Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/293043390

Si 3 N 4 as a biomaterial and its tribo-


characterization under water lubrication

Article in Lubrication Science · February 2016


DOI: 10.1002/ls.1329

CITATIONS READS

0 30

1 author:

Yilmaz Ozmen
45 PUBLICATIONS 68 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Graphene materials for biomedical applications and its bio-tribo-characterizations View project

Water Lubrication of Ceramics under harsh conditions View project

All content following this page was uploaded by Yilmaz Ozmen on 10 February 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
LUBRICATION SCIENCE
Lubrication Science (2016)
Published online in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/ls.1329

Si3N4 as a biomaterial and its tribo-characterization under water


lubrication

Yılmaz Özmen*,†
Pamukkale University, Biomedical Engineering Dept., 20070 Denizli, Turkey

ABSTRACT
Hip implant wear is recognised as the main cause of hip implant failure therefore has been widely inves-
tigated both experimentally and clinically, demonstrating the coexistence of abrasive, adhesive, fatigue and
corrosive wear. Many clinical in vivo and bulk material wear rate data from published literature have been
presented for non-oxide ceramic implants. Several studies have shown that the coefficient of friction of self-
mated silicon nitride in water decreases from an initially high value to about 0.002 after a certain run-in
period. Since the worn surfaces become extremely smooth, the low friction is attributed to the initiation
of hydrodynamic lubrication by a thin water film at the interface. The possibility of mixed lubrication,
i.e. hydrodynamic lubrication by water and boundary lubrication due to the presence of colloidal silica on
the wearing surfaces, has also been proposed.
Influence of load, speed and surface roughness on the duration of the run-in period of silicon nitride
under water lubrication was investigated in this study. The results confirmed that a low coefficient of fric-
tion is obtained following a run-in period when a wear scar of sufficient size is developed to reduce the
contact stress. The run-in period, during which the coefficient of friction is fairly high, is shorter for
smoother surfaces and at higher loads and speeds. The striations that appeared to be associated with the
high-friction spikes can be formed as a result of surface film breakdown. Although the results are consis-
tent with the proposed mechanisms of hydrodynamic lubrication or mixed lubrication, it is proposed that
the low-friction behaviour may also be related to fundamental interactions between two hard and elasti-
cally deforming surfaces covered with hydrogen-terminated oxide films. Copyright © 2016 John Wiley
& Sons, Ltd.

Received 7 May 2015; Revised 29 August 2015; Accepted 6 December 2015

KEY WORDS: silicon nitride; water lubrication; biomaterial; total joint replacements; friction; wear

INTRODUCTION

Industrial and commercial use of silicon nitride ceramics is common due to the fact that silicon nitride
has intrinsic mechanobiochemical properties.1 It has been used in engines, cutting tools and ball bear-
ings for its high mechanical performance, and their extreme strength has been validated using a

*Correspondence to: Yilmaz Özmen, Pamukkale University, Biomedical Engineering Dept., 20070 Denizli, Turkey.

E-mail: yilmaz.ozmen@gmail.com

Copyright © 2016 John Wiley & Sons, Ltd.


Y. ÖZMEN

number of techniques due to attractive tribological properties in extraordinary working environments,


i.e. high temperature or biochemical activities. During the last decades, advantages of this material
have led many investigators to hypothesise that silicon nitride may also have applicability in biomed-
ical fields.2–4 The choice of silicon nitride as an implant material is persuaded by its excellent biocom-
patibility, low wear rates as well as relatively high fracture toughness and strength.5–8
It can be outlined that silicon nitride could be used for possible prosthetic implant applications as
components and devices: total joint replacements,9–12 mini-osteofixation systems (plates, screws,
etc.),13,14 multiwall drug-release devices,15 micro-electromechanical systems (Bio-MEMS),16 3D
micro-electrode arrays,17 intervertebral spacers and spinal surgery,18 implantations in otorhinolaryn-
gology and traumatology,13,14 etc. Silicon nitride is of interest, because it is predicted to show high
wear resistance in water as well as produce soluble wear particles in vivo. This combination would
minimise the risk of osteolysis and implant loosening and consequently prolong the implant lifespan.19
Nowadays, about 200 000 and 80 000 interventions/year are performed in the USA and in the UK,
respectively, and they are estimated to increase of about 170% by 2030.20 Although hip arthroplasty
is considered one of the greatest achievements in orthopaedic surgery in the last decades,21 from an
engineering point of view, hip implants are not a complete success and still need further
developments.
When it comes to total hip joint replacements, it is the whole system, rather than the individual ma-
terials, that have to be biocompatible. It will be more realistic if we use ‘system approach for tribolog-
ical behaviours’ in this kind of comprehensive mechanisms. In addition, one must consider some
controversy in the literature about the biocompatibility of silicon nitride.5 Furthermore, published fric-
tion and wear data for different types of silicon nitride show a wide scatter due to different test condi-
tions used in various studies.
Many of the failures of total joint replacements are related to tribology, i.e. wear at the cup, head and
liner interface. Accumulation of wear particles at the adjacent bone can lead to bone loss and aseptic
implant loosening. Therefore, it is highly desirable to reduce the generation of wear particles from the
implant surfaces.21,22
It is claimed that the wear particles produced, when Si3N4 slides against Si3N4 in water often con-
sists of silica (SiO2) and are mainly amorphous,23 and small contents of Si4+ ions in extracts after
solubility tests of Si3N4 is found.24 The behaviour of ion release into biofluid is governed by the
electrochemical rule. Released ions do not always combine with biomolecules to appear toxic because
active ion immediately combine with a water molecule or an anion near the ion to form an oxide,
hydroxide or inorganic salt. Thus, there is only a small chance that the ion will combine with
biomolecules to cause cytotoxicity, allergy and other biological influences.25 Implant failures
can be due to several factors, but one of the most critical is release of wear particles from
the bearing surface of the implant. Accumulation of wear particles at the implant has been
linked to osteolysis that leads to bone loss and eventually aseptic implant loosening. The chem-
istry and particle size have been found to be of high importance to the inflammation response.26–28
Generation of wear particles and sources of friction coefficient considerably alters before and
after transition mode. While it is generally adhesive and abrasive contact before the transition pe-
riod (unsteady state), it turns in to adhesive, tribochemical and hydrodynamic contact after transi-
tion period (steady state or running in). It is better for any tribo-system reach steady state and/or
running in situation sooner. Transition (change from one form, state or style to another) could
be understood as the reaching of the tribo-parameter, i.e. friction, to a new value that could be higher
or lower (in our case, it reaches lower friction). The wear mode changes from mechanically

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
SI3N4 AS A BIOMATERIAL AND ITS TRIBO-CHARACTERIZATION

dominated wear, including transgranular and intergranular fracture-induced wear, to tribochemically


dominated wear, including the formation and delamination of the reaction film of silicon oxide or
hydrated silicon oxide.29
Scientists have dedicated significant efforts to characterise biomaterials in order to unveil their
structure and better understand their properties, i.e. tribological properties of silicon
nitride.1,4,30,31 Reaction products and wear particles of silicon nitride dissolve in aqueous
media,32–34 which may reduce the negative body response from the debris, and potentially
increase the longevity of the implant. The suitability of silicon nitride for hip and knee bearings
has been debated in the literature.4,35 Most authors agree that in the absence of material oxidation
in vivo, silicon nitride has the lowest friction, necessary to articulate against itself, even when
water is the only lubricant.5,35,36
The number of factors influences tribological performance of total joint systems. These factors in-
clude type of materials, contact stresses, surface hardness and roughness, number of cycles, lubri-
cants, coatings, solution particle count and distribution, oxidation of materials, surface abrasions
of both materials, etc.22,37 Aqueous solutions are drawing attention because of their ecological ad-
vantages, high thermal conductivity, non-toxicity, biodegradability, environmental friendliness and
good solvency. However, water-based solutions have several disadvantages as lubricants, such as
corrosiveness, vaporization and non-viscosity in case it is used against oxide ceramics or
metals.38–40 Besides, the mechanisms of aqueous lubrication are still not well known because of
the diversity and complexity of aqueous solutions. Some traditional lubrication theories are no lon-
ger applicable to aqueous lubrication because of the complex physical and chemical properties of
water.
The objective of this paper is to discuss the tribochemical reaction of self-mated silicon ni-
tride (non-oxide ceramic) sliding in water. Non-oxide ceramics in Total Joint Replacement
(TJR) have rarely been studied and published about especially self-mated materials with water
lubrication. Such materials, sometimes referred to as functional materials, are of fundamental
scientific interest and potentially useful in countless applications that range very wide area. It
is aimed at further extending our understanding of silicon nitride as a biomaterial and develop
new technologies in the field of biomaterials.

EXPERIMENTAL DETAILS

Materials
A bearing-grade silicon nitride (NBD-200, Norton Advanced Ceramics) was selected for these exper-
iments. The materials used in this investigation are characterised and tribochemically polished. Some
of the mechanical and chemical properties of the samples were given in the Tables I and II.

Tribo-characterizations of silicon nitride


Friction and wear characterizations of the Si3N4 samples were carried out by CSEM pin-on-disk
system with the parameters given in the Table III. The pin is mounted on a stiff lever, designed as a
frictionless force transducer. The friction coefficient is determined during the test by measuring the de-
flection of the elastic arm. Wear coefficients for the pin and disk materials are calculated from the
volume of material lost during the test.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
Y. ÖZMEN

Table I. Some of the mechanical properties of typical NBD-200 as received.41


Property Units Value
3
Density gm c 3.16
Flexural strength MPa 800
Elastic modulus GPa 320
Poisson’s ratio — 0.26
Compressive strength GPa 3
Hertz compressive strength GPa 28
Hardness, HV GPa 16.6
Tensile strength MPa 400
Fracture toughness K (Ic) Mpa m1/2 4.1

Bearing balls (12.7 mm diameter) in the as-received condition with a roughness (Ra) of 2 nm were
used for the pins. While the surface roughness of the lapped disks was about 50 nm (will be called rough
(R) surface), the roughness of the polished disks was nearly 2 nm (will be called fine (F) surface).
According to the manufacturer, the hardness of this material is 16.6 GPa. Disks (70 mm OD) with two
different (R and F) surface finishes were used in the tests.
The water used in the experiments was distilled, deionised and filtered. Both of the samples were
fully immersed in the water. Before tests, the samples were cleaned ultrasonically by ethanol (5 min)
and acetone (1 min) and then washed by the processed water thoroughly. Experimental conditions
were chosen in order to reveal the complete transition period and make comparison with other re-
searchers’ results.29,33 Friction force was recorded throughout the experiments. Then, friction coeffi-
cient was plotted versus time. In order to obtain wear rates, both samples were characterised by
optical microscopy and surface profilometer. All test conditions were conducted twice at every level.
The balls and the disks were also characterised by atomic force microscope (AFM) to assess the sur-
face topography within the wear scars and the wear tracks perpendicular to the sliding direction and by
scanning electron microscopy (SEM) to assess the condition of the worn surfaces. The surface rough-
ness of the balls and disks were measured with an AFM scanning several areas 50 × 50 μm in dimen-
sions before and after the experiments in order to understand surface contact conditions. Besides,
optical photographs of both ball and disk samples were taken.

RESULTS AND DISCUSSION

Severe stick slip occurs at the start of the test. We note that the observed friction force response dur-
ing stick slip is strongly dependent on the compliance and speed control characteristics of the
tribometer. The data are shown without filtering and with filtering using a 100-point moving average.
A typical friction–time plot is shown in Figure 1. The friction curve is smoother, and the friction force
trend is more apparent after filtering. Filtering has little effect on the plotted data once low-friction

Table II. Chemical composition of typical NBD-200 as received.41


Al C Ca Fe Mg O Si3N4
≤0.5 ≤0.88 ≤0.04 ≤0.17 0.6–1.0 2.3–3.3 97.1–94.11

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
SI3N4 AS A BIOMATERIAL AND ITS TRIBO-CHARACTERIZATION

Table III. Tribo-test conditions.


Sliding speeds (mm s1) 60, 120
Normal loads (N) 3, 5
Time (h) 24
Lubricants Purified water

sliding starts. In general, the friction test results were repeatable as seen in Figure 2. Note the stick slip
behaviour at the start of the test and a smoother behaviour after filtering the data using a 100-point
moving average.
It has been explained an iterated in detail42 about the plot of friction characteristics depending on
load, speed, surface roughness, etc. To summarise, on the lapped surface, the transition occurs more
quickly at the higher load; for the polished surfaces, the load effect is not as pronounced as for the
lapped surfaces. The transition period is shorter at the higher speed for both loads.
Both the wear scars and wear tracks are examined in the SEM/AFM/surface profilometer. Under
some circumstances, it contains striations exhibiting evidence of plastic deformation, delamination
and fracture (Figure 3). These striations are similar to what is normally observed for scuffing on metal
surfaces that happens when localised welding (or adhesion) occurs due to the breakdown of a lubricant
film. The observed high-friction spikes are probably associated with the striations.
Surface profilometer views of the wear tracks on the surfaces of the disks are presented in Figure 4.
Balls created very shallow wear tracks (less than 0.5μm depth) in water under a test load of 5 N, while
the depths of the wear tracks under a test load of 3 N were approximately 0.25μm depth. In the present
study, the results of the wear tests were quantified in terms of the volume of material removed over the
entire testing period. Wear rate of the disks and balls are presented in the unit of [mm3 Nm1] and cal-
culated as the volume of the material removed normalised by the test load and the sliding distance; for
disk samples, wear scar area are obtained by profilometer and wear volumes are calculated multiplying
by sliding distance in one circle. For ball samples, wear scar diameter are measured and wear volume
are calculated by using (πd2v.d2p)/64R equation (where dv and dp are vertical and parallel wear scar di-
ameters, respectively, and R is the radius of the ball).
It seems that dominating factor for transition period is the reaching conformity of two surfaces. Stick
slip phenomenon is observed in every test condition. However, transition takes place after a while. In

Figure 1. Friction coefficient versus time under 5 N load and 120 mm s1 sliding speed for rough surface.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
Y. ÖZMEN

Figure 2. Repeatability of the frictional behaviour shown for two tests at (a) 5 N and 120 mm s1 on a fine-
polished disk and (b) 5 N and 60 mm s1 on a rough lapped disk.

Figure 3. Typical images showing striations on the polished disk wear track. (a) SEM, (b) AFM and (c)
optical images.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
SI3N4 AS A BIOMATERIAL AND ITS TRIBO-CHARACTERIZATION

Figure 4. Typical surface profilometer measurements on the polished and lapped disks wear track. (1) R,
60 mm s1, 3–5 N; (2) R, 120 mm s1, 3–5 N; (3) F, 60 mm s1, 3–5 N; (4) F, 120 mm s1, 3–5 N.

the rough surface, it is shorter, and certain amount of wear may be necessary to create a gel-like struc-
ture or chemical reaction in the contact point of the surfaces. Wear particles are dissolved in the water
and cannot precipitate even after a very long waiting time. This reminds that dissolved wear particles
maintain low-friction coefficient and wear rate. If the initial surface roughness of the disk sample is
larger, transition period is longer and wear rate is higher. Sliding speed should also be high enough
to maintain elastohydrodynamic lubrication regime. At low speed, contact of asperities takes place in-
tensively causing the disturbance of the friction coefficient and leading high wear. Roughness and slid-
ing speed effects are much clearer, than force effect (Figures 5 and 6).
When the ball and the disk are smooth, run-in probably is necessary to produce a flat on the ball and
reduces the contact stress. Both the ball and the rougher disk must wear-in prior to the transition to the
low-friction condition. Wear rates occurred almost same in both experiments (i.e. test A series and test
B series) with some exception.
In general, the wear rates were higher for the balls than the disks. Calculation of the wear rates based
on total sliding distance is misleading, because of the transition phenomena and the possibility of much

Figure 5. The effect of sliding speed and load on the run-in period for a rough lapped surface depicted by
the filtered friction data. (1) 120 mm s1, 3 N; (2) 120 mm s1, 5 N; (3) 60 mm s1, 3 N; (4) 60 mm s1, 5 N.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
Y. ÖZMEN

Figure 6. The effect of sliding speed and the roughness on the run-in period for 5 N loads depicted by the
filtered friction data. (1) F, 60 mm s1; (2) F, 120 mm s1; (3) R, 120 mm s1; (4) R, 60 mm s1.
higher wear rate during the high-friction period. Therefore, modified wear rates (the wear in the high-
friction regime, before transition period) calculated are plotted in Figure 7 in the unit of [mm3 Nm1].
We can see the grain pullout and massive wear particles or groves on the ball wear track (Figure 8).
These phenomena may be the reason of random rise of friction coefficient after transition period.
The tribological tests performed in this study shows that the coefficient of friction does not directly
correlate to the wear of the components. Material properties should be interpreted intensively and
widely, in order to explain the differences. Local differentiation of materials’ bulk structure may be
main cause of this.
Wear creates a smooth and flat surface on the Si3N4 ball so that a flat-on-flat friction is established.
The result is a reduction of the local stresses responsible for the mechanical wear. When the area of the
wear scar is large enough, the friction may drop gradually. For the low-friction cases, the contact

Figure 7. Modified wear rates of the balls and disks at the duration of run-in period.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
SI3N4 AS A BIOMATERIAL AND ITS TRIBO-CHARACTERIZATION

Figure 8. Typical images showing striations on the polished ball wear tracks (F, 3 N, 60 mm s1). (a) AFM
and (b) optical images.

surfaces become very smooth and flat after different running in periods. For Si3N4/Si3N4 tribo-elements, it
was attributed the smoothing process to tribochemical wear.11,31,43 Smooth surface is necessary to gener-
ate hydrodynamic lubrication that maintains a low-friction coefficient.44,45 No wear particles are found on
the pin or on the disk. The particles removed from sliding solids are dissolved in water. Any particles, if
present, would have to be smaller than 70 nm; otherwise, they would prevent hydrodynamic lubrication.46
Silicon nitride, working against itself, or metal, or polyethylene seems to satisfy tribological require-
ments since under test conditions, the contact surface of silicon nitride becomes ultra-smooth due to
tribochemical polishing, and the friction becomes very low at increasing sliding distances. These
chemical and physical properties work effectively to reduce wear and to form smooth wear surface,
which is necessary for good lubrication and low friction. Thus, it should be an excellent material for
total hip bearings,5 especially in light of very favourable tribological properties in water.
The wear of Si3N4 is a tribochemical dissolution of the particles via the formation of SiO2. Then, its
dissolution in the form of silicic acid in water produces NH3 confirming the possible reactions30,31,47
between silicon nitride and water, namely are

Si3 N 4 þ 6H 2 ↔3SiO2 4NH 3


SiO2 þ 2H 2 O↔SiðOH Þ4

This dissolution of the particles occurs only under friction; it is therefore a tribochemical form of
wear in which they are removed molecule by molecule instead of massive particles. In addition, disso-
lution of the reaction products in water leaves only a very thin film or hydrogen-terminated surface and
a much lower coefficient of friction. During the process of tribochemical wear or polishing, oxidation
reaction and hydration occur. That allows material to be removed at the atomic level at very low
stresses, and no deformation or fracture occurs. As a result, nanometer-thick films are formed that
protect the surfaces and are almost instantly self-repairing. A thin film of silicon oxide (SiO) on
the order of a few nanometers is formed23,48 with the surface terminated with hydroxyl groups
(i.e. O–H bonds). Sliding then takes place between two silicon nitride surfaces with thin

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
Y. ÖZMEN

hydrogen-terminated SiO films. In order to achieve such a low-friction condition, however, the
surfaces must wear to develop a high degree of conformity and the wear scar of sufficient size
must develop to reduce the contact pressure. Further research, e.g. molecular dynamic simulation,
is needed to examine the validity of the proposed hypothesis.
It is considered that the wear particles generated on the surface, the remaining place behave
like micro-cracks (refer to Figures 3 and 8). Consequently, both hydrodynamic lubrication effect
and the change of surface property by polishing or tribo-film formation could bring lower
friction.49
Si3N4 ceramics are characterised by high hardness, thermal shock resistance and unusually high
strength and fracture toughness,50 which make them suitable for TJR applications.51 However, a
controversial discussion is taking place in the literature about their superficial oxidation, which results
in a silicon oxide (SiO2) rich layer several nanometer thick.52 Those oxidised layers have been found
on Si3N4 surfaces, and concerns have been expressed that they may chip off over time, resulting in a
significantly increased third-body wear (increased wear due to hard particles between softer articulat-
ing surfaces).53 However, the role of SiO2 flaking still needs to be clarified before clinical follow-up
studies can be planned.
The tests could be carried out in a synthetic body fluid, which has ion concentrations close to human
blood plasma. However, it is a hard task to keep on the testing environment as its fresh and original
composition. Tribological systems inevitably change with wear. Rather than trying to avoid these
changes, a better strategy can be to design systems that, after the expected changes, achieve the desired
friction and wear properties.
Future work should focus on understanding not only the wear but also the wear particles, since
dissolution of wear particles may reduce a possible negative body response. The amount, size,
shape, composition, bonding structure of the particles as well as pH etc. could affect the dissolu-
tion rate.

CONCLUSIONS

During sliding, high-friction spikes were observed occasionally. The observed high-friction spikes
may be associated with the surface striations that could form due to lubricant film breakdown. The sur-
faces of the wear tracks and wear scars contained a series of striations parallel to the sliding direction
and exhibiting plastic deformation, delamination and fracture. It is proposed that the low friction could
also be related to fundamental interactions between two hard and elastically deforming surfaces cov-
ered with hydrogen-terminated oxide films. In order to achieve such a low-friction condition, however,
the surfaces must wear to develop a high degree of conformity and the wear scar of sufficient size must
develop to reduce the contact pressure. Thus, low friction can only be obtained after a much smoother
worn surface achieved. The smoothing process may be attributed to tribochemical wear. Mechanical
wear could be the dominating factor at the high-friction case.

ACKNOWLEDGEMENTS

The author is grateful to Dr. Said Jahanmir, who guided and supported this project, for his enthusiastic encourage-
ment and his tremendous contribution to tribology. The tribochemically polished disks were supplied by Professor
T.E. Fischer at Stevens Institute of Technology. This project was supported by DARPA.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
SI3N4 AS A BIOMATERIAL AND ITS TRIBO-CHARACTERIZATION

REFERENCES
1. Muratov VA, Luangvaranunt T, Fischer TE. The tribochemistry of silicon nitride: effects of friction, temperature and sliding
velocity. Tribology International 1998; 31(10):601–611.
2. Howlett CR, McCartney E, Ching W. The effect of silicon-nitride ceramic on rabbit skeletal cells and tissue—an in vitro and
in vivo investigation. Clinical Orthopaedics and Related Research 1989; 244:293–304.
3. Neumann A, Jahnke K, Maier HR, Ragoß C. Biocompatibilty of silicon nitride ceramic in vitro. A comparative
fluorescence-microscopic and scanning electron-microscopic study. Laryngo-Rhino-Otologie 2004; 83(12):845–851.
4. Bal BS, Rahaman MN. Orthopedic applications of silicon nitride ceramics. Acta Biomaterialia 2012; 8(8):2889–2898.
5. Mazzocchi M, Bellosi A. On the possibility of silicon nitride as a ceramic for structural orthopaedic implants. Part I:
processing, microstructure, mechanical properties, cytotoxicity. Journal of Materials Science-Materials in Medicine 2008;
19(8):2881–2887.
6. Bal BS, Khandkar A, Lakshminarayanan R, Clarke I, Hoffman AA, Rahaman MN. Testing of silicon nitride ceramic bearings
for total hip arthroplasty. Journal of Biomedical Materials Research Part B-Applied Biomaterials 2008; 87B(2):447–454.
7. Bal BS, Khandkar A, Lakshminarayanan R, Clarke I, Hoffman AA, Rahaman MN. Fabrication and testing of silicon nitride
bearings in total hip arthroplasty winner of the 2007 "HAP" PAUL Award. Journal of Arthroplasty 2009; 24(1):110–116.
8. Pettersson M, Tkachenko S, Schmidt S, Berlind T, Jacobson S, Hultman L, Engqvist H, Persson C. Mechanical and tribo-
logical behavior of silicon nitride and silicon carbon nitride coatings for total joint replacements. Journal of the Mechanical
Behavior of Biomedical Materials 2013; 25:41–47.
9. Zhou YS, Ohashi M, Tomita N, Ikeuchi K, Takashima K. Study on the possibility of silicon nitride silicon nitride as a ma-
terial for hip prostheses. Materials Science & Engineering C-Biomimetic Materials Sensors and Systems 1997; 5(2):125–129.
10. Akazawa M, Kato K. Wear properties of Si3N4 in rolling-sliding contact. Wear 1988; 124(2):123–132.
11. Jahanmir S, Fischer TE. Friction and wear of silicon-nitride lubricated by humid air, water, hexadecane and hexadecane
+ 0.5-percent stearic-acıd. Tribology Transactions 1988; 31(1):32–43.
12. Zhou YS, Ikeuchi K, Ohashi M. Comparison of the friction properties of four ceramic materials for joint replacements. Wear
1997; 210(1–2):171–177.
13. Neumann A, Unkel C, Werry C, Herborn CU, Maier HR, Ragoß C, Jahnke K. Prototype of a silicon nitride ceramic-based
miniplate osteofixation system for the midface. Otolaryngology - Head and Neck Surgery 2006; 134(6):923–930.
14. Neumann A, Unkel C, Werry C, Herborn CU, Maier HR, Ragoss C, Jahnke K. Osteosynthesis in facial bones.
Silicon nitride ceramic as material. HNO 2006; 54(12):937–942.
15. LaVan DA, McGuire T, Langer R. Small-scale systems for in vivo drug delivery. Nature Biotechnology 2003; 21(10):1184–1191.
16. Kotzar G, Freas M, Abel P, Fleischman A, Roy S, Zorman C, Moran JM, Melzak J. Evaluation of MEMS materials of con-
struction for implantable medical devices. Biomaterials 2002; 23(13):2737–2750.
17. Kristensen BW, Noraberg J, Thiébaud P, Koudelka-Hep M, Zimmer J. Biocompatibility of silicon-based arrays of electrodes
coupled to organotypic hippocampal brain slice cultures. Brain Research 2001; 896(1–2):1–17.
18. Sorrell CC, Hardcastle PH, Druitt RK, Howlett CR, McCartney ER. Results of 15-year clinical study of reaction bonded silicon
nitride intervertebral spacers. In Transactions - 7th World Biomaterials Congress, 2004.
19. Oloffson J. Friction and wear mechanisms of ceramic surfaces, in Department of Engineering Science, Ph.D. Thesis,
Uppsala University: Uppsala, Sweden, 2011; 66.
20. Kurtz S, Ong K, Lau E, Mowat F, Halpern M. Projections of primary and revision hip and knee arthroplasty in the United
States from 2005 to 2030. Journal of Bone and Joint Surgery - Series A 2007; 89(4):780–785.
21. Mattei L, Di Puccio F, Piccigallo B, Ciulli E. Lubrication and wear modelling of artificial hip joints: a review. Tribology Inter-
national 2011; 44(5):532–549.
22. Buford A, Goswami T. Review of wear mechanisms in hip implants: paper I—general. Materials and Design 2004;
25(5):385–393.
23. Xu J, Kato K. Formation of tribochemical layer of ceramics sliding in water and its role for low friction. Wear 2000;
245(1–2):61–75.
24. Guedes e Silva CC, Higa OZ, Bressiani JC. Cytotoxic evaluation of silicon nitride-based ceramics. Materials Science and
Engineering: C 2004; 24(5):643–646.
25. Hanawa T. Metal ion release from metal implants. Materials Science and Engineering: C 2004; 24(6–8 SPEC.
Iss.):745–752.
26. Sargeant A, Goswami T. Hip implants: paper V. Physiological effects. Materials & Design 2006; 27(4):287–307.
27. Landgraeber S, Von Knoch M, Löer F, Wegner A, Tsokos M, Hussmann B, Totsch M. Extrinsic and intrinsic pathways of
apoptosis in aseptic loosening after total hip replacement. Biomaterials 2008; 29(24–25):3444–3450.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls
Y. ÖZMEN

28. Sargeant A, Goswami T. Hip implants—paper VI—ion concentrations. Materials & Design 2007; 28(1):155–171.
29. Xu JG, Kato K, Hirayama T. The transition of wear mode during the running-in process of silicon nitride sliding in water.
Wear 1997; 205(1–2):55–63.
30. Takadoum J, Houmid-Bennani H, Mairey D. The wear characteristics of silicon nitride. Journal of the European Ceramic
Society 1998; 18(5):553–556.
31. Kato K, Adachi K. Wear of advanced ceramics. Wear 2002; 253(11–12):1097–1104.
32. Laarz E, Zhmud BV, Bergstrom L. Dissolution and deagglomeration of silicon nitride in aqueous medium. Journal of the
American Ceramic Society 2000; 83(10):2394–2400.
33. Chen M, Kato KJ, Adachi K. The comparisons of sliding speed and normal load effect on friction coefficients of self-mated
Si3N4 and SiC under water lubrication. Tribology International 2002; 35(3):129–135.
34. Saito T, Hosoe T, Honda F. Chemical wear of sintered Si3N4, hBN and Si3N4-hBN composites by water lubrication. Wear
2001; 247(2):223–230.
35. Sonntag R, Reinders J, Kretzer JP. What’s next? Alternative materials for articulation in total joint replacement. Acta
Biomaterialia 2012; 8(7):2434–2441.
36. Bal BS, Rahaman M. (2011). The Rationale for Silicon Nitride Bearings in Orthopaedic Applications, Advances in Ce-
ramics - Electric and Magnetic Ceramics, Bioceramics, Ceramics and Environment, Prof. Costas Sikalidis (Ed.), DOI:
10.5772/19381. ISBN: 978-953-307-350-7, InTech, Available from: http://www.intechopen.com/books/advances-in-ce-
ramics-electric-and-magnetic-ceramics-bioceramics-ceramics-and-environment/the-rationale-for-silicon-nitride-bearings-in-
orthopaedic-applications
37. Chen CY, Chung C-J, Wu B-H, Li W-L, Chien C-W, Wu P-H, Cheng C-W. Microstructure and lubricating property
of ultra-fast laser pulse textured silicon carbide seals. Applied Physics A: Materials Science and Processing 2012;
107(2):345–350.
38. Ma LR, Zhang CH, Liu SH. Progress in experimental study of aqueous lubrication. Chinese Science Bulletin 2012;
57(17):2062–2069.
39. Ratoi M, Spikes HA. Lubricating properties of aqueous surfactant solutions. Tribology Transactions 1999; 42(3):479–486.
40. Wong PL, Huang P, Meng Y. The effect of the electric double layer on a very thin water lubricating film. Tribology Letters
2003; 14(3):197–203.
41. Hah SR, Fischer TE, Burk C. Ceramic Bearing Development, in Tribochemical Finishing of Silicon Nitride, Vol. 4. Wright
Labarotory 1995.
42. Jahanmir S, Ozmen Y, Ives LK. Water lubrication of silicon nitride in sliding. Tribology Letters 2004; 17(3):409–417.
43. Jahanmir S. Wear transitions and tribochemical reactions in ceramics. Proceedings of the Institution of Mechanical Engi-
neers, Part J: Journal of Engineering Tribology 2002; 216(6):371–385.
44. Huang W, Xu Y, Zheng Y, Wang X. The tribological performance of Ti(C,N)-based cermet sliding against Si3N4 in
water. Wear 2011; 270(9–10):682–687.
45. Litwin W. Experimental research on water lubricated three layer sliding bearing with lubrication grooves in the upper part of
the bush and its comparison with a rubber bearing. Tribology International 2015; 82(PA):153–161.
46. Fischer TE, Tomizawa H. Interaction of tribochemistry and microfracture in the friction and wear of silicon-nitride. Wear
1985; 105(1):29–45.
47. Ferreira V, Yoshimura HN, Sinatora A. Ultra-low friction coefficient in alumina-silicon nitride pair lubricated with water.
Wear 2012; 296:656–659.
48. Chen CY, Wu BH, Chung C-J, Li W-L, Chien C-W, Wu P-H, Cheng C-W. Low-friction characteristics of nanostructured sur-
faces on silicon carbide for water-lubricated seals. Tribology Letters 2013; 51(1):127–133.
49. Tokoro M, Aiyama Y, Masuko M, Suzuki A, Ito H, Yamamoto K. Improvement of tribological characteristics under water
lubrication of DLC-coatings by surface polishing. Wear 2009; 267(12):2167–2172.
50. Adachi K, Kato K. Formation of smooth wear surfaces on alumina ceramics by embedding and tribo-sintering of fine wear
particles. Wear 2000; 245(1–2):84–91.
51. Olofsson J, Grehk TM, Berlind T, Persson C, Jacobson S, Engqvist H. Evaluation of silicon nitride as a wear resistant and
resorbable alternative for total hip joint replacement. Biomaterials 2012; 2(2):94–102.
52. Sasaki S. The effects of the surrounding atmosphere on the friction and wear of alumina, zirconia, silicon carbide and silicon
nitride. Wear 1989; 134(1):185–200.
53. Rahaman MN, Yao A, Bal BS, Garino JP, Ries MD. Ceramics for prosthetic hip and knee joint replacement. Journal of the
American Ceramic Society 2007; 90(7):1965–1988.

Copyright © 2016 John Wiley & Sons, Ltd. Lubrication Science (2016)
DOI: 10.1002/ls

View publication stats

You might also like