less
robust to external damage than thick wall metallic sheaths. Pipe and duct applications are
better protected against external damage.

Foil laminate cables are vulnerable to internal damage by the radial expansion and sidewall
forces pressing a) the extruded insulation and shields into the wires, risking electrical failure
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-43
and b) the wires into the foil laminate and PE jacket thereby risking water ingress.
Semiconducting cushioning tapes should be laid over the extruded core and over the ground
wires.
The sheath type that is most vulnerable to reduction in fatigue life is extruded lead. As listed
in Table 13-15 it will be seen that some lead alloys have a higher fatigue resistance.

It would be unwise to select a lead sheathed cable for a pipe system as this inherently has a
large clearance and will readily form large amplitude wave patterns of small radius and high
sheath strain. Extruded corrugated aluminum sheaths possess the best fatigue resistance, but
will be stiffer and will increase the cable diameter. Thin wall welded sheaths (i.e. less than
2mm thickness) are expected to have the second best fatigue resistance. Foil laminate sheaths
bonded to the PE jacket have the third best fatigue resistance, but the cable will be larger in
diameter than one protected by a foil laminate sheath. Foil laminate sheaths have a fatigue
resistance significantly higher than that for a lead alloy sheath, but they depend on the
bending support of the jacket, which reduces at a typical operating temperature of 75-80
o
C. A
copper foil laminate is likely to have a better fatigue resistance than an aluminum foil.
If doubt exists on the design limit for sheath strain, test results should be obtained from the
cable supplier for the particular type of sheath. This is particularly true for designs of thin
welded sheaths and foil laminate coverings, as their fatigue performance will have a greater
dependence on the particular material type, cable dimensions, and manufacturing method.
A range of conductor constructions is available. The thermomechanical studies were based
on test results from stranded copper conductors with light-medium compacted round wires.
As part of this work a study was undertaken on the sensitivity of axial force to conductor
material and to conductor construction (stiffness):

Stranded copper conductors are preferred as these generate the lowest axial force: a) the
coefficient of thermal expansion is 40% less than that of an aluminum conductor and b) the
stranded wire construction has higher flexibility and so will form wave patterns at lower axial
force.

In the absence of experimental data on the axial and bending stiffnesses, EA and EI, the
conductor constructions that should be avoided for duct and pipe systems are a) solid
aluminum conductors and b) circular copper conductors formed of shaped segments. These
conductor constructions are a) less flexible and will generate higher thrusts and b) do not
have the benefit of force alleviation by the mechanism of conductor relaxation.
For pipe and duct applications it is important that the cable external diameter is circular and
without a flattened surface. This is to ensure that the clamp cleats used in the manhole
constraining system can achieve their specified withstand strength without crushing the
cable. It is desirable for pipe and duct applications that cables be specified to be circular. An
ovality specification of 2% maximum is recommended.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-44
Bending Radius of Preformed Bends
The minimum bending radii for the preformed pipe and duct bends are given in Table 13-14. The
thick wall sheaths are the extruded types with thicknesses of 2mm and above. The thin wall
sheaths are the welded types and foil laminate types. As with the limits for fatigue resistance
published information was not generally not available for thin wall sheaths. It is recommended
that when doubt exists, supporting test evidence be obtained from the cable supplier for the
particular pipe or cable application.
Table 13-11
Minimum bend radii for preformed pipe and duct bends
Minimum Bending Radius [m] Cable
Thick Wall
(1)
Sheaths Thin Wall Sheaths
69-161kV 25D 30D
220-345kV 30D 35D
(1): 2mm and greater radial thickness of metallic component
Bending Radius in Manholes
Table 13-12
Bend radii in joint manholes
Minimum Bending Radius:
Cable in Close Cleated Bend [m]
Cable
Thick
(1)
Wall Sheaths Thin Wall Sheaths
69-161kV 10D 15D
220-345kV 12D 19D
(1): 2mm and greater radial thickness of metallic component
Sidewall Pressure on XLPE Insulation and PE Jacket
The transverse pressure acting on the insulation is that which limits reduction of insulation/shield
thickness to less than 5%.
c b
R . R . 2
F
P = Equation 13-12
c s
R . P 2 F = Equation 13-13
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-45
b
s
R
F
F = Equation 13-14
P: pressure on insulation [Nm
-2
]
F: axial force into bend [N]
F
s
: sidewall force per unit length acting on cable [Nm
-1
]
R
b
: radius of bend [m]
R
c
: radius of conductor [m]

Table 13-13
Pressure Limits on XLPE insulation and PE Jacket
Pressure [bar] Cable
90
o
C 105
o
C
69-138kV 6 bar 2 bar
161-345kV 6 bar 2 bar
1bar: 10 x 10
5
N.m
-2
(14.5lbf.in
-2
)
Cable Axial Force
The magnitudes of compressive and tensile axial forces do not directly imperil the cable
construction in pipe and duct applications. The cable in direct buried applications experiences
higher forces. It is recommended that the cable manufacturer supplies evidence of satisfactory
service operation.
Thermomechanical movement of the cable within the pipe occurs in service resulting in
reduction in axial thrust and generation of curvature in the cable. The sidewall force will be
reduced in the preset pipe bends and in the regions within the main pipe span containing cable
deformation patterns.
The joints and cable within a manhole are required to withstand a) the compressive axial force
without buckling and b) the differential axial force without movement.
The maximum value of acceptable axial force is based on the cable insulation pressure limit
given in Table 13-13.
F 2R
b
R
c
P Equation 13-15
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-46
Where:
F: limiting value of cable axial force [N]
R
b
: radius of the cable bend [m]
R
c
: radius of the cable conductor [m]
P: limiting value of lateral pressure on insulation [Nm
-2
]
Cable Radius of Curvature
The values given in Table 13-14 are for limits in checking the output of the thermomechanical
design calculations for both the FEA and NS methods.
In the bends the cable curvature will be less than the radius of the preformed bend, because the
cable crosses the longitudinal axis of the bend at the entrance and exit. It is important that the
axial thrust does not force the cable into a small radius at this position.
In the pattern regions the cable will be curved. In some designs of system, the axial thrust will
rise at the higher operating temperatures when the pattern can no longer absorb more strain, it is
important that the cable within the patterns is not over bent.
Table 13-14
Minimum radii of curvature for free cable in ducts and pipes
Minimum Bending Radius [m] Cable
Thick Wall
(1)
Sheaths Thin Wall Sheaths
69-161kV 15D 20D
220-345kV 25D 33D
(1): 2mm and greater radial thickness of metallic component
Sheath Fatigue Strain
The FEA and NS calculation methods give the magnitude of strain on the outer surface of the
cable jacket. This is converted to the strain on the outer surface of the metallic sheath using
Equation 13-16. Alternatively, if the radius of curvature R is obtained from Section 13.15.6, the
sheath strain can be calculated using Equation 13-17.
j
j
s
s
D
D
= Equation 13-16
R 2
D
s
s
= Equation 13-17
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-47
Where:

s
: strain in sheath [per unit]

j
: strain in jacket [per unit]
D
s
: outer diameter of metallic sheath or foil [m]
D
j
: outer diameter over jacket [m]
R: radius of curvature of cable [m]

Table 13-15
Sheath fatigue limits
Sheath Type Limiting Strain [%]

Category
(1)
Construction Material 0.146 x10
5

cycles (40
years)
5 x10
5

cycles
Thick wall Corrugated continuous
extrusion
Corrugated aluminum
sheath
0.5 0.25
Thin wall Welded strip Stainless steel, copper or
aluminum
0.25
(4)
-
Thin wall Foil laminate +PE Copper + copolymer 0.15
(2,3)
0.01
(2)

Thin wall Foil laminate +PE Aluminum + copolymer 0.15
(2)
0.01
(2)

Thick wall Continuous extrusion Lead alloy: copper-
tellurium: (0.06% Cu +
0.04% Te)
0.048 -
Thick wall Continuous extrusion Lead alloy E: (0.4% Sn +
0.2% Sb)
0.037 -
Thick wall Continuous extrusion Lead alloy C: (0.2% Sn
+ 0.075% Cd)
0.024 -
(1): Thick wall is 2mm and greater, thin wall is less than 2mm.
(2): A factor of 4 taken to extrapolate results from 20
o
C to 80
o
C
(3): Copper expected to have equal or higher fatigue strength than Al depending upon condition
(4): Provisional limits based on 50% of aluminum thick wall sheath, to be confirmed by test
Design Parameters for Joint Manholes in Pipe and Duct Systems
Joints are recognized as being the weakest part of HV and EHV XLPE cable systems in terms of
electrical reliability.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-48
During the first series of Berlin Electricitys long term prequalification tests in the mid 1990s on
400kV XLPE systems, the majority of joint designs failed due to thermomechanical disturbance.
The cables and joints were installed within representative sections comprising a duct, a tunnel
and a trench. The second test series had more success as most manufacturers had developed more
robust joint designs using prefabricated insulation.
In direct buried systems, the joints and cables are constrained from longitudinal movement by
distributed pressure and friction from the compacted backfill. The forces on either side of the
joint are largely balanced and differential forces are minimized. Where a direct buried section
emerges into a tunnel, an anchor joint is used to withstand the differential force.
In waved systems in tunnels, the cable is able to expand and contract laterally, under the control
of uniformly spaced cleats. If necessary, the cleats are provided with curved saddles to control
the minimum bending radius and sheath strain (for example with foil laminate cables). The joints
experience low compressive axial force and no differential force.
In a duct or pipe system, the cable is free to move and is only constrained by the terminations
and manhole constraints. Although the cable is partially restrained by frictional forces, the
asymmetry in a) span geometries and b) thermomechanical wave patterns, results in the
application of force and movement to the joint. The joint is the most vulnerable component in the
XLPE cable system and the least able to withstand disturbance, for example:
Buckling from compressive axial force
Internal damage to insulation and shields from differential longitudinal force
External damage due to longitudinal movement
The purpose of the manhole constraint system is to provide complete protection to the joint and
to the cable exposed in the manhole.
The following topics in manhole design are considered:
Choice of joint
Cleat types
Traffic vibration and short circuit forces
In-line joint constraint in a pipe system
Offset joint constraint in a pipe system
Clamped offset constraint in a duct system
Expansion cleat and S-bend constraint in a duct system
Choice of Joint
The types of joint in the range 69kV to 345kV are:
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-49
Prefabricated Premolded Joint
Also called the one-piece joint, Figure 13-24. During assembly the outer extruded shields are
removed from the cable to expose the XLPE core. The XLPE surface is polished and cleaned. On
one side, a rubber one-piece molding is stretched, pushed back and temporarily parked, either
directly onto the cable surface, or onto a disposable mandrel. The molding is manufactured from
a stretchable elastomer, either ethylene propylene rubber (EPR), or silicone rubber. The
conductor connections are made and covered with a cylindrical metallic shield. The one-piece
molding is then pushed along the cable, or pushed off the mandrel, to symmetrically cover the
conductor connection and the exposed cable insulation.
One variant is the plug-in design in which the conductor connection uses a plug and socket
principle. The male and female components are first assembled and then the molding is slid onto
one cable. This has the advantage that the parking distance and the overall length of the joint is
reduced.

Figure 13-24
A differential force and slip interface in a 138kV one-piece straight joint
The safe operation of the joint depends on the degree of stretch of the molding achieving and
maintaining a high level of pressure onto the cable insulation, such that air voids cannot exist.
Advantages are that the molding reduces a) site assembly time and b) the number of assembly
stages requiring skill. One disadvantage is that the interface is vulnerable to disturbance by
differential longitudinal force. The joint is inherently unsuited to be an anchor joint and within a
manhole requires to be protected against differential force.
Figure 13-24 shows the cable thrusting into the joint at force F
L
from the left and F
R
from the
right, with the differential force F = F
L
-F
R
:
When the adjacent spans and cable thermomechanical behavior is symmetrical, F
L
=F
R
and
F =0. The joint is safe from longitudinal movement and the interface cannot be disturbed.
The joint still experiences the high compressive axial forces F
L
and F
R
and requires to be
laterally supported to prevent them from either rotating the joint, or causing it to buckle, as
shown in Figure 13-24.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-50
When the forces are unbalanced, the differential force F moves the complete joint and cable
together from left to right until the molding contacts some part of the joint casing, as shown.
This stops the movement, but imposes a shear force F on the interface between the joint and
cable insulation.
A design study on one particular design of molding on a 138kV 1500kcmil cable calculated
that the interface would start to slip if the differential force exceeded F = 2,000N.
In addition to longitudinal movement at the interface, a void forms in the interface on the
XLPE cable insulation; thus interface slip risks electrical stress enhancement and failure.
The calculated withstand of 2,000N is associated with a comparatively high radial stretch of
the particular molding of 27%. Different combinations of cable and molding diameters may
have lower stretch and this will reduce, or eliminate the withstand capability.
Although the withstand capability provides a design margin to the safe operation of the
molding, it should not be relied on as part of the manhole constraint system.
Prefabricated Composite Anchor Joint
Figure 13-25 shows a design of prefabricated composite anchor joint of the type that has been
used from 66kV to 500kV with XLPE cables with conductor sizes up to 5,000kcmil.The cable
insulation is prepared as above. A premolded rubber stress cone is pushed back onto each side of
the joint. An epoxy resin casting is then pushed back and parked on one side. The conductor
connections are completed. The epoxy resin casting is slid into position and mechanically locked
onto the conductor ferrule. The two stress cones are then pushed into position and are held
permanently in compression throughout the service life by two banks of metallic helical coil
springs. The epoxy resin casting is finally bolted to supporting steel work to directly anchor the
conductor, via the rigid epoxy resin insulation, to the manhole wall. The joint anchor can
withstand completely unbalanced differential loads with typical withstand levels up to 60,000N.

Figure 13-25
400kV anchor straight joint of the prefabricated composite type, showing differential force
taken by anchor
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-51
Field Molded Joint
The cable XLPE insulation is reconstituted, but to a larger diameter by heat consolidated XLPE
insulation. This is applied either by tape molding or by extrusion molding. The joint has no
interface between the cable and joint insulation and therefore is immune to disturbance by
differential force. Unfortunately the joint requires high cleanliness, sophisticated tooling, high
skill and a very long assembly time, this being a major disadvantage in repair situations. It is
now almost completely superseded by prefabricated designs of joint
Each of the above joint designs is required to be held in straight axial alignment, such that lateral
movement and buckling cannot occur. If an outer casing or joint shell is fitted, then the cable and
joint inside must similarly be protected from longitudinal and lateral movement, this is
sometimes achieved by filling the casing with a thermosetting resin.
Cleat Types
Cleats are selected for use in each type of manhole constraint. There are three types of cleat:
Clamp Cleat
This is the main type required for use in pipe and duct manholes constraints. The role of the
clamp cleat, is to grip the cable firmly and achieve and maintain a specific withstand force
throughout the circuit life.

Figure 13-26
Generic design of clamp cleat for XLPE cable
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-52
A typical design is shown in Figure 13-26. The cleat is lined with a medium to soft rubber (e.g.
60
o
Shore hardness or less) to be able to conform to the cable shape without crushing it. It is
noted that few cables are circular. The larger cables may have 5mm of ovality and flattening. To
achieve a firm grip it is necessary for pipe and duct applications that cables be specified to be
circular to less then 2% of ovality.
The rubber liner is usually a high performance, high temperature elastomer and may be custom
molded with a special gripping surface, to be suitable for foil laminate cables.
For the particular type of cleat shown in Figure 13-26 , the length L of the cleat is calculated in
Equation 13-18.
L=F.k Equation 13-18
Where:
L: cleat length [mm]
F: cleat axial withstand force [N]
k: cleat holding constant for the particular cable diameter [mmN
-1
]
k=0.037 mmN
-1
for a 125mm diameter PE jacketed lead sheathed XLPE cable
k = 0.016 mmN
-1
for a 83mm PE foil laminate jacketed cable with a copper tape ground
conductor
The contact pressure is kept below the threshold limit for the particular cable construction. This
is important for foil laminate cables with ground wire conductors as these components can be
damaged by the cleat and, in the extreme, pushed through the extruded shield into the XLPE
insulation. It is important that the withstand strength of the cleat is validated for the particular
design of cable by a test at 70
o
C for a period of 24 hours or longer.
The cleat is provided with helical coil springs to accommodate the radial thermal expansion of
the cable. They prevent the heated cable from suffering permanent compression set of the
jacket and internal components. When the cable contracts during cooling, the cleat will lose its
grip and will allow the cable to slide through. It is good practice to allow a cable diameter
increase of 2mm, whilst still maintaining the clamp pressure within the maximum and minimum
design limits for the particular cable construction. Thus the selected spring will require to be long
in length.
Guide Cleat
The guide cleat is only required in special applications in pipe and duct systems. For example, to
protect the joint from experiencing internal differential force by allowing it to move
longitudinally and without deflecting sideways.
Traditionally the role of this cleat was to guide the cable axial thrust around a particular bend in a
transition section between either a) two rigidly constrained or b) two flexibly constrained
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-53
systems, whilst preventing lateral movement. Guide cleats divide the cable into very short span
lengths, typically in straight sections of less than 1m and less than 0.6 around bends. The short
spacing is selected to prevent the cable from buckling. The cleat is rubber lined and may employ
a hard rubber to minimize friction grip (80 or 90
o
Shore hardness). It is tightened to lightly
contact the cable.
The cleat is sometimes used indirectly in a clamping application by holding the cable in tight
geometrical wave patterns (vertical), or snake patterns (horizontal). The longitudinal force is
withstood by the combination of the stiffness of the cable and the cleat reaction force.
Expansion Cleat
This is shown diagrammatically in Figure 13-31. It comprises a clamp cleat with a very high
holding strength, which is allowed to travel longitudinally into the manhole under the constraint
of metallic coil springs. The objective is to free some of the locked-in thermal strain from the
cable in the duct. The axial force that has to be withstood by the rest of the constraining system
within the manhole is significantly reduced. Upon cooling, the cable within the duct contracts.
The longitudinal compressive force of the springs helps to push the cleat and cable back into the
duct.
To design the spring it is necessary to construct the equilibrium diagram to calculate the point
of intersection of the force-extension characteristic of a) the cable and b) the expansion cleat, as
shown in Chapter 12 of the EPRI project report. Springs with different force-extension
characteristics will be selected to suit different manhole locations, span lengths and span angles
of inclination.
It is vital that the holding strength of the cleat has a design margin above the force that can be
generated by the cable in any particular duct location, to prevent the expansion length of the
cable from sliding through the cleat into the manhole.
The expansion cleat is sometimes inappropriately named a ratchet cleat.
Short Circuit Cleat
The design of this type of cleat is outside the scope of the guide. The purpose of the cleats is to
prevent the cables from flying apart in the transverse direction under the electromagnetic forces
that occur during a system through fault.
It is usual to design the thermomechanical cleating system to also withstand the short circuit
forces. In some circumstances short circuit straps may also be applied mid span between the
cleats.
The short circuit force formulae given in Chapter 10 of the EPRI project report may be used to
calculate the lateral force. The FEA modelling technique was developed and is available to
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-54
simulate the complex forces and movements that occur real time, during a short circuit
throughout the whole duct or pipe system, including the manholes.
Traffic Vibration and Short Circuit Forces
Chapters 9 and 10 of the EPRI project report show that the effect of these disturbances is to
suddenly release a high proportion of the locked-in thermal strain from the cable within the main
duct or pipe span. The consequence is that there is a sudden longitudinal movement of cable
along the pipe or duct into the wave patterns.
After these events, the cable in the majority of the span length is converted into wave patterns.
Providing that the shock can be withstood, this is not harmful to the cable, but this is not so for
the contents of the manhole. The momentum of the cable movement will apply a longitudinal
impulsive force to the manhole constraining system. In the case of a short circuit, electrodynamic
forces occur within the manhole at the same time.
It is recommended that a design margin of 20% be allowed in the selection of the cleat holding
strength, particularly for the expansion cleat design to ensure that it cannot lose grip onto the
cable.
In-line Joint Constraint in a Pipe System
Figure 13-27 shows that the joint is held in three positions either side of the centre line. The
objectives are to prevent the occurance of the lateral, longitudinal and rotational defelctions
shown in Figure 13-22.
Trifurcating Clamp Cleat. One is positioned either end of the casing such that together they
withstand the unbalanced axial force. The differential force is calculated using the FEA or NS
methods. The cleat length is selected from Equation 13-18.
L=F.k Equation 13-18
Cable Spider and Binder Assembly. These a) centralize the cables within the casing, b) prevent
them from bowing and deflecting and c) form the cables into a structure with greater bending
stiffness.
Joint Spider and Binder Assembly. This centers the joints and binds them together and to the
central spider, to keep them in parallel axial alignment. In the preferred Style B version shown,
the support cradle is designed to be able to slide longitudinally to prevent differential force
appearing on the electrically stressed joint interface.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-55

Figure 13-27
Pipe System: In-line joint casing
Offset Joint Constraint in a Pipe System
Figure 13-28 shows the joints and manhole in plan, elevation and end view as modeled for the
FEA analysis. Figure 13-29 is a plan view showing the details of the offset and cleat positions.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-56

Figure 13-28
Pipe System: Joints offset on manhole wall: plan, elevation and end view
The joints are constrained by:
Trifurcating Clamp Cleats.
These are positioned at the pipe entrances to keep the cables in axial alignment
Single Clamp Cleats.
Four cleats are positioned around the offset bend.
Cable Offsets, as shown in Figure 13-29.
Two 30
o
x 15D radius offsets are selected which, in combination with 10 x 100mm long clamp
cleats, have a differential holding strength of 9,400N. This is greater than required in a 138kV
1,500kcmil cable system with a 123mm pipe clearance. The maximum compressive force
calculated to be present at the particular manhole is 9,000N, thus in principle the design has an
ability to withstand a completely unbalanced axial force.
Joint Guide Cleats.
These constrain the joint against lateral movement, but allow it to move lengthwise. The
lengthwise movement is recommended to eliminate any differential load occurring. For example,
the clamp cleat rubber liners will elastically deform up to 4mm. A design allowance of 8mm is
recommended for joint movement.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-57

Figure 13-29
Pipe System: Plan view of manhole showing cable offset angle and cleat positions
Clamped Offset Constraint in a Duct System
Figure 13-30 is the plan view of three 230kV 2500kcmil joints mounted on the manhole wall,
one above the other. They are constrained by:
Clamp Cleats.
One clamp cleat is mounted at each duct entry and 7 x 200mm clamp cleats are mounted around
each offset, making a total of 16 clamp cleats.
Cable Offset.
Two 60
o
x 12D offsets are selected which, in combination with 14 x 200mm long clamp cleats,
have a differential holding strength of 10,500N. (The two cleats at the duct exits are included to
provide a nominal back load of 2,700N, representing a low level of cable back thrust, for the
purpose of design calculation.)
Joint Guide Cleat.
A conventional straight joint is selected if the force differential is limited to less than 10,000N by
the cleated offsets. The joint is allowed to slide longitudinally within the guide cleat by a
maximum of 8mm, to eliminate longitudinal force at the joint/cable interface.
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-58
If the force differential at a particular location is greater than 10,000N, then either a) the force is
reduced by increasing the duct clearance from to 50-75mm or b) an anchor joint is selected.

Figure 13-30
Duct System: Offset clamped cleat constraining system
Expansion Cleat and S-Bend Constraint in a Duct System
Figure 13-31 is the plan view of three 230kV 2500kcmil joints mounted on the manhole wall,
one above the other. They are constrained by:
Expansion Clamp Cleat.
One expansion cleat is mounted at the duct entry. This allows the cable to move from the duct
into the manhole by up to 250mm. The typical movement is 150mm. This system can be
effective in reducing the axial force to 10% of the rigid-bar value.
Flexible Cable S Bend.
The S-bend is a semi-flexible structural cable shape designed to accommodate 150mm
movement. In the example shown, an 8-10D radius S bend is required to move cyclically
outwards by a total movement of 200mm. Because of the elastic bending stiffness and memory
of an XLPE cable it is difficult to form it permanently into a tight radius. It is important that the
magnitude of cyclic strain within the metallic sheath is below the values given in Table 13-15.
Cable Clamp Cleat.
One clamp cleat is provided either side of the joint to keep the cable and joint in axial alignment
and to prevent longitudinal movement.
Anchor Joint.
The straight joint shown is of the prefabricated composite type with the joint casing anchored
EPRI Licensed Material

Application Guide for the Design of Duct-Manhole Systems and Pipe Systems Containing Underground HV XLPE
Cables
13-59
by joint clamp cleats. In certain circumstances it would be possible to use a conventional straight
joint without an anchor, as a properly balanced expansion cleat system reduces the differential
load to low level.
Cable Stocking.
At the opposite end of the manhole, the cable exits into a downhill duct. A cable stocking is
provided to limit its downhill movement. At the next manhole downhill the cable is constrained
by another expansion cleat. The designer has to balance the thermomechanical loads,
gravitational loads and duct frictional loads so that uniform cable movement occurs at each
manholes at the ends of a span. In the example in Chapter 12 of the EPRI project report, the
cable expands into the manhole also by 150mm, but is constrained by the weight of the downhill
cable, instead of the springs in an expansion cleat.
If the force differential at a particular location is greater than 10,000N then either a) the force can
be reduced by increasing the duct clearance to 50-75mm or b) an anchor joint can be selected.

Figure 13-31
Duct System: Expansion cleat and S bend constraining system
EPRI Licensed Material
14-1
14
CONCLUSIONS
The following are the overall project conclusions. Detailed conclusions are given at the end of
each Chapter:
The mechanisms of how thermomechanical, gravitational, short circuit and traffic vibration
forces effect XLPE cables in pipe and duct systems have been thoroughly studied and are
now understood. This is the first time in the history of pipe and duct systems that this
knowledge has been gained. This is a major advance in power cable engineering in the design
of reliable circuits free of thermomechanical problems.
The technology that made the progress possible was the development of a) a pipe and duct
modelling technique and b) the use of a specialist type of finite element analysis (FEA)
software. The technique permitted the dynamic behavior of pipe and duct systems to be
replicated, using the explicit solution method. This method overcame the problems of
modelling a) long and complex route geometries, b) dynamically moving cable patterns, c)
distributed frictional contact and e) high speed events. The case of cables in pipes is
particularly complex as the contact interactions occur between the three cables as well as the
pipe.
The thermomechanical behavior was based on measurements made at the EPRIsolutions
laboratory at Haslet. The cable measurements were taken over a range of temperatures from
ambient to 105
o
C; the parameters being a) axial stiffness in compression and tension, b)
bending stiffness, c) torsional stiffness, d) thermally constrained thrust and e) coefficient of
thermal expansion. All of the tests were performed on the 138kV XLPE 1500kcmil foil
laminate cable. Selected tests were performed on the 230kV XLPE 2500kcmil lead alloy
sheathed cable. The most useful tests were: a) axial stiffness in bending, b) thermally
constrained thrust and c) coefficient of thermal expansion.
The manhole design studies were based on measurements made at the EPRIsolutions
laboratory at Haslet. Tests were undertaken to measure a) the cleat pull-through strength at
ambient and operating temperature and b) the XLPE core pull-through strength and internal
coefficient of friction. The test samples were a) 138kV 1500kcmil foil laminate cable and b)
230kV 2500kcmil lead alloy sheathed cable
Two methods of calculating the forces, movements and geometries have been developed,
proven and are now available for engineering use:
The FEA method is recommended for the detailed design of new circuits and the
major re-appraisal of existing circuits. The advantage is that it closely models the
behavior of the complete cable system in the real world, under a variety of operating
conditions. This method is non-trivial and requires the use of specialist software, a
workstation computer and expert personnel. It is now fully adapted to model cable
EPRI Licensed Material

Conclusions
14-2
systems and has been successfully used to analyse the performance of four circuits
comprising cables and manholes. The method has a demonstrated single run
capability of either a) 16km lengths, or b) 40 loading cycles.
The NS method (Normalized Span) is recommended for the outline design of
systems, feasibility studies, preparation for requests for tender etc. The advantage is
that when transferred to run on standard PCs, results are obtained quickly. The
method compares particular parts of a service span with the performance of a
normalized reference span. The effects are interpolated from the12 parameters needed
to characterize the cable design and span geometry. The method uses large sets of
data analyzed for a sensitivity study using the FEA technique.
The parameters that limit the design and performance of duct and pipe systems have been
studied and quantified. The primary parameters are a) sidewall pressure in bends acting to
move the conductor through the XLPE insulation and shields, b) fatigue strain in the metallic
foil or sheath and c) the axial force acting in manholes. The secondary parameters are a)
sidewall force in bends, b) cable radius of curvature and c) radii of preformed pipe and duct
bends.
A guide to the application of the technology to the design of duct and pipe systems has been
written. The guide covers all aspects of thermomechanical design including a) selection of
cable and joints, b) calculation of axial force and pipe clearance, c) calculation of the limiting
parameters, d) comparison with quantified design values, e) location of manholes within the
route, f) selection and design of manhole constraints and g) proposals of how the
thermomechanical design of the system can be manipulated by novel techniques, (such as the
incorporation of small kinks into the pipe and cable, extra bends and variations in the pipe-
cable coefficient of friction).
In thermomechanical behavior the clearance between one single cable and the pipe is found
to be the key controlling parameter, for both pipe and duct applications. The prime-mover is
the axial force generated by thermal expansion.
The knowledge gained now permits pipe and duct systems to be designed with an optimum
clearance that a) permits reductions in the magnitudes of axial thrust, sidewall forces and
bending curvature, whilst b) ensuring that the fatigue strain does not exceed the limits for the
metallic sheath of foil. In this respect pipe and duct systems now have a unique advantage
over other types of cable installation methods, e.g. buried direct in the ground and waved or
snaked in tunnels:
Pipe systems have a built-in design advantage, as the presence of three cables ensures
that the key clearance parameter always is large and the axial thrust always is low.
Typical magnitudes of axial force as low as 25-30% of the rigid-bar force are
obtained.
Duct systems do not have this advantage and have traditionally been installed with
small, tight clearances. However, even with the smallest clearance of 38mm in the
study, the force was reduced to 80% of the rigid-bar value. The opportunity exists to
reduce the axial force to 50% by a moderate increase in clearance to 50-75mm.
EPRI Licensed Material

Conclusions
14-3
The effects of thermomechanical, gravitational, short circuit and traffic induced vibration
forces are understood and quantified. Each study is a world first in cable engineering. The
effect of traffic induced vibration has been named the Avalanche effect.
Thermomechanical patterns are initiated by the changes in cable curvature that occur at
bends in the route. It has been shown that a) bend angles as small as 5
o
and b) lateral
displacements as small as 20mm (kinks), are capable of initiating wave patterns.
The thermomechanical effects that occur when a circuit is heated are:
Cable first moves into the available space within a bend.
Axial force rises at the maximum rate as though the cable is a rigid-bar
Incipient half waves slowly form at the entrance to bends
At a minimum energy condition, large amplitude waves suddenly form and cause the
axial force either to fall, or the rate of rise to diminish. The clipped value of force is
named the turn-over force and occurs at the turn-over temperature.
Cylindrical sinusoidal (sine) and helical loop waves occur in duct systems, the
percentage depends upon the minimizing energy condition. The number of helices
increases with a) decreasing clearance, b) decreasing cable weight and c) increasing
temperature.
Only cylindrical sinusoidal waves can form in pipe systems.
The longitudinal distribution of axial force is controlled by frictional and gravitational
forces.
Longitudinal slip of the cable occurs when the thermomechanical force differential
exceeds the cable frictional constraint force.
A second turn-over force and turn-over temperature event occurs at the end of the
straight section of the span, when frictional slip into the approaching wave patterns
causes the axial force to be clipped and fall.
The gravitational influences on thermomechanical performance are:
Waves form at a lower temperature at a bend at the bottom of a hill. The locked-in
strain is released downhill more readily, but the length of the wave pattern is reduced.
Waves form at a higher temperature at a bend at the top of a hill, the locked-in strain
is not easily released in the uphill direction. However, the length of the wave pattern
is significantly increased and may extend over the complete section at lower
temperature.
The short circuit effect on thermomechanical performance is:
The electrodynamic forces between the three cables violently propel them to be
compressed against the outside of the pipe or duct group.
At the end of the short circuit, the cables spring back and are momentarily freed of
both the short circuit force and frictional force.
EPRI Licensed Material

Conclusions
14-4
The locked-in strain is instantly freed throughout the cable system and permits the
patterns to extend over the majority of the span length (studies were performed at the
operating temperature of 90
o
C).
The traffic-induced vibration effect on thermomechanical performance occurs when a heavy
vehicle drives along, or crosses part of, a pipe route containing cable with locked-in thermal
strain:
The vertical vibration momentarily lifts the affected cable off the pipe floor.
The locked-in strain is released and produces a sudden slip over a long length of cable
e.g. 50-200m.
If the vehicle drives on an uneven road surface parallel to the route, then the strain
can be completely released.
The performance has been analysed of three double spans of a 138kV 1500kcmil, 123mm
clearance, pipe system with in-line casings containing one-piece pre-molded joints. It has
been shown that:
The cable and joints within the casings start to deflect laterally under compressive
axial force and initiate wave patterns within the first 15
o
C temperature rise.
The joints experience a differential force and move longitudinally according to the
asymmetry of their position in relation to pipe bends and straight sections.
The mechanical performance of one particular size of one-piece molding with 27%
diameter assembly stretch has been analyzed. It has been shown that:
The molding has a longitudinal cable pull-through withstand of 2,000N before
unacceptable disturbance to the electrically stressed cable-joint interface occurs,
namely: a) longitudinal slip and b) transverse lift of the molding off the cable to form
a void. It is recommended that the pull-through withstand should not be relied upon in
the design of the manhole constraint system.
The molding and adjacent cable has the ability to withstand the 2m bending radius
that can occur following thermomechanical displacement within the casing.
The electrical performance of the molding is most sensitive to the degree of assembly
pre-stretch, with lower values of stretch being detrimental.
Three spans and four joint manholes have been analyzed in a 230kV 2500kcmil duct system
with a 38mm clearance. The manhole constraint system comprises a) an expansion cleat and
a flexible cable S-bend at one end and, at the other end, b) the downhill gravitational pull of
the cable and an identical S-bend. It was found that the expansion cleat allows a cyclic
150mm cable movement into the manhole, that the S-bend deflects outwards by 200mm and
that the axial compressive force is significantly reduced.
EPRI Licensed Material
15-1
15
FURTHER WORK
Detailed items of further work have been recorded at the end of each Chapter, the following are
the key items of work:
To develop further the proposed novel design techniques, with the objective of manipulating
and further optimizing the thermomechanical performance of duct and pipe systems.
To update the data-set used by the Normalized Span calculation method to be specific to duct
systems.
To transfer the Normalized Span calculation method for pipe and duct systems to run on a
standard personal computer.
To introduce internal friction and plastic behavior into the bending stiffness parameter of the
FEA model cable (primarily to better represent conductor behavior). This will permit a) cable
equilibrium conditions to be quantified after load cycling over a larger number of cycles and
b) global cable movements and changes in axial force and sheath strain to be quantified.
To measure parameters from a wider range of XLPE cable constructions and build up a data
bank to supplement the Design Guide.
To measure friction in a long pipe run after the lubricant has dispersed.
To develop a range of high grip strength clamp cleats specifically to hold XLPE cables in
manholes, with the objectives of a) accommodating variations in cable diameter and ovality,
b) accommodating thermal expansion and contraction through the range of ambient and
operating temperatures expected in North America and c) reducing risk of damage to foil
laminate sheath cables.
To verify the cable pull-through strength of a selection of one-piece joints using the
techniques recommended, with the objective of evaluating when anchor joints need to be
selected.
To verify the withstand strengths of the improved designs of manhole constraints.

2004 Electric Power Research Institute (EPRI), Inc. All rights
reserved. Electric Power Research Institute and EPRI are registered
service marks of the Electric Power Research Institute, Inc.
EPRI. ELECTRIFY THE WORLD is a service mark of the Electric
Power Research Institute, Inc.
Printed on recycled paper in the United States of America
1001849
Program:
Underground Transmission Systems
EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com
About EPRI
EPRI creates science and technology solutions for
the global energy and energy services industry.
U.S. electric utilities established the Electric Power
Research Institute in 1973 as a nonprofit research
consortium for the benefit of utility members, their
customers, and society. Now known simply as EPRI,
the company provides a wide range of innovative
products and services to more than 1000 energy-
related organizations in 40 countries. EPRIs
multidisciplinary team of scientists and engineers
draws on a worldwide network of technical and
business expertise to help solve todays toughest
energy and environmental problems.
EPRI. Electrify the World
Export Control Restrictions
Access to and use of EPRI Intellectual Property is granted with
the specific understanding and requirement that responsibility
for ensuring full compliance with all applicable U.S. and foreign
export laws and regulations is being undertaken by you and
your company. This includes an obligation to ensure that any
individual receiving access hereunder who is not a U.S. citizen
or permanent U.S. resident is permitted access under applica-
ble U.S. and foreign export laws and regulations. In the event
you are uncertain whether you or your company may lawfully
obtain access to this EPRI Intellectual Property, you acknowl-
edge that it is your obligation to consult with your company's
legal counsel to determine whether this access is lawful.
Although EPRI may make available on a case by case basis an
informal assessment of the applicable U.S. export classification
for specific EPRI Intellectual Property, you and your company
acknowledge that this assessment is solely for informational
purposes and not for reliance purposes. You and your compa-
ny acknowledge that it is still the obligation of you and your
company to make your own assessment of the applicable U.S.
export classification and ensure compliance accordingly. You
and your company understand and acknowledge your obliga-
tions to make a prompt report to EPRI and the appropriate
authorities regarding any access to or use of EPRI Intellectual
Property hereunder that may be in violation of applicable U.S.
or foreign export laws or regulations.
SINGLE USER LICENSE AGREEMENT
THIS IS A LEGALLY BINDING AGREEMENT BETWEEN YOU AND THE ELECTRIC POWER
RESEARCH INSTITUTE, INC. (EPRI). PLEASE READ IT CAREFULLY BEFORE REMOVING THE
WRAPPING MATERIAL.
BY OPENING THIS SEALED PACKAGE YOU ARE AGREEING TO THE TERMS OF THIS AGREEMENT. IF YOU DO
NOT AGREE TO THE TERMS OF THIS AGREEMENT, PROMPTLY RETURN THE UNOPENED PACKAGE TO EPRI
AND THE PURCHASE PRICE WILL BE REFUNDED.
1. GRANT OF LICENSE
EPRI grants you the nonexclusive and nontransferable right during the term of this agreement to use this package only for
your own benefit and the benefit of your organization.This means that the following may use this package: (I) your compa-
ny (at any site owned or operated by your company); (II) its subsidiaries or other related entities; and (III) a consultant to
your company or related entities, if the consultant has entered into a contract agreeing not to disclose the package outside
of its organization or to use the package for its own benefit or the benefit of any party other than your company.
This shrink-wrap license agreement is subordinate to the terms of the Master Utility License Agreement between most U.S.
EPRI member utilities and EPRI.Any EPRI member utility that does not have a Master Utility License Agreement may get one
on request.
2. COPYRIGHT
This package, including the information contained in it, is either licensed to EPRI or owned by EPRI and is protected by United
States and international copyright laws.You may not, without the prior written permission of EPRI, reproduce, translate or
modify this package, in any form, in whole or in part, or prepare any derivative work based on this package.
3. RESTRICTIONS
You may not rent, lease, license, disclose or give this package to any person or organization, or use the information contained
in this package, for the benefit of any third party or for any purpose other than as specified above unless such use is with
the prior written permission of EPRI.You agree to take all reasonable steps to prevent unauthorized disclosure or use of
this package. Except as specified above, this agreement does not grant you any right to patents, copyrights, trade secrets,
trade names, trademarks or any other intellectual property, rights or licenses in respect of this package.
4.TERM AND TERMINATION
This license and this agreement are effective until terminated.You may terminate them at any time by destroying this pack-
age. EPRI has the right to terminate the license and this agreement immediately if you fail to comply with any term or con-
dition of this agreement. Upon any termination you may destroy this package, but all obligations of nondisclosure will remain
in effect.
5. DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
NEITHER EPRI,ANY MEMBER OF EPRI,ANY COSPONSOR, NOR ANY PERSON OR ORGANIZATION ACTING ON
BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH RESPECT TO
THE USE OF ANY INFORMATION,APPARATUS, METHOD, PROCESS OR SIMILAR ITEM DISCLOSED IN THIS PACK-
AGE, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES
NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTYS INTELLECTU-
AL PROPERTY, OR (III) THAT THIS PACKAGE IS SUITABLE TO ANY PARTICULAR USERS CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING ANY CON-
SEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED OF THE POSSIBILITY
OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS PACKAGE OR ANY INFORMATION,
APPARATUS, METHOD, PROCESS OR SIMILAR ITEM DISCLOSED IN THIS PACKAGE.
6. EXPORT
The laws and regulations of the United States restrict the export and re-export of any portion of this package, and you agree
not to export or re-export this package or any related technical data in any form without the appropriate United States and
foreign government approvals.
7. CHOICE OF LAW
This agreement will be governed by the laws of the State of California as applied to transactions taking place entirely in
California between California residents.
8. INTEGRATION
You have read and understand this agreement, and acknowledge that it is the final, complete and exclusive agreement
between you and EPRI concerning its subject matter, superseding any prior related understanding or agreement. No waiv-
er, variation or different terms of this agreement will be enforceable against EPRI unless EPRI gives its prior written consent,
signed by an officer of EPRI.

Você também pode gostar