Você está na página 1de 8

786

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 2, APRIL 2014

Dynamic Thermal Modeling of


MV/LV Prefabricated Substations
M. Z. Degefa, Member, IEEE, R. J. Millar, Member, IEEE, M. Lehtonen, Member, IEEE, and P. Hyvnen

AbstractWith the expansion and infilling of urban areas, the


demand for electric power is driving the design and capacity of
distribution substations to their thermal limits. Distribution transformer substations are increasingly required to be compact, reliable, safe, and intelligent. To efficiently utilize city space and to support the intermittent load flows imposed by smart-grid features,
such as distributed generation, the transformers are expected to
operate close to or occasionally over their ratings, with stalled or
little air circulation inside the safety enclosure. Dynamic thermal
models with physically validated convection and radiation heattransfer components are essential for the real-time thermal rating
of substations. Natural convection via the air inside the cabin to
the outside ambient air plays the major role in cooling down a
transformer. In this study a scale model of a prefabricated substation is examined to draft a numerical solution which is based
on stack ventilation principles. A clear and expandable first principle approach is used to quantify heat transfer through ventilation
openings. Measurements from actual cabins and 3-D finite element
method simulations are used to validate the numerical model.
Index TermsHot-spot temperature, indoor substation, natural
ventilation, online monitoring, prefabricated substation, thermal
rating.

I. INTRODUCTION

HERMAL constraints under intermittent loading conditions are becoming more significant with the expansion
of distributed generation (DG) and demand-side management.
Knowing the true capacity of the grid in real time is essential,
especially if expensive network assets are to be used efficiently and safely. Exceeding the steady-state ratings of power
transformers may be necessary in open electricity markets for
economic reasons or simply to ensure a continuous energy
supply [1]. Besides, short-time peak overloads, without significantly decreasing their life expectancy, are very often requested
from distribution transformers installed in prefabricated substations [2]. The true capacity varies with the power system
dynamics as well as with varying environmental conditions.
Dynamic thermal ratings utilize these factors to optimize and
better manage the grids power transfer capacity in real time.

Manuscript received March 20, 2013; revised June 06, 2013; accepted July
25, 2013. Date of publication August 23, 2013; date of current version March
20, 2014. This work was supported by Aalto energy efficiency program through
the SAGA project. Paper no. TPWRD-00329-2013.
The authors are with the Department of Electrical Engineering, Aalto
University, Aalto FI-00076, Finland (e-mail: merkebu.degefa@aalto.fi;
john.millar@aalto.fi; matti.lehtonen@aalto.fi; petri.hyvonen@aalto.fi).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TPWRD.2013.2276941

Hence, dynamic top-oil temperature models should be accurate


enough to be considered for application in online monitoring
and a real-time rating system for power transformers.
The thermal-electric analogy model of a prefabricated
MV/LV substation proposed in this paper works in conjunction
with the top-oil thermal model presented in [3]. In this transformer model, Susa considers the changes in the oil viscosity
and winding losses with temperature. Although some complex
constant parameters are necessary in Susas model, it has been
shown that it yields good results [4]. The proposed model in
this paper considers thermal resistance and thermal capacitance
of the immediate environment of the distribution transformer
to be dynamically dependent on the transformer loading and
ambient temperature. It also includes the effects of geometry
and the orientation of components on the natural convection
and thermal radiation heat transfers.
The two dominant components of heat transfer are from the
transformer cooling surfaces to the surrounding air and by naturally streaming air through the inlet and outlet ventilation holes.
However, heat transfer by natural ventilation cannot be established based on theoretical considerations only [5], which is why
a one-third scale model of an actual cabin, together with FEM
simulations, was used to establish a valid numerical model.
In this study, a working dynamic thermal model for prefabricated secondary substations is developed. The most dominant
but difficult to solve heat-transfer mechanism, natural convection, is modeled by introducing the stack effect principle into
the problem [6]. The contribution of this paper lies specifically
in the quantification of the ventilation thermal resistance representing the natural convection heat transfer. The proposed
lumped thermal ladder network not only models the installation
more accurately by closely following the inlet and outlet vent air
temperature, it is also ready to incorporate the effects of wind
and solar radiation.
In Section II, the theory of lumped parameter thermal
modeling is reviewed to pave the way for the subsequently
proposed model. Inclusion of the cabin environment with the
existing top-oil distribution transformer model is discussed
in Section III. The core theory of stack ventilation is then
explained in Section IV. After the proposed model is presented
in Section V, validations with measurements are discussed in
Section VI. Section VII summarizes the main findings of this
study.
II. THERMAL-ELECTRIC ANALOGY
The thermal-electric analogy network is a widely used and
accepted way of modeling dynamic heat transfer from a heat

0885-8977 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

DEGEFA et al.: DYNAMIC THERMAL MODELING OF MV/LV PREFABRICATED SUBSTATIONS

Fig. 1.

RC thermal circuit.

source to the heat sink. Many heat-transfer problems can be reduced to an equivalent RC circuit, where the distributed parameters can be lumped to a suitable combination of thermal capacitances and resistances. In some cases, these elements must be
nonlinear, such as the thermal resistance shown in Fig. 1. Generally, most transformer top-oil or hot-spot thermal models go
through the following steps:
Step 1) Identify all heat-transfer components (i.e., convection, conduction, and radiation).
Step 2) Evaluate the thermal resistances and thermal capacitances of each medium between the heat source and
heat sink (Fig. 1).
Step 3) Produce a simplified RC network and relevant firstorder linear differential equation to solve the temperature at the required point.
(1)
, analogous to electric
The thermal resistance
resistance, obstructs the flow of heat
, which itself is
analogous to electric current in electric circuits
(2)
is the heat-transfer coefficient
where
and
is the area
.
The presence of thermal resistances and capacitances
in the network requires iterative numerical methods,
such as the Euler and Runge Kutta methods, to solve differential equations like (1). This approach is used in almost all
of the literature, for example, the IEEE guide for loading
mineral-oil-immersed transformers [3][5], [7], [8].
The heat-transfer coefficient in (2) depends on the physical
parameters of the medium which, in turn, might depend on temperature, hence giving nonlinearity to the thermal resistance.
The coefficient may include convection and radiation when the
medium is air.
For convection [9]

(3)

787

where
is the Nusselt number;
is the Prandtle number;
is the Grashof number; is the characteristic dimension, length,
width, or diameter
; is the gravitational constant
;
is the fluid thermal conductivity
;
is the fluid
density
; and are empirical constants; is the fluid
thermal expansion coefficient
;
is the specific heat
of the fluid
; is the dynamic fluid viscosity
;
is the fluid temperature gradient
; and
is the convection heat-transfer coefficient
.
The radiative heat transfer from a hot surface in air could
be surface_to_ambient and/or surface_to_surface radiation,
where the radiation heat-transfer coefficient
is calculated
as either:
a) Surface_To_Ambient: The ambient surrounding behaves
as a blackbody. This means that the emissivity and absorptivity are equal to 1, and zero reflectivity
(4)
is the surface emissivity and
is the
where
StefanBoltzmanns constant
.
b) Surface_To_Surface (from both ambient and surrounding
surfaces)
(5)
is the mutual irradiation, going to other boundwhere
aries;
is an ambient view factor whose value is equal
to the fraction of the field of view that is not covered by
other boundaries; and
is the assumed far-away temperature in the directions included in
The total heat-transfer coefficient for a heat source body in a
fluid medium
is given by
(6)
Although various combinations of empirical correlations and
modified solution mechanisms could be implemented, the general heat-transfer theories explained before are the basis of the
modeling in this paper.
III. INCLUSION OF CABIN ENVIRONMENT
A typical prefabricated substation (cabin) is comprised of an
MV/LV transformer, LV panel and MV switchboard, connections, and auxiliary equipment in an enclosure to supply the LV
energy from the MV system [2] (see Fig. 2). Due to the thermal
limits imposed by the enclosure, there is only a slight possibility
of exceeding the nominal rating of such transformers. Natural
convection plays the dominant role in the heat transfer, and is
mainly influenced by factors, such as the entry and exit grill
shapes, size, and position. The grills should be designed in such
a way as to reduce the slowing down of air circulation within
the substation or loss of air pressure differential [9]. Ventilation
openings are also arranged to prevent any undesired condensation on electrical equipment and inner wall surfaces.
Short-term peak loads can be handled by distribution transformers inside kiosk substations for economic reasons without
significantly decreasing their life expectancy [1], [2]. Hot-spot
temperatures considerably higher than 98 C can be carried for

788

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 2, APRIL 2014

the lower edge of the ventilation-outlet hole, and


is the
temperature rise of the outlet air at the outlet ventilation holes
with respect to the air at the inlet ventilation holes.
The thermal resistance by definition will be
(8)
The problem with these approaches is the unclear usage of
hydraulic resistance. Hydraulic resistance is an analogous term
in fluid flow to the electrical resistance in electrical circuits. The
hydraulic resistance in (9) is taken from the HagenPoiseuille
equation [12]
(9)

Fig. 2. Prefabricated distribution transformer cabin layout.

short periods of time without decreasing normal life expectancy,


if this is offset by extended operation below 98 C. However,
an algorithm for calculating the hot-spot temperature should be
dynamic and represent the dominant heat-transfer components
accurately. The two most important heat-transfer components
are:
1) natural convection and thermal radiation from the transformer surfaces to the surrounding inside cabin air.
2) natural convection by air flow through the inlet and outlet
ventilation holes [5].
We assume that the criteria to be fulfilled by transformer
thermal models integrated in an online monitoring system, such
as [1], include the overall thermal environment including, where
appropriate, the cabin. The criteria are:
1) the model should be as independent as possible from the
prefabricated substation, which means fewer design-dependent variables should have to be adapted;
2) the model should be sufficiently accurate;
3) it must be possible to retrofit the sensors, which deliver
the input values for the model, to transformers already installed in the substations.
There are nonlinear thermal models developed for MV/LV
prefabricated substations [5], [10] and [11]. The thermal-electrical analogy-based model in [11] considers the thermal resistance and thermal capacitance of the top oil to be dynamic and
dependent on the temperature variation. Also, the thermal resistance and thermal capacitances of the cabinet air ventilation
and of different parts of the cabinet (ceiling, walls, and door)
are considered. The aforementioned studies agree that the most
significant and, at the same time, difficult-to-represent parameter is the thermal resistance of the ventilation openings. These
studies mainly use the Hoppner formula, which defines the size
of the ventilation holes for certain heat power to be removed by
natural convection [5], [11]
(7)
where is the surface of the input, that is, outlet ventilation
holes (m ),
is the power transferred by natural ventilation
(W), is the hydraulic resistance of the air circulation,
is
the distance from the middle high point of the transformer to

is the pressure drop (Pa), is the length of pipe (in


where
meters), is the dynamic viscosity
,
is the volumetric flow rate (m /s), and is the radius (in meters).
The HagenPoiseuille equation, however, assumes laminar
flow of fluid through a pipe of constant circular cross-section
that is substantially longer than its diameter. This would seem
quite far from the geometry and nature of prefabricated distribution transformer cabins.
The thermal models for indoor transformer stations in [5] and
[11] claim good agreement with measurements. However, their
approach requires various empirical values in addition to the
already complicated nature of natural convection heat transfer.
Besides, they neglect the effect of solar radiation and external
wind. These approaches lack two important aspectsthe first
is the assumption of laminar isothermal air flow through the
inlet as well as the outlet vents and the second is the neglection
of the effect of grill types. We believe the natural ventilation
driven through ventilation holes is due to the stack effect, which
is also influenced by external wind. The stack effect approach
aids in better understanding the natural convection heat transfer
inside the cabin which, in turn, leads to greener design of the
cabins themselves. More important, it leads to a more accurate
hot-spot temperature estimation of such installations. The stack
effect approach will be discussed in Section IV.
Natural convection is a complex thermodynamic phenomenon involving laminar and turbulent air flows. The 3-D FEM
simulations show the presence of local air circulations at the
outlet vent openings in addition to the major air flow from the
downside inlet vent to the top-level outlet vents (see Fig. 3)
[13]. In addition, the air movement at the side-vent openings
and transformer front-side openings are significantly different,
requiring a closer look at the orientation of the transformer with
respect to the ventilation openings.
Transformer cabin ventilation openings are covered with
grills to avoid direct access to animals, moisture, or water, and
for safety reasons. However, the grill types and, consequently,
the vent openings size and shape vary from one substation to
another. This significantly affects the natural heat convection.
To investigate and solve the aforementioned two shortcomings of previous models, we conducted a series of experiments
on a transformer scale model, which is then compared with FEM
simulations and actual measurements to quantify a more reliable
thermal resistance for natural convection.

DEGEFA et al.: DYNAMIC THERMAL MODELING OF MV/LV PREFABRICATED SUBSTATIONS

789

Fig. 4. Potential for flow due to pressure variation with height.


Fig. 3. FEM simulation for the transformer cabin scale model for 150-W heat
supply from 14 ceramic resistors with a connected resistance of 31 . (a) Geometry of the scaled cabin model (see Fig. 5). (b) Transformer cabin cross section
along the length showing airflows and velocity. (c) Transformer cabin ventilation air flow with temperature ( C) (slice) and air velocity (arrow) [13].

In this study, we propose a more reliable and accurate


thermal model of indoor MV/LV transformer substations to be
used for dynamic thermal rating on a real-time basis. Besides,
the relationship between ventilation grill types and transformer
maximum rating is explained, which is important for cabin
manufacturers.
The top-oil thermal model presented by Susa et al. in [3],
[14] is extended to include the cabin environment while implementing the proposed model. Susas model considers the
changes of oil viscosity and winding losses with temperature.
The top-oil and hot-spot temperature formulation ready for the
implementation of the third-order RungeKutta method as well
as for forward Euler approximation is

(10)
In Susas model, besides the nameplate-rating information,
results of the heat-run test of the transformer conducted during
commissioning are among the input parameters.
IV. STACK EFFECT
Inside a cabin, after the air is heated at the surface of the transformer by natural convection, the hot air will rise to the roof. The
accumulation of hot air at the top creates positive pressure while
the abandoned floor area of the cabin experiences negative pressure (see Fig. 4). With the developed pressure difference and
the presence of inlet and outlet openings at the bottom and top
areas of the cabin, respectively, a sustained circulation of air will
occur. This phenomena is called the stack effect and continues

Fig. 5. Transformer scale model, where dimensions are scaled by one-third


from a typical MV/LV transformer substation. (a) The aluminum cabin and
(b) the model transformer filled with mineral oil.

to act as long as the inside air is hotter than the ambient environment. This form of natural ventilation is so effective that, for
instance, without the need to introduce forced ventilation, computational fluid dynamics (CFD) techniques applied to partially
ventilated cable troughs have shown an increase in the continuous rating of up to 31% [15].
The air streaming out of the outlet is quantified by the volumetric flow rate
and is given in (11) [6]

(11)
where
is the ventilation rate (m /s);
is a discharge coefficient;
is the free area of inlet opening
(m ), which is equal to the area of the outlet opening;
(m/s ) is acceleration due to gravity; is the vertical distance
between inlet and outlet midpoints (in meters);
is the temperature of indoor air (in Kelvin); and
is the temperature of
outdoor air (in Kelvin).

790

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 2, APRIL 2014

The heat loss due to the ventilating moving air is given [6]
(12)
is the specific heat
where is the heat loss (in kilowatts),
capacity air1.005 (kJ/kg C), and is the density of air1.2
(kg/m ).
Equation (11) consists of a discharge coefficient which directly incorporates the effect of the grill type and a covering
panel just behind the vent openings. The discharge coefficient
is a value describing the aerodynamic channeling of the airflow
on discharge. It represents the ratio between the actual air flows
compared with the theoretical airflow. The coefficient, ranging
from 0 to 0.65, can easily be attained experimentally and is
given in the manufacturers specification of the specific grill or
louver. The thermal resistance associated with ventilation openings will be as shown in (13), which is based on (11) and (12)
(13)
The temperature dependence of the air density and air viscosity could easily be incorporated in (11) to accurately account
for the pressure-driven ventilation and air humidity effects.
V. SCALED MODEL OF THE PREFABRICATED CABIN
Before crafting the overall thermal circuit, we closely studied
a scaled down model of the cabins as well as a 3-D FEM simulation. The model transformer has steel walls and wall-attached
ventilating fins, and was filled with mineral oil. The heating load
was supplied with 14 ceramic-walled hollow resistive rods with
a connected resistance of about 31 (see Fig. 5). Thermocouples were placed at different locations in the cabin, including
on the surfaces of the resistive rods and inside the model transformer top oil. The placement of the thermocouples was guided
by results from a 3-D FEM simulation shown in Fig. 3 and an
analytical heat-loss calculation of the parallel-connected resistive loads.
The first measurement was conducted by loading the resistors
with 150 W and measuring the incoming and outgoing air temperature. The measurements, shown in Fig. 6, reveal that the hot
air leaves the cabin at the very top of the upper windows, indicating that the remaining openings are less relevant. The measurements also show that after a certain limit of opening area,
further increasing the size of the ventilation opening does not
have an impact on the heat transfer. It is primarily the height
of the placement of the outlet opening with respect to the inlet
opening that will affect the top-oil temperature of the distribution transformer. This observation supports the logic behind the
stack effect.
The FEM simulation in Fig. 4 supports the fact that the hot air
requires only a small area at the very top of the outlet opening to
stream out, as explained before. Besides, the simulations show
locally circulating air at the nonventilated walls of the cabin.
The heat transfer from the transformer surface to the nonventilated cabin wall side, therefore, depends on surface-to-surface
radiation and convection at the inside cabin wall surface.
The thermal network shown in Fig. 7 is our proposal. The network considers that the surface-to-surface radiation between the

Fig. 6. Air temperature at different levels of the scaled down cabin model
loaded with a 150-W heat supply. (a) Right-side ventilation openings. (b) Frontside ventilation openings (see Fig. 5).

distribution transformer surface and the inner cabin wall surface


is negligible.
The thermal resistance for the heat transfer from oil to air
is given in (14), which is elaborated in the top-oil
thermal model of the ONAN transformer in [3].

(14)
where is the characteristic dimension, length, width, or diameter (in meters); is the gravitational constant (m/s ); is the oil
thermal conductivity (W/mK);
is the oil density (kg/m );
is the oil thermal expansion coefficient (1/K);
is the specific
heat of oil (J/kg K); is the oil viscosity
;
and
is the oil temperature gradient.
The empirical values for
and are given in Table III.
The thermal resistances
and
are computed
based on the natural convection Nusselt numbers for vertical
and horizontal planes facing either upwards or downwards and
either inside or outside, as shown in the Appendix. The thermal
resistance
is given in (13). The thermal ladder network
is implemented with an extended Matlab Simulink tool called

DEGEFA et al.: DYNAMIC THERMAL MODELING OF MV/LV PREFABRICATED SUBSTATIONS

791

Fig. 8. Step response of outgoing air temperature through the outlet ventilation
of a prefabricated distribution transformer cabin supplied with 3200 W followed
by 4780 W.
Fig. 7. Overall thermal ladder network for the cabin installation. (a) The
detailed network consisting of the major thermal resistances and capacitance.
(b) The simplified network.

Simscape, which is capable of authoring physical modeling


components [16].
VI. MEASUREMENT RESULTS AND DISCUSSION
The output of the proposed thermal ladder network is compared with measurements from a 1000-kVA ONAN distribution transformer installed inside a prefabricated cabin of dimension (2.5 1.525 2) m. During the transformer full load test
(28.5 A for our transformer), the transformer reached the critical
temperature after 5 h. This increment is due to the cabin environment, which causes the distribution transformer to perform
below its rated current. During the test, we decreased the total
full-load losses from 11600 to 9185 W on the 5th h and to 8500
W after 1 h, 45 min and, finally, to 7500 W after 2 h, as can be
seen in Fig. 9. For the 1000-kVA distribution transformer inside
the prefabricated cabin to stay under the upper thermal limit, it
had to be loaded significantly less than when installed in free
air.
The DormandPrince-based method solver ODE45 in Matlab
is used to solve the step response for the ladder network shown
in Fig. 10 [17]. The Simscape physical modeling enables the
modeling of physical property components by the users, while
enabling real-time dynamism for viscosity, density, air pressure,
etc.
Natural convection through the stack effect closely represents
the ventilation scenario of the prefabricated distribution transformer cabins. The proposed model estimated the top-oil temperature and temperature of hot air streaming out of the top ventilation openings very closely, as shown in Figs. 8 and 9.
Temperature differences and wind pressure differences are
the driving forces in natural ventilation. The requirement for
safety and compactness forces the ventilation inlet and outlet
openings to be single sided, meaning on the same wall. In addition, there are safety requirements to not allow openings that
permit a direct view of the transformer from the outside, which
makes temperature differences the dominant driving force for
natural ventilation. In this thermal model, we explained nat-

Fig. 9. Top-oil and transformer room air temperature comparisons for a fully
loaded 1000-kVA distribution transformer inside the cabin.

TABLE I
RAYLEIGH AND NUSSELT NUMBERS FOR HORIZONTAL PLATES [9]

TABLE II
RAYLEIGH AND NUSSELT NUMBERS FOR VERTICAL PLATES [9]

ural convection driven by temperature differences. However,


wind-driven ventilation on prefabricated substations should
be studied, especially for those cabins with a cross-ventilation
setup and significant openings in the direction of wind flow.
VII. CONCLUSION
A first principles approach is used to model the thermal environment of prefabricated distribution substations. Stack ventila-

792

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 29, NO. 2, APRIL 2014

2) Vertical Plates:
Characteristic length
3) Vertical Wall With Ventilating Fins: The Rayleigh number
is expressed as
TABLE III
EMPIRICAL VALUES FOR CONSTANTS

AND

[3]

ACKNOWLEDGMENT
The authors would like to thank Dejan Susa (Ph.D.) for explaining the transformer thermal model and to N. Tatu for assistance in the experimental setup.
REFERENCES

Fig. 10. Simscape physical model.

tion is considered to be the driving force of natural ventilation.


Consequently, a clearer thermal model with the open possibility
for expansion to consider external factors, such as wind and
solar irradiation, is proposed. The proposed model shows good
agreement with measurements in a controlled environment. To
be a fully deployable thermal model, the external factors mentioned before should be included, which the authors of this paper
are currently working on.
APPENDIX A
NUSSELT NUMBERS [9]
1) Horizontal Plates:
Surface area perimeter

Characteristic length

[1] S. Tenbohlen, T. Stirl, and M. Stach, Assessment of overload capacity


of power transformers by on-line monitoring systems, presented at the
IEEE Power Eng. Soc. Winter Meeting, Columbus, OH, USA, 2001.
[2] A. Kalam and S. Corhodzic, Performances of distribution transformers installed in metallic enclosuresAn Australian experience,
IEEE Trans. Power Del., vol. 20, no. 3, pp. 19701975, Jul. 2005.
[3] D. Susa, Dynamic thermal modeling of power transformer, Ph.D. dissertation, Dept. Elect. Commun. Eng., Helsinki Univ. Technol., Espoo,
Finland, 2005.
[4] R. Vilaithong, S. Tenbohlen, and T. Stirl, Investigation of different
top-oil temperature models for on-line monitoring system of power
transformer, presented at the IEEE Int. Conf. Condition Monitor. and
Diagnosis, Changwong, Korea, 2006.
[5] Z. Radakovic and S. Maksimovic, Non-stationary thermal model of
indoor transformer stations, Elect. Eng., vol. 84, pp. 109117, 2002.
[6] T. S. Larsen, Natural Ventilation Driven by Wind and Temperature
Difference, Ph.D. Thesis, Department of Civil Engineering, Aalborg
University, Aalborg, Denmark, June 2006.
[7] IEEE Guide for Loading Mineral-Oil-Immersed Transformers, IEEE
Standard C57.91, 1995.
[8] M. Z. Degefa, M. Lehtonen, and R. J. Millar, Comparison of air-gap
thermal models for MV power cables inside unfilled conduit, IEEE
Trans. Power Del., vol. 27, no. 3, pp. 16621669, Jul. 2012.
[9] F. P. Incropera and D. P. DeWitt, Fundamentals of Heat and Mass
Transfer. Hoboken, NJ: Wiley, 2006.
[10] P. Lepretre and D. Jacquier, Thermal mastering of MV/LV prefabricated substation, presented at the 20th Int. Conf. Elect. Distrib.,
Prague, Czech Republic, Jun. 811, 2009.
[11] I. Iskender and A. Mamizadeh, An improved nonlinear thermal
model for MV/LV prefabricated oil-immersed power transformer
substations, Elect. Eng., vol. 93, no. 1, Nov. 21, 2011.
[12] N. A. Mortensen, F. Okkels, and H. Bruus, Reexamination of HagenPoiseuille flow: Shape dependence of the hydraulic resistance in microchannels, Phys. Rev. E, vol. 71, no. 5, 2005.
[13] COMSOL, Heat transfer module. 2013. [Online]. Available: http://
www.comsol.fi
[14] D. Susa, M. Lehtonen, and H. Nordman, Dynamic thermal modelling
of power transformers, IEEE Trans. Power Del., vol. 20, no. 1, pp.
197204, Jan. 2005.
[15] J. A. Pilgrim, D. J. Swaffield, P. L. Lewin, and D. Payne, Rating
methods for cables installed in unventilated and partially ventilated surface troughs, presented at the Elect. Insul. Conf., Annapolis, MD, Jun.
58, 2011.
[16] Simscape Users Guide, Mathworks, 2013. [Online]. Available:
http://www.mathworks.se/help/physmod/simscape/
[17] Mathworks,, Ordinary Differential Equations. 2013. [Online].
Available: https://www.mathworks.se/help/matlab/ordinary-differential-equations.html

DEGEFA et al.: DYNAMIC THERMAL MODELING OF MV/LV PREFABRICATED SUBSTATIONS

M. Z. Degefa (M09) was born in 1984. He received


the B.Sc. degree in electrical engineering from
Mekelle University, Mekelle, Ethiopia, in 2007
and the M.Sc. degree in electronics and electrical
engineering (Hons.) from Aalto University, Aalto,
Finland, in 2010, where he is currently pursuing the
Ph.D. degree in power systems and high voltage
engineering.
His main research interests are power system
state estimation, dynamic thermal rating, and load
modeling.

R. J. Millar (M12) was born in 1963. He received


the B.Eng. degree in mechanical engineering
from the University of Auckland, Auckland, New
Zealand, in 1984, and the M.Sc. and D.Sc. (Tech)
degrees in electrical engineering from the Helsinki
University of Technology, Espoo, Finland, in 2002
and 2006, respectively.
Currently, he is a Researcher and Lecturer with
Aalto University , which is now part of Helsinki
University of Technology. His main research
interests are distribution network planning and
underground cable rating.

793

M. Lehtonen (M11) received the M.S. and Licentiate degrees in electrical engineering from
Aalto University (formerly Helsinki University of
Technology), Espoo, Finland, in 1984 and 1989,
respectively, and the D.Sc. degree from the Tampere
University of Technology, Tampere, Finland, in
1992.
Since 1987, he has been with VTT Energy, Espoo,
and since 1999, he has been with the Department of
Electrical Engineering, Aalto University, where he is
a Professor of IT applications in power systems. His
main activities include earth fault problems, and harmonic-related issues and
applications of information technology in distribution automation and distribution energy management.

P. Hyvnen was born in 1969. He received the


M.S. degree in electrical engineering from Tampere
University of Technology, Tampere, Finland, in
1997 and the Licentiate and D.Tech. degrees from
Helsinki University of Technology, Espoo, Finland,
in 2003 and 2008, respectively.
Currently, he is a Senior Researcher at the High
Voltage Laboratory, Helsinki University of Technology, where he has been since 2000 and Head of
the High Voltage Laboratory since 2008. Since 2012,
he has been a Visiting Professor in the Department
of Electrical Power Engineering, Tallinn University of Technology, Tallinn,
Estonia. His research interests include modern condition monitoring techniques
of high-voltage insulation, and the testing of high-voltage (HV) apparatuses
and HV measuring techniques.

Você também pode gostar