You are on page 1of 7

Carbon 143 (2019) 890e896

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Nanodiamonds for improving lubrication of titanium surfaces in


simulated body fluid
Asghar Shirani a, Nicholas Nunn b, Olga Shenderova b, Eiji Osawa c, Diana Berman a, *
a
Materials Science and Engineering, University of North Texas, Denton, TX, 76203, USA
b mas Nanotechnologies, Inc., Raleigh, NC, 27617, USA
Ada
c
NanoCarbon Research Institute, Asama Research Extension Centre, Shinshu University, Ueda, Nagano, 386-8567, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Hip implants are often made from titanium or titanium-based alloys. However, wear during the oper-
Received 11 October 2018 ation inside the human body is a key source of implant failure and adverse health effects. We propose
Received in revised form new insight on the lubrication of titanium components. Addition of small amounts (less than 0.2 wt%) of
18 November 2018
nanodiamonds (NDs) to simulated body fluid promotes a substantial improvement in friction (3 times
Accepted 1 December 2018
Available online 3 December 2018
reduction) and wear (up to 2 orders of magnitude wear reduction) behavior of the titanium surfaces.
Interestingly, the amount of NDs needed for improvement of friction and wear characteristics is critically
dependent on the applied loads. With higher contact loads, larger concentrations of NDs are needed for
better friction and wear reduction. Analysis of the wear track formed during sliding indicates the for-
mation of a carbon-rich tribolayer which improves tribological properties of the contacting surfaces. Our
results suggest that the carbon layer is formed from the nanodiamonds embedding in the top layer of
titanium.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction high-molecular-weight polyethylene (UHMWPE) was used in


earlier years [8], but raised concerns with regard to adverse tissue
Premature natural joint degeneration is a common problem in reactions [9]. UHMWPE was replaced with stainless steel and then
the population over the age of 40 as a result of excessive loading with Co-Cr-Mo and alumina [10], which demonstrated good wear
conditions as well as failure of normal repair processes [1]. Artificial resistance but lead to inflammation and pain in long-term. So far,
joints made from metal, ceramic, or plastic materials have become titanium remains the most favorable materials for artificial joints.
the only long-term solution for relief from pain, mobility, or other This has led to extensive research on titanium-based alloys for
adverse health effects related to joint degradation and failure [2]. In biomedical applications, such as Ti-Nb-Ta-Zr or TNZT [11], Ti-6Al-
recent years, the number of orthopaedic surgeries substantially 7Nb [12], Ti-6Al-4V [13], and Ti-5Al-2.5Fe [14], among others. Ti-
increased, though reliability and lifetime of the artificial joints tanium and titanium-based alloys are the preferred materials used
remain a major issue. for hip cup shells due to their high corrosion resistance and
During operation, artificial joints are exposed to a complex biocompatibility over other materials, such as conventional stain-
environment and subjected to mechanical degradation [3,4]. less steels and cobalt-based alloys [15]. However, high wear of the
Additionally, biocompatibility of the materials, or their ability not titanium components during exposure to normal and shear stresses
to cause an inflammatory or toxic response, is an important aspect is a major cause for their failure [16]. As a result, degradation of the
to consider [5]. metal implants during movement of the joints limits their lifetime
The search for biocompatible, tribologically efficient materials [17].
led to the exploration of different ceramic and metal alloy com- An alternative to modification of the alloy composition is to
ponents [6]. Ideally, the joint replacement material should exhibit modify the lubrication media with biocompatible additives
an identical performance to the bone when in operation [7]. Ultra- [18e21]. Nanodiamonds have proven to be excellent friction and
wear modifiers in various sliding systems [22e25]. Specifically,
adding small amounts of nanodiamonds to oils [26] or water [27,28]
* Corresponding author. resulted in substantial decrease in friction and wear of sliding metal
E-mail address: Diana.Berman@unt.edu (D. Berman).

https://doi.org/10.1016/j.carbon.2018.12.005
0008-6223/© 2018 Elsevier Ltd. All rights reserved.
A. Shirani et al. / Carbon 143 (2019) 890e896 891

surfaces [24] or silicon\sapphire surfaces [28]. Based on encour- analysis (EDX) to analyze the surface modification changes in the
aging results of improved lubricity of aqueous ND suspensions, it wear track.
was hypothesized that NDs can potentially provide lubrication for To estimate the wear rate after the tests, we calculated the wear
artificial joints [28]. In terms of biocompatibility, different forms of volume of the ball side as follows:
carbon were already considered in various in-vivo and in-vitro
studies [29]. Previous studies demonstrated biocompatibility of ph 3d2
V ¼ð Þð þ h2 Þ (1)
nanodiamonds [30,31]. Naturally occurring graphitic layers have 6 4
been detected in metal-on-metal hip replacements [32]. Therefore,
nanodiamonds could be an ideal solution for improving the where d is the wear scar diameter, r is the radius of the ball, and h is
longevity of titanium components in the body. the wear scar depth:
In this paper, we propose the use of nanodiamonds for friction sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
and wear reduction in titanium-based components of hip implants. d2
h ¼ r  r2  (2)
Addition of small amounts of nanodiamonds to simulated body 4
fluid (SBF), we demonstrate a substantial reduction in friction and
wear of the surfaces.

2. Experimental procedure 3. Results and discussion

To reproduce the joint environment in the body, tribological In the first test, we monitored changes in the coefficient of
testing was performed in a simulated body fluid. The SBF was friction (COF) and wear values of the titanium surfaces during
prepared using a standard protocol [33]. Sodium chloride, sodium sliding under 0.25 N applied load. Our results demonstrate that in
bicarbonate, potassium chloride, potassium phosphate dibasic tri- contrast to pure SBF, the addition of 0.05 wt% of the nanodiamond
hydrate, magnesium chloride hexahydrate, calcium chloride dihy- leads to a two times reduction of the COF value. As demonstrated in
drate, and sodium sulfate, purchased from Sigma Aldrich, were Fig. 1, 0.01 wt% shows almost negligible changes. Therefore, further
dissolved in the right proportion in distilled water and held in an analyses focused on 0.05, 0.1, and 0.15 wt% concentrations.
incubator at a constant temperature of 37  C. The solutions were Increasing the applied load necessitates an increase in the
used within a week to avoid any agglomeration and degradation of amount of nanodiamond for enhanced performance. As demon-
the mixture. The fresh solution was prepared for replicate tests. strated in Fig. 2, 0.1 wt% works better at 0.5 N applied load, while
Grade 5 purity (99.999% titanium) titanium balls (diameter 0.15 wt% is optimal for 1 N applied load. This behavior is attributed
6 mm) and titanium flats (RMS roughness measured with Veeco to higher contact loads requiring a more uniform ND tribolayer,
Dektak 150 Surface Profiler Rq ¼ 50e60 nm) purchased from which may be achieved with higher ND content.
American Elements were used during testing. The hardness of the Analysis of the wear tracks further supports the benefits of
balls and flats measured with a Shimadzu Microhardness Tester different ND concentrations for each specific applied load (Fig. 3).
were 4900 MPa and 3300 MPa, correspondingly. Tests were per- Nevertheless, on average, the presence of NDs in SBF resulted in
formed using the Anton Paar pin-on-disk macroscale tribometer in lower wear than for the SBF alone, and the resulting wear of the
reciprocating mode. The length of the wear track was kept at 5 mm titanium surfaces depends on both parameters: ND concentration
with reciprocal motion at 1 Hz. The samples were immersed in SBF and applied load. In Fig. 3, we highlight the wear scar measure-
during the tests and the temperature was kept at 37  C to ments for the optimal concentrations of NDs at each load.
approximate conditions inside the body a closely as possible. Dur- Further understanding of the mechanism of lubrication
ing the tests, the applied load was varied from 0.25 N up to 1 N improvement in the presence of NDs is possible through detailed
(maximum Hertzian contact pressure of 0.28e0.45 GPa). The con- characterization of the wear tracks corresponding to the tests in
tact pressures experienced by the surfaces during sliding were pure SBF and in SBF with an optimal concentration of NDs. For these
selected based on the previously reported contact pressure values analyses, we focused on the 0.5 N applied load.
for hip replacement contacts [34]. Analysis of the tracks formed during the sliding in pure SBF
Small amounts of carboxylated detonation nanodiamonds with (Fig. 4) and SBF with 0.1 wt% NDs at 0.5 N load (Fig. 5) was per-
5 nm average volumetric size (see supplementary information, formed by acquiring energy dispersive x-ray spectroscopy (EDS)
Fig. S1) and zeta potential of 35 mV dispersed in DI water at spectra from multiple points inside and outside of the wear tracks.
10 mg/mL concentration (NDH2O5N, Ad amas Nanotechnologies) Our tests indicate that for the pure SBF, titanium oxide is dominant
were introduced directly to the SBF and the resulting solution was within the wear track. In the case of pure SBF, the titanium surfaces
sonicated for at least 30 min. Previous studies indicate good are exposed to sliding contact and their wear increases oxidation of
colloidal stability of nanodiamonds with negative surface potential the wear track region. When NDs are introduced to the SBF, the
[35] as was confirmed in the present study for a period of one wear track oxidation is substantially lower. Interestingly, the EDS
month. The added amounts of 1, 5, 10, 15, and 20 vol% of the spectra inside the wear track clearly indicate the presence of car-
nanodiamond solution (NDS) correspond to 0.01, 0.05, 0.1, 0.15, and bon. Therefore, formation of the carbon rich layer in the wear track
0.2 wt% concentrations of nanodiamonds in SBF, correspondingly. provides better tribological performance of the sliding system and
To confirm reproducibility of the results, at least three replicate protects the underlying titanium substrate from extensive wear.
tribotests were performed for each concentration of nanodiamonds Similar wear and corrosion protection characteristics of carbon
in SBF. films were observed in case of graphitic layers [36,37]. As a result
After tests, the SBF was removed and the samples were rinsed less titanium surface is exposed to the oxidation in SBF environ-
using DI water and the wear tracks were further analyzed. Optical ment rich of corrosion promoting ions.
images of the wear tracks were acquired using a Zeiss Optical Mi- Further evaluation of the uniformity of the carbon coverage
croscope. Raman analysis was performed Nicolet Almega XR inside the wear track formed during sliding in the SBFþ0.1 wt% NDs
Dispersive Raman spectrometer with 532 nm green laser. The is shown in Fig. 6. A detailed EDS map of the SBFþ0.1 wt% NDs wear
samples were further characterized using an FEI Quanta 200 track (Fig. 6d) indicates uniformity of the carbon presence inside
Scanning Electron Microscope (SEM) with energy dispersion x-ray the wear track. Higher concentrations of oxygen (Fig. 6e) correlates
892 A. Shirani et al. / Carbon 143 (2019) 890e896

Fig. 1. (a) Coefficient of friction values and (b) corresponding wear tracks formed during sliding at 0.25 N load in different concentrations of ND solution in SBF. (A colour version of
this figure can be viewed online.)

Fig. 2. The coefficient of friction values during sliding in different concentration of ND in SBF under different applied load. (A colour version of this figure can be viewed online.)

with regions of lower carbon presence. Thus, the presence of car- track. In contrast, analysis of the wear track formed during sliding
bon plays a critical role in reducing oxidation and minimizing wear in pure SBF (Fig. 7d) demonstrates the presence of representative
of the sliding surfaces. titanium oxide peaks at ~416 and ~650 cm1 [38]. Previous studies
The nature of the carbon-based tribolayer formed inside the demonstrated a tendency for graphitic carbon formation in artifi-
wear track was evaluated using Raman spectroscopy (Fig. 7). Our cial joints during sliding [32]. Here, in contrast, the formed tribo-
results indicate that carbon shows a typical nanodiamond signa- layer demonstrates a high-intensity carbon D peak (at ~1370 cm1),
ture, suppressing formation of the titanium oxide inside the wear suggesting that nanodiamonds are not completely graphitized
A. Shirani et al. / Carbon 143 (2019) 890e896 893

Fig. 3. (a) Wear of the titanium surfaces during the tests at different loads. At 0.25 N load (Hertzian contact pressure ~ 0.28 GPa), the minimum wear is observed for 0.05 wt% of NDs;
at 0.5 N load (Hertzian contact pressure ~ 0.35 GPa), the minimum wear is for 0.1 wt% NDs; and at 1 N load (Hertzian contact pressure ~ 0.45 GPa), the minimum wear is for 0.15 wt%
NDs. (b) Optical images of the ball wear scar for the pure SBF and SBF with optimal concentrations of NDs. (A colour version of this figure can be viewed online.)

Fig. 4. SEM-EDS analysis of the wear track formed during the tests in SBF at 0.5 N load (Hertzian contact pressure ~ 0.35 GPa). Analysis indicates a high concentration of oxygen
inside the wear track. The Na, F, and Cl peaks originate from the residue of the SBF on the surface and are intentionally excluded from the further analysis of the oxidation amount.
(A colour version of this figure can be viewed online.)
894 A. Shirani et al. / Carbon 143 (2019) 890e896

Fig. 5. SEM-EDS analysis of the wear track formed during the tests in SBFþ0.1 wt% NDs at 0.5 N load (Hertzian contact pressure ~ 0.35 GPa). Inside the wear track, the presence of
carbon is observed. The Na, F, and Cl peaks originate from the residue of the SBF on the surface and are intentionally excluded from the further analysis of the oxidation amount. (A
colour version of this figure can be viewed online.)

Fig. 6. SEM-EDS map of the wear track formed during the test in SBFþ0.1 wt% NDs at 0.5 N load (Hertzian contact pressure ~ 0.35 GPa). (a) The area of the wear track with (b) higher
magnification image is highlighted. (c) Overlay map and detailed (d) carbon, (e) oxygen and (f) titanium map demonstrate uniformity of the titanium and oxygen concentration
inside and outside wear track while carbon is found inside the wear track only. (A colour version of this figure can be viewed online.)

during the sliding, enabling support of higher normal and shear as was observed earlier for carbon-based tribological films [39e42].
stresses. An additional peak at ~1560 cm1 is indicative of the sp2 Raman results confirm the formation of a carbon-rich tribolayer
bonded carbon peak, typical for graphitic carbon [32]. Presence of capable of minimizing friction and wear of sliding surfaces. Com-
the carbon layer results in substantial friction and wear reduction parison of the Raman spectra of dry NDs before the test (see
A. Shirani et al. / Carbon 143 (2019) 890e896 895

Fig. 7. Raman 2D map of the sp3 bonded carbon peak (at ~1330 cm1) of the wear track formed during the test with SBFþ0.1 wt% NDs at 0.5 N load (Hertzian contact
pressure ~ 0.35 GPa). (a) Overview of the wear track selected for the Raman 2D map analysis. (b) Raman spectra inside and outside the wear track for the tests performed in (c)
SBFþ0.1 wt% NDs and (d) pure SBF. (A colour version of this figure can be viewed online.)

supplementary information, Fig. S2) and of the formed tribofilm wear reduction in titanium tribopairs in the presence of nano-
suggests that NDs experience high degree of amorphization during diamond additives. Specifically, we show that adding nano-
sliding and embedding into the titanium substrate. diamonds into simulated body fluid, closely reproducing usual
Based on these observations, we conclude that the mechanism operating environment for hip implants, results in a more than
of the wear and friction reduction in presence of the NDs is as two-times decrease in friction and leads to 3e4 orders of magni-
follows: nanodiamonds embed inside a soft titanium surface layer tude reduction in wear of titanium surfaces. Notably, the optimal
and minimize wear of the titanium surface. With an increase in concentration of nanodiamonds is largely dependent on the
contact load, a larger amount of nanodiamonds is needed to form a applied load. While at the applied loads of 0.25 N, only ~0.05 wt% of
supportive tribolayer. At lower loads, an excess of nanodiamonds nanodiamonds are needed to observe improvements. Increasing
instigates continuous wear of the protective tribolayer and tita- the contact load to 1 N requires the concentration of nanodiamonds
nium surfaces via polishing. Therefore, the optimal concentration to be increased up to ~0.15 wt% to achieve a similar benefit. We
of nanodiamonds adjusts correspondingly to the applied load. attribute such an observation to the uniformity of the diamond-
However, the addition of NDs into SBF demonstrates reduced wear based tribofilm formation necessary for supporting contact pres-
and friction as compared to pristine SBF. sure and shear loads.
The trend observed for optimal concentration can be related to The presented results suggest the potential for nanodiamonds to
the thickness of the tribofilm needed to supply the necessary serve as effective biocompatible nanoadditives in joint re-
support under load. Since the higher applied load results in higher placements. By injecting small concentrations in the regions of high
contact pressure deformation and leads to higher friction force load and shear stresses, the lifetime of the artificial joints can be
experienced by the sliding surfaces, formation of a thicker protec- substantially increased.
tive film is critically important. The availability of NDs enables
formation of a thicker and more uniform tribofilm. Additionally, the Acknowledgments
optimal concentration of NDs replenishes the tribofilm damaged
during wear. However, with further concentration increase, excess This work was performed in part at the University of North
of NDs may result in large material build up, eventually delami- Texas' Materials Research Facility. Support from Advanced Mate-
nating from the titanium substrate and instigating abrasion during rials and Manufacturing Processes Institute (AMMPI) at the Uni-
sliding [43]. Though the optimal concentration of the nano- versity of North Texas is acknowledged.
diamonds changes, our results clearly indicate improvement in
friction and wear characteristics of titanium surfaces in the pres- Appendix A. Supplementary data
ence of very small amounts of nanodiamonds in contrast to SBF
only. Subsequent experiments will focus on evaluation of nano- Supplementary data to this article can be found online at
diamond performance in actual synovial joint fluid providing https://doi.org/10.1016/j.carbon.2018.12.005.
lubrication in the hip joints.
References
4. Conclusions
[1] W.W. Buchanan, W. Kean, Osteoarthritis I: epidemiological risk factors and
historical considerations, Inflammopharmacology 10 (2002) 5e21.
We demonstrated a new mechanism responsible for friction and [2] V.C. Mow, R. Huiskes, Basic Orthopaedic Biomechanics & Mechano-biology,
896 A. Shirani et al. / Carbon 143 (2019) 890e896

Lippincott Williams & Wilkins, 2005. [24] M. Ivanov, O. Shenderova, Nanodiamond-based nanolubricants for motor oils,
[3] P.S. Walker, B.L. Gold, The tribology (friction, lubrication and wear) of all- Curr. Opin. Solid State Mater. Sci. 21 (2017) 17e24.
metal artificial hip joints, Wear 17 (1971) 285e299. [25] W. Zhai, N. Srikanth, L.B. Kong, K. Zhou, Carbon nanomaterials in tribology,
[4] I.M. Hutchings, Friction, Lubrication and Wear of Artificial Joints, John Wiley & Carbon 119 (2017) 150e171.
Sons, 2003. [26] M.G. Ivanov, S.V. Pavlyshko, D.M. Ivanov, I. Petrov, O. Shenderova, Synergistic
[5] J. Black, Biological Performance of Materials: Fundamentals of Biocompati- compositions of colloidal nanodiamond as lubricant-additive, J. Vac. Sci.
bility, fourth ed., Taylor & Francis, 2005. Technol. B 28 (2010) 869e877.
[6] H. McKellop, I. Clarke, K. Markolf, H. Amstutz, Friction and wear properties of [27] Z. Liu, D. Leininger, A. Koolivand, A.I. Smirnov, O. Shenderova, D.W. Brenner,
polymer, metal, and ceramic prosthetic joint materials evaluated on a J. Krim, Tribological properties of nanodiamonds in aqueous suspensions:
multichannel screening device, J. Biomed. Mater. Res. 15 (1981) 619e653. effect of the surface charge, RSC Adv. 5 (2015) 78933e78940.
[7] F. Di Puccio, L. Mattei, Biotribology of artificial hip joints, World J. Orthoped. 6 [28] E. Osawa, Nanodiamonddan Emerging Nano-carbon Material, Handbook of
(2015) 77. Advanced Ceramics: Materials, Applications, Processing, and Properties,
[8] S.M. Kurtz, UHMWPE Biomaterials Handbook: Ultra High Molecular Weight Elsevier Inc, 2013, pp. 89e102.
Polyethylene in Total Joint Replacement and Medical Devices, Elsevier Science, [29] A.M. Schrand, L. Dai, J.J. Schlager, S.M. Hussain, E. Osawa, Differential
2009. biocompatibility of carbon nanotubes and nanodiamonds, Diam. Relat. Mater.
[9] V. Premnath, W.H. Harris, M. Jasty, E.W. Merrill, Gamma sterilization of 16 (2007) 2118e2123.
UHMWPE articular implants: an analysis of the oxidation problem, Bio- [30] Y. Zhu, J. Li, W. Li, Y. Zhang, X. Yang, N. Chen, Y. Sun, Y. Zhao, C. Fan, Q. Huang,
materials 17 (1996) 1741e1753. The biocompatibility of nanodiamonds and their application in drug delivery
[10] J.A. Davidson, F.S. Georgette, E. Society of Manufacturing, State of the Art systems, Theranostics 2 (2012) 302e312.
Materials for Orthopedic Prosthetic Devices, Society of Manufacturing Engi- [31] X. Zhang, S. Wang, M. Liu, J. Hui, B. Yang, L. Tao, Y. Wei, Surfactant-dispersed
neers, Dearborn, Mich., 1987. nanodiamond: biocompatibility evaluation and drug delivery applications,
[11] X. Liu, P.K. Chu, C. Ding, Surface modification of titanium, titanium alloys, and Toxicology Research 2 (2013) 335e342.
related materials for biomedical applications, Mater. Sci. Eng. R Rep. 47 (2004) [32] Y. Liao, R. Pourzal, M.A. Wimmer, J.J. Jacobs, A. Fischer, L.D. Marks, Graphitic
49e121. tribological layers in metal-on-metal hip replacements, Science 334 (2011)
[12] M.F. Semlitsch, H. Weber, R.M. Streicher, R. Scho €n, Joint replacement com- 1687e1690.
ponents made of hot-forged and surface-treated Ti-6Al-7Nb alloy, Bio- [33] T. Kokubo, H. Kushitani, S. Sakka, T. Kitsugi, T. Yamamuro, Solutions able to
materials 13 (1992) 781e788. reproduce in vivo surface-structure changes in bioactive glass-ceramic A-W3,
[13] Y. Okazaki, Y. Ito, A. Ito, T. Tateishi, Effect of alloying Elements on mechanical J. Biomed. Mater. Res. 24 (1990) 721e734.
properties of titanium alloys for medical implants, Mater. Trans., JIM 34 [34] F. Liu, S. Williams, J. Fisher, Effect of microseparation on contact mechanics in
(1993) 1217e1222. metal-on-metal hip replacementsda finite element analysis, J. Biomed. Mater.
[14] D. Kuroda, M. Niinomi, M. Morinaga, Y. Kato, T. Yashiro, Design and me- Res. B Appl. Biomater. 103 (2015) 1312e1319.
chanical properties of new b type titanium alloys for implant materials, Ma- [35] N. Gibson, O. Shenderova, T. Luo, S. Moseenkov, V. Bondar, A. Puzyr, K. Purtov,
terials Science and Engineering: A 243 (1998) 244e249. Z. Fitzgerald, D. Brenner, Colloidal stability of modified nanodiamond parti-
[15] M. Long, H.J. Rack, Titanium alloys in total joint replacementda materials cles, Diam. Relat. Mater. 18 (2009) 620e626.
science perspective, Biomaterials 19 (1998) 1621e1639. [36] D. Berman, A. Erdemir, A.V. Sumant, Few layer graphene to reduce wear and
[16] J.J. Jacobs, A. Shanbhag, T.T. Glant, J. Black, J.O. Galante, Wear debris in total friction on sliding steel surfaces, Carbon 54 (2013) 454e459.
joint replacements, JAAOS - Journal of the American Academy of Orthopaedic [37] D. Berman, A. Erdemir, A.V. Sumant, Reduced wear and friction enabled by
Surgeons 2 (1994) 212e220. graphene layers on sliding steel surfaces in dry nitrogen, Carbon 59 (2013)
[17] J.H. Dumbleton, The clinical significance of wear in total hip and knee pros- 167e175.
theses, J. Biomater. Appl. 3 (1988) 3e32. [38] N.T. Nolan, M.K. Seery, S.C. Pillai, Spectroscopic investigation of the anatase-
[18] L. Mattei, F. Di Puccio, B. Piccigallo, E. Ciulli, Lubrication and wear modelling of to-rutile transformation of sol gel-synthesized TiO2 photocatalysts, J. Phys.
artificial hip joints: a review, Tribol. Int. 44 (2011) 532e549. Chem. C 113 (2009) 16151e16157.
[19] T. Sun, Y. Sun, H. Zhang, Phospholipid-coated mesoporous silica nanoparticles [39] D. Berman, S.A. Deshmukh, S.K.R.S. Sankaranarayanan, A. Erdemir,
acting as lubricating drug nanocarriers, Polymers 10 (2018) 513. A.V. Sumant, Extraordinary macroscale wear resistance of one atom thick
[20] G. Liu, M. Cai, F. Zhou, W. Liu, Charged polymer brushes-grafted hollow silica graphene layer, Adv. Funct. Mater. 24 (2014) 6640e6646.
nanoparticles as a novel promising material for simultaneous joint lubrication [40] D. Berman, A. Erdemir, A.V. Sumant, Graphene: a new emerging lubricant,
and treatment, J. Phys. Chem. B 118 (2014) 4920e4931. Mater. Today 17 (2014) 31e42.
[21] G. Liu, Z. Liu, N. Li, X. Wang, F. Zhou, W. Liu, Hairy polyelectrolyte brushes- [41] D. Berman, B. Narayanan, M.J. Cherukara, S.K.R.S. Sankaranarayanan,
grafted thermosensitive microgels as artificial synovial fluid for simulta- A. Erdemir, A. Zinovev, A.V. Sumant, Operando tribochemical formation of
neous biomimetic lubrication and arthritis treatment, ACS Appl. Mater. In- onion-like-carbon leads to macroscale superlubricity, Nat. Commun. 9 (2018)
terfaces 6 (2014) 20452e20463. 1164.
[22] V.N. Mochalin, O. Shenderova, D. Ho, Y. Gogotsi, The properties and applica- [42] B. Wang, W. Tang, H. Lu, Z. Huang, Ionic liquid capped carbon dots as a high-
tions of nanodiamonds, Nat. Nanotechnol. 7 (2012) 11e23. performance friction-reducing and antiwear additive for poly (ethylene gly-
[23] D. Berman, S.A. Deshmukh, S.K.R.S. Sankaranarayanan, A. Erdemir, col), J. Mater. Chem. 4 (2016) 7257e7265.
A.V. Sumant, Macroscale superlubricity enabled by graphene nanoscroll for- [43] L. Bai, S. Narasimalu, G. Kang, K. Zhou, Influence of Third Particle on the
mation, Science 348 (2015) 1118e1122. Tribological Behaviors of Diamond-like Carbon Films, 2016.

You might also like