Você está na página 1de 33

6.

09
Erosion/Corrosion
R. J. K. WOOD
University of Southampton, Southampton, UK
6.09.1 INTRODUCTION AND DEFINITIONS 397
6.09.1.1 High-Temperature ErosionCorrosion 398
6.09.1.2 General Models 399
6.09.2 EROSIONCORROSION FUNDAMENTALS 399
6.09.2.1 Factors Affecting ErosionCorrosion 399
6.09.2.1.1 Material parameters 399
6.09.2.1.2 Coatings 401
6.09.2.1.3 Environment factors 401
6.09.2.1.4 Geometry factors 403
6.09.2.1.5 Galvanic interactions 405
6.09.2.2 Electrochemical Reaction Kinetics 406
6.09.2.2.1 Kinetics under polarization 406
6.09.2.2.2 Mass-transfer coefficient 408
6.09.2.3 Fluid Flow Fields 408
6.09.2.3.1 Turbulence 408
6.09.2.3.2 Wall shear stresses 410
6.09.2.4 Erosion Fundamentals 410
6.09.2.4.1 Solid-particle erosion 410
6.09.3 EROSIONCORROSION MECHANISMS 412
6.09.3.1 Flow-Accelerated Corrosion 412
6.09.3.1.1 KouteckyLevich equation 412
6.09.3.2 ParticleSurface Interactions 413
6.09.3.2.1 Depassivation and repassivation kinetics 413
6.09.3.3 Synergy 415
6.09.3.3.1 Models and mapping 415
6.09.4 EROSIONCORROSION TESTING 419
6.09.4.1 Gravimetric and Coupon Techniques 419
6.09.4.2 Rotating Systems 420
6.09.4.3 Flow Systems 421
6.09.5 EROSIONCORROSION DETECTION 421
6.09.5.1 Advanced Electrochemical Monitoring 421
6.09.5.1.1 Electrochemical noise 421
6.09.6 EROSIONCORROSION-RESISTANT MATERIAL AND SURFACE SELECTION 422
6.09.7 CONCLUSIONS 423
6.09.8 REFERENCES 425
NOMENCLATURE
a scaling constant
A real contact area (m
2
)
A
a
affected area (m
2
)
a
j
jump distance (m)
A
nom
nominal contact area (m
2
)
395
B migration kinetic constant (czFa
j
/
RT)
b
a
anodic Tafel slope (mV decade
1
)
b
c
cathodic Tafel slope (mV decade
1
)
C pure corrosion (mg or mg s
1
)
C9 static corrosion rate (mg or mg s
1
)
C0 corrosion rate under erosioncorro-
sion (mg or mg s
1
)
C
0
concentration of species at electrode
surface (M, mM, mol cm
3
)
C
1
constant
C

concentration of soluble ferrous


ions in the bulk solution (M, mM,
mol cm
3
)
C
b
bulk concentration of species (M,
mM, mol cm
3
)
C
eq
concentration of soluble ferrous
ions at equilibrium (M, mM,
mol cm
3
)
C
i
concentration of species i (M, mM,
mol cm
3
)
C
i
bulk
bulk concentration of species i (M,
mM, mol cm
3
)
C
k
cutting characteristic velocity
(m s
1
)
C
s
system constant
C
v
solids volume fraction
D diffusion coefficient (cm
2
s
1
)
D
b
bend bore diameter (m)
D
i
diffusion coefficient of species i
(cm
2
s
1
)
E
applied
applied potential (V)
E pure erosion (mg or mg s
1
)
E
corr
freely corroding potential (open cir-
cuit potential) (V)
E
eq
equilibrium potential (V)
E
k
kinetic energy of impacting particle
(J)
E
m
Youngs modulus (Pa)
E
r
Erosion rate (mg kg
1
)
F Faraday constant (C mol
1
)
f
a
stripping coefficient
F
y
characteristic function
H hardness (Vickers number)
H
p
particle hardness (Pa)
H
t
target hardness (Pa)
I anodic or cathodic current (A)
I
0
exchange current (A)
i
a
activation-controlled current den-
sity (A m
2
)
I
a
anodic current (A)
i
aff
affected current density (A m
2
)
i
c
cathodic current density (A m
2
)
I
c
cathodic current (A)
i
corr
corrosion current density (A m
2
)
i
d
diffusion-controlled current density
(A m
2
)
i
L
limiting corrosion current density
(A m
2
)
I
lim
limiting current (A)
I
reaction
reaction current (A)
i
u
unaffected current density (A m
2
)
J
i
flux of species i (mol s
1
cm
2
)
K constant
k
1
forward reaction rate
K
1
constant
k
1
backward reaction rate
k
m
mass-transport coefficient (cm s
1
)
k
a
anodic rate constant
k
B
rate constant
L characteristic length (m)
L
ox
oxide layer thickness (m)
m mass of oxide removal (kg)
M
m
equivalent molar mass
M
ox
mass of oxide (kg)
_
M
p
mass of sand impacting the surface
per unit time (kg s
1
)
n velocity ratio exponent or bulk con-
centration exponent
N number of erodent impacts
N
o,u
total number of surface sites
available
N
u
number of moles of oxide formed
q charge passed =
_
I dt
R universal molar gas con-
stant = 8.31451 (J K
1
mol
1
)
R
a
surface roughness (mm)
R
c
bend radius (m)
R
f
roundness factor for particle (value
01)
r
p
particle radius (m)
Re Reynolds number
S synergy (mg or mg s
1
)
Sc Schmidt number
Sh Sherwood number
T total wear or erosion loss (mg or
mg s
1
)
T
y
temperature (

C)
U free stream velocity (m s
1
)
U
cr
critical velocity for plastic deforma-
tion (m s
1
)
U
f
velocity component normal to sur-
face (m s
1
)
U
p
particle velocity (m s
1
)
v* friction velocity (=
_
(t
w
/r))
(m s
1
)
V
cr
critical particle impact velocity for
plastic flow (m s
1
)
v
s
volume fraction of abrasives
V
u
erosion rate (mm
3
impact
1
)
W erosion mass loss per unit time
(mg s
1
)
y vertical height (m)
y

dimensionless scale
z number of electrons
z
i
valence
a angle of impact (

)
a
1
symmetry constant
b charge-transfer barrier coefficient
d boundary layer thickness (m)
396 Erosion/Corrosion
DC change in concentration of species
(=(C
b
C
0
)) (M, mM, mol cm
3
)
DC
i
change in concentration of species i
(M, mM, mol cm
3
)
DC
e
enhancement of C due to the pre-
sence of erosion (mg or mg s
1
)
DC9
e
enhancement of C9 due to flow and
erosion (mg or mg s
1
)
DE
c
synergistic effect or enhanced ero-
sion rate due to corrosion (mg or
mg s
1
)
DH change in hardness (Hv or Pa)
Dw mass loss (mg)
Dx change in distance in x direction (m)
Df potential difference across oxide (V)
e
c
critical plastic strain (mm m
1
)
Z turbulent eddy viscosity (Pa s)
Z
c
cathodic overpotential (mV)
Z
a
anodic overpotential (mV)
Z
act
activation overpotential
(mV) = E
applied
E
eq
y surface coverage fraction
l particle flux (impact m
2
s
1
)
l
i
molar conductivity (S m
2
mol
1
)
r
f
density of fluid (kg m
3
)
r
ox
density of oxide (kg m
3
)
r
p
density of particle (kg m
3
)
r
t
density of target material (kg m
3
)
o dynamic plastic flow stress for target
(N m
2
)
o
C
standard deviation of flow corrosion
current (A)
o
T
standard deviation of erosioncor-
rosion current (A)
t passive film recovery time (s)
t
w
wall shear stress (N m
2
)
i kinematic viscosity (m
2
s
1
)
m dynamic viscosity (Pa s)
ABBREVIATIONS
AISI American Iron and Steel Institute
ASTM American Society for Testing and
Materials
CE counter electrode
CFD computation fluid dynamics
CRA corrosion-resistant alloy
CVD chemical vapor deposition
ECN electrochemical current noise
ENA electrochemical noise analysis
EPN electrochemical potential noise
EPS extracellular polymeric substances
FBC fluidized bed combustors
HVOF high-velocity oxy-fuel
KE kinetic energy
MDPE medium-density polyethylene
NAB nickel aluminum bronze
NACE National Association of Corrosion
Engineers
OCP open circuit potential
PEO plasma electrolytic oxidation
RCE rotating cylinder electrodes
RDE rotating disc electrodes
RE reference electrode
SCE saturated calomel electrode
UNS Unified National System
WC tungsten carbide
6.09.1 INTRODUCTION AND
DEFINITIONS
The derivations of both erosion and corro-
sion have similar origins in the Latin verb rodere
and corrodere which mean to gnaw away and
intensely gnaw away, respectively (Oxford
English Dictionary Online, 2006). The term ero-
sion is used in a diverse number of technologies
ranging from geological loss of rock, soil, and
beach to loss of wall thickness in engineering
systems. Corrosion can be defined as the chemi-
cal or electrochemical reaction between a
material, typically electrically conducting (i.e.,
metallic), and its environment which produces
material deterioration and degradation of mate-
rial properties. Corrosion has an equally wide
use, covering electrochemical dissolution of
metal surfaces to staining of metallic architec-
tural fittings on buildings to the degradation of
plastics. In the context of this chapter, the term
relates to surface degradation involving electro-
chemical processes only.
Erosion is a process by which discrete small
solid particles, with inertia, strike the surface of
a material causing damage or material loss to its
surface. Erosion can be a problem for compo-
nents such as turbine blades, propulsors,
pipelines, and fluidized bed combustion sys-
tems. Erosion does have its beneficial
applications, notably for cleaning and prepara-
tion of surfaces for subsequent coating/painting
by grit-blasting, descaling/dewaxing produc-
tion risers in the oil and gas industries, and
cutting through rocks or subsea steel oil struc-
tures using abrasive water jets.
Erosioncorrosion is a combination of erosion
and corrosion processes and has been used to
describe component wall dissolution rates under
flow-assisted corrosion as well as wall loss rates
under solid-particle erosion and corrosion or
cavitation erosion and corrosion conditions.
Within this chapter, the term erosioncorrosion
is restricted to describe conditions where both
mechanical erosion and electrochemical corro-
sion processes are acting together and where
Introduction and Definitions 397
mechanically induced material loss is equal to or
about 10 times greater than the corrosion-
induced material loss. The term corrosion
erosion relates to the opposite condition, where
corrosion-induced material loss is equal to or
about 10 times greater than the erosion-induced
material loss. The chapter uses flow-assisted cor-
rosion to describe enhanced corrosion where
mechanical processes are not involved such as
single-phase flows relative to component wall
surfaces. However, to fully discuss erosioncorro-
sion, the chapter initially sets out to detail some of
the current understanding of flow-assisted or
enhanced corrosion prior to introducing the con-
sequence of adding mechanical erosive processes
on material removal rates. It should be noted that
previous literature has sometimes used a much
more liberal definition of erosioncorrosion.
The need to minimize costs associated with
modern fluid handling and propulsion equip-
ment demands increasing flow rates with the
inherent risk of flow-dependent corrosion and,
if solids are entrained or cavitation occurs, ero-
sioncorrosion. This is especially true for
industries that transport slurries and other par-
ticle-laden fluids in pipes or seawater
propulsion systems such as offshore and marine
technologies. These industries expend the
equivalent of millions of pounds every year on
maintenance costs and on costs associated with
loss of productivity caused by solid-particle
impingement and cavitation erosioncorrosion
damage. Typical examples of this kind of mate-
rial destruction are erosioncorrosion damage
to pump impellers (see Figure 1), propellers,
valves, heat exchanger tubes, pneumatic and
hydrotransport systems, fluidized bed combus-
tors (FBCs), and many other types of fluid-
handling equipment. In a recent survey,
erosioncorrosion was rated in the top five
most prevalent forms of corrosion damage in
the oil and gas industry (McIntyre, 1999). The
erosioncorrosion data and models published
in the open literature aspire to allow informed
surface selection for combined erosion and cor-
rosion resistance. However, this information is
very patchy and incomplete, therefore badly
hampering any improvement in the perfor-
mance of machines in aggressive environments.
This chapter mainly focuses on aqueous-based
examples of erosioncorrosion. An excellent
reviewof solid-particle erosioncorrosion caused
by gas streams at roomtemperature and elevated
temperatures can be found in Levy (1995). A
brief overview of high-gas-temperature solid-
particle erosioncorrosion is given below to
introduce some of the complexities of erosion
corrosion and to highlight some important mate-
rial properties required for resisting erosion
corrosion. The main surface characteristic to
resist erosioncorrosion conditions, not for
high-temperature gassolid flow systems only,
is the ability to form a protection surface layer
or passive film. These surface films can restrict
charge transfer across liquid/solid interfaces,
reduce the contact stresses induced by low-
energy mechanical impacts by erosive processes,
and reform (repassivate or oxidize) after being
damaged to reduce corrosion loss rates.
6.09.1.1 High-Temperature
ErosionCorrosion
High-temperature erosioncorrosion can be
induced in such cases as heat exchanger tubing
in both FBCs and pulverized coal-fired burners,
where surfaces are exposed to impingements by
hot oxidizing gas-particle mixtures. The perfor-
mances of chromium steels under such
environments rely on the formation of a surface
oxide layer. Such oxides are sensitive to the
silicon content of the parent steel, with increas-
ing levels of Si resulting in retardation of oxide
formation under the transition temperature
(,450

C). For temperatures above transition,
a tenacious and ductile oxide scale is formed,
which protects the parent material from ero-
sioncorrosion under low solid-particle
impingement angles (30

) and impact energies


(particle kinetic energy ,34 nJ).
Generally for steels, various material proper-
ties are important in controlling erosion
corrosion induced by gassolid and liquidsolid
flows. Ductility is thought to be a controlling
parameter as it determines the contact stresses
induced by solid-particle impingements and
thus reduces the erosivity. Thermal conductivity
is also thought to influence contact heating when
solid particles impact steel surfaces. Calculations
show that temperatures close to annealing values
can be generated with significant consequences
for the near-surface oxide formation, microstruc-
ture, and mechanical properties of eroded steels.
The strain rate on impact is thought to affect the
ability of cracks to heal within oxide scales. The
Figure 1 Erosioncorrosion of a 150 mm diameter
offshore downhole centrifugal pump impeller.
398 Erosion/Corrosion
presence of cracks is important for the integrity
of the scale and its ability to protect the parent
steel from erosion and further oxidation.
6.09.1.2 General Models
Erosion and erosioncorrosion models are
very system specific, typically being related to
a unique combination of flow field and mate-
rial, and therefore no real generic models are
available. The published erosioncorrosion
data are sporadic in diverse applications/areas
and hamper robust guidance for surface selec-
tion. Therefore, this chapter points to what is
currently understood and aims to discuss issues
related to structural integrity.
6.09.2 EROSIONCORROSION
FUNDAMENTALS
6.09.2.1 Factors Affecting ErosionCorrosion
6.09.2.1.1 Material parameters
A review of available solid-particle erosion
models shows the diversity of material proper-
ties that have been included in published
predictive erosion equations. Meng and
Ludema (1995) have reviewed 28 erosion mod-
els and found that, on average, only five
parameters are used per equation, but in total
33 different parameters, not all material proper-
ties, are quoted. These are tabulated in Table 1.
It is tempting to assume that sand-erosion rate
would be inversely dependent on the surface
hardness, H, as predicted by simple plastic defor-
mation erosion models, eqn [1] (Hutchings,
1992). However, this relationship rarely holds in
practice for engineering materials:
E
r
=
Kr
T
U
p
n
H
f (y) [1[
The erosion rates, expressed as volume loss
per impact (assuming all erodent particles
impinge and damage the target surface), V
u
,
obtained by testing at the University of
Southampton are plotted against Vickers hard-
ness values in Figure 2. Numerous material
surfaces were tested under watersand slurry
erosion conditions of 16.5 m s
1
free jet impin-
gement at 90

jet impingement angle with


Table 1 Parameters selected in erosion wear models (Meng and Ludema, 1995)
Erodent Target Fluid flow
Density Density Impact angle
Hardness Hardness Impact angle max. wear
Moment of inertia Flow stress KE transfer from particle to target
Roundness Youngs modulus Temperature
Single mass Fracture toughness
Size Critical plastic strain
Velocity Depth of deformation
Rebound velocity Incremental strain per impact
Kinetic energy (KE) of particle Thermal conductivity
Melting temperature
Enthalpy of melting
Cutting energy
Deformation energy
Erosion resistance
Heat capacity
Grain molecular weight
Weibull flaw parameter
Lame constant
Grain diameter
y= 5E + 08x
3.2201
y =3040.9x
1.597
R
2
= 0.6767
R
2
= 0.9084
0.000 01
0.000 1
0.001
0.01
0.1
1
100 1000 10 000
Surface Vickers hardness (Hv)
Ductile surfaces
Brittle surfaces
E
r
o
s
i
o
n

r
a
t
e
(

m
3
/
i
m
p
a
c
t
)
Figure 2 Erosion rates for a wide range of
engineering materials and coatings under water
sand slurry erosion conditions of 16.5 m s
1
free jet
impingement at 90

jet impingement angle with


135 mm sand at 2.1% w/w concentration over 5 h
test duration. From Wood, R. J. K. 2004a.
Erosioncorrosion interactions and their effects on
marine and offshore components. Keynote at
EuroCorr Conference, Nice, France and accepted
for publication in the Special Issue on
Tribocorrosion, Wear 2006.
ErosionCorrosion Fundamentals 399
135 mm sand at 2.1% w/w concentration over
5 h test duration. As these results are for nor-
mal impingement erosion, the failures of the
ductile surfaces should be related to accumu-
lated plastic strain or low-cycle fatigue induced
by cyclic plastic deformations from successive
impacts. This would lead to the erosion rate E
r
,
expressed in mass of material removed per mass
of impacting particles, being proportional to
H
1.5
, as given in eqn [2]. For tests where the
erodent size and concentration are held con-
stant, the erosion rate E
r
and V
u
are
equivalent. In eqns [1] and [2], K and K
1
are
constants, r
T
is the density of the material being
eroded, r
P
is the density of the erosive particle,
U
P
is the particle velocity, f(y) is the angle of
impact dependency function, and e
c
is the criti-
cal plastic strain at which detachment of wear
debris occurs. Ductile surfaces show a V
u
_
H
1.6
trend, which is nearly as predicted by
eqn [2]. The erosion dependency on hardness
for brittle surfaces is more complex with the-
ories predicting between V
u
_ H
0.25
and V
u
_
H
0.1
(Hutchings, 1992). Figure 2 shows that the
erosion rates of brittle engineering surfaces
have a much increased sensitivity to hardness
with erosion rate proportional to H
3.2
. This
trend, it should be noted, only has a correlation
function R
2
value of 0.67 and is heavily skewed
by chemical-vapor deposition (CVD) diamond
that has a hardness of 80 GPa (,8000 Hv):
E
r
=
K
1
r
T
r
p
1,2
U
p
3
e
c
2
H
3,2
[2[
For cases where the surface hardness
equals the erodent hardness (sand
hardness = 1100 Hv), considerable scatter in
the data is seen. The conclusion from this is
that the selection of surfaces with increased sur-
face hardness may result in increased erosion
resistance. However, the erosion mechanisms
may change from ductile to brittle ones and
extreme care is required when selecting surfaces
of around 1100 Hv to resist sand erosion. Such
systems result in highly uncertain erosion and
therefore erosioncorrosion rates. This intro-
duces another important ratio, that of relative
hardness between the erodent and target surface.
As discussed above, the hardness of the par-
ticle H
p
involved in erosion can have a major
influence on the erosion rate. If the particle
hardness H
p
is lower than that of the compo-
nent surface hardness H
t
(i.e., H
p
/H
t
< 1),
much less wear will result compared to harder
particles where H
p
/H
t
> 1. For cases where H
p
is much greater than H
t
, the exact value of H
p
matters far less as the rate of wear is relatively
insensitive to hardness ratios H
p
/H
t
> 10 (see
Hutchings, 1992). Simple contact mechanics
shows that high levels of wear can be expected
when particles are 1.25 times harder than the
surface. The hardness of sand, a typical abra-
sive particle, is about 10001200 Hv and is
significantly higher in hardness than most steel
alloys, thereby indicating that steels are vulner-
able to sand erosion. Even surfaces containing
hard phases (carbides), such as thermally
sprayed cermets, will have softer binder phases
which are susceptible to abrasion. Ceramics, on
the other hand, offer hard and homogeneous
surfaces to combat plastic deformation and
inhibit erosive wear. Unfortunately, they suffer
brittle fracture failures under high stresses due
to their inherently low fracture toughness.
Similar to reducing sliding wear, perhaps
consideration should be given to controlling
erosion by selecting surfaces with high H/E
m
(hardness/Youngs modulus) ratios. Such sur-
faces can absorb high levels of energy elastically
as well as have high resilience and toughness
(Matthews, 2005), although replotting the data
from Figure 2 as erosion rate versus H/E
m
shows little correlation.
Other research has pointed to similar trends of
erosion rate with hardness. For example, the
wear environment of steels used for containing,
transporting, and processing erosive mineral
slurries often encounter fluid-borne particle
layers that move at high speed across the wear-
ing surface (Wood et al., 2004; Wood and Jones,
2003). Information on the performance ranking
of such materials is limited, particularly with
respect to the influence of steel hardness and
microstructure on the resistance to erosion.
This is particularly important for the oil sands
industry of Northern Alberta, Canada, where
handling and processing of essentially silica-
based solids (tar sands) results in extremely
severe wear conditions. Clark and Llewellyn
(2001) present slurry erosion data obtained on
commercially available wear-resistant pipeline
steels with hardness varying from 220 to
740 Hv. They conclude that the erosion resis-
tance of a hard internal surface of AISI 1045
steel pipe (740 Hv) was more than 5 times
greater than standard, nonhardened, AISI 1020
carbon steel (CS) pipeline material (220 Hv).
However, they note that the best erosion perfor-
mance was obtained for steel with homogeneous
microstructure with a hardness approaching that
of the silica sand erosion medium, namely
H
p
= 1000 Hv and H
p
/H
t
= 0.75.
Corrosion-resistant alloys (CRAs) are widely
used to resist flow corrosion. For example,
chromium is added to improve the resistance
to flow-induced corrosion to CS. The effect of
the different chromium addition levels on per-
formance is reviewed by Poulson (1999).
400 Erosion/Corrosion
Figure 14 shows the trend for such a passive
system as a function of flow velocity and its
vulnerability to mechanical processes that
accelerate corrosion loss rates (passive means
the surface has a propensity to form a protec-
tive surface film or layer with high impedance
and thus suppresses charge transfer at the
metalliquid interface). In general, corrosion-
resistant alloys do not resist erosion well due
to their low hardness, and the interactions
(synergy and additive effects) that exist between
corrosion and erosion are not well understood.
This presents a dilemma for those seeking ero-
sioncorrosion-resistant surface selections.
Also, cost-reduction considerations favor
replacing expensive solid alloy components
with coatings on CS substrates, hence consider-
able effort is now being made to research
coating performance under erosioncorrosion.
6.09.2.1.2 Coatings
The relatively cheap option of organic coat-
ings tend to perform relatively poorly under
high-energy solid-particle impingement. They
can be used as corrosion-barrier coatings within
low-energy flow components (Puget, 1999),
whereas ceramic materials are usually too
expensive to use, except in particularly critical
applications. Sprayed metallic coatings are a
relatively unexplored possibility for this duty.
The most likely metallic coating materials for
thermal spray systems, on cost grounds, are
aluminum, copper, nickel, and zinc, their
alloys, and possibly composite materials based
on them. For highly aggressive (very energetic
flows), cemented carbides of WC-CoCr or WC-
NiCr or other carbide system (TiC or CrC)
cermets are used either in sintered form or ther-
mally sprayed by detonation type or high-
velocity oxyfuel (HVOF) thermal spray guns
(Percy, 1990; Wood et al., 1997; Wheeler
et al., 2005; Allen and Ball, 1996).
To achieve long-term corrosion resistance, it is
important to understand the consequence of
selecting anodic and cathodic coatings relative
to their substrate. For example, zinc and alumi-
num are both anodic to CS, that is, they protect
steel galvanically in seawater, and are used in
offshore applications for this purpose
(McIntyre, 1999; Puget, 1999; Percy, 1990).
Copper and nickel are both cathodic coatings,
that is, they are not protective to steel. Where
defects in the coating occur, exposed steel is likely
to be rapidly corroded, adding to erosioncorro-
sion attack and coating delamination. So, if
copper or nickel or their alloys are used as coat-
ings, they should be defect free. If not, a sealant
should be used to protect against electrolyte
ingress and permeation into the coating via
interconnecting defects. A cautionary note
should be made here as defect-free coatings
are rarely achievable and even if defects are
minimized and sealed, erosioncorrosion pro-
cesses can readily generate defect sites through
induced stress systems. Such conditions exploit
defects near to critical size for growth, inclu-
sions, and any anisotropy in the near or
subsurface microstructure to grow defect fields,
and eventually accelerate wall loss rates. Hence,
for erosion resistance, modern surface engineer-
ing research is looking into depositing
high-quality and high-density (<1% porosity)
coatings with a wide range of carbides along
with the above metal-binder elements with the
aim to achieve good adhesion between carbide
and binder. These should successfully combine
corrosion- and erosion-resistant elements
together. Recent research has shown that even
this solution has difficulties; as Souza and
Neville (2005) report, the performance of
WCCoCr systems can suffer localized corro-
sion of the carbide phases, thereby
compromising the performance of such surfaces.
There are a number of factors that affect the
behavior of materials when they are subjected
to erosive and corrosive flows. These can be
placed in four categories: (1) the nature of the
fluid transporting the particles (liquid or gas-
eous), (2) the nature of the particles, (3) the flow
field, and (4) target parameters. They are sum-
marized in Table 2.
6.09.2.1.3 Environment factors
Corrosion is a reaction process on metallic
surfaces that occurs as a result of interaction
with the environment. The rates and modes of
corrosion attack are related to the electrolyte
environmental conditions such as pH, dissolved
oxygen, salinity, temperature, flow structure,
and biofouling.
The formation and influence of biofilms on
corrosion and erosioncorrosion has, by and
large, been ignored to date but it can have a
significant impact on system performance. In
industrial processes, the presence of biofilms on
metallic surfaces can cause serious equipment
malfunction as well as lower the efficiency of
components such as heat exchangers. Biofilms
can also lower the end-product quality or safety
in the food industry (Carpentier and Cerf 1993;
Flemming, 2002). Knowledge of the fundamen-
tal mechanisms that govern the interaction of
biofilms with metals in marine, fresh-water,
and industrial environments is essential to
avoid the environmental problems caused by
biofouling and biocorrosion. Any surface
ErosionCorrosion Fundamentals 401
immersed into an aquatic environment will
quickly adsorb dissolved organic and inorganic
matter, which forms a discontinuous film of
variable thickness, thereby conditioning the sur-
face (Loeb and Neihof, 1977; Taylor et al.,
1997). The conditioned surfaces are then colo-
nized by pioneering microorganisms including
bacteria, microalgae, and fungi. Such attach-
ment and growth of microorganisms on
surfaces are generally defined as a biofilm or
called microfouling.
The initial bacterial attachment to a metal
surface and the subsequent formation of a bio-
film will be dependent on surface characteristics
of the substratum, including metal surface free
energy, roughness, and hydrobicity (Muller
et al., 1992), as well as metallurgical features
(Walsh et al., 1993). In addition, the biochem-
ical characteristics and the extracellular
polymeric substances (EPSs) have also been
reported to be crucial in biofilm formation
(Jucker et al., 1998). EPSs are primarily com-
posed of polysaccharides, uronic acid, and
proteins containing functional groups, such as
carboxylic acid and amino acid groups, which
could be capable of binding metal ions
(Geesey et al., 1988; Paradies, 1995). Thus, the
EPS can affect the electrochemical characteris-
tics of a metal surface and play an important
role in the corrosion and erosioncorrosion of
metals (Roe, 1996).
The extent of marine growth depends on a
number of different environmental factors such
as the geographic location, season of the year,
water chemistry, temperature, substratum type,
sunlight, distance from the shoreline, and condi-
tions of turbulence. Microbial attachment has
been observed at wall shear stresses in the
order of 100300 N m
2
(Duddridge et al.,
1982; Finlay et al., 2002). It has been suggested
that biofilms formed under high fluid velocities
are thinner and denser in structure or consist of
cell clusters which exhibit a greater resistance to
detachment than single cells (Finlay et al., 2002;
Melo and Vieira, 1999; Wijeyekoon et al., 2000).
Therefore, biofilms are very environmentally
sensitive and under low particle impingement,
erosion conditions may well play a role in deter-
mining erosioncorrosion surface loss rates.
The influences of other environmental fac-
tors, such as flow velocity, pH of the solution,
and the system temperature, on erosioncorro-
sion are likely to be very system dependent.
However, the sensitivity of wall wastage rates
to such environmental parameters can be found
in the literature. Kim et al. (2003) report how
these parameters effect the galvanic corrosion
of CS in alkaline chloride solutions. They
Table 2 Principal factors that affect solid-particle erosion of materials
Category Factor
Nature of the fluid Viscosity
Density
Corrosivity such as salinity, pH, dissolved oxygen, conductivity, solubility
of species, concentration of active species
Temperature
Nature of the particles Size
Density
Friability (crush strength)
Shape
Particletarget hardness ratio
Flow field Particle velocity
Particle kinetic energy
Particleparticle interactions
Impact angle
Particle flux
Particle-impingement efficiency
Particle dropout
Asymmetry of particle burden
Target parameters Hardness
Elastic modulus
Fracture toughness
Residual stress
Surface roughness
Surface treatment
Size and distribution of microstructural flaws
Coefficient of restitution
402 Erosion/Corrosion
studied the aqueous corrosion characteristics of
a CS coupled to type 304 (UNS 530400) stain-
less steel in deaerated solutions as a function of
velocity, pH, and temperature. They found that
the galvanic corrosion between CS and type 304
stainless steel was accelerated by the presence of
an oxide layer formed on the active CS. This
resulted in extensive corrosion of the anode,
CS, while the stainless steel was cathodically
protected from corrosion. The galvanic current
density increased with increasing flow velocity
(between 0 and 0.6 m s
1
) and temperature
(between 25 and 75

C), but decreased with
increasing pH from 8 to 10. Flow velocity
effects were only significant at 50 and 75

C.
This effect can be explained by the increased
solubility of magnetite (corrosion oxidation
product) at the higher temperatures.
A guide to system performance for solutions
with different pH can be found by reference to
the relevant potentialpH (Pourbaix) diagram.
This approach has been illustrated by Stack and
Jana (2005) in their regime-mapping approach
to erosioncorrosion.
The influence of the bulk oxygen concentra-
tion, C
b,O
2
, in aqueous environments on the
corrosion rate depends on the controlling cor-
rosion mechanisms, which will be discussed in
greater detail later in the chapter. However,
when the system is under mass-transport con-
trol, the corrosion rate can be found to be
directly proportional to C
b,O
2
, or C
b,O
2
n
, where
n is flow and system dependent.
Temperature is an important environmental
variable which can influence erosioncorrosion
rates. In aqueous solutions, temperature not
only influences pH, solubility of gases (i.e., oxy-
gen and carbon dioxide), and ionic species as
well as diffusion rates of reactant species, but
can also affect the kinetics of oxidation reaction
at the oxide/metal interface and the reduction
(dissolution) reaction of oxides at the oxide/
water interface. Added complications can occur
when elevated temperatures increase hydrogen
diffusion rates and if this effect is coupled with
a surface susceptibility to hydrogen embrittle-
ment, then increased wall loss may result.
Another major environmental factor with sig-
nificant influence on erosioncorrosion rates is
that of flow velocity but it should be set in con-
text of the overall flow field as other parameters
such as wall shear stress, wall surface roughness,
turbulent flow intensity, and mass-transport
coefficient (this determines the rate of movement
of reactant species to reaction sites and can thus
relate to corrosion wall wastage rates). For
example, a single value of flow velocity referred
to as the critical velocity is often quoted to repre-
sent a transition from flow-induced corrosion
to enhanced mechanical-corrosion-interactive
erosioncorrosion processes. It is also used to
indicate the resistance of the passive and protec-
tive films to mechanical breakdown (Hu and
Neville, 2005). However, it should be remem-
bered that a single value of free stream velocity
can result in widely different flow-field structures
that are dependent on the Reynolds number of
the flow geometry and type and thickness of
boundary layer induced at the liquid/solid inter-
face. These factors will directly affect the velocity
profiles and transverse momentum transfer close
to the solid/liquid interface, which will be shown
later to dictate wall shear stress levels and mass-
transport efficiencies. Therefore, critical velocity
values are very geometry or test-rig specific and
cannot be readily applied to predict component
service life in generic flow systems.
6.09.2.1.4 Geometry factors
The geometry of components exposed to ero-
sioncorrosion can have a significant influence
on the levels of surface wastage suffered. From
classical erosion studies, it is well understood
that ductile surfaces suffer least solid-particle
erosion damage under normal impact conditions
whereas they are vulnerable under 30

impinge-
ment in gassolid flows or between 20

and 30

in slurry flows. Brittle surfaces, on the other


hand, are more vulnerable under normal impact
conditions rather than lower angle impinge-
ments. This finding has major implications for
flow component geometry should erosion and
erosioncorrosion need to be minimized.
Flow-enhanced corrosion can be at its most
severe in the vicinity of flow disturbances. In the
past, flow-enhanced corrosion under disturbed
flow conditions has been studied experimentally
in flow loops and numerically by performing
flow simulations. Recent studies such as Nesic
et al. (2000) investigate flow corrosion under
disturbed flow condition using a modified rotat-
ing cylinder geometry with a sudden step. Large
variations in the wall mass-transfer rates, similar
in character to those obtained by a sudden pipe
expansion, were observed behind the step. Flow
simulations show that this flow geometry will
create a qualitatively similar mean flow pattern
to that of a sudden pipe expansion flow invol-
ving flow separation and reattachment.
To understand the wear of pipe bends, an
erosion model that relates particle impacts to
volume loss is required. However, slurry ero-
sion is a complex and under-researched area,
and thus robust models have yet to be devel-
oped. This is certainly true for erosion of
pipelines and is compounded by the fact that
most published literature on the wear of pipe
work relates to pneumatic conveying systems.
ErosionCorrosion Fundamentals 403
Understanding the particle trajectories
through bends is of great importance if wear
rates are to be predicted at specific locations.
Blanchard et al. (1984) attempted a two-
dimensional model of particle trajectories in
liquid flows within bends but with limited suc-
cess due to the inability of the model to predict
secondary flows. Forder (2000) has studied
solid-particle trajectories in liquid flows in
pipe-bend geometry before modeling more
complex flows in choke valves using a com-
mercial computation fluid dynamics (CFD)
code. He extended a commercial CFD pro-
gram to include a predictive erosion element
based on a combined deformation and cutting
erosion model, detailed later. The model was
tested by comparing the predicted wear rates
and wear locations in pipe bends with the
experimental air/sand erosion results of
Bourgoyne (1989), with excellent correlations.
Further experimental wear results for pipe
bends could be used to test this approach in
the future, as reported by King et al. (1991)
and Wiedenroth (1984), for a limited range of
bend geometries, materials, and solidliquid
flows. Chen et al. (2002, 2004) have included
a stochastic rebound model in a similar
approach to Forder to investigate the model-
ing of erosion in elbows and plugged tees. The
stochastic rebound model was found to have
little effect on the predicted erosion of elbows
subjected to sandwater mixtures while differ-
ences were seen in the plugged tee geometry
exposed to airsand flows.
The redistribution of solids inside horizontal
bends for multisized particulate slurries has
been experimentally studied by Ahmed et al.
(1993). Conclusions from this work point to
the redistribution of large particles outward as
the likely cause of rapid bend erosion. Knowing
the distribution of particles in flows (particle
burden) and the motion of particles in flow
systems is a major point to consider and the
recurrent theme in this chapter.
Experimental studies on pipe systems show
variation of the erosion mechanism and loca-
tion with flow regimes. Wood et al. (2001,
2002a, 2002b) show erosion damage of AISI
304L stainless steel pipe components with a
nominal wall thickness of 5 mm from a hori-
zontal pipe loop test. The pipes were of 80 mm
nominal bore, although the actual mean of a
series of measurements was 77.8 mm. The ero-
sion of the bend with the greatest curvature of
the whole loop (R
c
/D
b
= 1.2) was analyzed
along with its upstream straight.
Figure 3 shows the top internal surfaces (12
oclock position) of the straight pipe after 210 h
of slurry testing at 3 m s
1
with an asymmetric
slurry flow of 10% w/w sand of 5001400 mm in
size within water. Figure 4 shows the overall
component details for the straight and bent
sections investigated. The damage in Figure 3
is consistent with that of the as-manufactured
(scaled) surface finish. There appears to be no
significant erosion impact damage although a
few individual impact craters are seen due to
impacts from the relatively few sand particles
that are entrained into the energetic upper fluid
flow. The lack of damage seen in Figure 3 is in
contrast with the damage seen at the bottom of
the straight pipe section as shown in Figure 5.
This reflects the asymmetry of sand burden in
the flow with the majority of sand particles
traveling in the bottom half of the pipe with
only a few energetic particles saltating into the
flow stream in the top half of the pipe.
In Figure 5, the impact features reveal that
the erosion processes occurred by a simulta-
neous plastic deformation and microcutting of
(a) (b)
NONE SEI 15.0kV 1,200 10 m WD 19.2 mm
NONE SEI 15.0kV 100 10 m WD 19.2 mm
Figure 3 SEM micrographs at two magnifications showing surface morphology of straight position (top A
position; after 210 h). From Wood, R. J. K. and Jones, T. F. 2003. Investigations of sandwater induced erosive
wear of AISI 304L stainless steel pipes by pilot-scale and laboratory-scale testing. Wear 255, 206218.
404 Erosion/Corrosion
the stainless steel pipe surface, similar to type II
cutting mentioned by Hutchings when the ero-
dent rotates backward on impact and efficiently
machines the surface (Hutchings, 1992).
Erosion of the bend outer wall at section 2
(shown in Figure 4) is even more intense as
shown in Figure 6. The surface of the bend at
position C (3 oclock) has clearly been eroded
with evidence of low-angle impingements with
extensive plastic deformation and cutting (type
II) resulting in the removal of the as-manufac-
tured surface morphology. This level of damage
is a result of impingement by the larger and
energetic sand particles with the bend wall. It
reflects the repositioning of the particle burden
due to the bend geometry and resulting change
in flow direction.
6.09.2.1.5 Galvanic interactions
Galvanically driven corrosion typically
results from coupling dissimilar alloys with the
more anodic (electronegative) surface
dissolving at an enhanced rate compared with
uncoupled rates. The cathodic (electropositive)
surface typically becomes protected and is less
likely to corrode. The range of potentials
experienced by various metals and alloys can
be listed in the form of galvanic series based on
practical observations in a specific environ-
ment, such as seawater, as illustrated in
Table 3 (Trethewey and Chamberlain, 1995).
The extent of such galvanically coupled
corrosion relies on numerous factors such as
surface area ratios of the anodic surfaces to
cathodic surfaces, solution conductivity,
flow-field intensity, temperature, and the effi-
ciency of the electropositive surface to act as
the cathode (which in turn controls the rate
of reduction). Dissimilar alloys are found
coupled to each other in flow systems.
Typically, designers generate galvanic couples
unintentionally and rapid wall thinning can
result in areas close to these couples. Such
severe corrosion attack results from interac-
tions between the enhanced corrosion of
galvanic coupling and flow-assisted corrosion
(a) (b)
Figure 5 a, SEM micrographs showing the surface morphology of the straight pipe at the bottom (position E)
after 210 h; b, individual impact scars from sand microcutting at low-angle impingements with the pipe surface.
From Wood, R. J. K. and Jones, T. F. 2003. Investigations of sandwater induced erosive wear of AISI 304L
stainless steel pipes by pilot-scale and laboratory-scale testing. Wear 255, 206218.
3
1
2
1
A
B
C
D
E
F
G
H
Flow into page
Weld
Flow
Figure 4 Drawings of bend and straight sections which were mounted horizontally and subjected to asymmetric
slurry flow for 210 h. From Wood, R. J. K. and Jones, T. F. 2003. Investigations of sandwater induced erosive
wear of AISI 304L stainless steel pipes by pilot-scale and laboratory-scale testing. Wear 255, 206218.
ErosionCorrosion Fundamentals 405
or even erosioncorrosion. These types of
processes act to destabilize protective surface
films. The growth of calcareous films on the
cathodic areas of galvanic couples (known as
chalking) can be beneficial to suppress charge
transfer by increased electrode/solution
impedance.
Al-Hosani et al. (1997) have studied the
galvanic corrosion induced by coupling of
copper-based alloys to Mo-stainless steels in
static Arabian Gulf seawater under a variety
of conditions including temperatures, cathode-
to-anode area ratios, and stirring. For static
conditions, the corrosion rates were found to
be proportional to the cathode/anode area
ratio. This can be attributed to the fact that
the cathodic reaction (the reduction of oxygen)
controls the overall corrosion process, and this
is diffusion controlled. However, the galvanic
corrosion rate of the couples increased non-
linearly with a rise in temperature to reach a
maximum at 50

C, decreasing sharply at
higher temperatures due to the decrease in
dissolved oxygen at this temperature. The
effect of stirring the electrolyte shifted the
open circuit potential (OCP) of the stainless
steel to a more electronegative value. Thus,
the corrosion rates in the stirred solutions
were found to be between 1.33 and 3.0 times
above those measured in stagnant solutions.
Erosioncorrosion conditions can themselves
develop local microgalvanic cells between eroded
and uneroded surface areas. Such microgalvanic
cells have been reported between erosion-
affected regions (depassivated) of a surface and
noneroded regions (passive). For example,
enhanced material loss rates due to such cells
are reported by Hodgkiess et al. (1999).
6.09.2.2 Electrochemical Reaction Kinetics
6.09.2.2.1 Kinetics under polarization
(i) Activation polarization
The rate of charge transfer across a metal
solution interface and the mechanisms of
charge transfer and their control are clearly
important to understand if wall wastage rates
are to be minimized in erosioncorrosion.
Typically, these occur by coupled oxidation
and reduction reactions. The rate of charge
transfer can be represented by the anodic (oxi-
dation) and cathodic (reduction) currents
associated with the reactions occurring at
anode and cathode sites on a metallic surface.
These currents can be related to the potential of
the surface. Under uncoupled and equilibrium
conditions an electrode potential is referred to
as either the equilibrium potential E
eq
, OCP, or
freely corroding potential E
corr
. If the surface
potential is polarized, for example, by galvanic
coupling or applying a potential, the level of
polarization is given by the difference between
the electrode potential and E
eq
. Polarization
of an electrode supporting one reduction
oxidation system is given by the Butler
Volmer eqn [3]:
I = I
0
exp b
zF
RT
y
Z
act
_ _
exp (1b)
zF
RT
y
Z
act
_ _ _ _
[3[
where I is the anodic or cathodic current, I
0
is
the exchange current density, u is the charge-
transfer barrier coefficient for the anodic or
cathodic reaction (u is typically 0.5), z is the
Figure 6 Micrographs showing the surface morphology of bend erosion outer wall at section 2 and position C.
Both images show impact craters with extensive plastic deformation and cutting damage. Wood, R. J. K. and
Jones, T. F. 2003. Investigations of sandwater induced erosive wear of AISI 304L stainless steel pipes by
pilot-scale and laboratory-scale testing. Wear 255, 206218.
406 Erosion/Corrosion
number of electrons transferred,
j
act
= E
applied
E
eq
. For anodic overpotentials,
I
reaction
= I
a
= I
0
exp (1 b)
zF
RT
y
Z
reaction
_ _
[4[
So
Z
a
= b
a
log
10
I
a
[ [
I
0
_ _
[5[
where b
a
is the Tafel coefficient which can
sometimes be obtained from the slope of plot-
ting j versus log [I[ with the intercept yielding a
value for I
0
:
b
a
= 2.303
RT
y
(1 b)nF
[6[
For cathodic reactions,
I
c
= I
0
exp b
nF
RT
y
Z
c
_ _ _ _
[7[
leading to
Z
c
= b
c
log
10
I
c
I
0
_ _
[8[
b
c
= 2.303
RT
y
bnF
[9[
Figure 7 illustrates Tafel slope extrapolation to
allow the OCP (E
corr
) and corrosion current
density (i
corr
) to be evaluated.
The discussion so far has focused on the
kinetics of reactions occurring at the metal/
solution interface. However, the transport of
the corrosion reactants/products through the
solution is also important because of the time
it takes. Unlike the passage of electrons, that is,
where the current flow through a metal is
deemed to occur in a negligible time compared
with other processes, the current flow through a
solution is carried by species much more mas-
sive than electrons. Thus, mass transport
during an electrochemical process/corrosion is
carried out by diffusion, migration, and convec-
tion. Diffusion arises due to concentration
gradients and it must occur whenever there is
a chemical change at the electrode surface.
Migration is the movement of charged species
due to potential gradients while convection is
the movement of a species due to mechanical
E
corr
corr
log current density (A m
2
)
P
o
l
a
r
i
z
a
t
i
o
n

(
V
)

a
Figure 7 Schematic of Tafel plot (Pletcher, 1991).
From Pletcher, D. 1991. A First Course in Electrode
Processes. The Electrochemical Consultancy, Romsey.
Table 3 Galvanic series in seawater (Trethewey and
Chamberlain, 1995)
Metal
Potential (V vs SCE)
(approximate)
NiCrMo alloy C 0.030 to 0.090
Titanium 0.050 to 0.060
NiCrMoCuSi alloy B 0.020 to 0.040
NiFeCr alloy 825 0.020 to 0.040
Alloy 20 stainless steels,
cast and wrought
0.040 to 0.060
316, 317 stainless steel 0.000 to 0.100
(0.350 to 0.450)
Nickel copper alloys 400
and K-500
0.050 to 0.150
302, 304, 321, 347
stainless steel
0.040 to 0.100
(0.475 to 0.580)
Silver 0.100 to 0.150
Nickel 200 0.100 to 0.200
Silver braze alloys 0.100 to 0.200
Nickel chromium alloy
600
0.125 to 0.175
Nickel aluminum bronze 0.150 to 0.220
70/30 Copper nickel 0.180 to 0.225
Lead 0.200 to 0.240
430 Stainless steel 0.200 to 0.280
80/20 Copper Nickel 0.220 to 0.280
90/10 Copper nickel 0.225 to 0.290
Nickel silver 0.260 to 0.280
416 and 410 stainless steel 0.260 to 0.330
(0.460 to 0.560)
Tin bronze 0.265 to 0.320
Silicon bronze 0.270 to 0.290
Manganese bronze 0.290 to 0.340
Admiralty brass,
aluminum bronze
0.290 to 0.350
Lead tin solder 50/50 0.290 to 0.375
Copper 0.300 to 0.360
Tin 0.300 to 0.320
Naval brass, yellow brass,
and red brass
0.300 to 0.390
Aluminum bronze 0.320 to 0.410
Austenitic nickel cast iron 0.420 to 0.530
Low alloy steel 0.570 to 0.630
Mild steel, cast iron 0.600 to 0.720
Cadmium 0.700 to 0.730
Aluminum alloys 0.750 to 0.990
Beryllium 0.970 to 0.990
Zinc 0.990 to 1.020
Magnesium 1.600 to 1.640
Values in brackets indicate low-velocity or poorly aerated
water, and at shielded/occluded areas.
Reference electrode SCE = saturated calomel electrode.
ErosionCorrosion Fundamentals 407
forces (Pletcher, 1991). However, the
mass-transfer regime within the electrolyte is
not completely diffusion dependent.
(ii) Diffusion-controlled polarization
For diffusion-controlled mass transport, the
flux of species i is described by Ficks first law:
J
i
= D
i
C
i
x
_ _
[10[
where J
i
is the flux of species i (mol s
1
cm
2
),
D
i
is the diffusion coefficient of species i
(cm
2
s
1
), and DC
i
/Dx is the concentration gra-
dient of species i across the interface (mol cm
4
)
(see Figure 8).
The diffusion coefficient of ionic species in
dilute solutions can be estimated by the Nernst
Einstein eqn [11]:
D
i
=
RT
y
l
i
[z
i
[
2
F
2
[11[
where z
i
is the valence of the species, l
i
is the
molar conductivity.
For flowover the surface or moving electrodes,
the concentrationdistance profile at the electrode
surface canbe approximatedby a simple gradient.
Because the concentration gradient DC
i
/Dx is
maximum when the surface concentration of
species i is zero then
J
i
= D
i
C
bulk
i
d
_ _
[12[
where c is the diffusion layer thickness.
Converting the flux of species J
i
into an
equivalent current density with Faradays law
i
c
= i
L
= nFJ
i
= nFD
i
C
bulk
i
d
_ _
[13[
Here, i
L
is also the anodic current density when
cathodically controlled which is the case for
many engineering situations.
(iii) Limiting current density
By considering electrode reactions involving
only one species, the subscript i can be omitted.
Thus
i = nDF
C
x
= nDF
C
b
C
0
d
[14[
C
0
represents the reactant concentration at the
electrode. As the potential is further decreased,
ions are consumed faster by the electrode reac-
tion, resulting in a lower concentration at the
electrode. Because c, the boundary-layer thick-
ness, is determined by the flow, and the bulk
concentration is fixed, the decrease in C
0
brings
about a higher current. Because the concentra-
tion at the electrode cannot be less than zero,
there is a limit; this mass-transport-limiting cur-
rent is determined by setting C
0
= 0 in eqn [14]
to obtain eqn [15]:
I
lim
= zFD
C
b
d
[15[
The limiting current I
lim
is an important
design parameter because it represents the max-
imum rate at which an electrode reaction can
proceed. Moreover, it provides a convenient
way to measure transport rate, because mas-
s-transport coefficients can be readily and
accurately calculated from the experimentally
obtained current plateau.
6.09.2.2.2 Mass-transfer coefficient
Mass transfer can significantly affect the reac-
tions rates and therefore corrosion rates and are
particularly important in erosioncorrosion con-
ditions. A useful parameter used to develop
predictive models for flow-assisted corrosion and
erosioncorrosion is the mass-transfer coefficient.
The mass-transport coefficient, k
m
, is defined by
J = k
m
C = k
m
(C
b
C
0
) [16[
Representing the flux in terms of the current
k
m
=
I
lim
zF(C
b
C
0
)
[17[
The only parameter on the right that cannot be
readily measured is C
0
; however, at the limiting
current, C
0
is zero (Ra cz et al., 1986), providing
k
m
=
I
lim
zFC
b
[18[
6.09.2.3 Fluid Flow Fields
6.09.2.3.1 Turbulence
The relationships between fluid mechanics
and induced flow patterns for both open and
x =
x
C = C
i

Concentration
Distance from electrode surface
x = 0
Electrode
Bulk solution
Figure 8 Concentration profile at the electrode/
solution interface.
408 Erosion/Corrosion
closed fluid systems are covered extensively in
undergraduate text on fluids mechanics such as
Douglas et al. (2005) and therefore will not be
reproduced here. Such textbooks detail the dif-
ferences between laminar and turbulent flow
and their importance on near-solid-wall shear
stresses and near-wall flow structures. It is these
near-wall flow conditions that are important to
understand if erosioncorrosion rates and loca-
tions are to be accurately quantified for
engineering components, in particular, the
development of eddy structures in turbulent
flows through lateral momentum transfer
between fluid elements. These in turn disrupt
the turbulent boundary layers that are devel-
oped within near-wall flows (Cantwell, 1981;
Robinson, 1991; Panton, 1984) and require par-
ticular attention, as these can develop localized
damage patterns on solid surfaces that can lead
to accelerated loss of fluid containment and
thus decrease component life considerably.
Two examples are given below, namely, rip-
pling in pipe flows and turbulent bursting/
jetting flows over flat plates:
(i) Rippling of pipe surfaces by erosion
corrosion induced by near-wall turbulence
In the cases of multiphase flows
(solidliquid), it is assumed that initially the
turbulent eddy pattern in the flow is determined
by the surface roughness generated by the final
machining operation of the pipe (see Figure 3).
The ability of the sand particles to erode is
largely determined by the bulk flow but will
result in the modification of the surface rough-
ness. The initial surface roughness is thus
replaced by the roughness resulting from indi-
vidual erosion events and varying depths of
corrosion driven by the varying flow and
mass-transport conditions near the pipe wall
(Figure 9a). In a ductile material, such as stain-
less steels, and for erosion-dominated erosion
corrosion, this roughness will be in proportion
to the impact scar dimensions (Figure 9b).
These in turn give rise to a new eddy pattern
in the near-wall flow. During the erosion pro-
cess, the sand particles have to pass through
these eddies and are thereby deflected. The
eddy pattern determines the angle of impact
and areas of increased sand-particle impact
and also the flow corrosion rates over the
surface. The concentration of damage at spe-
cific points results in the establishment of a
ripple pattern, reflecting the turbulence pat-
tern. The surface waviness increases until a
steady state is reached at which point the
surface continues to wear but the wavelength
and shape remain constant (Figure 9c). The
final waviness is thus a reflection of the
conditions that the material is exposed to.
The more severe the condition the greater is
the turbulence in the boundary layer and
hence the larger is the surface waviness.
(ii) Turbulent burst and near-wall flow jetting
The near-wall region can be divided into
three distinct regions: (1) the viscous sublayer,
(2) the buffer layer, and (3) the fully turbulent
region (also known simply as the log region).
Turbulent features are often reported with
reference to the dimensionless scale y

= v*y/
i, where v* is the friction velocity (=t
w
/, or
wall shear stress over fluid density), y is the
vertical height above the surface, and i is the
kinematic viscosity of the fluid. The viscous
sublayer extends from the wall out to about
y

= 58. In this region, the mean velocity pro-


file is linear, as if the region is devoid of
turbulence. However, flow visualization has
revealed slow-moving oscillations known as
near-wall or sublayer streaks (Donohue et al.,
1972; Kline et al., 1967). These streaks are
about 100 y

units across and 1000 y

units
long and are believed to occur as a result of
elongated regions of alternating low- and
high-speed fluid produced by streamwise vor-
tices close to the wall. The near-wall streaks
slowly lift up into the buffer region, where
they undergo a distinct oscillation and finally
break up violently around y

= 40 (see
Figure 10) (Kim et al., 1971; Blackwelder,
1989). The majority of the turbulence produc-
tion in the entire boundary layer occurs in the
buffer region from y

= 8 to 3050 during vio-


lent outward ejections of low-speed fluid. For
continuity considerations, the ejections are fol-
lowed by an inrush of high-speed fluid at a
(a) (b)
(c)
Figure 9 Ripple formation due to turbulence eddies
and the relation between microroughness and eddy
size: a, initial stage with turbulence boundary layer,
microroughness determined by prior machining
operations; b, intermediate stage with larger eddies
due to changing surface roughness (microroughness
determined by the individual erosion event); c, final
stage with steady ripple pattern. After Karimi, A.
and Schmid, R. K. 1992. Ripple formation in solid
liquid erosion. Wear 156, 3347.
ErosionCorrosion Fundamentals 409
shallow angle toward the wall (a sweep
motion). An ejection transports low-momentum
fluid from the wall region into the main flow.
Conversely, sweep events transport
high-momentumfluid fromthe main flowtoward
the wall (Figure 10) at velocities close to 0.9U,
where U is the free stream velocity. The result of
these flows is fluctuating wall shear stresses (see
eqn [19]), and pressures which will influence ero-
sioncorrosion rates and mechanisms locally.
6.09.2.3.2 Wall shear stresses
The wall shear stress can be calculated from
t
w
= (m Z
e
)
dU
dy
[19[
where m is the dynamic viscosity of fluid, j
e
is
the turbulent eddy viscosity, and dU/dy is the
velocity gradient at the wall.
Wall shear stress is important to erosioncor-
rosion conditions as it relates to the possible
stress that could be applied to loosely adherent
surface deposits or passive films. However, it is
now thought that passive film removal and its
associated critical velocity is related to mass-
transport-induced mechanisms (surface dissolu-
tion and polarization) and not wall shear stress,
t
w
(see eqn [19]), as previously assumed.
However, violent near-wall turbulent flow struc-
tures, generated by streaming and bursting, could
remove or disrupt passive films and oxide layers.
6.09.2.4 Erosion Fundamentals
There are three categories of erosion depend-
ing on the nature of the erodent. They are
v solid-particle impingement,
v liquid-droplet impingement, and
v cavitation and aeration.
To allow a focus to this chapter, only solid-
particle erosion is discussed further.
6.09.2.4.1 Solid-particle erosion
Mechanical damage to a surface is possible
when particles are suspended and/or entrained
in energetic fluid flows impinged on surfaces.
Particle contact with a surface induces surface
and subsurface stresses, which often result in per-
manent plastic deformation or removal of
material from the surface. The flow conditions
along with the properties of the target material
or any film, if present, as well as the fluid and
erodent properties, will influence the contact
mechanics of the erodent to component surface
interactions and thus influence erosioncorrosion
rates. Such material properties are hardness, frac-
ture toughness, coefficient of restitution, and
elasticity or the degree of surface roughness. The
general expression for erosion rate, W, has been
established empirically and can take the form
W _
_
M
p
Kf (c)U
p
n
[20[
where
_
M
p
is the mass of sand impacting the
surface per unit of time, U
p
is the particle velo-
city on impact, c is the particle impact angle, K
and n (typically between 2 and 3; Faddick,
1982) are constants assumed to be dependent
on characteristics of the erodent/target materi-
als involved. The functional relationship for the
dependence of the erosion rate on the impact
angle is given by f(c) (Haugen et al., 1995).
This can be rewritten for constant
temperature:
V
u
= E
k
m
f (c)C
v
A [21[
where V
u
is the erosion rate (volume loss per
impact or m
3
/impact), E
k
is the particle kinetic
energy, m is the energy exponent with m = 0.5n
assuming W _ d
3
(where d is the particle dia-
meter), C
v
is the volume fraction, Ais a constant,
and f(c) is as before. Fromeqn [20] or [21], it can
be seen that the erosion rate will be strongly
dependent on the kinetic energy of the impacting
sand particles, the number of impacting parti-
cles, and the impact angle. All three of these
factors vary for most industrial components
exposed to sand-laden flows and therefore, to
sustain long service life, the internal surfaces
must perform over a wide range of solid impact
conditions. The sand-to-wall impact conditions
are likely to depend on the flow regime present,
the orientation and/or geometry of the compo-
nent, the ability of the flow to keep the particles
in suspension (Turchaninov, 1973), and the rela-
tive size of erodents with respect to boundary-
layer thickness (small particles can form
protective clouds within subviscous layers). For
v
2 1
4 3
Sweep
Ejection
y
+
= 3050
y
+
= 58
y
+
= 0
u
u < 0
v > 0
u > 0
v <0
1
1 Near-wall streak
2 Liftup into buffer region
3 Violent breakup of ejection into outer region
4 Inrush of high-velocity fluid to wall (sweep)
2 4
Flow
3
Figure 10 Schematic showing the near-wall
turbulence structures within the viscous sublayer,
buffer layer, and fully turbulent region. From
Wharton, J. A. and Wood, R. J. K. 2004. Influence
of flow conditions on the corrosion of AISI 304L
stainless steel. Wear 256, 526536.
410 Erosion/Corrosion
suspended solids traveling in a horizontal liquid
flow, both the solid-particle impact velocities
(<3.0 m s
1
) and angles are relatively low for
most practical purposes (Shimoda and
Yukawa, 1983). Typical impact erosion craters
are shown in Figure 11.
Plots of V
u
(mm
3
impact
1
) against kinetic
energy of impact, E
k
(mJ), can be used to gen-
erate an erosion map. These maps enable
comparisons to be made between materials at
a wide range of impact energies independent of
particle size. E
k
is defined as
E
k
= 2,3pr
p
r
p
2
U
p
2
[22[
where r
p
is particle radius, ,
p
the erodent particle
density, and U
p
the impact velocity. Care should
be taken as experimental evidence shows erosion
rates can be dependent on U
p
n
r
p
y
where n and y
are far removed from 2 and 3 assumed in the
simple energy approach above (Hashish, 1998).
The main parameters of concern for erosion
relate to the solid-particletarget interactions
and thus the number of particles impinging,
individual particle energies, particle impinge-
ment angles, particle-to-target hardness ratios,
and the shape of the particles. Near-wall parti-
cleparticle interactions can also severely
influence erosion rates when the volume con-
centration of solid particles present is high.
Finnie (1972) developed an erosion model
based on cutting wear mechanisms of the form:
V
u
= C
1
M
p
U
p
2
4o
f (c) [23[
where C
1
is an arbitrary constant denoting the
number of particles that cut the surface and o is
the target flow stress. Gane and Murray (1979)
found that a value of C
1
= 0.5 gave reasonable
predictions. Keating and Nesic (2001), in
numerically predicting erosioncorrosion in
bends and sudden expansions by two-phase
flows (liquidsolid), used a modified Finnie
approach based on an earlier work by Bergevin
(1984). This approach incorporated the concept
of a critical velocity for plastic deformation, U
cr
.
They substituted (U
p
sinc U
cr
) for the impact
velocity in eqn [23] to give eqns [24] and [25].
For low angles (c _ 18.5

)
V
u
=
M
p
U
p
sinc U
cr
_ _
2o
U
p
cosc
3
2
U
p
sinc U
cr
_ _
_ _
[24[
For higher angles of impingement
V
u
=
M
p
U
p
sinc U
cr
_ _
2
12o
cos
2
c
sin
2
c
[25[
Bitter (1963) quotes a value of
U
cr
= 0.668 m s
1
for steel. Keating and Nesic
use this value to successfully predict erosion
rates in a sudden expansion and found the ori-
ginal Finnie model to be not so accurate.
However, modeling of erosioncorrosion
damage in a U-bend using the original Finnie
model as the modified version yielded no ero-
sion due to the low particle velocities involved.
Keating and Nesic conclude that their modeling
needs more experimental validation before
further refinements can be made.
Erosion models typically recognize that two
erosion mechanisms act. The Neilson and
Gilchrist erosion model, used for rocket motor
nozzles (Neilson and Gilchrist, 1968), incorpo-
rate cutting and deformation erosion, with
discrete models representing each. Such models
have been successfully used by Forder et al.
(1998) and Wood et al. (2001, 2002a, 2002b)
to predict erosion of internal components
within choke valves and slurry ducts. The cut-
ting erosion model for low impact angles was
(a) (b)
Acc.V Spot Magn Det WD 10 m
SE 16.4 20.0 kV 5.0 2000x
Acc.V Spot Magn Det WD 10m
SE 16.9 20.0 kV 5.0 2000x
Figure 11 SEM micrographs of erosion damage on the surface of samples tested at a, 7 m s
1
and b, 12 m s
1
,
1wt.% silica (296 mm) sand in 0.1 M NaOH at 40

C, test time = 30 min. The eroded region observed was
located approximately normal to the flow and in the center of the sample (Harvey et al., 2006).
ErosionCorrosion Fundamentals 411
first proposed by Finnie (1972) and later mod-
ified by Hashish (1998). The deformation
model was proposed by Bitter (1963) and is
thought applicable at higher impact angles
(30

90

). Particle shape and material proper-


ties for both particle and target have been
included which the earlier simpler models have
not considered. As the particle impingement
angles are predicted to be below 10

for critical
components such as straights and bends (see
Wood et al., 2001, 2002a, 2002b), the contribu-
tion to the overall wear rate from deformation
mechanisms can be ignored. The volumetric
erosion per impact can therefore be given by
the modified Hashish model only:
V
u
=
100
2

29
_ r
p
3
U
p
C
k
_ _
n
sin2c

sinc
_
_ _
[26[
where n = 2.54 and
C
k
=

3oR
f
0.6
r
p

[27[
6.09.3 EROSIONCORROSION
MECHANISMS
Published research has tried to deal with the
synergistic effects between erosion and corrosion
processes which result in accelerated material
loss and in some cases actually decelerate mate-
rial loss (Wood, 2004a, 1992; Neville and Hu,
2001; Wang and Stack, 2000; Wood and Hutton,
1990). A wide range of corrosion-resistant mate-
rials rely on a relatively thin surface film to
provide a barrier (of high impedance) to charge
transfer (and thus corrosion) between the rela-
tively active bulk material and the corrosive
environment. This film renders the surface pas-
sive, but for fluid machinery handling flows
where solid particles have been entrained or
cavitation is induced; the passive film will be
removed by mechanical wear or bubble col-
lapse/microjet/shock wave impingement
processes. Where the film is mechanically
removed, charge transfer can occur at the inter-
face without retardation from the barrier film
(Stemp et al., 2003). This interaction between
tribological and electrochemical corrosive effects
causes materials to corrode at a substantially
higher rate than those experienced under static
or quiescent conditions. The possible regions of
erosioncorrosion interactions between different
flow regimes and corrosion are illustrated in
Figure 12. The most significant erosioncorro-
sion regions are where turbulent slurry flow and
corrosion overlap in region A, hydrodynami-
cally induced cavitation overlaps with
corrosion in region B, and turbulent flow over-
laps with corrosion causing flow-accelerated
corrosion, although a fourth case can be envi-
saged when solids entrained into laminar flow
could induce erosioncorrosion if their impact
energy is sufficient to cause plastic deformation
of the surface. All of these have loosely been
referred to in the literature as erosioncorrosion
although in this chapter erosioncorrosion refers
to only regions A and B where mechanical and
electrochemical interactions are involved. Thus,
region C is referenced as turbulent flow corro-
sion or flow-accelerated corrosion rather than
erosioncorrosion.
6.09.3.1 Flow-Accelerated Corrosion
Mass transfer of reactants to and from the
electrode surfaces plays an important role during
flow-accelerated corrosion. Typically, dissolved
oxygen or the dissolved metal ions have to dif-
fuse from the solution to the metal or from the
metal to the solution, respectively (Sedamed
et al., 1998). Mass transport has been studied
by several authors in different hydrodynamic
systems that use convection to enhance the rate
of mass transfer to the electrode. Such devices
include free and submerged jet impingement
geometries, rotating disk electrodes (RDEs),
and rotating cylinder electrodes (RCEs).
6.09.3.1.1 KouteckyLevich equation
The standard model used to describe flow cor-
rosion current densities for mixed controlled
reactions (where activation and diffusion pro-
cesses are present) is the modified Koutecky
Levich model (Wood et al., 1990) given in eqn
[28]. In this equation, k
m
(the mass-transport
coefficient) can be derived from eqns [29][31].
Equation [28] is the summation of activation cur-
rent density, i
a
, which is controlled by the charge-
transfer kinetics and the diffusion-controlled
current density, i
d
, which is related to the
Laminar
flow
Corrosion
Turbulent
flow
Slurry erosion
Cavitation erosion
C
B
A
A = slurry erosioncorrosion
B = cavitation erosioncorrosion
C = turbulent flow corrosion
Figure 12 Venn diagram illustrating the possible
regions of erosioncorrosion interactions between
different flow regimes and corrosion: A slurry
erosioncorrosion, B cavitation erosion
corrosion, and C turbulent flow or flow-enhanced
corrosion (Wood et al., 2002c).
412 Erosion/Corrosion
mass-transport conditions of species and ions to
and from the reacting surface:
1
i
corr
=
1
i
a

1
i
d
=
1
nFk
l

k
1
k
l
nFk
m
[28[
where i
corr
= corrosion current density,
n = number of electrons, F = Faradays con-
stant, k
1
= forward reaction rate,
k
1
= backward reaction rate.
The values of k
m
are normally obtained from
nondimensional correlations between the
Sherwood number Sh = (k
m
L/D), the Schmidt
number Sc = (i/D), and the Reynolds number
Re = (UL/i):
Sh = a Re
x
Sc
y
[29[
giving
k
m
= aD
1y
L
x1
i
yx
U
x
[30[
Taking typical values for the exponents as
x = 0.6 and y = 0.33 (Poulson, 1999) gives
k
m
= aU
0.6
D
0.66
i
0.27
L
0.4
_ _
[31[
where D = diffusion coefficient, U = flow velo-
city, i = kinematic viscosity, L = characteristic
length, a = scaling constant.
Flow-enhanced corrosion rates induced by
solids-free flows, as suggested by eqn [28],
have been shown to depend on k
m
n
where n
varies depending on the surface/environment
combinations and the controlling corrosion
mechanism. For example, corrosion rates for
copper alloys in seawater under mixed activa-
tion and diffusion control are shown to have
n < 1. Steels in sulfuric acid and under simple
mass-transfer control have n = 1, whereas for
steels in water, where flow polarizes the cor-
rosion potential and also influences the
oxides solubility, n lies between 1 and 3 (see
Poulson, 1999). Mass-transport processes also
increase with surface roughness creating large
n values until an upper-bound value is
reached given by
Sh = 0.01 Re Sc
0.33
[32[
However, a power-law relationship between
erosioncorrosion rate and k
m
is not always
seen as in the case of steel in nitrates where
surface films are removed above a critical k
m
resulting in a sudden increase in erosioncorro-
sion rate (Poulson, 1999).
6.09.3.2 ParticleSurface Interactions
When corrosion rate is partially or wholly
controlled by mass transfer of reactant to or
product from the surface, local conditions
under erosion may well influence the
mass-transfer kinetics (measured by the mass-
transfer coefficient k
m
). Under such conditions,
the corrosion will be controlled by the mass
transfer (typically diffusion processes) and the
driving concentration gradient (relative concen-
trations of active species near surface compared
to free stream concentrations) (Silverman,
2004). Both the mass transfer and concentra-
tion gradient will be affected by solid-particle
impingement, which influence the local fluid
flow field and increase surface roughness.
Erosion will also increase the surface area
wetted by the corrosive electrolyte and could
establish microgalvanic cells on the surface
with damaged areas being anodic to the passive
and (cathodic) unaffected areas. Erosion will
also increase the dislocation density in affected
surface areas which can lead to potential differ-
ences between eroded and less eroded areas
being established and the formation of anodic
and cathodic sites generating a microgalvanic
effect enhancing material loss from the surface.
It is also likely that the corrosion kinetics will be
altered on eroded areas.
6.09.3.2.1 Depassivation and repassivation
kinetics
As suggested already, depassivation of pas-
sive or film surfaces is associated with flow
velocity. Figure 13 shows the trend for such a
passive system as a function of flow velocity
and its vulnerability to mechanical processes
which accelerate corrosion loss rates due to
depassivation.
(i) Depassivation
Depassivation in erosioncorrosion pro-
cesses is assumed to be rapid and associated
with the mechanical removal or stripping of
the passive surface layers through particle, cavi-
tation bubble, or liquid droplet impact.
Figure 14 illustrates the level of damage to pas-
sive films as a function of solid-particle
impingement angle. Mechanical removal and/
or rupture of the passive layer enables charge
transfer to proceed at varying rates and parent
metal dissolution is likely. Recovery or repas-
sivation aspects of the passive film are therefore
important but are very system dependent.
Figure 15 shows the current response over
time for individual solid-particle impacts on
the naturally passivating system of stainless
steel. Some impacts will only result in partial
passive layer removal or cracking which will
influence repassivation kinetics and possibly
the composition of the regrown layer.
ErosionCorrosion Mechanisms 413
(ii) Repassivation
Surface coverage models for passive film
growth/recovery relate to the number of moles
of oxide formed N
u
as a ratio of the total num-
ber of surface sites available, N
o,u
in units of
mol cm
2
(from Jemmely et al., 2000):
y =
N
u
N
o.u
[33[
The rate of lateral surface coverage, assum-
ing complete oxidation of the parent metal to
oxide, can be given by
dy
dt
= i
a
M
ox
L
ox
r
ox
nF
_ _
[34[
where i
a
is the anodic current density, n is the
charge number, L
ox
is the oxide film thickness,
,
ox
is the oxide density, and M
ox
is the molecu-
lar weight of the oxide.
If Tafel kinetics applies, then the anodic cur-
rent density i
a
can be given by
i
a
= k
a
(1 y)exp
E
applied
b
a
_ _
[35[
where k
a
is the rate constant and b
a
is the Tafel
slope (=dE/dln(i)). This assumes that oxidation
occurs exclusively on bare parent metal.
Film growth models may explain the thicken-
ing of these oxide layers but assume growth
occurs by high field conduction which is
thought to apply over the longer term but not
in the early stages of repassivation of nascent
surface sites:
i = k
B
exp
Bf
L
ox
_ _
[36[
where k
B
is the rate constant and the migration
kinetic constant for a single ionic species
B = czFa
j
/RT. Here c
1
is a symmetry constant,
0
25
50
75
100
125
150
0 2 4 6 8 10 12 14
Velocity (m s
1
)
Intact
passive
film
provides
barrier to
charge
transfer
Depassivation
and/or
transformation
of film and
increased
mass
transport
Complete
depassivation
and rate
limited
Mechanical
enhancement
(cavitation or
sand erosion)
Critical velocity
C
o
r
r
o
s
i
o
n

r
a
t
e
Figure 13 Possible corrosion trends (with arbitrary units) for a passive metallic surface. From Wood, R. J. K.
2004a. Erosioncorrosion interactions and their effects on marine and offshore components. Keynote at EuroCorr
Conference, Nice, France and accepted for publication in the Special Issue on Tribocorrosion, Wear 2006.
Passive layer or film
Substrate
(a)
(b)
(c)
(d)
Figure 14 Schematic showing that the level of passive
layer damage is dependent on particle impact angle.
6
5
4
t (ms)
l

(

A
)
3
2
270 280 290 300 310
Figure 15 Current transients in erosioncorrosion
due to depassivation/repassivation. From Sasaki, K.
and Burstein, G. T. 2000. Observation of a threshold
impact energy required to cause passive film rupture
during slurry erosion of stainless steel. Phil. Mag.
Lett. 80(7), 489493.
414 Erosion/Corrosion
z is the charge on the migrating ionic species,
and a
j
is the jump distance. The term Dc is the
potential difference across the oxide layer and
taken to be the difference between the electrode
potential and the minimum potential for initia-
tion of film growth. The other term in the
equation is the oxide layer thickness, L
ox
.
This can be linked to the currenttime tran-
sients by setting
L
ox
=
M
ox
q
nFr
ox
[37[
where q is the charge passed or
_
i dt.
The use of such models to predict wall loss
rates depends on the above assumptions and
knowing the oxidation process and the asso-
ciated charge number as well as the area of
depassivation due to the erosion processes.
Figure 16 is an attempt to map surface
responses to the whole range of erosion and
corrosion mechanisms from erosion-dominated
through erosioncorrosion, corrosionerosion
zones to corrosion-dominated mechanisms. To
identify what zone a particular application may
be operating under is difficult without some
data on erosion and corrosion performance of
the surface in question. To plot such a map,
three independent tests to determine pure ero-
sion E, pure corrosion C, and combined
erosioncorrosion gravimetric levels T are
required. However, Table 4 shows the possible
sand-particle energy of various applications
that are directly related to the likely mechanical
erosion damage rates E. This will define a posi-
tion on the TC axis of Figure 16 and the rate of
depassivation as this is likely to be linked to E
k
.
Surface repassivation, on the other hand, is
materialelectrolyteflow-field specific, and
this dictates the corrosion rate C. Three systems
are plotted on this mechanical erosion versus
corrosion map. Two are passivating systems
based on cast and coated nickel aluminum
bronze (NAB) surfaces in saline solution and
have a positive slope (increases with increasing
mechanical erosion effects). This increase is
related to repassivation kinetics being unable
to repassivate the eroded surface under increas-
ing erosion. The third is CS in saline solution
which is not passivating and has a slope near to
zero or slightly negative. Here, the corrosion
products formed on the surface are nonadher-
ent and easily removed and the trend is
dominated by erosion rate while the corrosion
rate remains relatively constant. These results
were obtained using a free jet impingement
erosioncorrosion rig (Figure 19).
6.09.3.3 Synergy
Synergism is the enhancement of material
loss encountered when both erosion and corro-
sion processes work together. This synergy
often leads to significant increase in material
loss rates over and above those expected by
adding the separate erosion and corrosion
rates together. Synergy is defined as the differ-
ence between erosioncorrosion and the
summation of its two parts.
6.09.3.3.1 Models and mapping
Ferng et al. (1999, 2000) modeled erosion
corrosion in pipe flow by adding the erosion
and corrosion contributions together. They
assumed erosioncorrosion of piping is a degra-
dation mechanism that is a coupled
phenomenon between chemical corrosion and
mechanical erosion, which is dominated by pip-
ing layout, fitting geometry, and local flow
structure. They used local flow models, includ-
ing the multidimensional, two-phase fluid
models to simulate flow characteristics within
piping. Erosioncorrosion models were used to
predict two-dimensional distributions of
erosioncorrosion locations in steel pipework.
The model showed satisfactory agreement
with distributions of wear sites measured in
practice.
The erosion model was based on droplet
impingement but used an oxide removal based
on eqn [38]:
m = C
s
NF
y
(y)
r
f
U
2
f
H
v
[38[
where C
s
is a system constant, N the frequency
of impingement, F
0
a characteristic function, 0
0.1
1
10
100
0.1 1 10 100
Mechanical erosion TC (mg)
E
l
e
c
t
r
o
c
h
e
m
i
c
a
l

c
o
r
r
o
s
i
o
n

C

(
m
g
)
Electrochemically
controlled zone
Mechanically
controlled zone
Corrosionerosion
zone
Erosioncorrosion
zone
(TC/C) = 10
(TC/C) = 1
(TC/C) = 0.1
Steel 1020
HVOF NAB
Cast NAB
Figure 16 Mapping electrochemical material loss
against mechanical erosion rates for a
nonpassivating surface CS (AISI 1020) along with
two potentally passivating surfaces of NAB, one
that has been thermally sprayed by HVOF
deposition as a coating on CS and the other which
has been cast. The results have been obtained by jet
impingement erosioncorrosion tests.
ErosionCorrosion Mechanisms 415
the impact angle, ,
f
the fluid density, U
f
the
normal velocity, and H
v
the pipe wall hardness.
The corrosion model was based on a mixed
control corrosion process of dissolution of
magnetite and the mass transfer of ferrous
ions Fe
2
giving
i =
C
eq
C

1,2k
a
( ) 1,k
m
( )
[39[
where k
a
is the reaction rate constant, C
eq
is the
soluble ferrous ion concentration at equili-
brium with magnetite, C

is the soluble
ferrous ion concentration in the bulk solution,
k
m
is the mass-transfer coefficient.
Erosion damage by solid-particle impact or
cavitation bubble collapse to an oxide or pas-
sive film will reveal the underlying nascent
surface. This induces a higher activity (higher
corrosion current), for a limited duration, than
for the intact oxide surface. Bozzini et al. (2003)
employed a simple approximate model using a
recovering target concept. This erosioncorro-
sion model has the advantage in that it can be
applied to both passivating and actively corrod-
ing conditions. The impacting particles are
modeled with rigid monodisperse spheres of
radius r
p
. The particle impact process is
assumed to be Poissonian with parameter l
(impacts m
2
s
1
). It is assumed that each
impact gives rise to an alteration of the corro-
sion rate through a localized change in
corrosion current density for a period of time,
relating to a recovery to the unaffected state.
The effective corrosion current density, i
corr
(nA cm
2
), at a given electrode potential (typi-
cally the corrosion potential) can be related to
the mechanically affected corrosion component
of the synergistic damage through a coefficient,
f
a
, (such that 0 _ f
a
_ 1). The fraction of the
corroding surface which is affected by the ero-
sive action of impinging particles is given by
eqn [40]:
i
corr
= f
a
i
a
(1 f
a
)i
u
[40[
where the subscripts a and u stand for
affected and unaffected, respectively. The
current densities i
a
and i
u
are characteristic for
the corroding material in the absence and in the
presence of the erosive action and can be mea-
sured separately by means of suitable
experiments. In general, the coefficient f
a
can
be defined in eqn [41]:
f
a
=
no. of impacts
control area
_ _

damaged area
impact
_ _
recovery time
= lA
a
t [41[
where A
a
is the affected surface area and t is the
passive recovery time.
Equation [40] indicates an accelerating ele-
ment to corrosion caused by film damage or
removal and is one element of the complex
interactions between erosion and corrosion
that needs to be understood.
These interactions can be defined as follows.
The total damage under erosioncorrosion, T,
can be represented as
T = E C S [42a[
where E is the pure erosion material loss, C is
the solids-free flow corrosion rate (possibly
derived from the corrosion current density i in
eqn [40]). The synergistic effect (interactive
term), S, is referred to as DE
c
or (DC
e
DE
c
),
depending on the literature source, where DE
c
is
the enhanced erosion loss due to corrosion and
DC
e
is the enhanced corrosion due to erosion.
The term DC
e
can be partially expressed by eqn
[40] but additional terms are required relating
to the effect of the erodent deforming the sur-
face leading to increased corrosion activity. The
S terms and how they should be measured are
given by the ASTM G119-93 standard which is
a useful guide to evaluate synergy (ASTM
G119-93, 1993). It is also important to note
that the synergistic term S can be positive or
negative. Bozzini et al. (2003) shows that
annealed CS has a more active corrosion poten-
tial before it is work-hardened, resulting in
i
u
> i
a
, and by using eqn [40] shows a reduction
(i.e., a negative synergy) in overall corrosion
rate with erosion present.
Recent literature in the erosioncorrosion
area has defined synergy in different ways and
thus a cautious note is raised here that care be
taken to determine the definition used when
Table 4 Potential particle kinetic energies and flow velocities for marine components
Marine component
Pipe Valve Pump Propulsor
Velocity (m s
1
) 03 5 10 25
Particle diameter (mm) 100 100 100 100
Particle kinetic energy (mJ) 00.007 0.02 0.07 0.44
416 Erosion/Corrosion
comparing synergistic levels from different
sources in the literature. For example, DE
c
in
some cases is defined as the synergy term and
DC
e
as the additive term. Equations [42b][42d]
all relate to erosioncorrosion but generate dif-
ferent synergy values:
T = E C E
c
C
e
[42b[
T = E C9 E
c
C
e
9 [42c[
T = E C0 E
c
[42d[
where C9 is the static corrosion rate, C is the
solids-free flow corrosion rate, C0 is the corro-
sion rate under erosioncorrosion conditions,
DC
e
is the enhancement of C due to the pre-
sence of erosion, DC9
e
is the enhancement of C9
due to the presence of flow and erosion, and
DE
c
is still the effect of corrosion on erosion.
The synergistic effect, S, is therefore referred to
as DE
c
or (DC
e
DE
c
) or (C0 DE
c
), depend-
ing on the literature source, and thus readers
must take care in extracting synergistic levels
for different materials when using multiple
sources of literature.
Neville et al. (2002) have shown that the
corrosion current density increases with
increasing solid loading (200, 400, and
600 mg l
1
) for UNS S31603 under an imping-
ing jet of 3.5% NaCl at 17 m s
1
. It has also
been shown that synergistic effects, which result
in the damage due to separate corrosion and
erosion processes, are normally greater than the
sum of the individual damage processes and can
accelerate material removal significantly (i.e.,
T = 50 (C E); see Wood and Hutton,
1990). Synergistic effects have also been shown
to be a function of impinging solid-particle
energy for various stainless steels (UNS
S31603, UNS S32100, UNS S32250, and UNS
S32760; Wood, 2004a). Such results indicate
that synergy is more pronounced at lower par-
ticles energies. For example, under an erosion
energy of 0.02 mJ per impact, the volume loss
under erosioncorrosion for UNS S32250
(Ferralium 255) is 10 times that for only erosion
(see Figure 17). However, at higher energies
around 7.5 mJ, the erosioncorrosion rate is
only 1.4 times that for only erosion. It is also
clear that the erosion rate is not simply related
to solid-particle impingement energy but has
two regimes dependent on the magnitude of
the energy: for the lower energy range, V
u
_
U
4
, and for the higher energy range, V
u
_ U
2
.
There is also potentially conflicting data
reported in the literature with certain erosion
corrosion test conditions producing negative
synergy whereas others producing a positive
synergy. Wood (1992) reports negative syner-
gies for a superduplex steel UNS S32750 in
3.5% NaCl solutions at E
k
= 0.05 mJ and 90

impingement angle. However, Neville and Hu


observed a positive synergy on two superaus-
tentic stainless steels (UNS S31254 and UNS
S32654) and a superduplex steel UNS S32750
but these tests were conducted at a higher E
k
of
4 mJ (Neville and Hu, 2001). This suggests a
complex relationship with synergy as a function
of E
k
as this affects the crater size, crater shape,
levels of strain rate, and passive film or oxide
layer recovery time. Table 5 is an attempt to list
the mechanisms that may influence the sign of
the interactive mechanisms, namely, the DC
e
and DE
c
components.
Another process that could influence the sign
of synergistic interactions and levels of synergy
is that of near-surface hardness changes
0.000 01
0.000 1
0.001
0.01
0.1
1
10
1.E09 1.E08 1.E07 1.E06 1.E05 1.E04
Sand particle energy E
k
(J)
E
r
o
s
i
o
n

r
a
t
e

V
u

(

m
3
/
i
m
p
a
c
t
)
V
u


U
4
V
u


U
2
UNSS32250 + tap water
UNSS32250 + 3%NaCl
UNSS32760 + tap water
UNSS32760 + 3%NaCl
UNSS31603 + 3%NaCl
UNSS32100 + 3%NaCl
Figure 17 Erosion and erosioncorrosion wear rates of duplex and austenitic stainless steels as a function of
sand energy (Wood, 2004). This material has been reproduced from Challenges of living with erosion
corrosion by R. J. K. Woods pp. 113132 of the Second Symposium on Advanced Materials for Fluid
Machinery ISBN 1 86058 441 1 published by Professional Engineering Publishing. Permission is granted by
the Council of the Institution of Mechanical Engineers.
ErosionCorrosion Mechanisms 417
Table 5 Synergy overview of processes that could lead to positive and negative interactive effects between
mechanical and electrochemical processes present under erosioncorrosion conditions
Mechanism factor Positive interaction Negative interaction
Critical impact
energy
Above-critical impact energy to penetrate/
damage passive/product film/coverage.
Results in increased charge transfer at
the liquid/metal interface
Below-critical impact energy. Results in
reduced charge transfer at the liquid/
metal interface as film could be
influenced in composition by impact
Surface roughness,
R
a
Roughening effects on mass-transfer
coefficients unknown but k
m
is likely to
increase with increasing R
a
. (The
Silverman review ( Silverman, 2004)
suggests mass-transfer sensitivities with
Reynolds number are affected by
surface roughness as k
m
_ Re
0.9
with
solids compared to k
m
_ Re
0.65
solids
free flows for rotating cylinder work
due to roughening effects on the
surface. Roughness may promote
microturbulence affecting the local
double layer. Roughness could also
promote local microelectrode behavior
(potential field distortion) at the tips of
impact craters
Roughness influences the contact
mechanics of angular solid-particle
impingement. Increased roughness
could reduce contact stresses and
thereby the near- and far-field stress
distributions
Plastic
deformation/
strain
Plastically deformed and stressed surfaces
enhance corrosion processes. Corrosion
causes premature detachment of
plastically deformed or strain-hardened
impact carter lips
Strain-hardened surface increases
hardness and reduces erosion rate and
could reduce corrosion due to change in
microstructure/grain lattice distortion
Increased or
unsteady
hydrodynamics
or turbulence
Unstable double layers (nonsteady state)
and unsteady driving concentration
gradients of active species
High concentrations of solid particles
could block the surface from incoming
particles. Particleparticle interaction
alters impact conditions
Contact
temperature
Local surface (flash) temperatures could
be significantly higher around and
within impact craters which could
accelerate corrosion rates
Local surface (flash) temperatures could
influence oxide film composition and
microstructure and thus could be more
erosion resistant
Localized corrosion Pitting and microgalvanic corrosion cells
due to localized defects in the passive
layers induced by erosion or exposure of
inclusions or voids. Crack systems would
also be vulnerable to crevice corrosion
attack and accelerate crack propagation.
Microgalvanic corrosion cells could be
formed between erosion craters and
surrounding unaffected areas
Multiphase surfaces Corrosion of interface potential to lose
bond integrity between hard and binder
phases, binder receding due to corrosion
leaves hard phase loose to be removed
by mechanical action. These interactions
could also apply to splat boundaries
which exist after spraying
Substrate corrosion products can eject via
interconnecting pores onto the coating
surface, reducing erosivity
Passive film state Depassivation: removal of air-borne
oxides or oxide layers produced by
passivation. Oxide layer could increase
friction between impacting solid particle
and the bulk substrate material
Repassivation: overall recovery times and
adherent oxide layers formed by
passivation. Oxide layer could decrease
friction between impacting solid particle
and the bulk substrate material.
Alternatively, denser oxide could be
formed that reduce corrosion or erosion
or both
Partly taken from reviews by Wood, R. J. K. 2004a. Erosioncorrosion interactions and their effects on marine and offshore
components. Keynote at EuroCorr Conference, Nice, France and accepted for publication in the Special Issue on
Tribocorrosion, Wear 2006 and Wang, H. W. and Stack, M. M. 2000. The erosive wear of mild steel and stainless steels
under controlled corrosion in alkaline slurries containing alumina particles. J. Mater. Sci. 35, 52635273.
418 Erosion/Corrosion
induced by erosion strain hardening or corro-
sion softening. For example, the hardness of a
metallic electrode will decrease when anodic
dissolution is present on the surface and the
relative hardness degradation, DH/H, can be
correlated approximately to the anodic current
density present at the electrode surface, i
a
, as
follows:
H
H
= Blog
i
a
i
th
_ _
[43[
where DH = H
*
H and is the change of hard-
ness due to anodic dissolution at the surface, H
and H
*
the hardness without and with anodic
current present at the electrode surface, respec-
tively, B (>0) is an experimental constant, and
i
th
is the threshold anodic current density to
induce the synergistic electrochemicalmechan-
ical effect, that is, when i
a
_ i
th
, DH = 0.
The change in surface hardness induced by
erosion or corrosion can also establish micro-
galvanic cells due to work function differences
between erosion-hardened areas or corrosion-
softened areas and unaffected areas (Guo et al.,
2005). This is an additional microgalvanic pro-
cess to that of surface film removal.
(i) Empirical models
Wood and Hutton (1990), summarizing
experimental data published on erosion wear
over a wide range of solid particle and cavita-
tion erosion, found that the data fell within two
groups, the medium- and high-synergistic
systems. The erosioncorrosion enhancement,
S, was found to be closely linked to the pure
erosion component E. Two expressions were
derived by plotting S/C against E/C ratios to
obtain the following expressions.
For the medium-synergy group:
S
C
= exp 1.277 ln
E
C
_ _
1.9125
_ _
= 0.1477
E
C
_ _
1.277
[44[
For the high-synergy group:
S
C
= exp 0.755 ln
E
C
_ _
1.222
_ _
= 3.3940
E
C
_ _
0.755
[45[
By using the two expressions above and
experimental values of erosion and corrosion,
these can be compared to experimental deter-
mined erosioncorrosion values. This will
indicate the level of synergy occurring in the
system under test. The equations can also be
used to predict synergistic levels should pure
erosion and corrosion rates be known.
(ii) Mapping
(Stack and Pungiwiwat (2002) and Stack
et al. (1995, 2003) have used erosion and corro-
sion models with coupling to potential versus
pH (Pourbaix) diagrams and mechanisms asso-
ciated with synergy levels to predict the
performance of various systems using maps.
The maps typically plot particle or slurry velo-
city against potential and show regime regions
and have been presented for several material/
electrolyte systems with electrolytes of various
pHs. Work is underway to develop erosion
corrosion rate maps based on experimental
data and existing mapping approaches.
6.09.4 EROSIONCORROSION TESTING
6.09.4.1 Gravimetric and Coupon Techniques
The manner by which material wastage is
measured in erosioncorrosion is an important
consideration. Experience shows that the use of
weight loss of a tested specimen or coupon is
not a sufficiently accurate way to represent
material wastage. When surfaces are submitted
to both chemically aggressive fluid flows and
solid-particle impingement, several different
weight changes occur, over and above
erosioncorrosion losses. The surfaces of the
specimen, other than the eroded area, can add
or subtract weight from chemical reactions such
as oxidation. Weight can also be added when
erodent particles become embedded into the
surface or become mixed within the surface
oxide layers or scales. The preferred way to
measure erosioncorrosion of gas-particle-
laden flows is by direct depth measurement
using a scribed grid placed onto the ground
surface to be eroded. The grid area thickness
is measured at the grid intersections using a
micrometer before testing. Post testing, the spe-
cimen is cross-sectioned and measured using a
microscope micrometer, relating the thickness
at the location of the deepest penetration to the
pretest grid thickness at the same position. This
process is reproducible to 1 mm (Levy, 1995).
Szyprowski (2003) reviews H
2
S corrosion and
the methods used to determine the efficiency of
inhibitor performance, employed to suppress
such oilfield flow corrosion. Gravimetric techni-
ques based on coupons in pipe flow prepared to
ASTM G4-95 (1995) are reviewed. Gravimetric
techniques should include the rate of deposit for-
mation on the coupons, the corrosion rate, the
depth and rate of formation of corrosion pits, the
presence of hydrogen blowholes, the changes in
mechanical properties of the specimens (the effect
of hydrogen cracking), and investigations on
metal microstructure to find intercrystalline
ErosionCorrosion Testing 419
corrosion. It is also recommended to use electro-
chemical methods that provide information
about the mechanism of inhibition, time of
adsorption, type of adsorption, and so on.
Corrosimetric (resistance) polarization probe
methods for use in refineries are also reviewed.
De Bruyn (Johnston et al., 1996) has reviewed the
modern methods of monitoring corrosion in pet-
rochemical industries, including electrochemical
noise methods.
Mass loss can be assessed using Faradays laws
of electrolysis (eqn [46]), where q is the charge,
F is Faradays constant (96 485 C mol
1
), z is the
number of electrons involved in the reaction, and
Dw is the mass loss (Barik et al., 2004). For
example, for a NAB composition, the calculated
equivalent oxidation state of the cations pro-
duced by dissolution is 2.09 and the equivalent
molar mass (M
m
) is 59.5 g mol
1
:
q = zF
w
M
m
[46[
6.09.4.2 Rotating Systems
A variety of geometries are used with varying
velocity/flow-regime capabilities. Figure 18 illus-
trates some commonly used geometries to study
the flow effects on electrodes of redox reactions.
The working electrode (under polarization
control) here is illustrated by the NAB electrode.
In Figure 18c, the performance of NAB and the
other metal cylindrical electrodes can be exam-
ined. They are used to study corrosion under
controlled hydrodynamic conditions and are
typically designed to simulate pipe flow regimes
and fluid structures. The first three are, however,
unsuitable for solid-particle erosioncorrosion
studies as the solid-particle impingement condi-
tions are difficult to quantify and reproduce. The
wall-jet geometry may lend itself to single-parti-
cle studies of erosioncorrosion, as solution
resistance in such cells inhibits measurement of
large currents induced by slurry erosioncorro-
sion studies. Slurry erosioncorrosion is
therefore studied using a larger-scale free jet or
submerged-jet impingement rig.
Tests carried out using a slurry-jet impinge-
ment rig are typically in kinetic energies ranging
from 0.02 to 0.40 mJ. However, larger facilities
with both free and submerged jets have been
commissioned that operate up to energies of
1 mJ using recirculated slurries. Modifications
to jet impingement rigs can be made to accom-
modate a silver/silver chloride (Ag/AgCl)
reference electrode (RE) and a platinum coun-
ter electrode (CE) (see Figure 19). This design
has a once-through principle for the sand, and
the ejector assembly, used for sand-particle
intake, is located downstream to prevent ero-
sion damage on the counter and reference
electrodes. A valve situated near the ejector
allows the sand intake to be completely isolated
under flow corrosion conditions. Typically,
flow corrosion tests use an electrolyte of a
3.5% NaCl solution, while slurry tests use a
3.5% NaCl solution with a 3% w/w silica sand
concentration. Pure corrosion (C) experiments
were carried out in the absence of slurry,
whereas the erosioncorrosion (T) experiments
Insulating
polymer
sheath
(a)
(c)
(b)
(d)
NAB
electrode
NAB
electrode
NAB
electrode
NAB
electrode
Other metal
(CuNi or Cu)
Jet nozzle
Figure 18 Types of hydrodynamic working
electrodes. Rotating disc electrode (RDE); rotating
cylinder electrode (RCE); bimetallic rotating cylinder
electrode (BRCE); wall-jet disc electrode (WJDE).
From Wharton, J. A., Barik, R. C., Kear, G.,
Wood, R. J. K., Stokes, K. R., and Walsh, F. C.
2005. The corrosion of nickelaluminium bronze in
seawater. A Century of Tafels Equation: A
Commemorative Issue of Corrosion Science 47(12),
33363367.
Computer and
data logging
Electrolyte
Sand suction
Specimen
holder
Ag/AgCl RE
Jet diameter
5.5 mm
Specimen
Pt CE
Figure 19 Erosioncorrosion rig for low-kinetic-
energy-free jet impingements instrumented to make
electrochemical measurements. Exposed specimen
area typically 7 cm
2
. From Tan, K. S., Wharton J. A.,
and Wood, R. J. K. 2005. Solid particle erosion
corrosion behaviour of a novel HVOF nickel
aluminium bronze coating for marine
applications correlation between mass loss and
electrochemical measurements. Wear 258, 629640.
420 Erosion/Corrosion
were carried out with sand. A potentiostat is
used to monitor the sample potential. For pure
erosion (E), a 200 mV cathodic protection
potential was applied, based on the potentials
observed from erosioncorrosion conditions.
6.09.4.3 Flow Systems
Flow systems are typically based on pipe
loop systems with working sections equipped
for electrochemical measurements. Straight
pipe sections aim to reproduce service condi-
tions and can generate a variety of flow
conditions from fully developed laminar to tur-
bulent or can be used to study pipe entry
conditions where the flow is not fully developed
(see Figure 20). Flow loops clearly require asso-
ciated pipework, filters, and pumps, all of
which have to be selected to be inert to the
experiment. Hence, typically nonmetallic
uPVC pipe fittings and components are used.
A typical loop layout is shown in Figure 21.
6.09.5 EROSIONCORROSION
DETECTION
6.09.5.1 Advanced Electrochemical
Monitoring
6.09.5.1.1 Electrochemical noise
As stated before, Sasaki and Burstein (2000)
observed current transients for single-particle
impacts on passive stainless steels, and illu-
strated that monitoring such transients can
provide direct information on the erosion
corrosion processes. Therefore, techniques to
detect the noise level (small perturbations) on
electrochemical corrosion measurements are
being developed. This is called electrochemical
noise analysis or ENA. Noise can be measured
on both current (electrochemical current noise,
ECN) and potential (electrochemical potential
noise, EPN) outputs and subsequent analysis
can yield corrosion resistance details, assuming
both measurements are in phase with each other.
EPN measurements can be made with two elec-
trode cells while current noise typically requires
three electrodes. Further details of this technique
can be found in Cottis and Turgoose (in press).
Such measurements can be made under flow
corrosion and erosioncorrosion and are now
being analyzed by researchers for insight into
synergistic processes and surface performance
indicators. Figure 22 shows a corrosion flow
cell designed for ENA and Figure 23 shows typi-
cal EPNand ECNoutputs for stainless steel pipe
section electrodes seen in Figure 22 subjected to a
flowing NaCl solution at Re = 2000. The
Figure 20 Flow development for laminar and
turbulent flow regimes. From Wharton, J. A. and
Wood, R. J. K. 2004. Influence of flow conditions
on the corrosion of AISI 304L stainless steel. Wear
256, 526536.
Fluid
reservoir
Header
tank
Bypass
loop
Flowmeter
Filter
Test section
60 pipe diameters
Fluid
reservoir
P
P
Figure 21 Recirculating low loop schematic. The
fluid volume of the flow loop was ,180 l. From
Wharton, J. A. and Wood, R. J. K. 2004. Influence
of flow conditions on the corrosion of AISI 304L
stainless steel. Wear 256, 526536.
Working
electrodes
ZRA
ECN
EPN
Reference electrode
2
8

m
m
Data logging
Flow
Spacer
Assembly clamping rods
Figure 22 Schematic of electrochemical flow cell
and measurement setup. From Wharton, J. A. and
Wood, R. J. K. 2004. Influence of flow conditions on
the corrosion of AISI 304L stainless steel. Wear 256,
526536.
ErosionCorrosion Detection 421
features seen relate to metastable and stable pit-
ting activity on the wetted surfaces of the
stainless steel electrodes.
A simple way to analyze noise data is to
take the standard deviation of the current
traces under flow corrosion, o
c
, and under
erosioncorrosion, o
T
. Attempts are currently
being made to link the standard deviation
ratios of the ECN measurements to synergy
to gain a further understanding of the effects
of erosion-enhanced corrosion, DC
E
.
Although individual events are not likely to
be resolved, the result of multiple impacts as
a function of time may be fruitful
(ZRA = zero-resistant ammeter).
However, caution should be applied to any
electrochemical measurement under erosion
corrosion conditions. Other issues make elec-
trochemical analysis and its comparison to
synergy difficult. These include the possibility
of local film currents between anodes/cathodes
(Oltra et al., 1995), which will not be seen by
ECN measurements, and the effects of
charging/recharging double-layer currents due
to fluctuating local events.
Table 6 lists the possible influences on the
depassivation and repassivation kinetics and
thus the sources of electrochemical noise.
6.09.6 EROSIONCORROSION-
RESISTANT MATERIAL AND
SURFACE SELECTION
Material selection for erosioncorrosion
duty tends to be based on the systems of alumi-
num, copper, zinc, and nickel due to their
relative cheapness. More relatively expensive
options are stainless steels or polymers (where
applicable), Co-based alloys (e.g., Stellite), Ni-
based alloys (e.g., Inconel), hardmetals or cer-
mets (e.g., sintered WC), ceramics and coatings
of these systems unto substrates of CS or low-
corrosion-resistant alloys (Neville et al., 2002;
Wood and Speyer, 2004; Puget et al., 1999;
Scrivani et al., 2001; Levin et al., 1995; Gee
et al., 2005; Gant et al., 2004a, 2004b). Surface
degradation control techniques other than sur-
face selection may well include the use of
inhibitors as well as controlling fluid tempera-
ture, pH, and dissolved oxygen levels. For
example, Poulson (1999) suggests a high pH
(>9), and dissolved oxygen levels as low as
10 ppb have been successfully used to control
flow-induced (solids-free) flow corrosion of CS.
Process parameters should also be reviewed to
minimize erosion and the use of sand-removal
(screens and cyclone separators) techniques and
sand-monitoring systems (acoustic) should be
deployed, where practicable, to reduce the poten-
tial for erosioncorrosion (Madge et al., 2004;
Mylvaganam, 2003; Al-Shammaa et al., 2003).
An overview of material selection for ero-
sioncorrosion was presented by the author at
the 2nd International Symposium on Advanced
Materials for Fluid Machinery at the IMechE
in February 2004 (Wood, 2004).
0.09
0.06
0.03
0
0.03
2
0
2
4
5 5.5 6
Time (h)
C
u
r
r
e
n
t

(

A
)
Re 2000
P
o
t
e
n
t
i
a
l

(
V
)
6.5 7
Figure 23 Electrochemical current and potential
noise for AISI 304L in 3.5 wt.% NaCl solution at a
transition Reynolds number of 2000. From
Wharton, J. A. and Wood, R. J. K. 2004. Influence
of flow conditions on the corrosion of AISI 304L
stainless steel. Wear 256, 526536.
Table 6 Possible influences on the depassivation and repassivation kinetics
Environmental Factor Depassivation Repassivation
Solid-particle impact Removes passive film Activates repassivation
Surface/liquid
interface
High velocities/wall shear
stresses
Film growth kinetics (monolayer coverage plus
lateral film growth)
Occluded geometries/
pits
Local environment changes
such as low pH
Metastable pits
Temperature Local temperature Local temperature
Turbulence bulk flow
structure
Turbulent flow features Rougher/larger surface area
Wood, R. J. K. 2004. Challenges of living with erosioncorrosion. Second International Symposium on Advanced Materials
for Fluid Machinery, IMechE Conference Transations (1), paper S965/008/2004, 113132, ISBN 1 86058 441 1. Professional
Engineering Publishing, Bury St Edmunds.
422 Erosion/Corrosion
Recent research into the erosioncorrosion
of polymer coatings and HVOF aluminum
and NAB coatings are reviewed in Moore and
Wood (1992) and Tan et al. (2005). This sug-
gests that medium-density polyethylene
(MDPE) surfaces should be considered, where
appropriate, as erosioncorrosion surfaces.
Figure 24 shows the erosioncorrosion rates
(T) of the cast NAB, 90/10 Cu/Ni, HVOF
NAB, plasma electrolytic oxidation (PEO) alu-
mina coatings, MDPE, and the 4360 steel
substrate plotted against particle energy, E
k
.
For comparison, data for MDPE pipe mate-
rial, taken from (Moore and Wood, 1992), are
also plotted on Figure 24. These results showthat
MDPE outperforms all the other surfaces tested
and is an attractive option for resisting slurry
erosion in corrosive media where fluid tempera-
tures permit selection. However, it has a high
sensitivity to impact energy with an energy expo-
nent over 1.5 (i.e., V
u
c U
3
) (see Table 7).
For low-energy solid-particle flows, materials
selection could be based on the degree of passivity
a surface offers under the design environmental
conditions it will be subjected to. For example,
alloys that strongly passivate with very adherent
films, such as titanium and some Ni/Cr/Mo
alloys, could be considered for some erosion
corrosion duties. Alloys such as steels and some
copper-based alloys that form protective or
semiprotective films that are easily removed by
flow or particle impingement may be selected for
very low-particle-energy flows but will be prone
to severe material wastage as the particle energies
and/or flow velocities are increased. However, it
is not only the dynamic service conditions that
should also be used for material selection.
Shutdown periods can allow solids to sediment
out of suspension, and accumulation of these on
surfaces such as some stainless steels and nickel-
based alloys can promote localized corrosion
such as pitting or crevice corrosion.
6.09.7 CONCLUSIONS
As demonstrated within this chapter, mate-
rial performance is highly system dependent
and modest changes in environmental condi-
tions can have a significant impact on wall
wastage rates. The discussion throughout this
chapter also confirms that our level of under-
standing of the mechanism involved and
controlling parameters is too low to enable
informed material or coating selection for
such aggressive duties. Therefore, selection
must be accompanied by either relevant experi-
ence of good material performance under
similar erosioncorrosion conditions or rele-
vant experimental results from carefully
constructed laboratory testing that has proved
to simulate field conditions to allow material
and coating screening. For a review of flow-
assisted corrosion, the reader is also referred
to a recent National Association of Corrosion
Engineers (NACE) book by Roberge (2004)
and a review chapter by Craig (1991).
This chapter has detailed the fundamental
aspects currently known to be of importance
to erosioncorrosion. The level of understand-
ing is incomplete and some of the complexities
of erosioncorrosion processes have been high-
lighted; there is a need for serious research
effort to gain full understanding. Therefore,
the overall conclusion, at present, is that it is
difficult to give generic statements of how ero-
sioncorrosion impacts upon structural
0.000 1
0.001
0.01
0.1
1
10
0.01 0.1 1
E
k
(J)
V
u

(

m
3
/
i
m
p
a
c
t
)
Vu alumina
Vu PEO
Vu steel
Vu HVOF NAB
Vu NAB
Vu 9010 CuNi
Vu MDPE
Figure 24 Relationship between erosioncorrosion
rate V
u
and kinetic energy E
k
under sand and 3.5%
NaCl solution slurry jet impingement at room
temperature and 90

free jet impingement angle.


Table 7 Erosion rate V
u
dependencies on impact energy and correlation coefficients for the various surfaces
tested under slurry jet erosioncorrosion conditions at 90

impact angle, as shown in Figure 24


Material/coating Power law Correlation coefficient R
2
Carbon steel AISI 1020 V
u
= 4.06E
k
0.87
0.93
Cast NAB V
u
= 1.06E
k
0.93
0.97
HVOF NAB coating on steel V
u
= 4.11E
k
1.18
0.99
PEO alumina coating on aluminum V
u
= 2.66E
k
1.44
NA
MDPE V
u
= 0.59E
k
1.52
0.99
90/10 Cu/Ni V
u
= 1.48E
k
1.12
0.99
Pipe-grade alumina V
u
= 0.157E
k
1.91
NA
Conclusions 423
integrity. Erosioncorrosion is extremely sys-
tem-process sensitive and can result in having
no influence at all on the component life, but
more often it can be the prime cause of prema-
ture component failure, loss of containment, or
system functionality. There are no erosioncor-
rosion models available, although some basic
corrosion enhancement models have been
embedded into commercial CFD particle-track-
ing codes to predict pipe wall wastage rates.
Such models are also system specific and need
to be calibrated (experimentally) before results
are used by designers.
A good example of the above is seen in pre-
dicting the structural integrity associated with
pipe systems handling corrosive fluids and cor-
rosive slurries. Most available data from the
literature are specific to the test rig geometry
used. This results in a deficiency of data to
allow accurate life prediction of specific pipe
systems operating under specific process condi-
tions. Thus, at present, the designers and
operators of specific systems cannot avoid
doing experimental tests on pipe loop systems
to gain accurate wear rates. This is primarily due
to the number of parameters to consider and the
complexity of the wear mechanisms and the lack
of comprehensive computational tools to esti-
mate wear. Knowledge from the literature can
be used to design experiments to simulate pro-
cess damage; however, scaling up wear rates to
full scale is problematic. Therefore, full-scale
tests are recommended where possible.
Computational evaluation of wear without
experimental calibration is not recommended.
Experimentally validated subscale wear predic-
tions increase the confidence of full-scale wear
rate estimation but are not currently robust.
Such sentiments are echoed by recent reports
for projects trying to assess the erosioncorro-
sion levels of pipe systems that are being
designed for handling and processing radioac-
tive slurries of nuclear waste (Duignan and Lee,
2002; Wood and Wharton, 2002).
Online techniques applicable for structural
integrity monitoring of components exposed to
erosioncorrosion, such as the use of test cou-
pons or electrochemical analysis, are growing in
popularity but are far from robust and standa-
lone. Considerable expertise is required to
analyze the results or output signatures and
some debate surrounds the interpretation of
electrochemical responses. Offshore oil and gas
process streams often use acoustic sand detectors
to alert operators to potential erosioncorrosion
issues. These sensors can also evaluate the colli-
sion energy (erosion effectiveness of the solid
particles known to be in the flow). But confi-
dence in such techniques has been undermined
by multiple false alarms and unexpected loss of
contaminant due to sand production that was
not detected by such sensors.
The selection of new materials and coatings
for erosioncorrosion control is also hampered
by incomplete wear rate data and again the test
conditions and geometry used to generate such
test data will probably not match those of the
application. Where data do match the material
composition under consideration, care should
be taken that the manufacturing and processing
routes are also identical (i.e., heat treatments,
cast or forged, descaled or not). Performance
mapping of surfaces over a wide range of ero-
sioncorrosion conditions is also helpful and
can be used to guide selection of materials and
process conditions to increase component life.
The effects of erosioncorrosion can be miti-
gated if hard and nonconducting surfaces can
be used. Ceramic and polymer surfaces, for
example, could be used if compatible with the
processes involved. Chemical treatments with
corrosion inhibitors, which coat wetted sur-
faces, can be used for low-energy flow systems
but they are not optimized for inhibiting ero-
sioncorrosion. They can be readily stripped
from the surface when solid impact occurs.
However, certain points for guidance can be
made to reduce or minimize erosioncorrosion
and increase structural integrity:
1. The galvanic compatibility of coupled dis-
similar materials should be checked for liquid
handling systems.
2. The compatibility of component materials
with the process environment for both operat-
ing and nonoperational conditions should be
checked.
3. For solid-handling systems (i.e., slurry
systems), the trajectory, energy, and concentra-
tion of solid particles must be considered for
accurate predictions of wear rates and
locations.
4. Wherever possible, component materials
should be tested using full-scale experimental
facilities that accurately simulate the service
conditions of the system under consideration.
5. Quantification of the synergistic interac-
tions between mechanical erosive wear
processes and electrochemical corrosive pro-
cesses is recommended. This should allow
correct wall wastage rates to be predicted and
the safety factors associated with component
wall thicknesses to be adjusted.
6. Inherently passive surfaces should be
selected, where appropriate, with knowledge
of their repassivation kinetics and properties.
7. Sacrificial components (bends, tees,
elbows) can be used in systems that allow easy
replacement and are relatively cheap.
424 Erosion/Corrosion
Special geometries, such as plugged tees or
flow expansion sections, can be used that mini-
mize erosion by reducing the energy of solid-
particle impingement by decelerating the flow
field or the impingement angle.
6.09.8 REFERENCES
Ahmed, M., Singh, S. N., and Seshadri, V. 1993.
Distribution of solid particles in multisized particulate
slurry flow through a 90

pipe bend in horizontal plane.


Bulk Solids Handl. 13(2), 379385.
Al-Hosani, H. I., Saber, T. M. H., Mohammed, R. A., and
Shams El Din, A. M. 1997. Galvanic corrosion of copper
base alloys in contact with molybdenum-containing
stainless steels in Arabian Gulf water. Desalination 109,
2537.
Allen, C. and Ball, A. 1996. A review of the performance of
engineering materials under prevalent tribological and
wear situations in South African industries. Tribol. Int.
29(2), 105116.
Al-Shammaa, A. I., Tanner, R., Shaw, A., and Lucas,
J. 2003. On line EM wave sand monitoring sensor
for oil industry. In: 33rd European Microwave
Conference, IEEE Conference Proceedings, vols.
13, pp. 535538.
ASTM G 119-93. 1993. Standard Guide for Determining
Synergism between Wear and Corrosion. American
Society for Testing and Materials, West Conshohocken,
PA.
ASTM G4-95. 1995. Standard Guide for conducting corro-
sion coupon tests in field applications. American Society
for Testing and Materials, West Conshohocken, PA.
Barik, R. C., Wharton, J. A., Wood, R. J. K., and
Stokes, K. R. 2004. Galvanic corrosion of nickel
aluminium bronze coupled to titanium or Cu15Ni alloy
in brackish seawater. Eurocorr 2004, Nice, paper 330.
Bergevin, K. 1984. Effect of slurry velocity on the mechan-
ical and electrochemical components of erosion
corrosion in vertical pipes. Masters thesis, University of
Saskatchewan.
Bitter, J. G. A. 1963. A study of erosion phenomena. Wear 6,
521.
Blackwelder, R. F. 1989. Some ideas on the control of near-
wall eddies. AIAA, paper 891009.
Blanchard, D. J., Griffith, P., and Rabinowicz, E. 1984.
Erosion of a pipe bend by solid particles entrained into
water. J. Eng. Ind. 106, 213217.
Bourgoyne, A. T. 1989. Experimental study of erosion in
diverter systems due to sand production. SPE/IADC
Drilling Conference, p. 807.
Bozzini, B., Ricotti, M. E., Boniardi, M., and Mele, C.
2003. Evaluation of erosioncorrosion in multiphase
flow via CFD and experimental analysis. Wear 255,
237245.
Cantwell, B. J. 1981. Organized notion in the turbulent
boundary layer. Annu. Rev. Fluid Mech. 13, 457515.
Carpentier, B. and Cerf, O. 1993. Biofilms and their con-
sequences, with particular reference to hygiene in the
food industry. J. Appl. Bacteriol. 75, 499511.
Chen, X., McLaury, B. S., and Shirazi, S. A. 2002.
Proceedings of ASME 2002 Fluids Engineering Division
Summer Meeting. Montreal, QC, paper FEDSM2002-
31289.
Chen, X. H., McLaury, B. S., and Shirazi, S. A. 2004.
Application and experimental validation of a computa-
tional fluid dynamics (CFD)-based erosion prediction
model in elbows and plugged tees. Comput. Fluids
33(10), 12511272.
Clark, H. M. and Llewellyn, R. J. 2001. Assessment of the
erosion resistance of steels used for slurry handling and
transport in mineral processing applications. Wear 250,
3244.
Craig, B. D. 1991. Fundamental Aspects of Corrosion Films
in Corrosion Science, Chapter 7, pp. 165190. Plenum,
New York.
Cottis, R. and Turgoose, S. 1999. Electrochemical
Impedance and Noise, ISBN 1-57590-093-9. National
Association of Corrosion Engineers, Houston.
Donohue, G. L., Tiedermann, W. G., and Reischman,
M. M. 1972. Flow visualization of the near-wall region
in a drag reducing channel flow. Fluid Mech. 56, 559575.
Douglas, J. F., Gasiorek, J. M., Swaffield, J. A., and
Jack, L. B. 2005. Fluid Mechanics, 5th ed. Pearson,
Harlow.
Duddridge, J. E., Kent, C. A., and Laws, J. F. 1982. Effect
of surface shear stress on the attachment of Pseudomonas
fluorescens to stainless steel under defined flow condi-
tions. Biotechnol. Bioeng. 24, 153164.
Duignan, M. R. and Lee, S. Y. 2002. RPP-WTP Slurry
Wear Evaluation: Literature Review, WSRC-TR-2001-
00156. Westinghouse Savannah River Company, Aiken,
SC. http://sti.srs.gov/fulltext.
Faddick, R. R. 1982. Wear in Pipes: Short Course on Slurry
Pipelining Technology, pp. 115. Camborne School of
Mines, University of Exeter, Devon.
Forder, A. A. 2000. Computational Fluid Dynamics
Investigation into the Particulate Erosion of Oilfield
Control Valves. Ph.D. thesis, School of Engineering
Sciences, University of Southampton.
Forder, A., Thew, M. T., and Harrison, D. 1998. A numer-
ical investigation of solid particle erosion experiences
within oilfield control valves. Wear 216, 184193.
Ferng, Y. M., Ma, Y. P., Ma, K. T., and Chung, N. M.
1999. A new approach for investigation of erosioncorro-
sion using local flow models. Corrosion 55(4), 332342.
Ferng, Y. M., Ma, Y. P., and Chung, N. M. 2000.
Application of local flow models in predicting distributions
of erosioncorrosion locations. Corrosion 56(2), 116126.
Finlay, J. A., Callow, M. E., Schultz, M. P., Swain, G. W.,
and Callow, J. A. 2002. Adhesion strength of settled
spores of the green alga Enteromorpha. Biofouling 18,
251256.
Finnie, I. 1972. Some observations on the erosion of ductile
metals.. Wear 19, 8190.
Flemming, H. C. 2002. Biofouling in water systems cases,
causes and countermeasures. Appl. Microbiol. Biot. 59,
629640.
Gane, N. and Murray, M. S. 1979. The transition from
ploughing to cutting in erosive wear. Proceedings of the
5th International Conference on Erosion by Solid and
Liquid Impact. Cambridge, UK, paper 40, pp. 18.
Gant, A. J., Gee, M. G., and May, A. T. 2004a.
Microabrasion of WC-Co hardmetals in corrosive
media. Wear 256(910), 500516.
Gant, A. J., Gee, M. G., and May, A. T. 2004b. The
evaluation of tribo-corrosion synergy for WC-Co hard-
metals in low stress abraision. Wear 256(5), 500516.
Gee, M. G., Phatak, C., and Darling, R. 2005.
Determination of wear mechanisms by stepwise erosion
and stereological analysis. Wear 258(104), 412425.
Geesey, G. G., Jang, L., Jolley, J. G., Hankins, M. R.,
Iwaoka, T., and Griffiths, P. R. 1988. Binding of metal
ions by extracellular polymers of biofilm bacteria. Water
Sci. Technol. 20, 161165.
Guo, H. X., Lu, B. T., and Luo, J. L. 2005. Interaction of
mechanical and electrochemical factors in erosioncorro-
sion of carbon steel. Electrochim. Acta 51(2), 315323.
Harvey, T. J., Wharton, J. A., and Wood, R. J. K. 2006.
Development of a model for predicting erosioncorro-
sion in a slurry pot. Report for NPL. Surface and
Coatings Technology (submitted).
Hashish, M. 1998. An improved model of erosion by solid
particles. Proceedings of the 7th International
References 425
Conference on Erosion by Liquid and Solid Impact,
paper 66. Cavendish Laboratory, Cambridge, UK.
Haugen, K., Kvernvold, O., Ronald, A., and Sandberg, R.
1995. Sand erosion of wear-resistant materials: Erosion in
choke valves. Wear 186187, 179188.
Hodgkiess, T., Neville, A., and Shrestha, S. 1999.
Electrochemical and mechanical interactions during ero-
sioncorrosion of a high-velocity oxy-fuel coating and a
stainless steel. Wear 235, 623634.
Hu, X. and Neville, A. 2005. The electrochemical response
of stainless steels in liquidsolid impingement. Wear
258(14), 641648.
Hutchings, I. M. 1992. Tribology Friction and Wear of
Engineering Materials. Arnold, London.
Jemmely, P., Mischler, S., and Landolt, D. 2000.
Electrochemical modelling of passivation phenomena in
tribocorrosion. Wear 237, 6376.
Johnston, J. M., Rathmell, S. K., Easterbrook-Smith, S. B.,
Wilson, M. R., and De Bruyn, H. J. 1996. Current corro-
sion monitoring trends in the petrochemical industry.
Int. J. Pressure Vessels Piping 66, 293303.
Jucker, B. A., Zehnder, A. J. B., and Harms, H. 1998.
Quantification of polymer interactions in bacteria adhe-
sion. Environ. Sci. Technol. 32, 29092915.
Karimi, A. and Schmid, R. K. 1992. Ripple formation in
solidliquid erosion. Wear 156, 3347.
Keating, A. and Nesic, S. 2001. Numerical prediction of
erosioncorrosion in bends. Corrosion 57(7), 621633.
King, R., Jacobs, B., and Jones, G. 1991. Factors affecting
the design of slurry transport systems for minimum wear.
Pipe Protection Conference, Cannes, p. 67.
Kim, H. T., Kline, S. J., and Reynolds, W. C. 1971. The
production of turbulence near a smooth wall in a turbu-
lent boundary layer. J. Fluid Mech. 50, 133160.
Kim, J. G., Choi, Y. S., Lee, H. D., and Chung, W. S. 2003.
Effects of flow velocity, pH, and temperature on galvanic
corrosion in alkaline-chloride solutions. Corrosion 59(2),
121129.
Kline, S. J., Reynolds, W. C., Schraub, F. A., and
Rundstrandler, P. 1967. The structure of turbulent
boundary layers. J. Fluid Mech. 30, 741773.
Levin, B. F., Dupont, J. N., and Marder, A. R. 1995. Weld
overlay coatings for erosion control. Wear 181(2),
810820.
Levy, A. V. 1995. Solid Particle Erosion And Erosion
Corrosion of Materials. ASM International, OH, ISBN:
0-87170-519-2.
Loeb, G. I. and Neihof, R. A. 1977. Adsorption of an
organic film at a platinum seawater interface. J. Mar.
Res. 35, 283291.
Madge, D. N., Romero, J., and Strand, W. L. 2004.
Hydrocarbon cyclones in hydrophilic oil sand environ-
ments. Miner. Eng. 17(5), 625636.
Matthews, A. 2005. The future of plasma-based surface
engineering techniques in tribology, 3rd World
Tribology Congress, Washington DC, paper WTC2005-
64180.
McIntyre, P. April 1999. Marine Corrosion Club Meeting,
Aberdeen.
Melo, L. F. and Vieira, M. J. 1999. Physical stability and
biological activity of biofilms under turbulent flow and
low substrate concentration. Bioprocess Eng. 20,
363368.
Meng, H. C. and Ludema, K. C. 1995. Wear models and
predictive equations: Their form and content. Wear
181183, 443457.
Moore, A. J. and Wood, R. J. K. 1992. Erosive wear map-
ping of pipeline materials. Plastic Pipes VIII Conference,
Koningshof, paper E1/4, pp. 110.
Muller, R. F., Characklis, W. G., Jones, W. L., and Sears,
J. T. 1992. Characterization of initial events in bacterial
surface colonization by two Pseudomonas species using
image analysis. Biotechnol. Bioeng. 39, 11611170.
Mylvaganam, S. 2003. Some applications of acoustic emis-
sion in particle science and technology. Particulate Sci.
Technology 21(3), 293301.
Neilson, J. H. and Gilchrist, A. 1968. An experimental
investigation into aspects of erosion of rocket motor tail
nozzles. Wear 11, 123143.
Nesic, S., Bienkowski, J., Bremhorst, K., and Yang, K. S.
2000. Testing for erosioncorrosion under disturbed flow
conditions using a rotating cylinder with a stepped sur-
face. Corrosion 56(10), 10051014.
Neville, A. and Hu, X. 2001. Mechanical and electrochemi-
cal interactions during liquidsolid impingement on high-
alloy stainless steels. Wear 251, 12841294.
Neville, A., Reyes, M., and Xu, H. 2002. Examining corro-
sion effects and corrosion/erosion interactions on
metallic materials in aqueous slurries. Tribol. Int.
35(10), 643650.
Oltra, R., Chapey, B., and Renuad, L. 1995. Abrasion
corrosion studies of passive stainless steels in acidic
media: Combination of acoustic emission and electroche-
mical techniques. Wear 186187, 533541.
Oxford English Dictionary Online. 2006. Oxford University
Press.
Paradies, H. H. 1995. Chemical and physicochemical
aspects of metal biofilm. In: Bioextraction and
Biodeterioration of Metals (eds. C. C. Gaylarde and
H. A. Videla), pp. 197269. Cambridge University
Press, Cambridge.
Panton, R. L. 1984. Incomprehensible Flows, p. 717. Wiley/
Interscience, New York.
Percy, R. J. 1990. The operational performance of BP choke
valves, BP engineering, risers and subsea equipment
group. BPE.90.ER152 (unpublished).
Pletcher, D. 1991. A First Course in Electrode Processes.
The Electrochemical Consultancy, Romsey.
Poulson, B. 1999. Complexities in predicting erosioncorro-
sion. Wear 233235, 497504.
Puget, Y., Wood, R. J. K., and Trethewey, K. R. 1999.
Electrochemical noise analysis of polyurethane coated steel
subjected to erosioncorrosion. Wear 233235, 552567.
Puget, Y., Wood, R. J. K., and Trethewey, K. R. 1999.
Electrochemical noise analysis of polyurethane coated steel
subjected to erosioncorrosion. Wear 233235, 552567.
Ra cz, I. G., Wassink, J. G., and Klaassen, R. 1986. Mass
transfer, fluid flow and membrane properties in flat and
corrugated plate hyperfiltration modules. Desalination
60, 213222.
Robinson, S. K. 1991. Coherent motions in the turbulent
boundary layer. Annu. Rev. Fluid Motion 23, 601639.
Roberge, P. 2004. ErosionCorrosion. Corrosion Testing
Made Easy Series. NACE International, Houston.
Roe, F. L., Lewandowski, Z., and Funk, T. 1996.
Simulating microbiologically influenced corrosion by
depositing extracellular biopolymer on mild steel sur-
faces. Corrosion 52, 744752.
Sasaki, K. and Burstein, G. T. 2000. Observation of a
threshold impact energy required to cause passive film
rupture during slurry erosion of stainless steel. Phil. Mag.
Lett. 80(7), 489493.
Scrivani, A., Ianelli, S., Rossi, A., Groppetti, R., Casadei,
F., and Rizzi, G. 2001. A contribution to the surface
analysis and characterisation of HVOF coatings for pet-
rochemical application. Wear 250(1), 107113.
Sedamed, G. H., Abdo, M. S. E., Amder, M., and El-Latif,
G. A. 1998. Mass transfer at a pipe inlet zone in relation to
impingement. Int. Comm. Heat Mass Transfer 443451.
Shimoda, K. and Yukawa, T. 1983. Erosion of pipe bend in
pneumatic conveyer. Proceedings of the 6th International
Conference on Erosion by Liquid and Solid Impact,
University of Cambridge, paper 59.
Silverman, D. C. 2004. The rotating cylinder electrode for
examining velocity-sensitive corrosion a review.
Corrosion 60(11), 10031023.
426 Erosion/Corrosion
Souza, V. A. D. and Neville, A. 2005. Corrosion
and synergy in a WC-Co-Cr HVOF thermal spray coat-
ing understanding their role in erosioncorrosion
degradation. Wear 259(16), 171180.
Stack, M. M. and Jana, B. D. 2005. Modelling particulate
erosioncorrosion regime transitions for Al/Al2O3 and
Cu/Al2O3 MMCs in aqueous conditions. Tribol. Int.
38(1112), 9951006.
Stack, M. M. and Pungiwiwat, N. 2002. Particulate ero-
sioncorrosion of Al in aqueous conditions: Some
perspectives on pH effects on the erosioncorrosion
map. Tribol. Int. 35(10), 651660.
Stack, M. M., Corlett, N., and Turgoose, S. 2003. Some
thoughts on modelling the effects of oxygen and particle
concentration on the erosioncorrosion of steels in aqu-
eous slurries. Wear (Part 1) 255, 225236.
Stack, M. M., Zhou, S., and Newman, R. C. 1995.
Identification of transitions in erosioncorrosion regimes
in aqueous environments. Wear 186187, 523532.
Stemp, M., Mischler, S., and Landolt, D. 2003. The effect of
mechanical and electrochemical parameters on the tribo-
corrosion rate of stainless steel in sulphuric acid. Wear
255(16), 466475.
Szyprowski, A. J. 2003. Methods of investigation on hydro-
gen sulphide corrosion of steel and its inhibitors.
Corrosion 59(1), 6881.
Tan, K. S., Wharton, J. A., and Wood, R. J. K. 2005. Solid
particle erosioncorrosion behaviour of a novel HVOF
nickel aluminium bronze coating for marine applica-
tionscorrelation between mass loss and
electrochemical measurements. Wear 258, 629640.
Taylor, G. T., Zheng, D., Lee, M., Troy, P. J., Gyananath,
G., and Sharma, S. K. 1997. Influence of surface proper-
ties on the accumulation of conditioning films and marine
bacteria on substrata exposed to oligotrophic waters.
Biofouling 11, 3157.
Trethewey, K. R. and Chamberlain, J. 1995. Corrosion for
Science and Engineering, 2nd edn. Addison Wesley
Longman, London.
Turchaninov, S. P. 1973. The Life of Hydrotransport
Pipelines. Nedra Press, Moscow.
Walsh, D., Pope, D., Danford, M., and Huff, T. 1993. The
effect of microstructure on microbiologically influenced
corrosion. J. Min. Met. Mater. Soc. 45, 2230.
Wang, H. W. and Stack, M. M. 2000. The erosive wear of
mild steel and stainless steels under controlled corrosion
in alkaline slurries containing alumina particles. J. Mater.
Sci. 35, 52635273.
Wharton, J. A. and Wood, R. J. K. 2004. Influence of flow
conditions on the corrosion of AISI 304L stainless steel.
Wear 256, 525536.
Wharton, J. A., Barik, R. C., Kear, G., Wood, R. J. K.,
Stokes, K. R., and Walsh, F. C. 2005. The corrosion of
nickelaluminium bronze in seawater. A Century of
Tafels Equation: A Commemorative Issue of Corrosion
Science 47(12), 33363367.
Wheeler, D. W. and Wood, R. J. K. 2005. Erosion of hard
surface coatings for use in offshore gate valves. Wear 258,
526536.
Wiedenroth, W. 1984. An experimental study of wear of
centrifugal pumps and pipeline components. J. Pipelines
4, 223228.
Wijeyekoon, S., Mino, T., Satoh, H., and Matsuo, T. 2000.
Growth and novel structural features of tubular biofilms
produced under different hydrodynamic conditions.
Water Sci. Technol. 41, 129138.
Wood, R. J. K. 1992. Erosioncorrosion synergism for
multi-phase flowline materials. La Houille Blanche 78,
605610.
Wood, R. J. K. 2004. Challenges of living with erosion
corrosion. Second International Symposium on
Advanced Materials for Fluid Machinery, IMechE
Conference Transactions (1), paper S965/008/2004,
113132, ISBN 1 86058 441 1. Professional Engineering
Publishing, Bury St Edmunds.
Wood, R. J. K. 2004a. Erosioncorrosion interactions and
their effects on marine and offshore components.
Keynote at EuroCorr Conference, Nice, France and
accepted for publication in the Special Issue on
Tribocorrosion, Wear 2006.
Wood, R. J. K. and Hutton, S. P. 1990. The synergistic
effect of erosion and corrosion: Trends in published
results. Wear 140, 387394.
Wood, R. J. K. and Jones, T. F. 2003. Investigations of
sandwater induced erosive wear of AISI 304L stainless
steel pipes by pilot-scale and laboratory-scale testing.
Wear 255, 206218.
Wood, R. J. K. and Wharton, J. A. 2002. The influence
of chloride concentrations on the erosioncorrosion of
AISI 304L stainless steel in alkaline slurries. 15th
International Conference on Hydrotransport, Banff,
vol. 1, 173186.
Wood, R. J. K. and Speyer, A. J. 2004. Erosioncorrosion
of candidate HVOF aluminium based marine coatings.
Wear 256(5), 545556.
Wood, R. J. K., Hutton, S. P., and Schiffrin, D. J. 1990.
Mass transfer effects of non-cavitating seawater on the
corrosion of Cu and 7030 CuNi. Corros. Sci. J. 30(12),
11771201.
Wood, R. J. K., Mellor, B. G., and Binfield, M. L. 1997.
Sand erosion performance of detonation gun applied
tungsten carbide/cobaltchromium coatings. Wear 211,
7083.
Wood, R. J. K., Jones, T. F., Miles, N. J., and
Ganeshalingam, J. 2001. Upstream swirl-induction for
reduction of erosion damage from slurries in pipeline
bends. Wear 250(112), 771779.
Wood, R. J. K., Jones, T. F., Ganeshalingam, J., and Wang,
M. 2002a. Erosion modelling of swirling and non-swir-
ling slurries in pipes. Hydrotransport 15, Banff, BHR
Group, 497.
Wood, R. J. K., Jones, T. F., and Ganeshalingam, J.
2002b. Erosion in swirl inducing pipes. ASME
Fluids Engineering Division Summer Meeting,
Montreal, paper FEDSM2002-31287, ASME
International.
Wood, R. J. K., Wharton, J. A., Speyer, A. J., and Tan,
K. S. 2002c. Investigation of erosioncorrosion pro-
cesses using electrochemical noise. Tribol. Int. 35,
631641.
Wood, R. J. K., Jones, T. F., Ganeshalingam, J., and Miles,
N. J. 2004. Comparison of predicted and experimental
erosion estimates in slurry ducts. Wear 256(910),
937947.
Copyright 2007, Elsevier Ltd. All Rights Reserved. Comprehensive Structural Integrity
No part of this publication may be reproduced, stored in any retrieval system or ISBN (set): 0-08-043749-4
transmitted in any form or by any means: electronic, electrostatic, magnetic tape,
mechanical, photocopying, recording or otherwise, without permission in writing Volume 6; (ISBN: 978-0-0804-3749-1); pg. 395427
from the publishers.
References 427

Você também pode gostar