You are on page 1of 15

8

Wear Debris
Classification

8.1 Introduction
8.2 How Wear Debris Is Generated
8.3 Collection of Wear Debris
8.4 Diagnostics with Wear Debris
Particle Counting • Biological Debris Sampling • Reworked
William A. Glaeser Wear Debris
Battelle 8.5 Conclusions

8.1 Introduction
Wear generates debris. The debris comes in a wide variety of sizes and shapes. Wear debris turns motor
oil black. You can see it on your hands if you shake hands with your garage mechanic. The black in used
oil is like a pigment — it is nanometer-size colloidal metal particles (and carbon) suspended in the oil.
If you take an oil filter apart, you will find other types of wear debris. Some are shiny metal particles
visible to the naked eye and in the millimeter size category. If you are in the surface mining business,
you will find even larger metal particles in the gravel produced by rock crushers. This is metal that is
gouged from the metal jaws in the crusher. In a factory grinding room, dust swept up from the floor
and examined in a scanning electron microscope will show perfect micron-size metal spheres from rapidly
solidifying drops of molten metal generated by the grinding wheel. Examples of these unique debris
particles will be found later in this chapter.
Wear debris represents loss of geometric accuracy of moving contacting parts. It can also foul orifices
and close spaced parts. Although the total material lost as wear debris in a machine is minute compared
to the volume and weight of the moving parts, it can signal failure of gears or bearings, and expensive
repairs or warranty payments.
Because wear debris can be carried in a circulating oil lubrication system, the condition of an engine
or a critical machine can be diagnosed by analyzing a sample of the oil. Debris can be removed from the
oil by centrifuging or by a magnetic process. The separated debris can then be examined in a microscope
to determine morphology, or it can be chemically analyzed. A standard test, the spectrographic oil analysis
program (SOAP), has been used for condition monitoring by railroads, truck fleets, and the military for
tanks and aircraft.
Until the Industrial Revolution and the development of steam and internal combustion engines, wear
was taken for granted. When wear of machinery and tools became a problem, the development of wear-
resistant materials and the study of wear itself increased rapidly. Now we have a large body of information
developed from experience and scientific investigation that can be used in the design and maintenance
of more reliable and economical machinery.

© 2001 by CRC Press LLC


8.2 How Wear Debris Is Generated
Early wear theory, as related by Bowden and Tabor (1964) and Holm (1946) indicated that real surfaces
contact at high spots in the surface microtopography (asperities) and wear was caused by the shearing
off of these contacts. The theory was not clear, however, on how the wear particles were separated from
contacting asperities. One obvious mechanism is cutting, as in machining. Under abrasive wear condi-
tions, it is presumed that hard particles embedded in one surface will plow through a softer counterface
and produce microchips. The debris thus generated looks like the example shown in Figure 8.1. The
curled form that the debris particles assume reminds one of chips from machining. In fact this signals
a severe type of abrasive wear. The attack angle, hardness, and depth of penetration of abrading particles
determine whether cutting-type abrasion will occur as shown by Mulhearn and Samuels (1962). A
frequency distribution of attack angles for abrasive particles on SiC bonded abrasive is shown in
Figure 8.2. A number of metals are listed in Table 8.1 together with their critical attack angles — or the

FIGURE 8.1 SEM micrograph of cutting-type debris from abrasive wear, M2 tool steel. (From Rigney, D.A. (1992),
The role of characterization in understanding debris generation, Tribology Series, 21, Wear Particles: From the Cradle
to the Grave, Elsevier. With permission.)

100

80
Critical
attack
60 angle
Frequency

40

20

Noncutting points Cutting points


0
0 20 40 60 80 100 120 140 160 180
Attack angle - α

FIGURE 8.2 Diagram showing the effect of cutting angle on hard particle cutting wear.

© 2001 by CRC Press LLC


TABLE 8.1
HV
Material kg/mm2 Critical Attack Angle

Lead 5 50° to 60°


α-brass 180 50° to 60°
Aluminum 35 80° to 90°
Aluminum alloy 145 ~35°
Copper 120 40° to 50°
Nickel 350 60° to 70°
1040 steel, annealed 180–260 40° to 90°
1082 steel, quenched and tempered 860 ~30°
Tool steel, quenched and tempered 760 ~30°

angle above which cutting will occur. Note that the lower the critical angle, the greater the percentage of
particles that will cut. Note also that hardened steel is quite susceptible to cutting abrasion. In general,
the higher the hardness of the metal, the more susceptible to cutting wear it is. In order for cutting
abrasion to take place, the hardness of the abrasive medium must be over 1.2 times that of the material
being abraded.
Contact load between sliding surfaces does not have to be large for cutting wear to take place. In fact,
Zhao and Bhushan (1998) have demonstrated that cutting abrasion can occur with normal loads of 40 to
80 µN using a diamond pyramidal point on single crystal silicon. The debris produced was similar in
morphology to metal-cutting debris (curly and stringy). The size of the chips was about 50 nm wide by
200 nm long.
Wear debris can also be generated by plastic deformation. Hard asperities, plowing over a softer surface
without cutting will produce ridges. Ridges can then be flattened by further contact. Extrusions or lips
are formed. These lips are broken off and become flat wear flakes. The process is illustrated in Figure 8.3.
Extrusions can also occur at the exit border of a wear scar formed by a curved surface rotating against
a flat surface. An example is shown in Figure 8.4. The example in Figure 8.4 is from a block-on-ring test
in which a ring rotates against a block.
Wear debris can also be generated by material transfer from one surface to another. An asperity plowing
through a surface will cold weld to the mating surface and the weld will shear off, leaving some of the
asperity stuck to the mating surface. An in situ scanning electron microscope (SEM) wear experiment by
Glaeser (1981) demonstrated this process. A SEM photomicrograph shows this in Figure 8.5. You can see
the track the asperity made as it penetrated the other surface. Then, suddenly, adhesion occurs and a piece
of the asperity breaks off and sticks to the surface as a lump. Figure 8.6 shows the lump after the rubbing
surface has passed over it many times. The lump has been squashed and is breaking up into smaller particles.
A unique type of wear debris is found in grinding swarf and from electric spark discharge damage to
metal surfaces. It is shown in Figure 8.7. Each particle is spherical. A dendritic microstructure is often
visible on the sphere surface. The particles are formed from molten droplets caused by frictional heating
of the grinding process. They can also be produced by the hot arc developed in electric discharge damage
to metal surfaces or from electric arc welding. Spherical debris can be found in oil samples from machinery
operating in the vicinity of grinding operations.
Rubber produces wear particles by a different process than metals. During abrasion, the rubber is
heated by frictional energy, softens, and sticks to the abrading surface. The rubber then stretches and
forms abrasion ridges oriented normal to the direction of sliding. An example is shown in Figure 8.8.
Debris comes from these ridges in the form of rolled up particles. A familiar example is the debris that
a rubber eraser produces.
Most of the wear debris formed by the above mechanisms is visible using light microscopy. Submicron
debris can be found in SEM analysis of material filtered from used oil. In a series of experiments designed
to observe the generation of wear debris under lubricated conditions, a block-on-ring wear tester was
used by Glaeser (1983). The block was Cu-3.2 wt% Al alloy and the ring was carburized steel. A drop of

© 2001 by CRC Press LLC


PLOWING MICRO CUTTING EXTRUSION

EXTRUSION WEDGING FORWARD EXTRUSION

GOUGING

FIGURE 8.3 Ways in which wear debris is formed by plowing action.

FIGURE 8.4 SEM micrograph of stainless steel extrusions.

silicone oil was applied to the ring so that it clung to the bottom of the ring. The configuration is shown
in Figure 8.9. The drop was illuminated with a light source and was observed through a microscope. As
the ring rotated, oil from the drop circulated through the block contact and back to the drop, carrying
any debris generated into the drop. At the beginning of rotation, the oil drop began to fill with metal
flakes that sparkled as they tumbled in the drop. Later in the test, the drop became cloudy. At this point,
the experiment was terminated and the drop removed for analysis. The large wear debris was removed
by centrifuging. The still cloudy oil was then filtered through a Nucleopore polycarbonate filter.

© 2001 by CRC Press LLC


FIGURE 8.5 SEM micrograph of in-SEM wear experiment, showing asperity plowing, adhesion, and shear off.
Transfer of iron to copper–nickel alloy.

FIGURE 8.6 SEM micrograph of transfer lump after being run over many times by the mating surface.

The filter was mounted on a carbon grid for an electron microscope and the filter dissolved with a
solvent. The remaining material was then examined in a transmission electron microscope and was found
to consist of equiaxed nanoparticles embedded in a gel. An example of the debris is shown in an STEM
micrograph, Figure 8.10. The size of the round solid particle indicated in Figure 8.10 is about 130 nm.
EDX analysis of the particle showed it to be 80% Cu, 1.6% Fe, and 2% Al. The balance was silicon. The

© 2001 by CRC Press LLC


FIGURE 8.7 Spherical wear debris from grinding swarf.

FIGURE 8.8 Rubber abrasion pattern.

silicone oil had formed a gel with the submicron metal particles. The large bronze flakes that appeared
initially were from the wear-in process as the roughness in the bronze surface was removed. Heavy surface
flow produced smeared layers. As these layers were subjected to increasing numbers of rubbing cycles,
particles broke off and were trapped in the surface. These particles were reduced to very small particles
by comminution, as happens in the ball milling process. SEM analysis revealed particle reduction in
progress on the aluminum bronze block as shown in Figure 8.11. The submicron particles produced
reacted with the oxidizing lubricant to form a gel which clouded the oil drop. This type of wear is often
found in boundary-lubricated or very thin film elastohydrodynamic-lubricated contacting surfaces.

8.3 Collection of Wear Debris


In circulating oil systems, the SOAP (spectrographic oil analysis program) process of debris monitoring
has proved valuable in certain applications. Instead of extracting the debris from the oil, the whole oil

© 2001 by CRC Press LLC


FIGURE 8.9 Experimental setup for viewing wear particles in an oil drop in a block-on-ring wear tester.

FIGURE 8.10 TEM micrograph of a submicron wear particle embedded in a gel.

sample is subjected to spectrographic analysis. A summary of this type of oil analysis was reported by
Lukas and Anderson (1998). Metal content is measured in ppm. The following potential failures can be
detected in this way:
• Broken piston rings, worn or scuffed cylinder walls, and worn valve guides
• Worn ball or roller bearings and/or retainers
• Worn journal bearings
• Worn spline couplings
• Worn cams and followers
• Scored pistons in hydraulic pumps

© 2001 by CRC Press LLC


FIGURE 8.11 Wear scar of aluminum bronze block showing the process of comminution of wear debris.

This method works well for colloidal suspensions of metals in oil. However, it is not practical for
analysis of larger pieces. By examining the shape, surface texture, and chemistry of larger particles, one
can obtain supplemental information on failure of mechanical components.
Wear debris can be collected by allowing it to fall on a glass slide. The debris can then be fixed to the
slide and examined under a microscope. In a circulating oil system, a filter in the line can be used to collect
debris. Debris can be washed out of the filter with solvent, or the filter element, if of an organic material,
can be dissolved by an appropriate solvent and the remaining metallic debris collected. Ferrography, devel-
oped in the 1970s and described by Seifert, Westcott, and Bowen (1972), uses a magnetic field to collect
ferromagnetic particles from a fluid. The fluid with suspended metal particles is allowed to flow over a glass
slide inclined at a low angle. The glass slide is held to a strong permanent magnet in such a way that a
magnetic gradient is developed along the length of the slide. In this way, the particles are separated by size.
Ferrous particles separate in strings, with the largest particles at one end. Weak magnetic particles deposit
randomly on the slide. A complete issue of the journal Wear (1983) was devoted to ferrography.
Centrifuging suspensions of wear debris in oil-solvent solutions will separate metal particles according
to mass. The particles will be distributed in layers. Once the centrifuging is complete, the fluid is decanted
from the vessel and the wear particles removed by washing out with solvent. The solvent volume can be
reduced by evaporation, the wear particles suspended by ultrasound and the remaining fluid poured out
on a glass slide. The wear particles will be well dispersed on the slide when the solvent evaporates. If one
wishes to separate the particles by mass, the layers can be removed by pipette and deposited on separate
slides.
Wear tests will often leave wear debris on one of the wear specimens. For instance in a fretting test in
which a ball is oscillated against a flat, debris will appear at the edges of the wear scar. The debris can
be picked up with a cellulose acetate film dampened with solvent. The film can then be dissolved, leaving
the debris free for examination.

8.4 Diagnostics with Wear Debris


As has been noted, wear debris comes in all sizes and shapes. The way in which the debris has been
formed can be deduced from size, shape, and surface texture. For instance, cutting wear debris is unique
in having a curly and stringy shape. The presence of cutting wear debris in an oil sample signals a severe
wear problem needing immediate attention.

© 2001 by CRC Press LLC


120
110
100
90
80
70 Iron

ppm
60
50
40
30
20 Chromium
10
0
10 20 30 40 50 60 70 80 90 100 110 120
Hours

FIGURE 8.12 Plot of iron and chromium content as a function of hours of engine operation. (From spectrographic
analyses of oil samples.)

Use of wear debris as a diagnostic tool for the health of operating machinery is difficult. Recognition
of significant morphological characteristics, color, and chemical make up of the variety of particles and
elements that turn up in a sample requires experience. The SOAP program has proven effective for truck,
tank, and railroad diesels, hydraulic systems, gear boxes, and compressors. The diagnosis is based on the
history of concentration of specific elements found in oil samples. The particles involved are under 10 µm
in size and more likely to be of submicron size. They are detected by spectrographic methods which give
quantitative results, usually in ppm. The quantitative results are plotted as a function of running hours.
An example of such a plot is shown in Figure 8.12. The trends for several metals are compared with the
original oil sample taken at hour one. By experience, these trends (for instance, sudden continuing
increase in iron content) can be related to pending wear failures of certain components.

8.4.1 Particle Counting


Particles from a given sample of oil separated by ferrograph or other methods can be classified as to size
and a size-frequency distribution plot made. Assuming that the larger particles are related to severe wear,
significant increases in their content compared with the running-in plot should signal shut down and
maintenance. A sample plot from a review by Lockwood and Dally (1992) is shown in Figure 8.13. It
shows the change in size distribution as running hours accumulate.
The shape, thickness, and chemistry of a wear particle can tell much about how it was formed.
Individual particles can be examined by SEM. An atlas of particle types has been assembled by Anderson
(1982) and contains a wide variety of wear particles collect by ferrography. Micrographs, some in color,
are presented together with the conditions that produced the wear particles. A page from the atlas is
shown in Figure 8.14. The figure shows the way in which the particles line up in strings in the ferrographs.
Individual particles can be seen in SEM micrographs in Figures 8.15 and 8.16. Some of the more easily
recognized characteristics of wear particles include:
• Wear-in conditions. Flat, fairly smooth particles about 10 µm or smaller formed from broken off
lips and extrusions produced by machining.
• Cutting. Curled strings and crumpled flakes resulting from abrasive wear process.
• Deformation. Flat flakes with parallel striations from breakout of deformed surface material caused
by heavy sliding contact.

© 2001 by CRC Press LLC


Catastrophic failure

Wear particle concentration.


Advanced failure mode
Onset of severe wear mode

particles/10 mL
Benign wear

e
m
Ti
0.1 1 10 100 103
Wear particle size, µm

FIGURE 8.13 Plot of wear particle size distribution from oil analysis.

• Contact fatigue. Chunk-like particles 50 µm in size resulting from contact fatigue spalls. Some
investigators, including Douglas Scott (1983) found spherical debris 1 to 5 µm in size having the
same composition as the steel of ball bearings in engine oil and attributed them to contact fatigue.
One must be careful in drawing conclusions from spherical debris. It has been found in hydraulic
systems and engines having no rolling contact bearing in the oil stream. J.E. Davies (1986) found
spherical particles in diesel engine oil. A dendritic pattern was shown on the particle surfaces as
shown in Figure 8.7. These were solidified liquid drops. The spherisurface features and their
composition should be determined to diagnose their source.
• Babbitt fatigue. Silver-colored flakes of babbitt analyzing for tin or lead (by EDS).
• Fretting. For iron base materials, orange powder with particle size on the order of 100 nm and
containing alpha Fe2O3 (hematite).
Under heavy load, surface features can be produced that are peculiar to the surface morphology. For
instance, Sarkar (1983) noted that the surface of an Al–Si alloy pin which had been run against a hardened
disk had developed “inclined shear plates” on the surface. A particle, extracted from the wear debris
showed the same morphology. Therefore, it was recognized as a piece of the pin surface broken out
during heavy sliding contact. The pin and particle are shown in Figure 8.15.
Ferrographs will also collect nonmagnetic and nonmetallic particles. They will tend to precipitate out
of the fluid as it moves over the glass slide. The same analysis can be performed on wear particles separated
from oil by other means (filters, centrifuging). Organic materials can be detected by viewing the slide
with transmitted light. Polarized light will show up crystalline minerals.

8.4.2 Biological Debris Sampling


Human joints are similar to lubricated mechanical components. The hip joint is a ball-in-socket config-
uration with cartilage sliding against cartilage, lubricated with a slippery lubricant called synovial fluid.
As these joints age, the cartilage is worn away until exposed bone slides on bone. This causes irritation
of the joint and inflammation and pain. In the last 25 to 30 years, worn joints have been replaced with
artificial components. One part of the joint prosthesis is made of a tough plastic compatible with body
tissue (ultra-high-molecular weight polyethylene). The other component is a metal alloy (titanium alloy
or cobalt base alloy). The combination operates with relatively low friction but is subject to wear. Both
the original human joint or the prosthesis can be tested for wear by using the ferrographic method as
shown by Mills and Hunter (1983). An example of extracted cartilage particles is shown in Figure 8.16.

© 2001 by CRC Press LLC


FIGURE 8.14 Page from Wear Particle Atlas.

In biological joints, samples of synovial fluid were extracted and treated with a magnetizing solution by
Evans and Bowen (1980). The particles recovered by ferrography can be identified by stains or X-ray
analysis.
Examination of tissues around failed total hip replacements for UHMWPE wear debris requires a
special process described by Campbell, Ma, Yeom et al. (1995). Tissue samples were processed and the
polymer wear debris extracted and deposited on Nucleopore polycarbonate filters. The filter plates

© 2001 by CRC Press LLC


FIGURE 8.15 SEM micrographs of a wear scar and a wear debris particle showing the same surface roughness
features.

FIGURE 8.16 SEM micrograph of cartilage particles extracted from a human joint and separated by ferrography.

containing the wear debris were sputter coated with gold palladium and examined in an SEM. A typical
SEM micrograph of the debris is shown in Figure 8.17.

8.4.3 Reworked Wear Debris


It must be assumed that debris collected from a wear process is as it was when separated from a surface
if one is to deduce how it was generated. This often is not the case. Wear debris can go through a contact
zone many times and in the process be altered in size and shape. This is true in roller bearings, recipro-
cating flat-on-flat sliders, splines, gear couplings, and wet clutches. Ductile metals can be mashed out
into extremely thin flakes — thin enough to transmit electrons. When these submicron particles are
examined in a transmitting electron microscope (TEM), they appear transparent. An example is shown
in Figure 8.18. In experiments to produce this type of particle, iron powders were ball milled under
lubricated conditions by Glaeser (1991) and similar submicron flakes produced as shown in Figure 8.19.

8.5 Conclusions
Wear debris comes in many sizes and shapes. Study of wear debris morphology and chemistry can provide
clues to the condition of various lubricated parts in a machine and the wear mechanisms extant — all

© 2001 by CRC Press LLC


FIGURE 8.17 SEM micrograph of UHMWPE wear debris from a human joint prosthesis.

FIGURE 8.18 TEM micrograph of a very thin submicron wear flake.

this without tearing an engine down. By monitoring the amounts of given metals in a circulating oil
system as a function of running time, one can detect the onset of a mechanical part (bearing, gear, piston
ring, or seal) failure. This program (SOAP) has been successfully used for a number of years for condition
monitoring of engines.
Individual wear particles can provide clues to the wear mechanism that produced them. For instance,
a sudden increase in bronze particles from an engine with bronze bearings may be the result of ingestion
of abrasive contaminants. Just monitoring the increase in copper in the oil would not indicate the change
in wear mode. Bronze particle size and shape would suggest this. Failure analyses can be enhanced by
the examination of wear particle size and shape.

© 2001 by CRC Press LLC


FIGURE 8.19 TEM micrograph of an iron particle subjected to ball milling to simulate wear debris subject to
repeated passage through a rolling contract.

Researchers can use wear particle analysis as a tool to follow the wear process as it changes with changes
in operating conditions (bearing load, change in environment, increased sliding velocity). Classification
of wear debris according to size, thickness, shape, color, and chemistry and relating debris types to wear
mechanisms or failure of mechanical components has been the goal of debris atlases. However, care must
be exercised in using this information since the character of wear debris is often influenced by factors
other than wear mode or parts failure. Diagnostics with wear debris requires experience in interpreting
and recognizing significant characteristics.

References
Anderson, D.P. (1982), Wear Particle Atlas, NAEC Report 92-163, Lakehurst, NJ.
Bowden, F.P. and Tabor, D. (1964), The friction and lubrication of solids, Part 1 (1950) and Part II,
Oxford Press, London.
Bowen, E.R., Scott, D., Seifert, W.W., and Westcott, V.C. (1964), Ferrography, Tribology International,
9(3), 109-115.
Campbell, P, Ma, S., Yeom, B., McKellop, H., Schmalzried, T.P., and Amstuz, H.C. (1995), Isolation of
predominantly submicron sized UHMWPE particles from periprosthetic tissue, J. Biomed Material
Res., 29, 127-131.
Davies, J.E. (1986), Spherical debris in engine oil, Wear, 107-189.
Evans, C.H., Bowen, E.R. et al. (1980), Synovial fluid analysis by ferrography, J. Biochem. Biophys. Meth.,
2, 11-18.
Glaeser, W.A. (1981), Wear experiments in the scanning electron microscope, Wear, 73, 371-386.
Glaeser, W.A. (1983), The nature of wear debris generated during lubricated wear, ASLE Transactions,
26, 517-522.
Glaeser, W.A. (1991), Ball mill simulation of wear debris attrition, Proceedings, 18th Leeds–Lyon Sym-
posium on Tribology, Tribology Series 21, Elsevier, 515-521.
Holm, R. (1946), Electric Contacts, Almquist and Wicksells, Uppsala, Sweden.
International Conference on Advances in Ferrography (1983), Wear, 90, 1-181.
Lockwood, F.F. and Dally (1992), Lubricant Analysis, 18th edition, Metals Handbook, Tribology, ASM, 19,
300-312.
Lukas, M. and Anderson, D.P. (1998), Laboratory used oil analysis methods, Lubrication Engineering,
STLE, 31-35.

© 2001 by CRC Press LLC


Mills, G.H. and Hunter, J.A. (1983), A preliminary use of ferrography in the study of arthritic diseases,
Wear, 90, 107-111.
Mulhearn, T.O. and Samuels, L.E. (1962), Abrasion of metals, Wear, 5, 478-498.
Rigney, D.A. (1992), The role of characterization in understanding debris generation, Tribology Series
21, Wear Particles: From the Cradle to the Grave, Elsevier, 405-412.
Sarkar, A., D. (1983), The role of wear debris in the study of wear, Wear, 90, 39-47.
Scott, D. (1983), The application of ferrography to the condition monitoring of gas turbines, Wear, 90,
21-29.
Seifert, W.W. and Westcott, V.C. (1972), A method for the study of wear particles in lubricating oil, Wear,
21, 22-42.
Zhao, X. and Bhushan, B. (1998), Material removal mechanisms of single-crystal silicon on nanoscale
and at ultralow loads, Wear, 223, 66-78.

© 2001 by CRC Press LLC

You might also like