Вы находитесь на странице: 1из 15

Downloaded from rspa.royalsocietypublishing.

org on November 22, 2012

On the thermodynamics of degradation


M.D Bryant, M.M Khonsari and F.F Ling
Proc. R. Soc. A 2008 464, 2001-2014
doi: 10.1098/rspa.2007.0371

References This article cites 42 articles, 1 of which can be accessed free


http://rspa.royalsocietypublishing.org/content/464/2096/2001.full.
html#ref-list-1

Article cited in:


http://rspa.royalsocietypublishing.org/content/464/2096/2001.full.html#re
lated-urls

Subject collections Articles on similar topics can be found in the following collections

mechanical engineering (144 articles)

Email alerting service Receive free email alerts when new articles cite this article - sign up in
the box at the top right-hand corner of the article or click here

To subscribe to Proc. R. Soc. A go to: http://rspa.royalsocietypublishing.org/subscriptions


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Proc. R. Soc. A (2008) 464, 2001–2014


doi:10.1098/rspa.2007.0371
Published online 8 April 2008

On the thermodynamics of degradation


B Y M. D. B RYANT 1 , M. M. K HONSARI 2, * AND F. F. L ING 1,3
1
Department of Mechanical Engineering, University of Texas,
Austin, TX 78712, USA
2
Department of Mechanical Engineering, Louisiana State University,
Baton Rouge, LA 70803, USA
3
Rensselaer Polytechnic Institute, 250 South End Avenue,
New York, NY 10280, USA

The science base that underlies modelling and analysis of machine reliability has remained
substantially unchanged for decades. Therefore, it is not surprising that a significant
gap exists between available machinery technology and science to capture degradation
dynamics for prediction of failure. Further, there is a lack of a systematic technique for the
development of accelerated failure testing of machinery components. This article develops a
thermodynamic characterization of degradation dynamics, which employs entropy, a
measure of thermodynamic disorder, as the fundamental measure of degradation; this
relates entropy generation to irreversible degradation and shows that components of
material degradation can be related to the production of corresponding thermodynamic
entropy by the irreversible dissipative processes that characterize the degradation. A
theorem that relates entropy generation to irreversible degradation, via generalized thermo-
dynamic forces and degradation forces, is constructed. This theorem provides the basis of a
structured method for formulating degradation models consistent with the laws of thermo-
dynamics. Applications of the theorem to problems involving sliding wear and fretting wear,
caused by effects of friction and associated with tribological components, are presented.
Keywords: thermodynamic entropy; degradation; ageing; wear; friction; tribology

1. Introduction

Manufacturing transforms nature’s raw materials into highly organized finished


components, reduces entropy and increases thermodynamic energies. Ageing or
degradation, which tends to return these components back to their natural state,
must increase entropy and reduce thermodynamic energies (Feinberg & Widom
1995, 1996, 2000), to be consistent with the laws of thermodynamics. This
recognition led to the creation of a thermodynamic degradation paradigm,
wherein experimentally it was shown that degradation in the form of wear
correlated with ‘entropy flow’ (Doelling et al. 2000) produced by accompanying
irreversible processes occurring at the wearing surface.
At present, most degradation models are collections of heuristic equations, with
little in common. As noted by Ludema (1995), tribology literature contains a pre-
ponderance of wear mechanisms each based on many different physical principles,
* Author for correspondence (khonsari@me.lsu.edu).

Received 17 December 2007


Accepted 13 March 2008 2001 This journal is q 2008 The Royal Society
Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2002 M. D. Bryant et al.

often expressed in terms of many different variables. These phenomenological


models are limited primarily to the specific system being examined. Additionally,
friction and wear—manifestations of the same interfacial physics—are often
treated as separate phenomena. Proper assessment of these phenomena, leading
to a fruitful generalization, requires a fundamental unifying principle.
In general, degradation processes involve different mechanisms with distinctive
features, types, rates and sequences of dissipative processes. Nevertheless, to elicit
permanent and irreversible changes to bodies, at least one of the processes must
be dissipative. The crux of this article is that all types of permanent degradation
are irreversible processes, which disorder a system and generate irreversible
entropy to agree with the second law of thermodynamics. Therefore, quite
naturally, entropy can be used in a fundamental way to quantify the behaviour of
irreversible degradation, including tribological processes such as friction and wear
(Klamecki 1980a,b, 1982, 1984; Zmitrowicz 1987a–c). For sliding wear, this
irreversible entropy is generated by irreversible dissipation associated with non-
conservative friction forces. Similarly, other forms of degradation such as fretting
(Dai et al. 2000) and fatigue (Bhattacharya & Ellingwood 1998) of machine
components are consequences of irreversible processes that tend to add disorder to
the system. Disorder increases with time until a critical stage is reached whereby
failure occurs. Simultaneous with the rise in disorder, entropy monotonically
increases. Indeed, the emerging field of damage mechanics tracks entropy and
thermodynamic energies (Lemaitre 1985; Bhattacharya & Ellingwood 1998, 1999;
Voyiadjis 1999). Thus, entropy and thermodynamic energies offer a natural
measure of component degradation. Dissipative processes can be directly linked to
thermodynamic entropy, or associated thermodynamic energies, e.g. plasticity,
dislocations (Weertman & Weertman 1964), wear, fracture (Knott 1973; Maugis
2000), fatigue (Izumi et al. 1981; Mura 1987), fretting, corrosion (Uhlig & Revie
1985), thermal degradation and associated failure of tribological components. We
note that Feinberg & Widom (1995, 1996) related material or component
parameter degradation to Gibbs free energy and predicted changes in the system
characteristic results with a log-time ageing behaviour versus time.
Examples of dissipative irreversible processes (Van Wylen & Sonntag 1968;
De Grote & Mazur 1984; Bejan 1988) include viscous dissipation in fluids,
unrestrained expansion of gases, mixing of different chemical species, fracture and
plastic flow in solids (Green 1955; Collins 1980; Fahrenthold & Venkataraman 1993,
1994) and thermally or chemically induced changes of materials. The change in
irreversible entropy dS 0 can be calculated via the theory equations (2.9a)–(2.9d )
with the first law equation (3.1), if the work, heat or energy dissipated is known. For
fluids, this is the dissipation function (Keller 1976; Bejan 1988). For plastic
dissipation, this is the plastic work (De Grote & Mazur 1984). For dry or poorly
lubricated friction and wear involving metals, work dissipated by rubbing involves
work of plastic deformation (Van Wylen & Sonntag 1968; Klamecki 1984) and
other dissipative effects (Robbins & Krim 1998). In fact, Tsuya (1976) observed for
copper sliding against steel, intense plastic deformations that occur in a 10–50 mm
thick skin layer on the softer copper, and attributed plastic dissipation in this
‘severely deformed region’ as a principal mechanism of friction force. Rigney & Hirth
(1979), Heilmann & Rigney (1981), Rigney et al. (1984) and Rigney & Hammerberg
(1998) equated the friction energy to the plastic work dissipated in this region, and
obtained good agreement with Tsuya’s measured friction force. Kennedy (1989)

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2003

measured the near-surface deformations due to sliding using microscopic


observations of the contact region and compared these values with values
predicted by his analytical model. In the analytical model, finite-element
viscoplasticity techniques were developed to model high rate plastic strains in
the vicinity of a moving contact. Kennedy (1989) suggested that the thickness
of the severely deformed region for the loading conditions in his experiment was
less than 300 mm. The thickness, however, is dependent on the load and speed in
the experiment. Suh (1986) estimated the friction force for sliding of metals via
mechanics of asperity adhesion, shearing and ploughing, and adhesive wear and
abrasive wear by ploughing. These mechanisms assumed irreversible plastic
deformations coupled friction forces and energy dissipation. On this, Klamecki’s
(1980a,b, 1982, 1984) thermodynamic analysis of friction and wear applied
conservation of energy, mass and entropy for bodies in sliding contact. Dai et al.
(2000) analysed the entropy production associated with fretting wear.
Zmitrowicz (1987a–c) balanced mass, momentum, energy and entropy between
bodies in contact, to predict friction and wear.
This paper develops a thermodynamic characterization of degradation
dynamics, which employs entropy, a measure of thermodynamic disorder, as
the fundamental measure of degradation. This assumption was implicit in
Feinberg & Widom’s (1996, 2000) analyses with thermodynamic energies, and
explicit in Doelling et al. (2000) and Ling et al. (2002). In this paper, we establish
a general relationship between the degradation of systems undergoing irreversible
dissipative processes, and the concomitant entropy produced. Specifically, we link
components of material degradation and production of thermodynamic entropy
to the irreversible dissipative processes that drive the degradation. Through the
dissipative processes, we relate the rate of degradation to production of irreversible
entropy, encapsulate the formulations into an equation, and finally summarize
the results into a degradation-entropy theorem. An example is presented that
illustrates the application of the theorem to a problem involving degradation by
wear caused by rubbing friction during sliding.

2. Formulations

(a ) Definitions
A dissipative degradation process pi is the minimum 0
group of physical
sub-processes for which the entropy production S_ i produced by the group is
0
non-negative, i.e. S_ i R 0.
A degradation mechanism is a sequence of one or more dissipative degradation
processes pi (iZ1, 2, ., m) that degrade or impair the functionality of a material
or material body.
A degradation measure wZw{p1, p2, ., pi , ., pm} is a non-negative, non-
decreasing continuous function, differentiable in one or more variables pi ,
associated with a degradation mechanism.

(b ) Thermodynamics of degradation
Suppose a degradation mechanism consists of iZ1, 2, ., n dissipative
processes pi , where each pi Z pi ðzji Þ describes an energy, work or heat characteristic

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2004 M. D. Bryant et al.

of the process, and depends on a set of time-dependent phenomenological


variables zji Z zji ðtÞ, jZ1, 2, ., mi. To accumulate the effects of the processes
on overall degradation or ageing, let a degradation measure
n  o
w Z w pi zji ; i Z 1; 2; .; n; j Z 1; 2; .; mi ; ð2:1Þ
which depends on phenomenological variables zji via the n processes pi. Any
dissipative process pi must produce an irreversible entropy S i0 Z S i0 fpi ðzji Þg,
characterized by the same set of variables zji . Here, the prime denotes irreversible
entropy, subscript i references specific process pi and superscript j indicates the
phenomenological variable zji of process pi.
In accordance with the second law, the degradation mechanism must generate
total irreversible entropy
n  o X
S 0 Z S 0 pi zji Z S i0 ; i Z 1; 2; .; n; j Z 1; 2; .; mi : ð2:2Þ
i
The second law mandates non-negative entropy generation and the sum over the
dissipative processes. The rate of degradation dw/dt can be determined by
applying the chain rule to equation (2.1),
!
dw X X vw vpi vzji X XX j j
w_ Z Z Z _
w i Z Yi Ji : ð2:3Þ
dt i j
vpi vzji vt i i j
The rate of entropy dS 0 /dt, which is entropy generation, via equation (2.2) and
the chain rule is
!
0 dS 0 X X vS 0 vpi vzj X 0 XX j j
S_ Z Z j
i
Z S_ i Z Xi J i : ð2:4Þ
dt i j
vp i vz i
vt i i j
0
In equations (2.3) and (2.4), S_ i and w_ i denote entropy generation and degradation
rate contributions arising from specific process pi. For stationary systems or
systems near equilibrium, irreversible thermodynamics (Prigogine 1967; P Bejan
0
1988; Kondepudi & Prigogine 1998) expresses entropy generation S i Z j Xij J ji as
_
the product of generalized
P j forces Xij and generalized rates or flows J ji . We note in
j
equation (2.3), w_ i Z j Yi J i . Comparing sums in equations (2.4) and (2.3),
  9
Xij Z vS 0 =vpi vpi =vzji ; > >
>
>
j j >
=
Ji Z vzi =vt
ð2:5Þ
and >
>
  >>
>
Y j Z vw=vp vp =vzj : ;
i i i i

Since the second law mandates non-negative entropy generation, signs in equation
(2.4) of multiplying factors must be identical, i.e. sgnðXij ÞZ sgnðJij Þ. Equations
(2.3) and (2.4) share rate factors J ji . Irreversible thermodynamics considers forces
Xij as drivers of flows J ji . Each J ji can depend on all forces (De Grote & Mazur 1984)
and intensive quantities (e.g. temperature T ) associated with the dissipative
process via (Bejan 1988)
X
J j Z J j ðX 1 ; X 2 ; .; TÞ z Lq j X q ; ð2:6Þ
q

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2005

where subscript i was dropped in equation (2.6) for clarity. For systems near
equilibrium or stationary, relation (2.6) is invertible (Hillert & Agren 2006) and
usually assumed linear (Bejan 1988). Non-negative entropy production demands
symmetric, positive definite Lq j. Applications of equations (2.4) and (2.5) have
explained diverse phenomena—thermoelectric effect, diffusion (Bejan 1988),
reactions (Kondepudi & Prigogine 1998) and phase changes, amongst others—
and have given an alternate derivation of Kirchhoff’s voltage law for resistive
networks (Županović et al. 2004).

(c ) Degradation force and degradation coefficient


Equation (2.4) can be constructed via principles of thermodynamics (Prigogine
1967; Bejan 1988; Kondepudi & Prigogine 1998). Since equations (2.3) and (2.4)
share Jij , an entropy production analysis (obtained by constructing equation
(2.4)) can elucidate the rates associated with degradation equation (2.3). In
analogy to generalized thermodynamic force Xij , we call Yij the ‘generalized
degradation force’ and define the degradation coefficient
 
j ðvw=vp Þ vp =vzj 
Yi i i i vw=vpi vw 
Bi Z j Z  Z 0 Z 0 ; ð2:7Þ
Xi ðvS 0 =vpi Þ vpi =vzji vS =vpi vS pi

which exists, since Xij s0 except when the system is not degrading. Equation (2.5)
was used in equation (2.7). The last term in equation (2.7) means vw/vS 0
with process pi active, which suggests Bi measures how entropy generation and
degradation interact on the level of processes pi , rather than process variables zji .
Since increments of degradation are non-negative, coefficients BiR0. Finally, if
equations (2.5) and (2.7) are applied to equations (2.3) and (2.4), then
! !
X j j X vw vpi vzj X vw=vpi vS 0 vpi vzj
i i
w_ i Z Yi Ji Z j Z 0 =vp j
j j
vpi vz i
vt j
vS i vp i vz i
vt
X 0
Z Bi Xij Jij Z Bi S_ i : ð2:8Þ
j

(d ) Degradation-entropy generation theorem


The preceding formulations can be summarized into the following theorem:
Degradation-entropy generation theorem. Given an irreversible material
degradation consisting of iZ1, 2, ., n dissipative processes pi Z pi ðzji Þ characterized
by a set of time-dependent variables zji , jZ1, 2, ., mi. Assume the degradation
can be described by a degradation measure, equation (2.1). Then
P P 0
(i) the degradation rate w_ Z i w_ i is a linear P combination w_ Z i Bi S_ i
0
of the components of entropy production S_ i Z j Xij Jij of the dissipative
processes pi, P
(ii) the degradation components w_ i Z j Yij J ji proceed at rates J ji determined
0 P
by the entropy production S_ i Z j Xij Jij of the dissipative processes pi,

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2006 M. D. Bryant et al.

(iii) the generalized ‘degradation forces’ Yij are linear functions Yij Z Bi Xij
of the generalized ‘thermodynamic forces’ Xij , and
(iv) the proportionality factors Bi are the degradation coefficients given by
equation (2.7).

Integration of item (i) of the previous theorem


P yields a corollary to the
theorem: the total amount of degradation wZ i Bi S i0 is a linear combination of
the total components of entropy S i0 produced by the dissipative processes pi.

(e ) Conservation of energy
All systems must obey conservation of energy, stated by the first law of
thermodynamics X
dE Z dQKdW C hk dNk ; ð2:9aÞ
where E is internal energy; Q and W are heat flow and work across the boundary of
the relevant control volume; and hk and Nk are chemical potential and number of
moles of species k. A change in entropy,
dS Z dS 0 C dSe ; ð2:9bÞ
consists of a reversible change dSe from entropy flow, and an irreversible change
from entropy generation dS 0 . We note
P that entropy flow arises from heat transfer
via heat flow dQ and matter flow h k d e Nk ,
X
T dSe Z dQ C h k d e Nk : ð2:9cÞ
A change in the number of moles,
dNk Z d 0 Nk C d e Nk ; ð2:9dÞ
consists of a chemical reaction term d 0 Nk and a matter transport term deNk . At
equilibrium, a system’s entropy is maximum and entropy production ceases:
dS 0 /dtZ0 (Kondepudi & Prigogine 1998). A system produces entropy (dS 0 /dtO0)
until equilibrium. Stationary systems have dEZ0 and dSZ0.
Our interest is the irreversible effects of dissipation caused by work of
non-conservative forces. This P dissipated work must eventually diffuse through
heat flow dQ and/or mass flow hk dNk , as an entropy flow dSe/dt. Via equation
(2.9a)–(2.9d ), the irreversible entropy produced can be linked to the work
dissipated (Frederick & Chang 1965; Klamecki 1984; Bejan 1988). Open systems
demand balancing flows of entropy, heat, work, energy and mass over a control
volume about the degrading body or system, to construct open system counterparts
of entropy generation equation (2.4), and possibly degradation equation (2.3).

(f ) Degradation analysis procedure


The preceding formulations suggest an approach for degradation analysis.

(i) From knowledge of the degradation mechanism, list the irreversible


processes pi Z pi ðzji Þ and variables zji . Often, the process pi can be energy
dissipated, or may be posed in terms of lost work, heat transferred or a
thermodynamic energy such as internal energy or Gibbs free energy.

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2007

(ii) Obtain entropy generation dS 0 /dt of equation (2.4) via irreversible


thermodynamics. Historically, this has involved applying laws of thermo-
dynamics to a control volume about the body in question. Klamecki
(1980a,b, 1982, 1984) presents examples relevant to tribology.
(iii) Using equation (2.3), obtain an expression for the degradation rate dw/dt.
(iv) Via equation (2.4), obtain process rates Jij , found in both equations (2.3)
and (2.4).
(v) Via equations (2.3), (2.4) and (2.7), get thermodynamic forces Xij and
degradation forces Yij . By measuring coefficients Bi of equation (2.7), Yij can
be related to Xij . For Xij and Bi , if process pi is an energy dissipated, then
definition of entropy suggests vS 0 /vpiZ1/Ti , where Ti is a temperature.

If needed, via equation (2.6), relate rates Jij to thermodynamic forces Xij .

3. Applications

(a ) Single variable systems


In some systems, degradation wZw{p(z)} involves principally a single process p(z)
with but one phenomenological variable z. Equations (2.3) and (2.4) yield
0
w_ Z YJ and S_ Z XJ ;
where JZdz/dt, XZdS 0 /dp(dp/dz) and YZdw/dp(dp/dz).
The degradation coefficient (see equation (2.7)) can be obtained from a ratio of
equations (2.3) and (2.4) to relate Y to X. If w and S 0 are measured at common
times t, then
Y =X Z ðdw=dtÞ=ðdS 0 =dtÞ Z B:
Then equation (2.3) gives
w_ Z YJ Z BXJ : ð3:1Þ
Besides z, irreversible thermodynamics suggest that B can depend on intensive
variables such as temperature T (Bejan 1988). For single variable systems,
therefore, equation (3.1) suggests a direct relation between degradation rate and
irreversible entropy production.

(b ) Degradation wear due to friction rubbing


Wear in bodies in sliding contact is measured as volume wv of material lost. A
counter body moves beneath and rubs against a fixed wearing body. The counter
body applies friction force Fm directed along the counter-body motion onto the
wearing body. Consider a ‘wear control volume’ enclosing the sliding surface and
near interior regions of the wearing body. The friction force Fm through distance dx
does work dW on the control volume, i.e.
KdW Z F dx Z N dx;
where m is the friction coefficient and N is the normal load. The minus sign on dW
indicates work done on the system from outside the control volume.

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2008 M. D. Bryant et al.

This work is dissipated within the control volume. For sliding of ductile metals,
Rigney & Hirth (1979) identified the dominant dissipative process p to be work of
plastic deformation. Let us assume that (i) rubbing is at steady speed and force,
and is a stationary process. In other words, for steady-state sliding, changes to the
internal energy dE and entropy dS contained within the small wear control volume
that surrounds the sliding surface and near interior regions of the wearing
P body are
small. (ii) Energy transport effects of material loss are small, hk d e Nk is
negligible compared with other terms in equation (2.9c), thus permitting us to treat
the system as closed. (iii) There are no chemical reactions occurring. (iv) All
friction work is dissipated within the control volume, i.e. dpZKdW. Applying
assumptions (i)–(iv) to equation (2.9a)–(9c) yields
dW Z dQ Z T dSe ZKT dS 0 :
The rate of irreversible entropy generated due to friction can be obtained as follows:
dS 0 vS 0 vp vx Fm dx
Z Z ; ð3:2aÞ
dt vp vx vt T dt
where contact temperature T arose in equation (3.2a) from vS 0 /vpZ(dS 0 /dW )
(dW/dp)Z1/T. For dS 0 /dtR0, since TR0, Fm and dx/dt must have the same signs
in the relevant free body diagram for the slider; this is consistent with the friction
force always opposing the direction of the sliding velocity.
Comparing equation (3.2a) with equation (2.4),
J Z dx=dt; ð3:2bÞ
zZx, dp/dxZFm and thus
X Z vS 0 =vpðvp=vzÞ Z F =T: ð3:2cÞ
Also, JZJ(X ) by equation (2.6) is consistent with the observed friction force
dependency on sliding speed and temperature.
Applying equations (3.1) or (2.3) yields
dw BFm dx Bm dx
w_ v Z v Z YJ Z BXJ Z Z N ; ð3:2dÞ
dt T dt T dt
where wv represents wear volume and YZBmN/T.

(c ) Archard’s wear law


One of the most commonly used relationships for assessing wear in a friction
pair is due to the work of Archard (1953, 1980). It states: wv Z ðk=H ÞNx, which
relates wv to N and x, via wear coefficient k and hardness H of the softer of the
material pair.
With load and temperature constant, the time rate of wear can be obtained by
differentiation of Archard’s law as follows:
dwv k dx
Z N : ð3:2eÞ
dt H dt
Comparing equation (3.2e) with equation (3.2d ),
k Z BmH =T: ð3:2f Þ

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2009

Constant B can be measured via equation (2.7), i.e. BZdw/dS 0 , using dS 0 ZKdSe
for a stationary process.
Doelling et al. (2000) conducted extensive experimental measurements of wear
in an apparatus where a stationary copper specimen was in boundary lubricated
contact with a rotating steel cylinder. They reported a series of measurements to
establish relationship between normalized wear as a function of normalized entropy
flow. The entropy flow, the negative
P of entropy production for a stationary process,
was calculated using SM Z M DQ ðlÞ =T ðlÞ , where DQ(l ) is the heat input to the
slider during the lth time interval and T (l ) is the corresponding average absolute
surface temperature of the stationary rider on the rotating cylinder. Both
temperature and wear rates were measured and the slope of normalized wear
plotted versus the normalized entropy flow was evaluated. This slope is precisely
the degradation ratio B defined in §2c, equation (2.7).
Having determined B, equation (3.2f ) was used to estimate the wear coefficient,
k. The average of a series of repeatable tests yielded kZ1.01!10K4. For
same metals under poor lubrication, Rabinowicz (1980) gives kZ1.0!10K4. Note
that the k estimated by equation (3.2f ) arose from measured wear, temperatures
and forces; Rabinowicz’s k, calculated via Archard’s wear law (1953), arose from
measured wear, forces and distance. Additional tests show similar results with
remarkable accuracy when compared to published wear coefficients. These
experiments reveal that wear (a measure of material degradation) and entropy
are intimately related, and the relationship between entropy and temperature can be
put to use for prediction of material degradation.

(d ) Application to fretting wear


Fretting involves rubbing between bodies with small, cyclic motions. Fretting
degrades components through surface wear and structural fatigue. Fretting wear of
metals involves intense plastic deformations near the surface of the wearing body,
sometimes accompanied with corrosion or oxidation of material. Rise of bulk
temperatures are usually moderate, of the order of tens of degrees Kelvin. The
fretting process is gradual, with the fretted component in a state of equilibrium or
quasi-equilibrium.
During the last two decades, Mohrbacher et al. (1994), Huq & Celis (2002), and
later Fouvry et al. (1997, 2003, 2007) and Fouvry & Kapsa (2001) developed an
energy-based model for fretting wear. Exhaustive measurements by both groups
showed the Archard wear law to be an inappropriate descriptor of their data.
Mohrbacher et al. (1994) found that wear volume wv (measured via geometry of a
wear scar), when plotted versus total friction energy dissipated Em, produced a
straight line. Fouvry et al. (1997, 2003, 2007), Fouvry & Kapsa (2001), Huq & Celis
(2002) and Lee et al. (2005) confirmed this linear relation for fretting wear of
various coatings on substrates.
Dai et al. (2000) treated fretting wear as an irreversible thermodynamic process,
in a critically stable state near equilibrium, but in the process of transitioning to a
new equilibrium state. At equilibrium, entropy maximizes and entropy production
ceases. Perturbations of the system about the equilibrium state reactivate the
dissipative processes and entropy production. For perturbations, thermodynamic
quantities including entropy production are expressed as variations about the
equilibrium state (Kondepudi & Prigogine 1998). For stable systems, the first

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2010 M. D. Bryant et al.

variation of entropy production about the equilibrium state vanishes, which


demands that the intensive thermodynamic states, including temperature, assume
equilibrium values. For a critically stable system, the rate of the second variation of
entropy vanishes. With this condition, Dai et al. (2000) equated the perturbed
entropy flow to the perturbed entropy production, and solved for wear as the mass
flux component of entropy flow. The entropy flow component due to heat
conduction was ignored in this equation, which limits the use of the formulation of
Dai et al. (2000). 0
Our theorem suggests that the rate of fretting wear can be written as w_ v Z B S_ .
Following Mohrbacher et al. (1994), Fouvry et al. (1997, 2003, 2007), Fouvry &
Kapsa (2001) and Huq & Celis (2002), we assume work of plastic deformation,
driven by friction force Fm, to be the dominant dissipative process p. Letting p
be the dissipation from the friction force, an analysis similar to that which led to
equation (3.2d ) suggests
 
dwv _ 0 Fm dx 2B dx dEm
Z B S Z B½2ðdXÞðdJ Þ Z 2B Z Fm Z Bf : ð3:3Þ
dt T dt T dt dt
The delta notation on X and J indicate perturbations from the current equilibrium
state, and the two arises from the second variation of entropy (Dai et al. 2000).
Near equilibrium, temperature T is very near the equilibrium temperature, as shown
by Szolwinski et al. (1999). The first and last terms of equation (3.3) suggest that
the volume of material lost wv should be proportional to the friction energy Em
dissipated, with coefficient BfZ2B/T. This shows that the energy-based models
for fretting wear are a consequence of the laws of thermodynamics and the
degradation-entropy generation theorem, when viewed through the formalism of
equations (2.1)–(2.9d ).

4. Discussion

Coefficients Bi , defined by the formulations of equation (2.7), are material constants


relating generalized degradation forces to generalized thermodynamic forces.
Material properties such as thermal conductivity, magnetic permeability, fluid
viscosity, elastic modulus or fracture toughness are defined by formulations of
physics, but of necessity, the constitutive relations of these formulations must involve
heuristic measurements. In §3c, coefficients Bi were measured, used and compared
with the results of Rabinowicz (1980) based on fundamentally different
measurements. Coefficients Bi can be geometrically interpreted as ratios of
components of gradients Vpi w and Vpi S 0 , for two hypersurfaces, wZw {pi} and
S 0 ZS 0 {pi}, in a space {pi , iZ1, 2, ., m} spanned by the processes pi (which serve as
coordinates in the space). For degradation by wear, the surface wZ wfpi ðzij Þg defines
a wear curve (Bayer 1994) or wear map (Bhushan 2002). Each Bi gives the ratio
between the local slopes vw=vS 0 Z ðvw=vpi Þ=ðvS 0 =vpi Þ of the surfaces, along the
direction pi.
A possible application of this procedure is in the development of a methodology
for accelerated testing of degradation. It deals with an enabling technology for
predicting behaviour of a system undergoing degradation and ultimate failure after
a long period of time, based on short-term laboratory experiments. As mentioned

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2011

in §2b, equations (2.3) and (2.4) have common rate factors Jij Z vzji =vt that depend
on system parameters zji . If the flow rates Jij are judiciously chosen, then the rate of
degradation dw/dt in equation (2.3) can be observed without waiting for long times.
Equations (2.3)–(2.7) suggest that one can conduct an accelerated failure testing
scheme by increasing process rates Jij while maintaining ‘equivalent’ forces Xij and
Yij to obtain the same sequence of physical processes, in identical proportions, but
with a higher rate. Simply increasing rates Jij may alter the physical processes (Ling
et al. 1997). For example, moderate heat hatches an egg; if heating is accelerated
without maintaining equivalent forces, the egg cooks. Via equation (2.5), altering
any Jij could change every Xij .
As an example, to accelerate wear testing, equations (3.2a) and (3.2b) suggest
increasing sliding (rubbing or slip) speed JZdx/dt while maintaining equivalent
thermodynamic force X and degradation force Y. Higher dx/dt will accelerate wear,
but higher temperatures from friction heat will affect constancy of XZFm/T via
contact temperature T, and YZBmN/T through dependencies of B and possibly
m, on T. This suggests adjusting the normal force N and temperature T, to
compensate and maintain equivalent Y.

5. Concluding remarks

The science that captures degradation dynamics in a modelling paradigm suitable


for early failure prediction, structured development of accelerated failure testing,
and control of machine and structure maintenance is of particular interest in many
disciplines. However, the modelling and analysis of machine and structure reliability
has remained substantially unchanged for decades. In this paper, we presented a
thermodynamic characterization of degradation dynamics, which employed entropy
(a measure of thermodynamic disorder) as the fundamental measure of degradation,
and suggested methods of measurement of state variables that are functionally
related to entropy, which is, necessarily, an implicit system variable. Degradation of
machinery components is a time-dependent phenomenon that arises because effects
of irreversible processes accumulate to disorder the component. Therefore, it follows
that thermodynamic entropy can be used to properly characterize the extent of
disorder, the rate at which it progresses, and measure the component degradation.
Hence, the second law of thermodynamics provided the framework for predicting
failure, based on the rate of production and accumulation of entropy. The theory
presented lays a thermodynamically consistent foundation for applying this concept
to a variety of failure problems, including the tribological applications (adhesive
wear and fretting wear) presented here as case examples. Application to tribology
systems involves placing a tribological control volume about the region where the
dissipative processes occur. Godet (1984, 1990) identified this region as the ‘third
body’. The work can be extended to determine the critical damage parameter in
processes involving fretting fatigue and to predict fretting fatigue life of machine
element (cf. Quraishi et al. 2005).
The authors would like to thank the Accenture Endowed Professorship in Manufacturing Systems
Engineering, University of Texas at Austin and the Dow Chemical Endowed Chair in Rotating
Machinery at Louisiana State University for support to conduct this work.

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2012 M. D. Bryant et al.

References
Archard, J. F. 1953 Contact and rubbing of flat surfaces. J. Appl. Phys. 24, 981–988. (doi:10.1063/1.
1721448)
Archard, J. F. 1980 Wear theory and mechanisms. In Wear control handbook (eds M. B. Peterson &
W. O. Winer), pp. 35–80. New York, NY: ASME.
Bayer, R. G. 1994 Mechanical wear prediction and prevention. New York, NY: Marcel-Dekker.
Bejan, A. 1988 Advanced engineering thermodynamics. New York, NY: Wiley.
Bhattacharya, B. & Ellingwood, B. 1998 Continuum damage mechanics analysis of fatigue crack
initiation. Int. J. Fatigue 20, 631–639. (doi:10.1016/S0142-1123(98)00032-2)
Bhattacharya, B. & Ellingwood, B. 1999 A new CDM based approach to structural deterioration. Int.
J. Solids Struct. 36, 1757–1779. (doi:10.1016/S0020-7683(98)00057-2)
Bhushan, B. 2002 Introduction to tribology. New York, NY: Wiley.
Collins, I. F. 1980 Geometrically self-similar deformations of a plastic wedge under combined shear
and compression loading by a rough flat die. Int. J. Mech. Sci. 22, 735–742. (doi:10.1016/0020-
7403(80)90058-2)
Dai, Z., Yang, S. & Xue, Q. 2000 Thermodynamic model of fretting wear. J. Nanjing Univ. Aeronaut.
Astronaut. 32, 125–131.
De Grote, S. R. & Mazur, P. 1984 Non-equilibrium thermodynamics. New York, NY: Dover.
Doelling, K. L., Ling, F. F., Bryant, M. D. & Heilman, B. P. 2000 An experimental study of the
correlation between wear and entropy flow in machinery components. J. Appl. Phys. 88,
2999–3003. (doi:10.1063/1.1287778)
Fahrenthold, E. P. & Venkataraman, M. 1993 Continuum damage modeling of impact fracture in
ceramic materials. Compos. Eng. 3, 1039–1050.
Fahrenthold, E. P. & Venkataraman, M. 1994 Anisotropic elastic-damage models of coupled solid
dynamics and heat diffusion. Compos. Eng. 4, 913–930.
Feinberg, A. A. & Widom, A. 1995 The reliability physics of thermodynamic aging. In Recent
advances in life-testing and reliability (ed. N. Balakrishnan), pp. 241–255. New York, NY: CRC
Press.
Feinberg, A. A. & Widom, A. 1996 Connecting parametric aging to catastrophic failure through
thermodynamics. IEEE Trans. Reliab. 45, 28–33. (doi:10.1109/24.488913)
Feinberg, A. A. & Widom, A. 2000 On thermodynamic reliability engineering. IEEE Trans. Reliab.
49, 136–146. (doi:10.1109/TR.2000.877330)
Fouvry, S. & Kapsa, P. 2001 An energy description of hard coating wear mechanisms. Surf. Coat.
Technol. 138, 141–148. (doi:10.1016/S0257-8972(00)01161-0)
Fouvry, S., Kapsa, P., Zahouani, H. & Vincent, L. 1997 Wear analysis in fretting of hard
coatings through a dissipated energy concept. Wear 203–204, 393–403. (doi:10.1016/S0043-
1648(96)07436-4)
Fouvry, S., Liskiewicz, T., Kapsa, P., Hannel, S. & Sauger, E. 2003 An energy description of wear
mechanisms and its applications to oscillating sliding contacts. Wear 255, 287–298. (doi:10.1016/
S0043-1648(03)00117-0)
Fouvry, S., Paulin, C. & Liskiewicz, T. 2007 Application of an energy wear approach to quantify
fretting contact durability: introduction of a wear energy capacity concept. Tribol. Int. 40,
1428–1440. (doi:10.1016/j.triboint.2007.02.011)
Frederick, D. & Chang, T. S. 1965 Continuum mechanics. Boston, MA: Allyn & Bacon.
Godet, M. 1984 The third body approach: a mechanical view of wear. Wear 100, 437–452. (doi:10.
1016/0043-1648(84)90025-5)
Godet, M. 1990 Third-bodies in tribology. Wear 136, 29–45. (doi:10.1016/0043-1648(90)90070-Q)
Green, A. P. 1955 Friction between unlubricated metals: a theoretical analysis of the junction model.
Proc. R. Soc. A 228, 191–204. (doi:10.1098/rspa.1955.0043)
Heilmann, B. P. & Rigney, D. A. 1981 An energy-based model of friction and its applications to
coated systems. Wear 72, 195–217. (doi:10.1016/0043-1648(81)90367-7)

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

Thermodynamics of degradation 2013

Hillert, M. & Agren, J. 2006 Extremum principles for irreversible processes. Acta Mater. 54,
2063–2066. (doi:10.1016/j.actamat.2005.12.033)
Huq, M. Z. & Celis, J. P. 2002 Expressing wear rate in sliding contacts based on dissipated energy.
Wear 252, 375–383. (doi:10.1016/S0043-1648(01)00867-5)
Izumi, Y., Fine, M. E. & Mura, T. 1981 Energy considerations in fatigue crack propagation. Int.
J. Fract. 17, 15–25. (doi:10.1007/BF00043118)
Keller, J. V. 1976 The fundamental inequality for thermodynamic systems with heat and mass
transfer. J. Non-Equil. Thermodyn. 1, 67–73.
Kennedy Jr, F. E. 1989 Plastic analysis of near-surface zones in sliding contact metals. Key Eng.
Mater. 33, 35–47.
Klamecki, B. E. 1980a Wear—an entropy production model. Wear 58, 325–330. (doi:10.1016/0043-
1648(80)90161-1)
Klamecki, B. E. 1980b Wear—a thermodynamic model of friction. Wear 63, 113–120. (doi:10.1016/
0043-1648(80)90078-2)
Klamecki, B. E. 1982 Energy dissipation in sliding. Wear 77, 115–128.
Klamecki, B. E. 1984 Wear—an entropy based model of plastic deformation energy dissipation in
sliding. Wear 96, 319–329. (doi:10.1016/0043-1648(84)90044-9)
Knott, J. F. 1973 Fundamentals of fracture mechanics. London, UK: Butterworths.
Kondepudi, D. & Prigogine, I. 1998 Modern thermodynamics: from heat engines to dissipative
structures. New York, NY: Wiley.
Lee, H., Mall, S., Sanders, J. H. & Sharma, S. K. 2005 Wear analysis of Cu–Al coating on Ti–6Al–4V
substrate under fretting condition. Tribol. Lett. 19, 239–248. (doi:10.1007/s11249-005-6151-7)
Lemaitre, J. 1985 Continuous damage mechanics model for ductile fracture. J. Eng. Mater. Technol.
107, 83–89.
Ling, F. F., Klutke, G. G. & Wortman, M. A. 1997 Workshop on maintenance science. In Proc.
Energy Week Conf. Book V(1), pp. 10–15.
Ling, F. F., Bryant, M. D. & Doelling, K. L. 2002 On irreversible thermodynamics for wear
prediction. Wear 253, 1165–1172. (doi:10.1016/S0043-1648(02)00241-7)
Ludema, K. C. 1995 The state of modeling in friction and wear and a proposal for improving that
state. J. Korean Soc. Tribol. Lubr. Eng. 11, 10–14.
Maugis, D. 2000 Contact, adhesion, and rupture of elastic solids. Berlin, Germany: Springer.
Mohrbacher, H., Blanpain, B., Celis, J. P., Roos, J. R., Stals, L. & Van Stappen, M. 1994 Oxidational
wear of TiN coatings on tool steel and nitrided tool steel in unlubricated fretting. Wear 188,
130–137. (doi:10.1016/0043-1648(95)06637-3)
Mura, T. 1987 Micromechanics of defects in solids, 2nd edn. Dordrecht, The Netherlands: Martinus
Nijhoff.
Prigogine, I. 1967 Introduction to thermodynamics of irreversible processes, 3rd edn. New York, NY:
Interscience Publishers, Wiley.
Quraishi, S. M., Khonsari, M. M. & Baek, D. K. 2005 A thermodynamic approach for predicting
fretting fatigue life. Tribol. Lett. 19, 169–175. (doi:10.1007/s11249-005-6143-7)
Rabinowicz, E. 1980 Wear control handbook, p. 486. New York, NY: ASME.
Rigney, D. A. & Hammerberg, J. E. 1998 Unlubricated sliding behavior of metals. MRS Bull. 23,
32–36.
Rigney, D. A. & Hirth, J. P. 1979 Plastic deformation and sliding friction of metals. Wear 53,
345–370. (doi:10.1016/0043-1648(79)90087-5)
Rigney, D. A., Chen, L. H., Naylor, M. G. S. & Rosenfield, A. R. 1984 Wear processes in sliding
systems. Wear 100, 195–219. (doi:10.1016/0043-1648(84)90013-9)
Robbins, M. O. & Krim, J. 1998 Energy dissipation in interfacial friction. MRS Bull. 23, 23–26.
Suh, N. 1986 Tribophysics. Englewood Cliffs, NJ: Prentice-Hall.
Szolwinski, M. P., Harish, G., Farris, T. N. & Sakagami, T. 1999 In-situ measurement of near-surface
fretting contact temperatures in an aluminium alloy. ASME J. Tribol. 121, 11–19. (doi:10.1115/
1.2833791)

Proc. R. Soc. A (2008)


Downloaded from rspa.royalsocietypublishing.org on November 22, 2012

2014 M. D. Bryant et al.

Tsuya, Y. 1976 Microstructures of wear, friction, and solid lubrication. Technical report 81,
Mechanical Engineering Laboratory, Igusa Suginami-ku, Tokyo.
Uhlig, H. H. & Revie, R. W. 1985 Corrosion and corrosion control, pp. 16–20, 3rd edn. New York,
NY: Wiley.
Van Wylen, G. & Sonntag, R. E. 1968 Fundamentals of classical thermodynamics, pp. 165–168. New
York, NY: Wiley.
Voyiadjis, G. 1999 Advances in damage mechanics. Amsterdam, The Netherlands: Elsevier.
Weertman, J. & Weertman, J. R. 1964 Elementary dislocation theory. London, UK: MacMillan.
Zmitrowicz, A. 1987a A thermodynamical model of contact, friction, and wear: I. Governing
equations. Wear 114, 135–168. (doi:10.1016/0043-1648(87)90086-X)
Zmitrowicz, A. 1987b A thermodynamical model of contact, friction, and wear: II. Constitutive
equations for materials and linearized theory. Wear 114, 169–197. (doi:10.1016/0043-
1648(87)90087-1)
Zmitrowicz, A. 1987c A thermodynamical model of contact, friction, and wear: III. Constitutive
equations for friction, wear and frictional heat. Wear 114, 199–221. (doi:10.1016/0043-
1648(87)90088-3)
Županović, P., Juretić, D. & Botrić, S. 2004 Kirchhoff’s loop law and the maximum entropy
production principle. Phys. Rev. E 70, 056 108. (doi:10.1103/PhysRevE.70.056108)

Proc. R. Soc. A (2008)

Вам также может понравиться