The Role of Hyaluronic Acid in Cartilage Boundary Lubrication

Weifeng Lin, Zhang Liu, Nir Kampf, Jacob Klein, Weifeng Lin, Zhang Liu, Nir Kampf, Jacob Klein

Abstract

Hydration lubrication has emerged as a new paradigm for lubrication in aqueous and biological media, accounting especially for the extremely low friction (friction coefficients down to 0.001) of articular cartilage lubrication in joints. Among the ensemble of molecules acting in the joint, phosphatidylcholine (PC) lipids have been proposed as the key molecules forming, in a complex with other molecules including hyaluronic acid (HA), a robust layer on the outer surface of the cartilage. HA, ubiquitous in synovial joints, is not in itself a good boundary lubricant, but binds the PC lipids at the cartilage surface; these, in turn, massively reduce the friction via hydration lubrication at their exposed, highly hydrated phosphocholine headgroups. An important unresolved issue in this scenario is why the free HA molecules in the synovial fluid do not suppress the lubricity by adsorbing simultaneously to the opposing lipid layers, i.e., forming an adhesive, dissipative bridge between them, as they slide past each other during joint articulation. To address this question, we directly examined the friction between two hydrogenated soy PC (HSPC) lipid layers (in the form of liposomes) immersed in HA solution or two palmitoyl-oleoyl PC (POPC) lipid layers across HA-POPC solution using a surface force balance (SFB). The results show, clearly and surprisingly, that HA addition does not affect the outstanding lubrication provided by the PC lipid layers. A possible mechanism indicated by our data that may account for this is that multiple lipid layers form on each cartilage surface, so that the slip plane may move from the midplane between the opposing surfaces, which is bridged by the HA, to an HA-free interface within a multilayer, where hydration lubrication is freely active. Another possibility suggested by our model experiments is that lipids in synovial fluid may complex with HA, thereby inhibiting the HA molecules from adhering to the lipids on the cartilage surfaces.

Keywords: cartilage lubrication; hyaluronic acid; hydration lubrication; phosphatidylcholine lipids; polymer bridging.

Conflict of interest statement

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Figures

Figure 1
Figure 1
Illustrating the proposed boundary layer on articular cartilage [33], where lubricin molecules anchor hyaluronic acid (HA) chains at the articular cartilage surface and the HA is complexed with phosphatidylcholine (PC) layers exposing their highly hydrated phosphocholine groups at the very outer surface (the slip plane). Reproduced with permission from [33]. Copyright 2016, Annual Review.
Figure 2
Figure 2
Size distribution of HSPC and POPC SUVs in water determined by dynamic light scattering (DLS). The hydrodynamic size and polydispersity index (PDI) for HSPC and POPC SUVs in water are 73 ± 6 nm (PDI = 0.071 ± 0.015) and 68 ± 5 nm (PDI = 0.056 ± 0.021), respectively.
Figure 3
Figure 3
Atomic force microscope (AFM) images of HSPC SUVs adsorbed on mica under water (a) and HA solution (b). The height distance cross section of the surface along the red line is presented below each image.
Figure 4
Figure 4
Schematic illustration of the surface force balance (SFB) setup, where normal and shear forces between two surfaces are directly measured via the bending of two orthogonal springs. White light undergoes multiple beam interference in passing through the half-silvered mica sheets, resulting in Fringes of Equal Chromatic Order (FECO, top right inset). These yield absolute separation between the surfaces to within ±0.2 nm. A sectored piezoelectric tube can move the surfaces normally and laterally relative to each other. The bending of normal and shear springs in response to applied motion, monitored via the change in an air gap capacitor or through changes in the interface fringes, directly reveals the shear and normal forces, respectively.
Figure 5
Figure 5
(a) Normal force profiles Fn(D)/R versus separation D between two mica surfaces across pure water (black symbols) following overnight incubation in HSPC small unilamellar vesicle (SUV) dispersion (0.3 mM) followed by washing, and thereafter across HA solution (red symbols), normalized in the Derjaguin approximation by the mean radius of curvature of the mica sheets (R). (b) Normalized force profiles Fn(D)/R versus separation D between two mica surfaces across POPC dispersion (0.3 mM, black symbols) following overnight incubation, and then across the HA–POPC mixture (0.5 mg/mL:0.3 mM, red symbols). The zero of separation (D = 0) is with respect to air contact between bare mica surfaces. The insets show the ‘hard wall’ behavior on a larger, linear–linear scale with the same data points as in the main figure; in (a) the hard wall in HA-free water (black symbols) is generally at ca. 30 nm (and occasionally at ca. 20 nm), and in HA solution it is consistently at ca. 20 nm; in (b) the hard wall is at ca. 10 nm in both POPC dispersion and in the HA–POPC mixture. Different shaped symbols correspond to different contact positions based on 5 independent experiments with first approaches (filled symbols), second approaches (empty symbols), and receding profiles (crossed symbols).
Figure 6
Figure 6
(a) Frictional force Fs versus Fn between two mica surfaces across pure water following overnight incubation with HSPC SUVs dispersion (black symbols), where μ ≈ 0.00028 ± 0.00014, and across HA solution (red symbols), where μ ≈ 0.00041 ± 0.00016. (b) Frictional force Fs versus Fn between two mica surfaces across POPC SUVs dispersion following overnight incubation with POPC SUVs dispersion (black symbols) and across POPC-HA mixture (red symbols), where both of them showed μ < 0.0001. The inset shows directly measured shear traces Fs(t) for POPC layer across HA–POPC (2nd trace) and HSPC layer across HA (3rd trace) with similar normal force as a function of applied sliding motion (top trace). The low values of Fs are determined by fast Fourier transform of the shear traces [12].
Figure 6
Figure 6
(a) Frictional force Fs versus Fn between two mica surfaces across pure water following overnight incubation with HSPC SUVs dispersion (black symbols), where μ ≈ 0.00028 ± 0.00014, and across HA solution (red symbols), where μ ≈ 0.00041 ± 0.00016. (b) Frictional force Fs versus Fn between two mica surfaces across POPC SUVs dispersion following overnight incubation with POPC SUVs dispersion (black symbols) and across POPC-HA mixture (red symbols), where both of them showed μ < 0.0001. The inset shows directly measured shear traces Fs(t) for POPC layer across HA–POPC (2nd trace) and HSPC layer across HA (3rd trace) with similar normal force as a function of applied sliding motion (top trace). The low values of Fs are determined by fast Fourier transform of the shear traces [12].
Figure 7
Figure 7
Illustrating the two mechanisms which are attributed to enable low friction hydration lubrication between PC-coated layers, even in the presence of HA. (a) In the case of the gel-phase lipid layers (HSPC) in HA-free water, the slip plane is the mid-plane between the vesicles. (b) When HA is added, it may adsorb on the vesicles’ surfaces to bridge the gap between them, leading to high friction at the mid-plane interface. As a result, the slip plane shifts, as shown, to the inner interfaces, where hydration lubrication is unhindered by the HA bridging. (c) In the case of the liquid-phase POPC, where the surfaces are immersed in a POPC dispersion (see text), the addition of HA leads to interactions between the lipids in solution and the polysaccharide, as indicated; this suppresses their adsorption on the surface-attached POPC bilayers and thus enables hydration lubrication between them, unhindered by any HA bridging.

References

    1. Forster H., Fisher J. The influence of loading time and lubricant on the friction of articular cartilage. Proc. Inst. Mech. Eng. H. 1996;210:109–119. doi: 10.1243/PIME_PROC_1996_210_399_02.
    1. Crockett R., Grubelnik A., Roos S., Dora C., Born W., Troxler H. Biochemical composition of the superficial layer of articular cartilage. J. Biomed. Mater. Res. A. 2007;82:958–964. doi: 10.1002/jbm.a.31248.
    1. Jahn S., Seror J., Klein J. Lubrication of articular cartilage. Annu. Rev. Biomed. Eng. 2016;18:235–258. doi: 10.1146/annurev-bioeng-081514-123305.
    1. Hodge W., Fijan R., Carlson K., Burgess R., Harris W., Mann R. Contact pressures in the human hip joint measured in vivo. Proc. Natl. Acad. Sci. USA. 1986;83:2879–2883. doi: 10.1073/pnas.83.9.2879.
    1. Afoke N., Byers P., Hutton W. Contact pressures in the human hip joint. J. Bone Joint Surg. Br. 1987;69:536–541. doi: 10.1302/0301-620X.69B4.3611154.
    1. Lewis P.R., McCutchen C.W. Mechanism of animal joints: Experimental evidence for weeping lubrication in mammalian joints. Nature. 1959;184:1285. doi: 10.1038/1841285a0.
    1. Ateshian G.A. The role of interstitial fluid pressurization in articular cartilage lubrication. J. Biomech. 2009;42:1163–1176. doi: 10.1016/j.jbiomech.2009.04.040.
    1. Murakami T., Higaki H., Sawae Y., Ohtsuki N., Moriyama S., Nakanishi Y. Adaptive multimode lubrication in natural synovial joints and artificial joints. Proc. Inst. Mech. Eng. H. 1998;212:23–35. doi: 10.1243/0954411981533791.
    1. Schmidt T.A., Sah R.L. Effect of synovial fluid on boundary lubrication of articular cartilage. Osteoarthr. Cartil. 2007;15:35–47. doi: 10.1016/j.joca.2006.06.005.
    1. Dowson D. Bio-tribology. Faraday Discuss. 2012;156:9–30. doi: 10.1039/c2fd20103h.
    1. Jay G.D., Waller K.A. The biology of Lubricin: Near frictionless joint motion. Matrix Biol. 2014;39:17–24. doi: 10.1016/j.matbio.2014.08.008.
    1. Raviv U., Klein J. Fluidity of bound hydration layers. Science. 2002;297:1540–1543. doi: 10.1126/science.1074481.
    1. Drobek T., Spencer N.D. Nanotribology of surface-grafted PEG layers in an aqueous environment. Langmuir. 2008;24:1484–1488. doi: 10.1021/la702289n.
    1. Ma L., Gaisinskaya-Kipnis A., Kampf N., Klein J. Origins of hydration lubrication. Nat. Commun. 2015;6:1–6. doi: 10.1038/ncomms7060.
    1. Rosenhek-Goldian I., Kampf N., Klein J. Trapped aqueous films lubricate highly hydrophobic surfaces. ACS Nano. 2018;12:10075–10083. doi: 10.1021/acsnano.8b04735.
    1. Lin W., Klein J. Control of surface forces through hydrated boundary layers. Curr. Opin. Colloid Interface Sci. 2019;44:94–106. doi: 10.1016/j.cocis.2019.10.001.
    1. Sarma A., Powell G., LaBerge M. Phospholipid composition of articular cartilage boundary lubricant. J. Orthop. Res. 2001;19:671–676. doi: 10.1016/S0736-0266(00)00064-4.
    1. Kosinska M.K., Liebisch G., Lochnit G., Wilhelm J., Klein H., Kaesser U., Lasczkowski G., Rickert M., Schmitz G., Steinmeyer J. A lipidomic study of phospholipid classes and species in human synovial fluid. Arthritis. Rheum. 2013;65:2323–2333. doi: 10.1002/art.38053.
    1. Trunfio-Sfarghiu A.-M., Berthier Y., Meurisse M.-H., Rieu J.-P. Role of nanomechanical properties in the tribological performance of phospholipid biomimetic surfaces. Langmuir. 2008;24:8765–8771. doi: 10.1021/la8005234.
    1. Goldberg R., Schroeder A., Silbert G., Turjeman K., Barenholz Y., Klein J. Boundary lubricants with exceptionally low friction coefficients based on 2D close-packed phosphatidylcholine liposomes. Adv. Mater. 2011;23:3517–3521. doi: 10.1002/adma.201101053.
    1. Sorkin R., Dror Y., Kampf N., Klein J. Mechanical stability and lubrication by phosphatidylcholine boundary layers in the vesicular and in the extended lamellar phases. Langmuir. 2014;30:5005–5014. doi: 10.1021/la500420u.
    1. Sorkin R., Kampf N., Dror Y., Shimoni E., Klein J. Origins of extreme boundary lubrication by phosphatidylcholine liposomes. Biomaterials. 2013;34:5465–5475. doi: 10.1016/j.biomaterials.2013.03.098.
    1. Sorkin R., Kampf N., Zhu L., Klein J. Hydration lubrication and shear-induced self-healing of lipid bilayer boundary lubricants in phosphatidylcholine dispersions. Soft Matter. 2016;12:2773–2784. doi: 10.1039/C5SM02475G.
    1. Walker P., Sikorski J., Dowson D., Longfield M., Wright V., Buckley T. Behaviour of synovial fluid on surfaces of articular cartilage. A scanning electron microscope study. Ann. Rheum. Dis. 1969;28:1. doi: 10.1136/ard.28.1.1.
    1. Seror J., Merkher Y., Kampf N., Collinson L., Day A.J., Maroudas A., Klein J. Articular cartilage proteoglycans as boundary lubricants: Structure and frictional interaction of surface-attached hyaluronan and hyaluronan–aggrecan complexes. Biomacromolecules. 2011;12:3432–3443. doi: 10.1021/bm2004912.
    1. Radin E.L., Swann D.A., Weisser P.A. Separation of a hyaluronate-free lubricating fraction from synovial fluid. Nature. 1970;228:377–378. doi: 10.1038/228377a0.
    1. Swann D.A., Radin E.L., Nazimiec M., Weisser P.A., Curran N., Lewinnek G. Role of hyaluronic acid in joint lubrication. Ann. Rheum. Dis. 1974;33:318. doi: 10.1136/ard.33.4.318.
    1. Ghosh P. The role of hyaluronic acid (hyaluronan) in health and disease: Interactions with cells, cartilage and components of synovial fluid. Clin. Exp. Rheumatol. 1994;12:75–82.
    1. Nitzan D., Nitzan U., Dan P., Yedgar S. The role of hyaluronic acid in protecting surface-active phospholipids from lysis by exogenous phospholipase A2. Rheumatology. 2001;40:336–340. doi: 10.1093/rheumatology/40.3.336.
    1. Brondello J.M., Djouad F., Jorgensen C. Where to Stand with Stromal Cells and Chronic Synovitis in Rheumatoid Arthritis? Cells. 2019;8:1257. doi: 10.3390/cells8101257.
    1. Benz M., Chen N., Israelachvili J. Lubrication and wear properties of grafted polyelectrolytes, hyaluronan and hylan, measured in the surface forces apparatus. J. Biomed. Mater. Res. A. 2004;71:6–15. doi: 10.1002/jbm.a.30123.
    1. Das S., Banquy X., Zappone B., Greene G.W., Jay G.D., Israelachvili J.N. Synergistic interactions between grafted hyaluronic acid and lubricin provide enhanced wear protection and lubrication. Biomacromolecules. 2013;14:1669–1677. doi: 10.1021/bm400327a.
    1. Seror J., Zhu L., Goldberg R., Day A.J., Klein J. Supramolecular synergy in the boundary lubrication of synovial joints. Nat. Commun. 2015;6:6497. doi: 10.1038/ncomms7497.
    1. Zhu L., Seror J., Day A.J., Kampf N., Klein J. Ultra-low friction between boundary layers of hyaluronan-phosphatidylcholine complexes. Acta Biomater. 2017;59:283–292. doi: 10.1016/j.actbio.2017.06.043.
    1. Pasquali-Ronchetti I., Quaglino D., Mori G., Bacchelli B., Ghosh P. Hyaluronan–phospholipid interactions. J. Struct. Biol. 1997;120:1–10. doi: 10.1006/jsbi.1997.3908.
    1. Crescenzi V., Taglienti A., Pasquali-Ronchetti I. Supramolecular structures prevailing in aqueous hyaluronic acid and phospholipid vesicles mixtures: An electron microscopy and rheometric study. Colloids Surf. A. 2004;245:133–135. doi: 10.1016/j.colsurfa.2004.06.030.
    1. Liu C., Wang M., An J., Thormann E., Dėdinaitė A. Hyaluronan and phospholipids in boundary lubrication. Soft Matter. 2012;8:10241. doi: 10.1039/c2sm26615f.
    1. Wang M., Liu C., Thormann E., Dedinaite A. Hyaluronan and phospholipid association in biolubrication. Biomacromolecules. 2013;14:4198–4206. doi: 10.1021/bm400947v.
    1. Jung S., Petelska A., Beldowski P., Augé W.K., Casey T., Walczak D., Lemke K., Gadomski A. Hyaluronic acid and phospholipid interactions useful for repaired articular cartilage surfaces—A mini review toward tribological surgical adjuvants. Colloid Polym. Sci. 2017;295:403–412. doi: 10.1007/s00396-017-4014-z.
    1. Siódmiak J., Bełdowski P., Augé W.K., Ledziński D., Śmigiel S., Gadomski A. Molecular dynamic analysis of hyaluronic acid and phospholipid interaction in tribological surgical adjuvant design for osteoarthritis. Molecules. 2017;22:1436. doi: 10.3390/molecules22091436.
    1. Dedinaite A., Wieland D.C.F., Beldowski P., Claesson P.M. Biolubrication synergy: Hyaluronan-phospholipid interactions at interfaces. Adv. Colloid Interface Sci. 2019;274:102050. doi: 10.1016/j.cis.2019.102050.
    1. Lin W., Mashiah R., Seror J., Kadar A., Dolkart O., Pritsch T., Goldberg R., Klein J. Lipid-hyaluronan synergy strongly reduces intrasynovial tissue boundary friction. Acta Biomater. 2019;83:314–321. doi: 10.1016/j.actbio.2018.11.015.
    1. Wang Z., Li J., Ge X., Liu Y., Luo J., Chetwynd D.G., Mao K. Investigation of the lubrication properties and synergistic interaction of biocompatible liposome-polymer complexes applicable to artificial joints. Colloids Surf. B. 2019;178:469–478. doi: 10.1016/j.colsurfb.2019.03.041.
    1. Fakhari A., Phan Q., Thakkar S.V., Middaugh C.R., Berkland C. Hyaluronic acid nanoparticles titrate the viscoelastic properties of viscosupplements. Langmuir. 2013;29:5123–5131. doi: 10.1021/la304575x.
    1. Banquy X., Lee D.W., Das S., Hogan J., Israelachvili J.N. Shear-Induced Aggregation of Mammalian Synovial Fluid Components under Boundary Lubrication Conditions. Adv. Funct. Mater. 2014;24:3152–3161. doi: 10.1002/adfm.201302959.
    1. Corvelli M., Che B., Saeui C., Singh A., Elisseeff J. Biodynamic performance of hyaluronic acid versus synovial fluid of the knee in osteoarthritis. Methods. 2015;84:90–98. doi: 10.1016/j.ymeth.2015.03.019.
    1. Subbotin A., Semenov A., Doi M. Friction in strongly confined polymer melts: Effect of polymer bridges. Phys. Rev. E. 1997;56:623. doi: 10.1103/PhysRevE.56.623.
    1. Raviv U., Tadmor R., Klein J. Shear and frictional interactions between adsorbed polymer layers in a good solvent. J. Phys. Chem. B. 2001;105:8125–8134. doi: 10.1021/jp0041860.
    1. Cao Y., Kampf N., Klein J. Boundary lubrication, hemifusion, and self-healing of binary saturated and monounsaturated phosphatidylcholine mixtures. Langmuir. 2019;35:15459–15468. doi: 10.1021/acs.langmuir.9b01660.
    1. Klein J., Kumacheva E. Simple liquids confined to molecularly thin layers. I. Confinement-induced liquid-to-solid phase transitions. J. Chem. Phys. 1998;108:6996–7009. doi: 10.1063/1.476114.
    1. Perez-Martinez C.S., Perkin S. Interfacial structure and boundary lubrication of a dicationic ionic liquid. Langmuir. 2019;35:15444–15450. doi: 10.1021/acs.langmuir.9b01415.
    1. Redeker C., Briscoe W.H. Interactions between mutant bacterial lipopolysaccharide (LPS-Ra) surface layers: Surface vesicles, membrane fusion, and effect of Ca2+ and temperature. Langmuir. 2019;35:15739–15750. doi: 10.1021/acs.langmuir.9b02609.
    1. Perkin S., Chai L., Kampf N., Raviv U., Briscoe W., Dunlop I., Titmuss S., Seo M., Kumacheva E., Klein J. Forces between mica surfaces, prepared in different ways, across aqueous and nonaqueous liquids confined to molecularly thin films. Langmuir. 2006;22:6142–6152. doi: 10.1021/la053097h.
    1. Lin W., Kampf N., Goldberg R., Driver M.J., Klein J. Poly-phosphocholinated liposomes form stable superlubrication vectors. Langmuir. 2019;35:6048–6054. doi: 10.1021/acs.langmuir.9b00610.
    1. Duan Y., Liu Y., Li J., Feng S., Wen S. AFM study on superlubricity between Ti6Al4V/polymer surfaces achieved with liposomes. Biomacromolecules. 2019;20:1522–1529. doi: 10.1021/acs.biomac.8b01683.
    1. Smith P., Ziolek R.M., Gazzarrini E., Owen D.M., Lorenz C.D. On the interaction of hyaluronic acid with synovial fluid lipid membranes. Phys. Chem. Chem. Phys. 2019;21:9845–9857. doi: 10.1039/C9CP01532A.
    1. Schmidt T.A., Gastelum N.S., Nguyen Q.T., Schumacher B.L., Sah R.L. Boundary lubrication of articular cartilage: Role of synovial fluid constituents. Arthritis Rheumatol. 2007;56:882–891. doi: 10.1002/art.22446.
    1. Klein J. Hydration lubrication. Friction. 2013;1:1–23. doi: 10.1007/s40544-013-0001-7.
    1. Moro-Oka T., Miura H., Mawatari T., Kawano T., Nakanishi Y., Higaki H., Iwamoto Y. Mixture of hyaluronic acid and phospholipid prevents adhesion formation on the injured flexor tendon in rabbits. J. Orthop. Res. 2000;18:835–840. doi: 10.1002/jor.1100180523.
    1. Raj A., Wang M., Zander T., Wieland D.C.F., Liu X., An J., Garamus V.M., Willumeit-Römer R., Fielden M., Claesson P.M., et al. Lubrication synergy: Mixture of hyaluronan and dipalmitoylphosphatidylcholine (DPPC) vesicles. J. Colloid Interface Sci. 2017;488:225–233. doi: 10.1016/j.jcis.2016.10.091.
    1. Veselack T., Aldebert G., Trunfio-Sfarghiu A.M., Schmid T.M., Laurent M.P., Wimmer M.A. Phospholipid Vesicles in Media for Tribological Studies against Live Cartilage. Lubricants. 2018;6:19. doi: 10.3390/lubricants6010019.
    1. Beldowski P., Yuvan S., Dedinaite A., Claesson P.M., Poschel T. Interactions of a short hyaluronan chain with a phospholipid membrane. Colloids Surf. B Biointerfaces. 2019;184:110539. doi: 10.1016/j.colsurfb.2019.110539.
    1. Dėdinaitė A., Claesson P.M. How synergistic aqueous lubrication is mediated by natural and synthetic molecular aggregates. IOP Conf. Ser. Mater. Sci. Eng. 2019;500:012030. doi: 10.1088/1757-899X/500/1/012030.
    1. Herzog M., Li L., Galla H.J., Winter R. Effect of hyaluronic acid on phospholipid model membranes. Colloids Surf. B Biointerfaces. 2019;173:327–334. doi: 10.1016/j.colsurfb.2018.10.006.
    1. Riley J.K., Tilton R.D. Sequential adsorption of nanoparticulate polymer brushes as a strategy to control adhesion and friction. Langmuir. 2016;32:11440–11447. doi: 10.1021/acs.langmuir.6b02963.
    1. Hills B. Surface-active phospholipid: A Pandora’s box of clinical applications. Part II. Barrier and lubricating properties. Intern. Med. J. 2002;32:242–251. doi: 10.1046/j.1445-5994.2002.00201.x.

Source: PubMed

3
Subskrybuj