ArticlePDF Available

Macrophage proteomic analysis of covalent immobilization of hyaluronic acid and graphene oxide on CoCr alloy in a tribocorrosive environment

Authors:

Abstract

In this work, a sequential covalent immobilization of graphene oxide (GO) and hyaluronic acid (HA) is performed to obtain a biocompatible wear‐resistant nanocoating on the surface of the biomedical grade cobalt‐chrome (CoCr) alloy. Nanocoated CoCr surfaces were characterized by Raman spectroscopy and electrochemical impedance spectroscopy (EIS) in 3 g/L HA electrolyte. Tribocorrosion tests of the nanocoated CoCr surfaces were carried out in a pin on flat tribometer. The biological response of covalently HA/GO biofunctionalized CoCr surfaces with and without wear‐corrosion processes was studied through the analysis of the proteome of macrophages. Raman spectra revealed characteristic bands of GO and HA on the functionalized CoCr surfaces. The electrochemical response by EIS showed a stable and protective behavior over 23 days in the simulated biological environment. HA/GO covalently immobilized on CoCr alloy is able to protect the surface and reduce the wear volume released under tribocorrosion tests. Unsupervised classification analysis of the macrophage proteome via hierarchical clustering and principal component analysis (PCA) revealed that the covalent functionalization on CoCr enhances the macrophage biocompatibility in vitro. On the other hand, disruption of the HA/GO nanocoating by tribocorrosion processes induced a macrophage proteome which was differently clustered and was distantly located in the PCA space. In addition, tribocorrosion induced an increase in the percentage of upregulated and downregulated proteins in the macrophage proteome, revealing that disruption of the covalent nanocoating impacts the macrophage proteome. Although macrophage inflammation induced by tribocorrosion of HA/GO/CoCr surfaces is observed, it is ameliorated by the covalently grafting of HA, which provides immunomodulation by eliciting downregulations in characteristic pro‐inflammatory signaling involved in inflammation and aseptic loosening of CoCr joint arthroplasties. Covalent HA/GO nanocoating on CoCr provides potential applications for in vivo joint prostheses led a reduced metal‐induced inflammation and degradation by wear‐corrosion.
RESEARCH ARTICLE
Macrophage proteomic analysis of covalent immobilization of
hyaluronic acid and graphene oxide on CoCr alloy in a
tribocorrosive environment
L. Sánchez-López
1,2,3
| B. Chico
2
| Maria Cristina García-Alonso
2
|
Rosa M. Lozano
1
1
Centro de Investigaciones Biológicas-Margarita Salas (CIB Margarita Salas), Consejo Superior de Investigaciones Científicas (CSIC), Madrid, Spain
2
Centro Nacional de Investigaciones Metalúrgicas (CENIM), Consejo Superior de Investigaciones Científicas (CSIC), Madrid, Spain
3
PhD Program in Advanced Materials and Nanotechnology, Doctoral School, Universidad Autónoma de Madrid, Madrid, Spain
Correspondence
Rosa M. Lozano, Centro de Investigaciones
Biológicas-Margarita Salas (CIB Margarita
Salas), Consejo Superior de Investigaciones
Científicas (CSIC), C/ Ramiro de Maeztu,
28040 Madrid, Spain.
Email: rlozano@cib.csic.es
Maria Cristina García-Alonso, Centro Nacional
de Investigaciones Metalúrgicas (CENIM),
Consejo Superior de Investigaciones
Científicas (CSIC), Avenida Gregorio del Amo,
8, 28040 Madrid, Spain.
Email: crisga@cenim.csic.es
Funding information
Ministerio de Ciencia, Innovación y
Universidades (MICIU/FEDER), Grant/Award
Numbers: PRE2019-090122,
RTI2018-101506-B-C31, RTI2018-101506-B-
C33
Abstract
In this work, a sequential covalent immobilization of graphene oxide (GO) and hya-
luronic acid (HA) is performed to obtain a biocompatible wear-resistant nanocoating
on the surface of the biomedical grade cobalt-chrome (CoCr) alloy. Nanocoated CoCr
surfaces were characterized by Raman spectroscopy and electrochemical impedance
spectroscopy (EIS) in 3 g/L HA electrolyte. Tribocorrosion tests of the nanocoated
CoCr surfaces were carried out in a pin on flat tribometer. The biological response of
covalently HA/GO biofunctionalized CoCr surfaces with and without wear-corrosion
processes was studied through the analysis of the proteome of macrophages. Raman
spectra revealed characteristic bands of GO and HA on the functionalized CoCr sur-
faces. The electrochemical response by EIS showed a stable and protective behavior
over 23 days in the simulated biological environment. HA/GO covalently immobilized
on CoCr alloy is able to protect the surface and reduce the wear volume released
under tribocorrosion tests. Unsupervised classification analysis of the macrophage
proteome via hierarchical clustering and principal component analysis (PCA) revealed
that the covalent functionalization on CoCr enhances the macrophage biocompatibil-
ity in vitro. On the other hand, disruption of the HA/GO nanocoating by tribocorro-
sion processes induced a macrophage proteome which was differently clustered and
was distantly located in the PCA space. In addition, tribocorrosion induced an
increase in the percentage of upregulated and downregulated proteins in the macro-
phage proteome, revealing that disruption of the covalent nanocoating impacts the
macrophage proteome. Although macrophage inflammation induced by tribocorro-
sion of HA/GO/CoCr surfaces is observed, it is ameliorated by the covalently grafting
of HA, which provides immunomodulation by eliciting downregulations in characteris-
tic pro-inflammatory signaling involved in inflammation and aseptic loosening of CoCr
joint arthroplasties. Covalent HA/GO nanocoating on CoCr provides potential
Received: 14 February 2024 Revised: 6 May 2024 Accepted: 8 May 2024
DOI: 10.1002/jbm.a.37751
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited.
© 2024 The Authors. Journal of Biomedical Materials Research Part A published by Wiley Periodicals LLC.
J Biomed Mater Res. 2024;119. wileyonlinelibrary.com/journal/jbma 1
applications for in vivo joint prostheses led a reduced metal-induced inflammation
and degradation by wear-corrosion.
KEYWORDS
CoCr, covalent immobilization, graphene oxide, hyaluronic acid, macrophage proteome,
tribocorrosion
1|INTRODUCTION
Cobalt-chrome (CoCr) implants used as joint prostheses suffer from
wear-corrosion degradation processes at some degree upon implanta-
tion, resulting in the release of metal ions and particulate debris into
the surrounding environments. Metal tribocorrosion intensifies the
foreign body response (FBR) causing CoCr-mediated inflammation,
osteolysis, and aseptic loosening. These adverse effects may ulti-
mately lead to the replacement of CoCr implants through revision sur-
geries due to implant failure.
13
Both innate and adaptive immune
responses play a role in the overall immune response. Nevertheless,
the severe inflammatory reaction localized around the implant is
mainly characterized by an innate response marked by a prominent
infiltration of macrophages.
46
The interaction between macrophages
and the biomaterial significantly influences the success and long-term
functionality of the implant, ultimately impacting the risk of implant
rejection. Hence, the development of immunomodulatory biomaterials
can be crucial for regulating macrophage response, a key factor in
ensuring the extended viability and effectiveness of implanted bio-
medical devices.
79
Regarding potential approaches for manufacturing immunomodu-
latory biomaterial, surface chemistry modifications can be employed
to establish a beneficial cell-biomaterial interface that effectively
modulates immune responses, all while preserving the required bulk
properties of metallic implants.
10
Given that metal debris is the primary trigger for adverse inflam-
matory reactions, it is crucial to explore surface chemistry modifica-
tions with the goal of creating a surface that not only reduces the
release of metal debris but also minimizes the biological FBR to the
greatest extent possible. In an effort to improve the tribocorrosion
performance and minimize the release of metallic debris, researchers
have proposed the use of graphene-based chemical modifications.
11
This approach seeks to replicate the presence of carbon-enriched tri-
bological layers observed in worn areas of in vivo retrieved prosthe-
ses, as their existence is associated with wear reduction. The concept
is to implement a comprehensive graphene-based surface modifica-
tion that acts as a barrier against the release of metal debris while pre-
serving the necessary bulk mechanical properties of the metal implant
and ensuring a favorable biological response.
On the other hand, the integration of biomolecules such as pro-
teins, peptides, glycosaminoglycans, or other organic compounds can
further address the modulation of the biological response through bio-
mimetic signaling. This approach involves the incorporation of biomol-
ecules on the biomaterial surface, allowing for the precise interaction
of cellular membrane receptors with these biomaterial motifs to influ-
ence and fine-tune desired immune responses.
The authors studied the electrochemical reduction of graphene
oxide (GO) on CoCr surfaces by cyclic voltammetry and posterior bio-
functionalization with HA 3 g/L. Reduction of GO on the CoCr sur-
faces and the posterior physical adsorption of hyaluronic acid
(HA) provided a decreased inflammatory macrophage response,
12
however, the reduced GO increased the wear-corrosion of the CoCr
surfaces.
13
The type of immobilization mechanism applied, based on
physical or chemical adsorption methods, can greatly affect the final
implant outcome. Molecules that are physically immobilized can
exhibit spontaneous desorption, typically leading to short-term thera-
peutic effects at the implant surface. Besides, desorbed biomolecules
may disseminate to other body sites, potentially eliciting undesired
biological effects in ectopic locations.
14
Conversely, biomolecules that
are covalently immobilized by chemical adsorption methods can effec-
tively prevent desorption and diffusion from the intended site of
action, thereby minimizing the potential adverse effects of these bio-
molecules. Importantly, biomolecules bound irreversibly can be
retained at the biomaterial site for extended periods, allowing them to
elicit the desired cellular responses at the cellbiomaterial interface
over prolonged periods.
15
Furthermore, the stability of the target bio-
molecule is enhanced through multi-point covalent immobilization,
resulting in improved resilience to variations in temperatures and
resistance to organic solvents.
1620
This immobilization also serves as
a barrier against proteases and proteolytic degradation, improving the
biomolecule resistance to biodegradation when it is attached to
organic or inorganic supports.
18
A wide variety of biomolecules has been covalently attached to
metal surfaces for this purpose.
2127
This strategy holds special signif-
icance in the field of biomedical applications, especially concerning
HA, one of the main components of the synovial fluids involved in
lubrication joints.
HA is prone to rapid degradation under physiological conditions,
due to enzymatic and chemical degradation mechanisms by hyaluroni-
dases, oxidation, and hydrolysis.
2830
The challenges involved in pre-
serving the durability and long-term effectiveness of HA emphasize
the critical importance of employing biomolecule immobilization on
the biomaterial surface as a strategic solution. Chemical modification
strategies, such as chemically anchored or cross-linked HA, aim to
improve the stability and resistance to degradation of HA.
31
Minimal
physicochemical alterations are introduced through one-point and
multi-point covalent immobilizations
17
while they can also improve
the degradation resistance of HA and other extracellular matrix (ECM)
2S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
components.
3235
Nevertheless, it is important to be careful not to
alter the chemical structure and cross-linking density of HA, since
excessive changes can potentially induce cytotoxicity and trigger
immunogenic reactions.
3638
In this work, the sequential covalent immobilization of HA and
GO nanosheets on the biomedical grade CoCr alloy is characterized in
terms of electrochemical response, tribocorrosion behavior, and the
analysis of the macrophage response to modified CoCr surfaces, both
with and without wear. Potential applications of covalently HA/GO
biofunctionalized CoCr surfaces for metal joint prostheses will be
analyzed in terms of the biological response of the proteome of mac-
rophages associated with metal-induced inflammation by wear-
corrosion processes.
2|MATERIALS AND METHODS
2.1 |Materials
Biomedical grade CoCr alloy was supplied by International Edge with
the nominal chemical composition (wt %): 27.25% Cr, 5.36% Mo,
0.69% Mn, 0.68% Si, 0.044% C, 0.02% W, 0.15% N, 0.002% Al,
0.001% S, 0.002% P, 0.002% B, 0.001% Ti, and balance Co. CoCr sur-
faces were polished with SiC abrasive papers from 600 to 2000 grain
size and then mirror-polished with 3 and 1 μm diamond paste (hereaf-
ter samples named CoCr) and rinsed with ultrapure water (Milli-Q
Direct 8 water purification system, Merck, Germany).
3-aminopropyl-triethoxysilane (APTES) agent, used as an interme-
diate compound for the covalent immobilization of GO on CoCr, was
purchased from Merck, SigmaAldrich. GO nanosheets were supplied
by GRAnPH. Biofunctionalization of GO nanocoating was provided by
using high molecular weight HA (HMW-HA, 15001800 kDa, refer-
ence 53,747, SigmaAldrich Chemie GmbH, Schnelldorf, Germany).
Adipic dihydrazide 98% (ADH) agent, used as intermediate compound
for the biofunctionalization of GO with HA, was purchased from
Thermo Fisher Scientific, Acros Organics.
2.2 |Sequential covalent immobilization of HA and
GO on CoCr surface
Covalent immobilization of GO on CoCr was obtained as previously
described.
39
Briefly, CoCr surfaces were hydroxylated by alkaliniza-
tion in NaOH 5 M for 2 h followed by self-assembly of silane mono-
layers using the amino-silane intermediate coupling agent APTES,
40
which was pre-mixed at 2% vol in isopropanol-water (200: 1 v/v) and
stirred for 1 h. After this time, hydroxylated CoCr surfaces were
immersed in the APTES solution at RT for 1 min and kept 24 h at the
curing temperature of 45C. Silanized CoCr surfaces were immersed
in GO suspension at 4 g/L during 24 h at 60C to trigger covalent
immobilization reaction of GO sheets on silanized CoCr (hereafter
samples named GO/CoCr).
Carboxylic groups of covalently immobilized GO in the
GO/CoCr samples were activated by immersing the GO/CoCr
surfaces in ultrapure water (50 mL) with N-(3-dimethylaminopro-
pyl)-N-ethylcarbodiimide hydrochloride (EDC, 250 mg) and N-
hydroxysuccinimide (NHS, 750 mg) during 24 h in mild orbital
agitation. After that, the GO/CoCr surfaces with activated carboxyl
groups were immersed in ultrapure water (100 mL) with dissolved
adipic dihydrazide (ADH, 1.04 g) and let to react for another 24 h in
mild orbital agitation (samples named ADH/GO/CoCr).
Further covalent grafting of high molecular weight HA was per-
formed on GO nanosheets on GO/CoCr samples through ADH as an
intermediate covalent GO-HA linker using EDC/NHS carbodiimide
chemistry.
41,42
Pre-activated HA-(EDC/NHS) solution was obtained by dissolving
HA (100 mg), EDC (253 mg), and NHS (754 mg) in 100-mL phosphate
buffer saline (PBS) and stirred for 24 h to activate the carboxylic acids
of HA. After that, ADH/GO/CoCr surfaces were immersed in pre-
activated HA-(EDC/NHS) solution and let to react for 24 h in mild
orbital motion (hereafter samples named HA/GO/CoCr). Surfaces
were rinsed with ultrapure water between each immobilization step
to remove excess reagents.
2.3 |Characterization of covalent
immobilizations on CoCr
GO/CoCr and HA/GO/CoCr surfaces were characterized by using a
field-emission Hitachi S-4800 equipment by scanning electron micros-
copy (SEM) and Raman spectroscopy. Secondary electron images at
SEM were taken at 15 kV. Raman spectra were collected with a
Raman spectrometer B&W TEK (BTC675N) using a 532 nm wave-
length laser. Exposure time was set at 100 ms per scan. Spectral data
were processed using BWSpec4 software with dark-frame subtrac-
tion. Spectra were collected from 50 to 4000 cm
1
with integration
times of 100 and 300 s.
The barrier effect of GO and HA/GO on the CoCr surfaces was
assessed through electrochemical impedance spectroscopy (EIS). EIS
data were recorded up to 23 days of immersion in HA aqueous elec-
trolyte at physiological concentration of 3 g/L simulating the biologi-
cal synovial fluid environment.
43
A three-electrode electrochemical
cell consisting of the GO/CoCr or HA/GO/CoCr surfaces as working
electrodes, a Pt wire as counter electrode and Ag/AgCl (3 M KCl) as
reference electrode were used. An Autolab32 potentiostat/
galvanostat was used to generate a sinusoidal wave of 10 mV ampli-
tude applied at the corrosion potential Ecorr in the 100 kHz1 mHz
frequency range logarithmically spaced by 5 points/decade. Fitting of
the experimental impedance data into equivalent electric circuits was
performed by using Z-view software based on a nonlinear least-
squares program. The criteria used for selecting the optimal equiva-
lent circuits were the lowest chi-square value and lowest relative
errors (in %). A good reproducibility was obtained for EIS tests per-
formed in duplicate.
S´
ANCHEZ-LÓPEZ ET AL.3
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2.4 |Wear-corrosion tests
To evaluate the tribocorrosion performance in the pair MoM used for
joint replacement, CoCr-CoCr tribosystems were assembled in a trib-
ometer (Microtest MT/30/NI) with a pin-on-disk (ball-on-plate) con-
figuration. CoCr balls of 8 mm diameter were used as counterpart
(pin) in contact with HA/GO/CoCr disks of 38 mm diameter. The use
of CoCr balls as counterpart allows a more suitable configuration in
the tribometer promoting a uniform contact between the surface
materials involved in tribocorrosion tests.
The applied load was 5 N and a rotation speed of 120 rpm. To
simulate physiologically significant tribological conditions, a total slid-
ing distance of 30,000 m was set in the tribometer to consider the
motion experienced by a hip prosthesis, with an average of one million
cycles per year.
44
Wear-corrosion tests were carried out in recirculating HA aque-
ous electrolyte at physiological concentration 3 g/L used as lubricant
in a recirculation pump system in the tribometer configuration. HA
constitutes one of the main components of synovial fluid, playing a
crucial role in joint lubrication.
45
This rationale supports the choice of
HA as an electrolyte for tribocorrosion tests, as it represents more
closely the physiological environment where the materials will be
implanted. Tribocorrosion tests were performed in triplicate.
The worn tracks were characterized with a 3D Profilm3D equip-
ment (FilmetricsKLV Company).
2.5 |Modified CoCr surfaces-macrophage
cell cultures
Mouse macrophage J774A.1 cell line was provided by DSMZ Human
and Animal Cell Bank, Braunschweig, Germany. Three different groups
were analyzed in cell-biomaterial assays. A control group where macro-
phages were directly seeded on the surface of cell-culture treated poly-
styrene petri dishes (BD, 1 Becton Drive, NJ, USA) without additional
biomaterial (Polystyrene control). A second group, where macrophages
were seeded on intact HA/GO/CoCr disks, that were previously incu-
bated 4 days in the HA solution as used in tribocorrosion tests; and a
third group where macrophages were seeded on worn HA/GO/CoCr
disks, with the worn disk surface in contact with cells.
All types of CoCr disks were UV-sterilized prior to cell cultures.
The macrophage culture was seeding at a cell density of 12,500
cells/cm
2
for 72 h and 10,000 cells/cm
2
for 96 h cell-biomaterial
cultures in complete cell culture medium (DMEM 41966, 10% heat-
inactivated FBS, 100 units/mL penicillin, and 100 μg/mL streptomy-
cin, Gibco ThermoFisher Scientific, USA) and incubated in a cell cul-
ture chamber at 37C and 5% CO
2
.
2.6 |Macrophage cell extracts for proteomics
analysis
Macrophage cell extracts were taken for proteomic analysis after
72 and 96 h of interaction with the three groups described. Firstly,
macrophages cultures were washed with ice-cold PBS and then incu-
bated with a cell lysis buffer supplemented with protease inhibitors to
prevent protein degradation (hereafter complete cell lysis buffer).
Complete cell lysis buffer contained 5 mL RIPA lysis buffer (name of a
commercial cell lysis buffer from Santa Cruz Biotechnology, Santa
Cruz, CA, USA), 1:100 protease inhibitor cocktail provided in the RIPA
lysis buffer system reagent.
Cells extracts were scraped and transferred to a 15 mL conical
tube, where lysates in complete cell lysis buffer were incubated on
ice for 15 min. Then, cellular lysates were sonicated three times dur-
ing 2 s-pulses with 1 min of rest in ice between sonication pulses.
Lysates were then incubated on ice for additionally 15 min to
increase protein extraction yields and centrifuged at 12,100 gfor
5 min at 4C, collecting the supernatants into Eppendorf microtubes
and storing in aliquots at 20C, avoiding multiple freeze/thaw
cycles.
2.7 |LCMS/MS
Preparation of samples for liquid chromatography tandem mass
spectrometry (LCMS/MS) analysis was performed as previously.
12
Briefly, the total protein concentration in purified cell lysates
supernatants was determined using Thermo ScientificPierce
660 nm protein assay reactive (Pierce Biotechnology, USA) and
quantified by absorbance spectroscopy at 660 nm in an iMark
microplate absorbance reader (Bio-Rad Laboratories, Inc., USA).
Then, 60 μg of protein from macrophage cell extracts was precipi-
tated in cold acetone and resuspended in loading buffer for sodium
dodecyl sulfate polyacrylamide gel electrophoresis (SDS/PAGE).
Identified proteome bands from SDS/PAGE gels were purified and
digested in trypsin (Pierce, Thermo Fisher Scientific, USA). The
digested peptides were purified and separated in a nLC 1000 nano-
system (Thermo Scientific, USA) using Acclaim PepMap 100 and
rapidseparationliquidchromatographyPepMapC18chromato-
graphic columns (Thermo Scientific, USA). Eluted peptides were
incorporated into a Q Exactive mass spectrometer (Thermo Scien-
tific, USA) at the previous described setting conditions
12
and label-
free quantification (LFQ) of three technical replicates. MS scans
were analyzed with the Proteome Discoverer software (Thermo
Fisher Scientific, USA) using standardized workflows. Protein IDs
from mass spectra *.raw files were searched using MaxQuant pro-
gram and the SwissProt Mus musculus database, generating the raw
proteome matrix data.
2.8 |Proteome data analyses
The proteome raw data matrix was analyzed using Perseus-MaxQuant
software. Histograms of total MS counts scatter plots displaying pro-
tein abundance were inspected previously to data analysis for each
sample. Processing of raw data matrix was performed by pre-filtering
of contaminants and redundant values followed by Log
2
transforma-
tion of the proteome matrix and multiple sample analysis of variance
4S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
(ANOVA, FDR 0.05). Normalization of the matrix was obtained by
computation of Z-scores. Hierarchical clustering was then performed
using the Euclidian distance and average linkage method, to obtain
the hierarchical dendrogram and heatMap of the clustered matrix.
Principal component analysis (PCA) was computed applying 14 princi-
pal components and delta of 1 10
5
with visualization of only the
first five principal components (PC).
Differentially expressed proteins (DEPs) were searched in t-test
pairwise group comparisons with a p-value <.05 and a fold change of
2, where significantly upregulated (FC > 2) and downregulated
(FC < 2) DEPs were displayed on the right and left upper panels of
volcano plots, respectively. The presence/absence of protein expres-
sion between paired groups was also included at the extremes of vol-
cano Xaxes. Gene-ontology enrichment analysis of upregulated and
downregulated DEPs was performed with ShinyGO 0.77 software
(South Dakota State University, FDR 0.05).
46
2.9 |Inflammatory response of J774A.1
macrophages
The cytokine secretion levels of murine tumor necrosis factor α(TNF-
α) and interleukin-10 (IL-10) were measured to analyze the inflamma-
tory behavior of materials under study. Macrophages were cultured
on both intact modified surfaces and those after wear-corrosion tests.
The modified surfaces included CoCr surfaces covalently immobilized
with HA and GO (HA/GO/CoCr). Polystyrene was included as a con-
trol. After 72 h of cell-biomaterial culture, cell medium supernatants
were collected and centrifuged (1024 g, 5 min) and the cytokine
secretion levels were quantified with commercial ELISA kits (Diaclone,
Besançon, France). The TNF-α/IL-10 ratios were the average of three
independent assays in triplicate.
2.10 |Statistical analysis
For proteome statistical analysis, two biological replicates were taken
at 72 h of exposure while three replicates were taken at 96 h of bio-
material exposure (n=5) in each of the three groups under study. In
each proteome replicate, the mean LFQ intensity value was obtained
from three technical measurements by LCMS/MS. One-way analysis
of variance (ANOVA) was applied for proteome data analysis of the
three different groups in Z-scoring and hierarchical clustering in Per-
seus software. T-test was applied for analysis of DEPs between paired
groups.
For statistical analysis of the inflammatory response, mean TNF-
α/IL-10 ratios were computed from biological replicates from three
independent assays in triplicates, taken after 72 h of biomaterial
exposure. Inflammatory ratios were statistically analyzed with
GraphPad Prism software via one-way ANOVA. If a significant
ANOVA p-value was obtained (p-value <.05), post-hoc multiple
Tukey test comparisons were performed to assess significance
between paired groups.
3|RESULTS
3.1 |Synthesis and Raman characterization
of HA/GO/CoCr surfaces
The covalent grafting of HA is performed on GO/silane nanocoating
through ADH that works as an intermediate covalent GO-HA linker
through EDC/NHS carbodiimide chemistry. When GO nanosheets are
covalently immobilized on silanized CoCr surfaces,
39
the GO carboxyl
groups are activated by EDC/NHS reagents and the intermediate
ADH linker is assembled via amidation reaction between reactive
ADH amine groups and the activated carboxyl groups of GO.
41,42
A
premixed HA solution containing EDC/NHS is covalently grafted to
the immobilized ADH layer via amidation reaction between activated
HA carboxyl groups
47
and the other free reactive amine group of
ADH to obtain the covalent immobilization forming the HA/GO nano-
coating on CoCr surfaces. A scheme of the complete nanocoating
assembly is depicted in Figure 1.
After completing the immobilization procedure of HA on GO, the
final HA/GO nanocoating on CoCr surfaces was characterized by
SEM and Raman spectroscopy. Figure 2A shows the secondary elec-
tron image of the HA/GO/CoCr surface where the dimensions of the
GO nanosheets of about 200-nm thickness and 1-μm length are
observed. Figure 2BDshow the Raman spectra of CoCr surfaces
with HA/GO and GO nanocoatings. GO and HA references have been
added for comparative purposes. GO reference was obtained by
drop-casting of a 4 mg/mL GO aqueous dispersion, allowing it to dry
to obtain a thick layer of GO on the CoCr surface. HA reference was
directly taken from powder HA.
CoCr surfaces with the GO and HA/GO nanocoatings (Figure 2B)
show the characteristic bands of GO compounds. The D band located
at 1339 cm
1
is associated with the disruption of the pristine aro-
matic ring lattice structure's symmetry. The G band sited at
1605 cm
1
is attributed to the in-plane bond-stretching motion of
pairs of sp2 carbon atoms. The 2D band and the D+G combination
band appear around 2700 cm and 2930 cm
1
, respectively. Upon
covalent immobilization of HA, the HA/GO/CoCr surfaces (Figure 2B)
also show the characteristic peaks of HA in Raman spectra
48
at higher
wavenumber, around 32503750 cm
1
(see HA reference in
Figure 2C). However, the two bands of high intensity at 2904 and
2933 cm
1
corresponding to the C H stretching band and N H
stretching in HA (Figure 2C) are not discernible, which might be
caused by the overlap with the D+G combination band of immobi-
lized GO on the surface, which locates around the same wavenumber.
It can be observed that the heights of the D and G bands are
comparable for both conditions: GO/CoCr and GO reference
(Figure 2D). However, upon functionalizing with HA (HA/GO/CoCr
spectrum), the heights of the D and G bands are lightly higher than
those in GO reference and GO/CoCr surfaces (Figure 2D). The D band
is usually associated with the concentration of defects whereas the G
band is related to the size of sp2 domains. Hence, the ID/IG intensity
ratio can be a good indicator of the degree of disorder for graphene
networks.
S´
ANCHEZ-LÓPEZ ET AL.5
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
FIGURE 1 Scheme of the sequential covalent immobilization on CoCr surfaces. CoCr, cobalt-chrome; EDC, N-(3-dimethylaminopropyl)-N-
ethylcarbodiimide; GO, graphene oxide; HA, hyaluronic acid; NHS, N-hydroxysuccinimide.
2500 3000 3500 4000
Intensity, a.u.
Raman shift, cm–1
HA reference
HA/GO/CoCr
(A) (C)
(B) (D)
1250 1500 1750
G
GO reference
GO/CoCr
HA/GO/CoCr
Intensity, a.u.
Raman shift, cm
–1
D
1000 1500 2000 2500 3000 3500
2933
G
HA reference
GO reference
GO/CoCr
HA/GO/CoCr
Intensity, a.u.
Raman shift, cm
–1
D
D+G HA peaks
2904
2D
FIGURE 2 (A) Secondary electron image of the HA/GO/CoCr surface; (B) Raman spectra of the covalent immobilizations of GO/CoCr and
GO/HA/CoCr surfaces, respectively. Reference spectra are added for powder HA and drop-casted GO; (C) magnification at high wave number of
Raman spectra of HA reference and HA/GO/CoCr surface; (D) magnification of D and G bands in the Raman spectra for GO reference, GO/CoCr,
and HA/GO/CoCr surfaces. CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid.
6S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
The estimation of D and G band intensity ratio, ID/IG ratios, for
GO/CoCr, HA/GO/CoCr, and GO reference are shown in Table 1.
The higher intensity ratio, the higher structural defects. On the
opposite way, lower intensity ratios suggest lower disorder of the
crystalline and symmetry graphene structure. The ID/IG ratio of GO
reference was 0.980. When GO is covalently immobilized on silanized
CoCr surfaces, the comparable I
D
/I
G
ratio (0.998 in GO/CoCr) indi-
cates that the covalent attachment of GO does not induce a substan-
tial degree of disorder within the GO layer, and the size of sp2
domains remains largely unchanged. On the other hand, following the
immobilization of HA, a moderate decrease in the I
D
/I
G
ratio is
observed (0.952 in HA/GO/CoCr), indicative of a lower disorder and
increase in the size of sp2 domains (Figure 2). It is tempting to suggest
that this observation can denote that structural changes in the GO
nanolayer are induced by the EDC/NHS treatment and posterior
immobilization of ADH and HA.
3.2 |Electrochemical characterization of HA/GO
nanocoating on CoCr
The barrier effect of the HA/GO nanocoating on CoCr surfaces was
analyzed by EIS. Experimental impedance data of CoCr surfaces func-
tionalized with the covalent HA/GO nanocoating are represented in
Nyquist diagrams and Bode plots for impedance modulus and phase
angle in Figure 3AC, respectively, when these surfaces were
immersed in HA aqueous electrolyte at 3 g/L. The electrochemical
behavior of the GO/CoCr surfaces at the day 7 of immersion in HA
aqueous electrolyte is also added for comparative purposes.
Figure 3shows that the impedance results for HA/GO/CoCr sur-
faces reveal a high stability with minor differences during immersion
in the HA electrolyte up to 10 immersion days. Nyquist diagrams
show arcs of wide amplitude (Figure 3A) and stabilized impedance
modulus (Figure 3B) and phase angle (Figure 3C) in Bode diagrams
over immersion time. On the other hand, the Nyquist diagrams for the
HA/GO/CoCr surfaces show lower capacitive arcs than those
obtained for the GO/CoCr surfaces. This result indicates that the pos-
terior immobilization of HA on GO nanocoating changed the electro-
chemical properties of the GO/CoCr surfaces.
The experimental impedance data were fitted to electric equiva-
lent circuits (EEC) which assign physical meaning to the parameters of
the circuit. An EEC with two time-constants in series provided a
proper fitting of the metal surface with a thin nanocoating.
43
The pas-
sive electrical components shown in Figure 3A forming the EEC were:
the electrolyte resistance (R
S
), the coating resistance (R
COAT
) due to
HA/GO or GO nanocoating in parallel with a constant phase element
(CPE
2
) and the oxide layer resistance (R
L
) in parallel with its constant
phase element (CPE
1
). CPE
1
and CPE
2
simulate the nonideal capaci-
tive behavior associated with distributed time constants of the modi-
fied surfaces. The fitting values for each EEC element for HA/GO/
CoCr surfaces along with the immersion times up to 23 days are
shown in Table 2. Fitting values for GO/CoCr surfaces at 0 days and
after 7 days of immersion are also added for comparative reasons.
Following the immobilization of HA on GO, a decrease in the
magnitude of the R
COAT
is seen at the first 24 h of immersion, in com-
parison to the R
COAT
values of the GO/CoCr surfaces (Table 2). This
decline in resistivity may be attributed to the enhanced conductivity
induced by HA immobilization in the nanocoating. Nevertheless, with
prolonged immersion, the R
COAT
value in HA/GO/CoCr surfaces
TABLE 1 I
D
/I
G
ratios of GO immobilized (GO/CoCr) and
immobilized with HA and GO (HA/GO/CoCr) in the nanocoatings on
CoCr surface and reference GO.
I
D
(a.u) I
G
(a.u) I
D
/I
G
ratio
GO reference 23,853 24,338 0.980
GO/CoCr 22,109 22,158 0.998
HA/GO/CoCr 41,034 43,122 0.952
Abbreviations: CoCr, cobalt-chrome; GO, graphene oxide; HA,
hyaluronic acid.
FIGURE 3 Experimental Nyquist diagrams (A) and Bode plots (B, C) of HA/GO/CoCr surfaces throughout immersion time in HA electrolyte.
Impedance data for GO/CoCr surfaces after 7 days of immersion is included for comparative reasons. Fitted lines were not included to enhance
the clarity of the figure. Replicate number n=2. CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid.
S´
ANCHEZ-LÓPEZ ET AL.7
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
increases up to 7 days, reaching values comparable to those in
GO/CoCr surfaces (Table 2). Notably, values rise significantly, up to
six times higher values at 23 days, compared with the initial days of
immersion. This behavior suggests the possibility of degradation
occurring in the HA chemically bonded to the GO. However, it is
noteworthy that the electrochemical response of the oxide layer (R
L
,
CPE
1
, and associated n values) remains nearly constant through the
immersion period indicating electrochemical stability of the oxide
layer. This stability may be attributed to the hindering effect of large
HA molecules anchored in the GO, limiting electrolyte entry. Notably,
the electrolyte used for electrochemical testing is a HA solution. The
adsorption of HA molecules onto the HA/GO nanocoating must not
be discarded, potentially increasing resistance in the HA/GO nano-
coating on the CoCr surfaces. The diffusion process of molecules
through the cross-linked HA/GO nanocoating is reflected in the
nvalues associated with the CPE
2
values in Table 2. The high CPE
2
values and an exponent nvalue between 0.5 and 1 denote that diffu-
sion plays a significant role in controlling the electrochemical process
of the system. Overall, the electrochemical response of the protective
nanocoating immersed in HA electrolyte demonstrated remarkable
stability over the chosen immersion period in this work.
The R
L
and CPE
1
values are associated with the oxide layer of the
CoCr in both HA/GO/CoCr and GO/CoCr samples. The HA/GO/CoCr
samples exhibit lower R
L
values compared with the GO/CoCr samples,
possibly due to hindered oxygen diffusion through the HA/GO nano-
coating into the CoCr oxide layer. However, R
L
and CPE
1
values remain
nearly constant for each sample, irrespective of the immersion time
(Table 2). This consistency highlights the electrochemical stability of the
modified CoCr surface. This sustained electrochemical stability on the
CoCr metal surface is also attributed to the strong anchoring of GO
with the silane layer and HA with GO/Si nanocoating. The formation of
a protective nanocoating based on GO anchoring imparts barrier prop-
erties, owing to the impermeable sp2 graphene network securely
adhered to the metal surface through GO/Si interfacial bonding,
39,43
further enhanced by HA immobilization.
3.3 |Characterization of worn surfaces
In Figure 4appears the profilometry of the worn tracks after wear-
corrosion tests of CoCr and HA/GO/CoCr disks against CoCr ball at
5 N load, 120 rpm for 30,000 m in 3 g/L HA medium.
The track length is practically the same on both surfaces, CoCr and
HA/GO/CoCr, as can be seen in the annexed table in Figure 4. However,
the depth of the track analyzed on CoCr surfaces is significantly higher
(24 microns) compared with the HA/GO/CoCr surface (about
16 microns). This difference in track depth results in a lower volume of
metal particles and ions released into the environment (annexed table in
Figure 4), indicating that the presence of the nanocoating exerts a signifi-
cant lubricating effect in the tribocorrosion process of the CoCr surface.
3.4 |Macrophage proteome characterization
Macrophage proteome characterization was carried out after 72 and
96 h interaction with polystyrene control surfaces, intact HA/GO/
CoCr and worn HA/GO/CoCr surfaces. Unsupervised classification
analyses were performed by hierarchical clustering and PCA to clas-
sify the macrophage proteomes expressed in response to the different
biomaterial surfaces.
3.4.1 | Hierarchical clustering
In an initial analysis aimed at offering a comprehensive overview of
trends, Figure 5shows the hierarchical clustering of J744A.1 macro-
phage proteome data when cultured on the polystyrene control, intact
HA/GO/CoCr, and worn HA/GO/CoCr surfaces. Exposure of macro-
phages to the different surfaces was maintained for 72 h (R1 and R2
replicates) and for 96 h (R3, R4, and R5 replicates).
The heatmap displays the replicates of each group within columns
(designates as R1R5 for each group) with their corresponding
TABLE 2 Fitting values of experimental impedance data from day 0 to day 23 of immersion of HA/GO/CoCr and GO/CoCr at 0 and 7 days
of immersion using the equivalent electric circuit shown in Figure 3a.
Time, d
R
S
± abs.
error, Ω
R
COAT
± abs.
error, Ω
CPE
2
± abs.
error, μSsn
n
2
± abs.
error
R
L
± abs.
error, MΩ
CPE
1
± abs.
error, μSsn
n
1
± abs.
error Chi
2
GO/
CoCr
0 3328 ± 7.78 3532 ± 281.44 73.3 ± 4.49 0.68 ± 0.02 66.5 ± 0.35 16.0 ± 0.05 0.94 ± 0.001 4.02 ± 10
5
7 3449 ± 8.63 4549 ± 346.76 61.6 ± 3.42 0.67 ± 0.01 81.9 ± 0.52 14.96 ± 0.05 0.94 ± 0.001 4.49 ± 10
5
HA/
GO/
CoCr
0 2170 ± 7.02 1307 ± 545.14 253.7 ± 51.3 0.73 ± 0.05 12.0 ± 0.22 15.8 ± 0.09 0.91 ± 0.002 8.91 ± 10
5
1 2646 ± 8.87 1803 ± 479.84 154.7 ± 24.0 0.74 ± 0.04 14.0 ± 0.31 16.3 ± 0.09 0.92 ± 0.002 1.00 ± 10
4
2 2527 ± 8.95 3055 ± 918.35 139.5 ± 18.0 0.81 ± 0.04 12.1 ± 0.27 16.1 ± 0.13 0.92 ± 0.003 1.36 ± 10
4
3 2162 ± 7.05 2938 ± 890.6 161.0 ± 19.5 0.77 ± 0.03 13.2 ± 0.29 16.2 ± 0.12 0.92 ± 0.002 1.06 ± 10
4
4 2078 ± 6.67 3547 ± 1117.4 163.1 ± 19.0 0.76 ± 0.03 13.4 ± 0.29 16.1 ± 0.13 0.92 ± 0.002 1.02 ± 10
4
7 2145 ± 7.71 4041 ± 1573.4 173.2 ± 22.5 0.72 ± 0.03 12.6 ± 0.28 15.9 ± 0.16 0.92 ± 0.003 1.19 ± 10
4
23 2240 ± 6.39 25,088 ± 5945.4 82.9 ± 7.6 0.74 ± 0.01 16.8 ± 0.51 16.6 ± 0.29 0.93 ± 0.004 8.37 ± 10
5
Note:R
S
electrolyte resistance; R
COAT
coating resistance, and CPE
2
constant phase element, associated with the HA/GO or GO nanocoating on CoCr
surfaces; R
L
oxide layer resistance and CPE
1
constant phase element related to the metal surface; absolute errors for each element and Chi-square (χ
2
)is
also added.
Abbreviations: CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid.
8S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
macrophage proteome expression profile within rows (comprising
2375 proteins IDs) clustered in the associated dendrogram shown in
Figure 5A. Two main ancestor branches clearly separate macrophage
proteome profiles. Within the first ancestor branch, proteome repli-
cates in response to worn HA/GO/CoCr surfaces are clustered as a
separate population of macrophage proteome (first left dendrogram
branch). In the second hierarchical branch, proteome replicates in
response to both polystyrene control and intact HA/GO/CoCr are
clustered (first right dendrogram branch). This finding highlights the
similarity of macrophage proteomes expressed in polystyrene con-
trol and intact HA/GO/CoCr surfaces. The heatmap also categorizes
proteome expression into two primary clusters (as shown in
Figure 5B), reflecting opposing upregulations and downregulations
observed in the main dendrogram branches. These clusters are
defined as Cluster 1, encompassing proteins with increased expres-
sion following macrophage exposure to the worn HA/GO/CoCr, and
Cluster 2, comprising proteins with increased expression following
macrophage exposure to the intact nanocoating on CoCr or to poly-
styrene. Expression changes as distances from the cluster center are
plottedinFigure5B.
3.4.2 | Principal component analysis
To assess the relative effect of the different surface chemistries on
the overall expression of the macrophage proteome, a multiparametric
space PCA was conducted, as depicted in Figure 6. PCA determines
the direction with the largest total variance between parameters
(or samples) in the multidimensional space determined by PC axes.
The PC 1 determines the new x-axis in the PCA space describing the
direction of maximum percentage of variance between samples
(23.3% of percentage of variance), followed by the rest of PC contrib-
uting with lower percentages of variances, contributing up to the total
percentage (100%) of variance between all replicates in the multidi-
mensional PCA space. Plots of the first five PC and their PCA scores
(%) are shown in Figure 6, revealing that proteome replicates from the
polystyrene control and intact HA/GO/CoCr surfaces were clustered
around the PC1 origin, whereas proteome replicates from worn
HA/GO/CoCr surfaces were located at further PCA distances along
the PC1 axis and spread away from the previous clustered population.
This result reinforces that the covalent nanocoating on CoCr pro-
motes the same protein patterns as polystyrene control while the
wear-corrosion degradation of the nanocoating induces a modification
in the macrophage protein expression.
3.4.3 | Protein abundance
Figure 7shows scatter pairwise density plots that compare protein
abundances between replicate groups. It is important to note that when
comparing macrophage protein expression for each protein ID (each ID
represented by a black dot), those in the control group (x-axis value)
and other surface groups (y-axis value) often exhibit similar protein
abundances. This is evident when black circles cluster closely around
the red line. However, there are fewer instances where certain protein
IDs show increased or decreased abundances between the paired
groups, which is indicated by black circles positioned above or below
the red lines, respectively. In cases where the protein abundances are
exactly the same in two different macrophage groups, they would align
perfectly along the red line (when black dots overlap the red line).
These plots reveal a close similarity in protein abundances
between the polystyrene control and the intact HA/GO/CoCr
FIGURE 4 Profilometry of the worn tracks of CoCr (A) and HA/GO/CoCr (B) surfaces, against CoCr ball at 5 N load, 120 rpm for 30,000 m in
3 g/L HA medium. CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid.
S´
ANCHEZ-LÓPEZ ET AL.9
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
FIGURE 6 Principal component analysis (PCA) showing distribution of sample groups under study along PCA axis: PC1 versus PC2, PC3, PC4,
and PC5. Replicate number n=5.
FIGURE 5 Hierarchical clustering of high-throughput proteome data from the J744A.1 macrophage when cultured on polystyrene control
surface, intact HA/GO/CoCr, and worn HA/GO/CoCr surfaces during 72 h (R1 and R2 samples) or 96 h (R3, R4, and R5 samples).
(A) Dendrogram showing clustered groups with heatmap color representation of corresponding macrophage proteome expressions (2375 total
protein IDs detected); (B) Cluster profile plot with distances to the cluster center of proteins classified with reduced/increased expression (1201
protein IDs) or increased/reduced expression (1174 protein ID) in cluster 1 or cluster 2, respectively. Replicate number n=5. CoCr, cobalt-
chrome; GO, graphene oxide; HA, hyaluronic acid.
10 S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
(Figure 7A), as indicated by the narrow linear slopes and high correla-
tion coefficients (R2 closer to 1, represented by the line y=x). On the
other hand, in worn HA/GO/CoCr surfaces, protein abundances
exhibit more significant differences in comparison to both the polysty-
rene control and intact HA/GO/CoCr surfaces. This is evident from a
greater number of data points which deviate from the y=xreference
line, resulting in smaller correlation coefficient (R2 values in
Figure 7B,C).
3.4.4 | Differentially expressed proteins and Gene-
ontology enrichment analysis
Figure 8shows the proteome profiling of J744A.1 macrophage in
response to each group surfaces from pairwise volcano plots upon
72 h (A) and 96 h (B) of exposure.
DEPs are shown in the context of pairwise surface group compari-
sons through the percentage of total upregulated (UP) and downregu-
lated(DOWN).TheanalysisofDEPswasalsoperformedwithGene-
ontology enrichment analysis. The intact HA/GO nanocoating on the
CoCr surface induced only marginal percentages of upregulated and
downregulated DEPs with respect to the polystyrene control. Specifically,
at 72 h, it resulted in 1.3% upregulated and 0.6% downregulated DEPs,
while at 96 h, it led to 0.04% upregulated and 0.3% downregulated DEPs
(Figure 8a.1, a.2). However, for worn HA/GO/CoCr surfaces, there is a
notable increase in the percentage of upregulated DEPs and, more signifi-
cantly, in the percentages of downregulated DEPs in comparison to the
polystyrene control (asdepictedinFigure8b.1, b.2). This trend is also
observed when comparing worn HA/GO/CoCr with the intact HA/GO/
CoCr surfaces (as shown in Figure 8c.1, c.2) at both 72 and 96 h.
When DEPs of intact HA/GO/CoCr and polystyrene control at
both exposure times are summed, the percentages were 0.3%
FIGURE 7 Scatter plots showing the comparison of proteome abundances in macrophages cultured on the three surface groups, (A) intact
HA/GO/CoCr vs control (B) worn HA/GO/CoCr vs control, and (C) worn HA/GO/CoCr vs intact HA/GO/CoCr. Normalized protein abundance of
sample replicates in x and y axes (Log2-normalization of LFQ values). To quantify the similarity/dispersion degree of the proteome abundances,
correlation coefficients (R2) are shown. Replicate number n=5. CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid; LFQ, label-free
quantification.
S´
ANCHEZ-LÓPEZ ET AL.11
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
upregulated and 0.7% downregulated (as illustrated in Figure 9A), with
respect to the entire macrophage proteome.
The same pattern of DEPs for worn HA/GO/CoCr versus polysty-
rene (Figure 9B) and versus intact HA/GO/CoCr (Figure 9C) is evident
when both exposure times are summed. The Gene-ontology enrich-
ment analysis of upregulated and downregulated DEPs can be
accessed in the Figures S1S3.
3.5 |Macrophage inflammatory balance
The macrophage inflammatory ratio was determined by measuring
secreted pro-inflammatory and anti-inflammatory cytokines TNF-α
and IL-10, respectively, in cell culture extracts, after 72 h. To
assess the implication of the covalent immobilization of the
HA/GOonCoCrsurfacesonthemacrophageinflammatory
FIGURE 8 Macrophage proteome profiling upon 72 h (a.1, b.1, c.1) and 96 h (a.2, b.2, c.3) of exposure showing percentage of total
upregulated (UP) and downregulated (DOWN) DEPs from pairwise volcano comparisons between intact HA/GO/CoCr vs polystyrene control
(a.1, a.2); worn HA/GO/CoCr vs polystyrene control (b.1, b.2), and worn HA/GO/CoCr vs intact HA/GO/CoCr (c.1, c.2). Some labeled proteins
were identified in the samples analyzed. Replicate number n=2 for 72 h of cell-biomaterial exposure; replicate number n=3 for 96 h of cell-
biomaterial exposure. CoCr, cobalt-chrome; DEP, differentially expressed protein; GO, graphene oxide; HA, hyaluronic acid.
12 S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
response, a comparative analysis of the TNF-α/IL-10 ratio was also
evaluated with different surface chemistry and using different
types of surface immobilization (covalent vs. physical). In
Figure 10, a comparative of modified CoCr surfaces with covalent
immobilization of HA and GO (HA/GO/CoCr) on intact modified
surfaces (Figure 10A) and after applying wear-corrosion
(Figure 10B), and physical immobilization of ErGO (ErGO/CoCr)
and HA (HA/ErGO/CoCr) is shown. Polystyrene is added as con-
trol. No significant differences were obtained in TNF-α/IL-10
ratios of macrophages cultures on intact HA/GO/CoCr versus
polystyrene control after 72 h (Figure 10A).
As it is shown in Figure 10A, the inflammatory response
decreases significantly upon the covalent immobilization of HA/GO
on CoCr in comparison with the physical immobilization of electro-
chemically reduced GO (ErGO) and further functionalized under
immersion with HA (ErGO/HA). Covalent immobilization of HA and
GO on CoCr surfaces provides the largest effect by recovering a TNF-
α/IL-10 ratio comparable with the basal ratios cultured on commercial
polystyrene surface, obtaining nonsignificant differences between
HA/GO/CoCr versus polystyrene control surfaces (Figure 10A). This
agrees with proteome analyses by PCA and hierarchical clustering,
which showed similar macrophage proteomes when they are cultured
FIGURE 9 Macrophage proteome profiling considering the data summed at 72 h and 96 h between (A) intact HA/GO/CoCr versus control
(B) worn HA/GO/CoCr versus control, and (C) worn HA/GO/CoCr versus intact HA/GO/CoCr. Some labeled proteins were identified in the
samples analyzed. Replicate number n=5. CoCr, cobalt-chrome; GO, graphene oxide; HA, hyaluronic acid.
S´
ANCHEZ-LÓPEZ ET AL.13
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
on treated polystyrene and on HA/ GO/CoCr surfaces (see sections
3.4.1,3.4.2).
It is known that wear-corrosion phenomena on CoCr surfaces fur-
ther exacerbates macrophage pro-inflammatory ratios and increases
TNF-αsecretion.
12,49,50
Likewise, a slight but no significant increase in
TNF-α/IL-10 ratio is observed upon wear-corrosion of HA/GO/CoCr
with respect to the polystyrene control (Figure 10B). The covalent
immobilization in HA/GO/CoCr surfaces is able to reduce wear-
corrosion-induced inflammation at the largest extent with respect to
other physical adsorption of ErGO on CoCr surfaces, such as ErGO/
CoCr, HA/ErGO/CoCr (Figure 10B).
4|DISCUSSION
The beneficial impact of HA in healthcare is evident, given its wide-
spread use in various biomedical applications. Among its remarkable
attributes, alongside its abundance in the human body, is its capacity
to reduce friction. These properties render it a vital component in
synovial fluid, playing a significant role in joint lubrication. Conversely,
graphene compounds, when employed as solid lubricants on metallic
substrates, exhibit lubrication properties that could effectively reduce
wear in joint prostheses. The challenge in this research has been to
characterize the CoCr surface immobilized with HA and GO assess
the macrophage response.
The use of ADH as an intermediate reagent to promote the cova-
lent immobilization of HA on the GO induces modifications in the GO
network. The functionalization of GO by ADH is reported to occur
through the amidation reaction between activated EDC/NHS carboxyl
groups of GO and the primary amine groups of the ADH agent.
41,42
This process results in the removal of an oxygen atom from GO and
the nitrogen functionalization of GO immobilized on CoCr. In this
step, the simultaneous reduction and functionalization of GO caused
by the removal of oxygen-containing groups is also well-known to
occur for other reagents.
5157
The functionalization of GO with ADH
did not introduce defects in the carbon network, as the ID/IG inten-
sity ratio for HA/GO/CoCr surfaces (0.952) is lower than that of
GO/CoCr surfaces (0.998), suggesting that ADH treatment may indi-
cate the mitigation of sp3 defects.
58
This decrease in the ID/IG ratio
suggests potential defect repair by ADH, particularly at the edges of
the GO lattice where functionalization takes place. Similar linear
amino-containing agent also reported not only avoid introducing
defects but also repair defects and enlargement of the sp2 domains
by a decrease in ID/IG ratios upon functionalization of GO with the
ethylene diamine (EDA).
59
The reduction of oxygenated groups of GO is known to be
directly proportional to an increase in the electric conductivity of GO
and a decrease in its hydrophilic character. The heightened conductiv-
ity in HA/GO nanocoating on CoCr surfaces can also be associated
with the functionalization with ADH, as functionalization with
diamines partially restore the aromatic structure in GO and the amine
groups donate electron density to the aromatic rings of GO.
55,60,61
This restoration likely constitutes another contributing factor to the
improved conductivity which correlates with the low R
COAT
values
estimated in the HA/GO/CoCr samples within the first 24 h of immer-
sion in HA electrolyte (Table 2). According to Ref. 62, besides the
reduction of terminal carboxylic groups of GO upon GO/ADH functio-
nalization, activation by EDC/NHS carbodiimide chemistry can also
contribute to a decrease in the R
COAT
of this nanocoating.
The GO-based encapsulation of CoCr by the HA/GO nanocoating
(see section 3.2) works hindering metal ion release.
43
This is corre-
lated with a reduction in the toxicities observed in graphene- and GO-
encapsulated metal implants, attributed to reduced metal ion
release.
6366
The electrochemical parameter of R
L
in the HA/GO/CoCr exhibits
remarkable stability in the aqueous electrolyte solution. This indicates
that the infiltration of the electrolyte into the oxide layer remains
unchanged. Furthermore, it implies that the covalent linkage of HA to
ADH can improve its stability against degradation processes. This is
especially valuable considering that HA degradation primarily proceeds
through nonenzymatic processes in synovial fluid, an environment
where hyaluronidase activity is barely detectable.
30
In this medium,
joints HA degradation is predominantly driven by oxidation and hydro-
lysis mechanisms under the physiological pH of the synovial fluid.
28,67
70
The covalent binding of HA to ADH represents a promising approach
to enhance resistance to degradation. Studies show that covalently
functionalizing HA with ADH leads to considerably slower degradation
kinetics compared with untreated HA.
71,72
This suggests that covalent
immobilization of HA and GO could provide added benefits in terms of
safeguarding against hydrolytic degradation after implantation.
FIGURE 10 Comparative TNF-α/IL-10 macrophage ratio on
modified CoCr surfaces with physical immobilization of ErGO (ErGO/
CoCr) and physical immobilization of HA and ErGO (HA/ErGO/CoCr)
or covalent immobilization of HA and GO (HA/GO/CoCr) on intact
modified surfaces (A) and after applying wear-corrosion (B) after 72 h.
Polystyrene is used as control. p-value <.05 (*), p-value <.01 (**), p-
value <.005 (***), p-value <.001 (****), nsnonsignificant. The TNF-
α/IL-10 ratios were the average of three independent assays in
triplicates. Replicate number n=3. CoCr, cobalt-chrome; GO,
graphene oxide; HA, hyaluronic acid; TNF-α, tumor necrosis factor α.
14 S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
However, given that significantly altered HA chemical structures
can potentially induce cytotoxicity, the study of the macrophage
response to the modified HA structure immobilized in the nanocoat-
ing is needed. The characterization of the macrophage proteome and
the analysis of the pro-inflammatory and anti-inflammatory molecule
secretion ratio is a key point in the proposal of this immobilization as
a suitable stable and protective bionanocoating on CoCr surfaces.
Unsupervised analyses through hierarchical clustering and PCA
analysis offer an unbiased classification of massive proteomic data.
The outcomes of hierarchical clustering and PCA revealed highly com-
parable macrophage proteomes on both cell-culture treated polysty-
rene and CoCr surfaces covalently immobilized with the HA/GO
nanocoating. Hierarchical clustering analysis classified the proteomes
of macrophages cultured on polystyrene or intact HA/GO/CoCr in
the same dendrogram branches, as they present similar gene expres-
sions in cluster 2 (Figure 5). In a similar way, PCA placed proteome
replicates from both polystyrene control and intact HA/GO/CoCr in a
mixed population around the PC1 origin (Figure 6). These results
strongly indicate a significant homology in proteome expressions.
Likewise, macrophages cultured on polystyrene control and intact
HA/GO/CoCr surfaces exhibit highly comparable protein abundances,
with R2 correlation coefficients near 1 (Figure 7A), indicating very low
percentages of DEPs and, consequently, very similar proteomes
(Figure 8a.1, a.2; Figure 9A). Gene-ontology enrichment analysis of
DEPs revealed upregulations in adhesion-related proteins (Parvb,
Itgb4bp, Ctnnd1) after 96 h. This suggests the formation of stable
adhesions between macrophages and the intact HA/GO/CoCr sur-
faces through these adhesion proteins (see Figure S2a). The increased
abundance of integrin beta 4, (Itgb4bp or Eif6) can facilitate the direct
binding of cells to material surfaces via integrins.
73
The increased
presence of beta parvin (Parvb), a focal adhesion protein connecting
ECM-binding integrins to the actin cytoskeleton, suggests the forma-
tion of focal adhesions. These structures facilitate cell spreading and
the establishment of stable adhesions on material substrates by link-
ing integrins to the actin cytoskeleton.
7476
Moreover, these focal
adhesions play a crucial role in in conveying physicochemical clues
from the substrate to the associated cytoskeletal network through
mechano-transduction at the focal adhesions.
77
Similarly, the
increased abundance of catenin (Ctnnd1) can also regulate integrin
cell-matrix adhesions.
78
Altogether, this indicates an enhanced cell
adhesion on the intact HA/GO/CoCr substrate. The covalent grafting
of HA on the nanocoating likely assumes a significant role in adhesion
on the functionalized metal surface, as HA enhances integrin-
mediated mechano-transduction and promotes focal adhesion forma-
tion.
79
Additionally, the deposition of anionic HMW-HA increases the
hydrophilicity at CoCr surface.
80
It is well-established that enhancing
the hydrophilicity of hydrophobic surfaces, such as metals, improves
cell adhesion on biomaterials.
81
The heightened response in the mac-
rophage proteome can be attributed to the increased surface hydro-
philicity provided by the HA/GO nanomodifications on CoCr. These
modifications are reminiscent of patented surface treatments
designed to improve cell adhesion on polystyrene surfaces. Polysty-
rene surfaces are typically highly hydrophobic, exhibiting low cellular
attachment in their unmodified state. Various surface treatments
employed on commercial cell-culturing plastic plates include plasma
treatments to increase surface hydrophilicity,
82,83
silica film
deposition,
84
and the application of ECM proteins and proteoglycans,
85
among other methods. The latter, involving ECM proteins and proteo-
glycans, may help explain the comparable outcomes observed with the
HA/GO nanocoating on CoCr surfaces.
In summary, these findings underscore the efficacy of the cova-
lently immobilized HA/GO nanocoating on CoCr surfaces in creating a
cell-culture-treated metallic surface with excellent biocompatibility
with macrophages.
However, the degradation of the HA/GO nanocoating during
wear-corrosion processes exposes certain bare CoCr areas to the cell
culture. The coexistence of both the intact HA/GO nanocoating and
the bare CoCr areas induces significant differences in the macrophage
proteome. Hierarchical clustering, in this context, classified the worn
HA/GO/CoCr surfaces as inducing proteome replicates in a distinct
cluster. Additionally, PCA positioned these surfaces in a different mul-
tiparametric space compared with the polystyrene control and intact
HA/GO/CoCr surfaces. Scatter plots depicting protein abundance,
characterized by R
2
correlation coefficients less than 1 (Figure 7B,C),
indicated more pronounced differences in protein abundance within
the wear-corrosion group compared with the other two groups. Con-
sequently, higher percentages of significantly upregulated and down-
regulated DEPs were found in the wear-corrosion group as opposed
to the other two groups. This suggests a distinctively expressed prote-
ome, particularly in the percentages of downregulated DEPs (Figure 8
b.1, b.2, c.1, c.2; Figure 9B,C). Notable downregulations in pro-
inflammatory pathways were evident in the macrophage proteome
after 72 and 96 h of exposure to the worn HA/GO/CoCr surfaces.
Significant enrichments were observed in the downregulated proteins
interacting with the TNF-αnF-kB signaling pathway (Pfdn2, Cdc37,
Psmd6, Hsp90aa1, Hsp90ab1, Smarca4, Smarcc1, Txlna Arhgef2, Skp1,
Ppp6c, Rack1, or Gnb2l1) (Figures S1c,S2b,c, and S3b,c).
86,87
Downre-
gulation enrichments in the NLRP3 inflammasome (Casp1, Pycard,
Sugt1), the IL-1 signaling pathway (Capn1, Casp1, Sirpa) and IFN sig-
naling (Ptpn6, Abce1) and IFN production (G3bp1, Tomm70a, Irf5,
Hmgb1, Pycard, Sirpa) were also observed (Figures S2b,c and 3b,c).
The downregulations stimulated by immunosuppressive HMW-
HA can effectively inhibit the induction of pro-inflammatory
responses in macrophages exposed to implant wear debris, potentially
leading to a reduction in inflammation. This hypothesis was further
corroborated by the analysis of pro- and anti-inflammatory TNF-α/IL-
10 cytokine secretion ratios. The findings support the idea that immu-
nosuppressive HMW-HA from the covalent nanocoating mitigates
TNF-α/IL-10 ratios during biomaterial tribocorrosion. Macrophages
present in different tissues are polarized according to changes in their
environment, resulting in distinct subtypes such as M1 and M2 macro-
phages.
88,89
M1 macrophages exhibit pro-inflammatory capabilities,
secreting factors like IL-6, IL-12, and TNF-α. Conversely, M2 macro-
phages demonstrate anti-inflammatory functions, aiding in tissue
repair and expressing cytokines as IL-10.
90,91
TNF-αis recognized as a
marker for the pro-inflammatory M1 phenotype, while IL-10 serves as
S´
ANCHEZ-LÓPEZ ET AL.15
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
a marker for the anti-inflammatory M2 phenotype. It is known that
CoCr biomaterial and specially CoCr particulate debris and ions fur-
ther exacerbates macrophage pro-inflammatory ratios and increases
TNF-αsecretion.
12,49,50
Previous studies have shown that macro-
phages cultured on bare CoCr display a highly inflammatory response,
inducing more than a 15-fold in the TNF-α/IL-10 ratios with respect
to the basal ratios on control cultures.
12
Chemical modifications of
the bare CoCr surface, achieved through the physical immobilization
of ErGO sheets
92
and HMW-HA,
93
partially alleviate the inflammation
typically associated with the surface. Nevertheless, the TNF-α/IL-10
ratios on these modified surfaces remain significantly elevated com-
pared to the basal level in macrophages (Figure 10A), indicative of a
shift toward the pro-inflammatory M1 macrophage phenotype.
Similarly, a marginal, though not statistically significant, rise in the
TNF-α/IL-10 ratio is noted when macrophages are cultured on CoCr
with the HA/GO nanocoating. This observation suggests a potential
presence of inflammation, possibly triggered by the CoCr metal implant
beneath the coating and/or by the release of metal ions or particles into
the cell culture. Nevertheless, under wear conditions, the covalent
immobilization of the HA/GO nanocoating significantly minimizes the
inflammatory ratio compared with the physical adsorption of ErGO and
HA on CoCr surfaces (Figure 10B). Therefore, covalent modification of
CoCr with GO/HA is more likely to induce macrophage polarization
toward the M2 phenotype anti-inflammatory compared with CoCr sur-
faces modified by the physical adsorption of ErGO and ErGO/HA.
The covalent immobilization of HA and GO on CoCr surface
exhibits the most significant biocompatible effect, leading to TNF-
α/IL-10 ratios that closely resemble the basal ratios of macrophages
cultured on commercial polystyrene control surfaces (ns differences,
Figure 10), suggesting compatibility with an anti-inflammatory M2
phenotype that promote tissue repair. This finding is consistent with
proteome analyses by PCA and hierarchical clustering, demonstrating
similar macrophage proteomes when cultured on treated polystyrene
and intact HA/GO/CoCr (see sections 3.3.1, 3.3.2).
The decrease in pro-inflammatory TNF-α/IL-10 ratios on the
worn HA/GO/CoCr surfaces correlates with observed downregula-
tions in pro-inflammatory signaling pathways within the macrophage
proteome. Notably, reductions were identified in TNF-α, NLRP3, IFN-
γ, and IL-1 signaling pathways, signifying the immunosuppression
effects of HMW-HA covalently grafted onto the functionalized metal
surface in response to tribocorrosion-induced inflammatory signals.
This modulation plays a role in suppressing the macrophage inflamma-
tion response.
In summary, the modification of the surface chemistry CoCr
through covalent immobilization of HA and GO shows both electro-
chemical stability and biocompatibility evidenced by analyses of mac-
rophage proteome expression and the TNF-α/IL-10 secretion ratio.
5|CONCLUSIONS
The covalent immobilization of HA and GO and does not incorpo-
rate additional structural disorder in the graphene network. Even
more, a slight enhancement in the size of sp2 domains is verified.
The electrochemical response corroborates a barrier effect of
HA/GO nanocoating on the CoCr surfaces.
The macrophages cultured on CoCr surfaces covalently modified
with the HA/GO nanocoating expressed a proteome highly equiva-
lent to those cultured on polystyrene control.
Disruption of covalent immobilization HA/GO nanocoating on
CoCr surfaces by tribocorrosion processes induces a significantly
different macrophage proteome. Nevertheless, the covalent
anchoring of the nanocoating on the metallic material represents
an advantage over the physical adsorption of the coating.
The immune response exerted by HA/GO covalently immobilized
on the metal surface was able to reduce the macrophage inflamma-
tory response, indicating that the covalent HA/GO nanocoating on
CoCr provides potential applications for metal implants and pros-
theses associated with metal-induced inflammation and degrada-
tion by wear-corrosion in vivo.
AUTHOR CONTRIBUTIONS
All authors contributed to the study conception and design. Material
preparation, data collection, and analysis were performed by all
authors. The first draft of the manuscript was written by L. Sánchez-
López and all authors commented on previous versions of the manu-
script. All authors read and approved the final manuscript.
ACKNOWLEDGMENTS
The authors wish to thank Francisco García Tabares (F.G.T.), Tamar San
HipólitoMarín(T.S.H.M.),andViviandelosRíos(V.d.l.R)oftheProteo-
mic and genomic facility at the Centro de Investigaciones Biológicas Mar-
garita Salas (CIB-MS, CSIC) for technical assistance in the preparation of
samples for proteomic analysis (F.G.T. and T.S.H.M) and technical sup-
port for the analysis of proteomic data (V.d.l.R). We acknowledge support
of the publication fee by the CSIC Open Access Publication Support Ini-
tiative through its Unit of Information Resources for Research (URICI).
FUNDING INFORMATION
This wok was supported by RTI2018-101506-B-C31 and
RTI2018-101506-B-C33 from the Ministerio de Ciencia, Innovación y
Universidades (MICIU/FEDER) in Spain. Author L. Sánchez-López has
received financial support from MICIU for the predoctoral contract
PRE2019-090122.
CONFLICT OF INTEREST STATEMENT
The authors declare no conflicts of interest.
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are openly available in
Digital CSIC at https://digital.csic.es/.
ORCID
L. Sánchez-López https://orcid.org/0000-0002-5065-9239
B. Chico https://orcid.org/0000-0001-8697-6298
Maria Cristina García-Alonso https://orcid.org/0000-0003-0275-
4626
Rosa M. Lozano https://orcid.org/0000-0003-2762-6938
16 S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
REFERENCES
1. Connors JP, Stelzer JW, Garvin PM, Wellington IJ, Solovyova O. The
role of the innate immune system in wear debris-induced inflamma-
tory peri-implant osteolysis in total joint arthroplasty. Bioengineering.
2022;9:764. doi:10.3390/bioengineering9120764
2. Hodges NA, Sussman EM, Stegemann JP. Aseptic and septic pros-
thetic joint loosening: impact of biomaterial wear on immune cell
function, inflammation, and infection. Biomaterials. 2021;278:121-
127. doi:10.1016/j.biomaterials.2021.121127
3. Ude CC, Esdaille CJ, Ogueri KS, et al. The mechanism of metallosis
after total hip arthroplasty. Regen Eng Transl Med. 2021;7:247-261.
doi:10.1007/s40883-021-00222-1
4. Nich C, Takakubo Y, Pajarinen J, et al. Macrophageskey cells in the
response to wear debris from joint replacements. J Biomed Mater Res
Part A. 2013;101:3033-3045. doi:10.1002/jbm.a.34599
5. Posada O, Tate R, Meek RM, Grant M. In vitro analyses of the toxicity,
immunological, and gene expression effects of cobalt-chromium alloy
wear debris and Co ions derived from metal-on-metal hip implants.
Lubricants. 2015;3:539-568. doi:10.3390/lubricants3030539
6. Ricciardi BF, Nocon AA, Jerabek SA, et al. Histopathological charac-
terization of corrosion product associated adverse local tissue reac-
tion in hip implants: a study of 285 cases. BMC Clin Pathol. 2016;16:3.
doi:10.1186/s12907-016-0025-9
7. Barbosa JN, Vasconcelos DP. Macrophage response to biomaterials.
In: Mozafari M, ed. Handbook of Biomaterials Biocompatibility. Wood-
head Publishing Series in Biomaterials; 2020:43-52. doi:10.1016/
B978-0-08-102967-1.00003-7
8. Davenport Huyer L, Pascual-Gil S, Wang Y, Mandla S, Yee B,
Radisic M. Advanced strategies for modulation of the material
macrophage interface. Adv Funct Mater. 2020;30:1909331. doi:10.
1002/adfm.201909331
9. Martin KE, García AJ. Macrophage phenotypes in tissue repair and
the foreign body response: implications for biomaterial-based regen-
erative medicine strategies. Acta Biomater. 2021;133:4-16. doi:10.
1016/j.actbio.2021.03.038
10. Roh S, Jang Y, Yoo J, Seong H. Surface modification strategies for bio-
medical applications: enhancing cellbiomaterial interfaces and
biochip performances. Biochip J. 2023;17:174-191. doi:10.1007/
s13206-023-00104-4
11. García-Argumánez A, Llorente I, Caballero-Calero O, et al. Electro-
chemical reduction of graphene oxide on biomedical grade CoCr alloy.
Appl Surf Sci. 2019;465:1028-1036. doi:10.1016/j.apsusc.2018.
09.188
12. Sánchez-López L, Ropero de Torres N, Chico B, et al. Effect of wear-
corrosion of reduced graphene oxide functionalized with hyaluronic
acid on inflammatory and proteomic response of J774A.1 macro-
phages. Metals. 2023;13:598. doi:10.3390/met13030598
13. García-Alonso MC, Chico B, Lozano RM, Escudero ML. Tribocorro-
sion behavior of graphene-based solid lubricants biofunctionalized
with hyaluronic acid on CoCr surfaces. Tribol Int. 2023;193:108420.
doi:10.1016/j.triboint.2023.108420
14. Masters KS. Covalent growth factor immobilization strategies for tis-
sue repair and regeneration. Macromol Biosci. 2011;11:1149-1163.
doi:10.1002/mabi.201000505
15. Ito Y. Covalently immobilized biosignal molecule materials for tissue
engineering. Soft Matter. 2008;4:46-56. doi:10.1039/B708359A
16. Bolivar JM, Woodley JM, Fernandez-Lafuente R. Is enzyme immobili-
zation a mature discipline? Some critical considerations to capitalize
on the benefits of immobilization. Chem Soc Rev. 2022;51:6251-
6290. doi:10.1039/D2CS00083K
17. Guisan JM, López-Gallego F, Bolivar JM, Rocha-Martín J, Fernandez-
Lorente G. The science of enzyme immobilization. In: Guisan JM,
Bolivar JM, López-Gallego F, et al., eds. Immobilization of Enzymes and
Cells: Methods and Protocols. 4th ed. Humana Press; 2020:1-26. doi:
10.1007/978-1-0716-0215-7_1
18. Rodrigues RC, Berenguer-Murcia ´
A, Carballares D, Morellon-
Sterling R, Fernandez-Lafuente R. Stabilization of enzymes via immo-
bilization: multipoint covalent attachment and other stabilization
strategies. Biotechnol Adv. 2021;52:107821. doi:10.1016/j.
biotechadv.2021.107821
19. Singh RK, Tiwari M, Singh R, Lee JK. From protein engineering to
immobilization: promising strategies for the upgrade of industrial
enzymes. Int J Mol Sci. 2013;14:1232-1277. doi:10.3390/
ijms14011232
20. Tripathi A, Melo JS. Immobilization strategies: biomedical, bioengineer-
ing and environmental applications. Springer; 2021. doi:10.1007/978-
981-15-7998-1
21. Castellanos MI, Mas-Moruno C, Grau A, et al. Functionalization of
CoCr surfaces with cell adhesive peptides to promote HUVECs adhe-
sion and proliferation. Appl Surf Sci. 2017;393:82-92. doi:10.1016/j.
apsusc.2016.09.107
22. Castellanos MI, Zenses A-S, Grau A, et al. Biofunctionalization of
REDV elastin-like recombinamers improves endothelialization on
CoCr alloy surfaces for cardiovascular applications. Colloids Surf B
Biointerfaces. 2015;127:22-32. doi:10.1016/j.colsurfb.2014.
12.056
23. Kang J, Sakuragi M, Shibata A, et al. Immobilization of epidermal
growth factor on titanium and stainless steel surfaces via dopamine
treatment. Mater Sci Eng C. 2012;32:2552-2561. doi:10.1016/j.msec.
2012.07.039
24. Li Y, Giesbers M, Gerth M, Zuilhof H. Generic top-functionalization of
patterned antifouling zwitterionic polymers on indium tin oxide. Lang-
muir. 2012;28:12509-12517. doi:10.1021/la3022563
25. Puleo DA, Kissling RA, Sheu M-S. A technique to immobilize bioactive
proteins, including bone morphogenetic protein-4 (BMP-4), on tita-
nium alloy. Biomaterials. 2002;23:2079-2087. doi:10.1016/S0142-
9612(01)00339-8
26. Santos M, Waterhouse A, Lee BSL, et al. Simple one-step covalent
immobilization of bioactive agents without use of chemicals on
plasma-activated low thrombogenic stent coatings. In: Wall JG,
Podbielska H, Wawrzyńska M, eds. Functionalised Cardiovascular
Stents. Woodhead Publishing; 2018:211-228. doi:10.1016/B978-0-
08-100496-8.00011-1
27. Sasaki M, Inoue M, Katada Y, Taguchi T. The effect of VEGF-
immobilized nickel-free high-nitrogen stainless steel on viability and
proliferation of vascular endothelial cells. Colloids Surf B Biointerfaces.
2012;92:1-8. doi:10.1016/j.colsurfb.2011.10.061
28. Stern R, Kogan G, Jedrzejas MJ, ˇ
Soltés L. The many ways to cleave
hyaluronan. Biotechnol Adv. 2007;25:537-557. doi:10.1016/j.
biotechadv.2007.07.001
29. Tamer TM. Hyaluronan and synovial joint: function, distribution and
healing. Interdiscip Toxicol. 2013;6:111-125. doi:10.2478/intox-2013-
0019
30. Volpi N, Schiller J, Stern R, Soltes L. Role, metabolism, chemical modi-
fications and applications of hyaluronan. Curr Med Chem. 2009;16:
1718-1745. doi:10.2174/092986709788186138
31. Hintze V, Schnabelrauch M, Rother S. Chemical modification of hya-
luronan and their biomedical applications. Front Chem. 2022;10:
830671. doi:10.3389/fchem.2022.830671
32. Ao H, Xie Y, Tan H, et al. Fabrication and in vitro evaluation of stable
collagen/hyaluronic acid biomimetic multilayer on titanium coatings.
J R Soc Interface. 2013;10:20130070. doi:10.1098/rsif.2013.0070
33. Ao H, Xie Y, Tan H, et al. Improved hMSC functions on titanium coat-
ings by type I collagen immobilization. J Biomed Mater Res A. 2014;
102:204-214. doi:10.1002/jbm.a.34682
34. Kotla NG, Mohd Isa IL, Larrañaga A, et al. Hyaluronic acid-based bio-
conjugate systems, scaffolds, and their therapeutic potential. Adv
Healthc Mater. 2023;12:2203104. doi:10.1002/adhm.202203104
35. Schanté C, Zuber G, Herlin C, Vandamme TF. Synthesis of N-alanyl-
hyaluronamide with high degree of substitution for enhanced
S´
ANCHEZ-LÓPEZ ET AL.17
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
resistance to hyaluronidase-mediated digestion. Carbohydr Polym.
2011;86:747-752. doi:10.1016/j.carbpol.2011.05.017
36. Edwards P, Fantasia JE. Review of long-term adverse effects associ-
ated with the use of chemically-modified animal and nonanimal
source hyaluronic acid dermal fillers. Clin Interv Aging. 2007;2:509-
519. doi:10.2147/cia.s382
37. Gallo N, Nasser H, Salvatore L, et al. Hyaluronic acid for advanced
therapies: promises and challenges. Eur Polym J. 2019;117:134-147.
doi:10.1016/j.eurpolymj.2019.05.007
38. Tiwari S, Bahadur P. Modified hyaluronic acid based materials for bio-
medical applications. Int J Biol Macromol. 2019;121:556-571. doi:10.
1016/j.ijbiomac.2018.10.049
39. Sánchez-López L, Chico B, Llorente I, Escudero ML, Lozano RM,
García-Alonso MC. Covalent immobilization of graphene oxide on
biomedical grade CoCr alloy by an improved multilayer system assem-
bly via silane/GO bonding. Mater Chem Phys. 2022;287:126296. doi:
10.1016/j.matchemphys.2022.126296
40. Pujari SP, Scheres L, Marcelis ATM, Zuilhof H. Covalent surface modi-
fication of oxide surfaces. Angew Chem Int ed. 2014;53:6322-6356.
doi:10.1002/anie.201306709
41. Wu H, Shi H, Wang Y, et al. Hyaluronic acid conjugated graphene
oxide for targeted drug delivery. Carbon. 2014;69:379-389. doi:10.
1016/j.carbon.2013.12.039
42. Zhang H, Li Y, Pan Z, et al. Multifunctional nanosystem based on gra-
phene oxide for synergistic multistage tumor-targeting and combined
chemo-photothermal therapy. Mol Pharm. 2019;16:1982-1998. doi:
10.1021/acs.molpharmaceut.8b01335
43. Sánchez-López L, Chico B, Escudero ML, Lozano RM, García-
Alonso MC. Barrier graphene oxide on a CoCr alloy via silane/GO
covalent bonding and its electrochemical behavior in a simulated
synovial fluid electrolyte. Metals. 2023;13:1331. doi:10.3390/
met13081331
44. Madl AK, Liong M, Kovochich M, et al. Toxicology of wear particles of
cobalt-chromium alloy metal-on-metal hip implants part I: physico-
chemical properties in patient and simulator studies. Nanomedicine:
nanotechnology. Biol Med. 2015;11:1201-1215. doi:10.1016/j.nano.
2014.12.005
45. Mazzucco D, Scott R, Spector M. Composition of joint fluid in
patients undergoing total knee replacement and revision arthroplasty:
correlation with flow properties. Biomaterials. 2004;25:4433-4445.
doi:10.1016/j.biomaterials.2003.11.023
46. Ge SX, Jung D, Yao R. ShinyGO: a graphical gene-set enrichment tool
for animals and plants. Bioinformatics. 2020;36(8):2628-2629. doi:10.
1093/bioinformatics/btz931
47. Huang G, Chen J. Preparation and applications of hyaluronic acid and
its derivatives. Int J Biol Macromol. 2019;125:478-484. doi:10.1016/j.
ijbiomac.2018.12.074
48. Essendoubi M, Gobinet C, Reynaud R, Angiboust JF, Manfait M,
Piot O. Human skin penetration of hyaluronic acid of different molec-
ular weights as probed by Raman spectroscopy. Skin Res Technol.
2016;22:55-62. doi:10.1111/srt.12228
49. Posada OM, Tate RJ, Grant MH. Effects of CoCr metal wear debris
generated from metal-on-metal hip implants and Co ions on human
monocyte-like U937 cells. Toxicol in Vitro. 2015;29:271-280. doi:10.
1016/j.tiv.2014.11.006
50. Samelko L, Caicedo M, McAllister K, Jacobs J, Hallab NJ. Metal-
induced delayed type hypersensitivity responses potentiate particle
induced osteolysis in a sex and age dependent manner. PLoS One.
2021;16:e0251885. doi:10.1371/journal.pone.0251885
51. Begum H, Ahmed MS, Cho S, Jeon S. Simultaneous reduction and
nitrogen functionalization of graphene oxide using lemon for metal-
free oxygen reduction reaction. J Power Sources. 2017;372:116-124.
doi:10.1016/j.jpowsour.2017.10.035
52. Belessi V, Petridis D, Steriotis T, et al. Simultaneous reduction and
surface functionalization of graphene oxide for highly conductive
and water dispersible graphene derivatives. SN Appl Sci. 2019;1:77.
doi:10.1007/s42452-018-0077-9
53. Li W, Tang X-Z, Zhang H-B, et al. Simultaneous surface
functionalization and reduction of graphene oxide with octadecyla-
mine for electrically conductive polystyrene composites. Carbon.
2011;49:4724-4730. doi:10.1016/j.carbon.2011.06.077
54. Lin P, Meng L, Huang Y, Liu L, Fan D. Simultaneously functionalization
and reduction of graphene oxide containing isocyanate groups. Appl
Surf Sci. 2015;324:784-790. doi:10.1016/j.apsusc.2014.11.038
55. Su Z, Wang H, Tian K, Xu F, Huang W, Tian X. Simultaneous reduc-
tion and surface functionalization of graphene oxide with wrinkled
structure by diethylenetriamine (DETA) and their reinforcing effects
in the flexible poly (2-ethylhexyl acrylate) (P2EHA) films. Compos A:
Appl Sci Manuf. 2016;84:64-75. doi:10.1016/j.compositesa.2015.
11.033
56. Wang J, Zhang K, Hao S, et al. Simultaneous reduction and surface
functionalization of graphene oxide by cystamine dihydrochloride for
rubber composites. Composites Part A: Appl Sci Manufacturing. 2019;
122:18-26. doi:10.1016/j.compositesa.2019.04.018
57. Wang J, Zhang K, Hao S, Xia H, Lavorgna M. Simultaneous reduction
and surface functionalization of graphene oxide and the application
for rubber composites. J Appl Polym Sci. 2019;136:47375. doi:10.
1002/app.47375
58. Sim HJ, Li Z, Xiao P, Lu H. The influence of lateral size and oxidation
of graphene oxide on its chemical reduction and electrical conductiv-
ity of reduced graphene oxide. Molecules. 2022;27:7840. doi:10.
3390/molecules27227840
59. Maslekar N, Mat Noor RA, Kuchel RP, Yao Y, Zetterlund PB,
Agarwal V. Synthesis of diamine functionalised graphene oxide and
its application in the fabrication of electrically conducting reduced
graphene oxide/polymer nanocomposite films. Nanoscale Adv. 2020;
2:4702-4712. doi:10.1039/D0NA00534G
60. Lai L, Chen L, Zhan D, et al. One-step synthesis of NH2-graphene
from in situ graphene-oxide reduction and its improved electrochemi-
cal properties. Carbon. 2011;49:3250-3257. doi:10.1016/j.carbon.
2011.03.051
61. Vrettos K, Karouta N, Loginos P, Donthula S, Gournis D,
Georgakilas V. The role of diamines in the formation of graphene
aerogels. Front Mater. 2018;5:20. doi:10.3389/fmats.2018.00020
62. Stankovi
cV,Đurđi
c S, Ognjanovi
c M, et al. Anti-human albumin
monoclonal antibody immobilized on EDC-NHS functionalized car-
boxylic graphene/AuNPs composite as promising electrochemical
HSA immunosensor. J Electroanal Chem. 2020;860:113928. doi:10.
1016/j.jelechem.2020.113928
63. Liu T, Rui S, Li S. Layer-by-layer self-assembly composite coatings for
improved corrosion and wear resistance of Mg alloy for biomedical
applications. Coatings. 2021;11:515. doi:10.3390/coatings11050515
64. Malhotra R, Han YM, Morin JLP, et al. Inhibiting corrosion of
biomedical-grade Ti-6Al-4V alloys with graphene nanocoating. J Dent
Res. 2020;99:285-292. doi:10.1177/0022034519897003
65. Zhang W, Lee S, McNear KL, et al. Use of graphene as protection film
in biological environments. Sci Rep. 2014;4:4097. doi:10.1038/
srep04097
66. Zhao S, Liu X, Xu Z, et al. Graphene encapsulated copper microwires
as highly MRI compatible neural electrodes. Nano Lett. 2016;16:
7731-7738. doi:10.1021/acs.nanolett.6b03829
67. Cowman MK. Hyaluronan and hyaluronan fragments. In: Baker DC,
ed. Advances in Carbohydrate Chemistry and Biochemistry. 1st ed. Aca-
demic Press, Elsevier; 2017:1-59. doi:10.1016/bs.accb.2017.10.001
68. Juranek I, Stern R, ˇ
Soltes L. Hyaluronan peroxidation is required for
normal synovial function: an hypothesis. Med Hypotheses. 2014;82:
662-666. doi:10.1016/j.mehy.2014.02.024
69. Tokita Y, Okamoto A. Hydrolytic degradation of hyaluronic acid.
Polym Degrad Stab. 1995;48:269-273. doi:10.1016/0141-3910(95)
00041-J
18 S´
ANCHEZ-LÓPEZ ET AL.
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
70. Tokita Y, Okamoto A. Degradation of hyaluronic acidkinetic study
and thermodynamics. Eur Polym J. 1996;32:1011-1014. doi:10.1016/
0014-3057(96)00019-5
71. França CG, Sacomani DP, Villalva DG, Nascimento VF, Dávila JL,
Santana MHA. Structural changes and crosslinking modulated func-
tional properties of oxi-HA/ADH hydrogels useful for regenerative
purposes. Eur Polym J. 2019;121:109288. doi:10.1016/j.eurpolymj.
2019.109288
72. Zhang L, Xiao Y, Jiang B, Fan H, Zhang X. Effect of adipic dihydrazide
modification on the performance of collagen/hyaluronic acid scaffold.
J Biomed Mater Res B Appl Biomater. 2010;92:307-316. doi:10.1002/
jbm.b.31516
73. Rogers TH, Babensee JE. The role of integrins in the recognition and
response of dendritic cells to biomaterials. Biomaterials. 2011;32:
1270-1279. doi:10.1016/j.biomaterials.2010.10.014
74. Case LB, Waterman CM. Integration of actin dynamics and cell adhe-
sion by a three-dimensional, mechanosensitive molecular clutch. Nat
Cell Biol. 2015;17:955-963. doi:10.1038/ncb3191
75. Geiger B, Bershadsky A, Pankov R, Yamada KM. Transmembrane
crosstalk between the extracellular matrix and the cytoskeleton. Nat
Rev Mol Cell Biol. 2001;2:793-805. doi:10.1038/35099066
76. Zamir E, Geiger B. Components of cell-matrix adhesions. J Cell Sci.
2001;114:3577-3579. doi:10.1242/jcs.114.20.3577
77. Geiger B, Spatz JP, Bershadsky AD. Environmental sensing through
focal adhesions. Nat Rev Mol Cell Biol. 2009;10:21-33. doi:10.1038/
nrm2593
78. Mukherjee A, Melamed S, Damouny-Khoury H, et al. α-Catenin links
integrin adhesions to F-actin to regulate ECM mechanosensing and
rigidity dependence. J Cell Biol. 2022;221:e202102121. doi:10.1083/
jcb.202102121
79. Chopra A, Murray ME, Byfield FJ, et al. Augmentation of integrin-
mediated mechanotransduction by hyaluronic acid. Biomaterials.
2014;35:71-82. doi:10.1016/j.biomaterials.2013.09.066
80. Chico B, Pérez-Maceda B, San-José S, Escudero M, García-Alonso M,
Lozano R. Wettability, corrosion resistance, and osteoblast response
to reduced graphene oxide on CoCr functionalized with hyaluronic
acid. Materials. 2022;15:2693. doi:10.3390/Fma15072693
81. Cai S, Wu C, Yang W, Liang W, Yu H, Liu L. Recent advance in surface
modification for regulating cell adhesion and behaviors. Nanotechnol
Rev. 2020;9:971-989. doi:10.1515/ntrev-2020-0076
82. Bryhan MD, Gagnon PE, LaChance OV, et al. Method for creating a
cell growth surface on a polymeric substrate. 2003 U.S Patent
No. US6,617,152 B2.
83. Van Kooten TG, Spijker HT, Busscher HJ. Plasma-treated polystyrene
surfaces: model surfaces for studying cellbiomaterial interactions.
Biomaterials. 2004;25:1735-1747. doi:10.1016/j.biomaterials.2003.
08.071
84. Wolcott CC. Colloidal silica films for cell culture. 1998 U.S Patent
No. 5,814,550.
85. Sanyal S, Saxena D, Xiuqui Qian S, et al. Defined cell culturing surfaces
and methods of use. 2015 U.S Patent No US2015/0024494 A1.
86. Bouwmeester T, Bauch A, Ruffner H, et al. A physical and functional
map of the human TNF-α/NF-κB signal transduction pathway. Nat
Cell Biol. 2004;6:97-105. doi:10.1038/ncb1086
87. Chen G, Cao P, Goeddel DV. TNF-induced recruitment and activation
of the IKK complex require Cdc37 and Hsp90. Mol Cell. 2002;9:401-
410. doi:10.1016/s1097-2765(02)00450-1
88. Martinez FO, Gordon S. The M1 and M2 paradigm of macrophage
activation: time for reassessment. F1000Prime Reports. 2014;6:13.
doi:10.12703/P6-13
89. Yunna C, Mengrua H, Leia W, et al. Macrophage M1/M2 polarization.
European J Pharma. 2020;877:173090. doi:10.1016/j.ejphar.2020.
173090
90. Murray PJ, Allen JE, Biswas SK, et al. Macrophage activation and
polarization: nomenclature and experimental guidelines. Immunity.
2014;41:14-20. doi:10.1016/j.immuni.2014.06
91. Ivashkiv LB. Epigenetic regulation of macrophage polarization and
function. Trends Immunol. 2013;34:216-223. doi:10.1016/j.it.2012.
11.001
92. Escudero ML, Llorente I, Pérez-Maceda BT, et al. Electrochemically
reduced graphene oxide on CoCr biomedical alloy: characterization,
macrophage biocompatibility and hemocompatibility in rats with gra-
phene and graphene oxide. Mater Sci Eng C Mater Biol Appl. 2020;
109:110522. doi:10.1016/j.msec.2019.110522
93. Chico B, Pérez-Maceda BT, San José S, Escudero ML,
García-Alonso MC, Lozano RM. Corrosion behaviour and J774A.1
macrophage response to hyaluronic acid functionalization of electro-
chemically reduced graphene oxide on biomedical grade CoCr. Metals.
2021;11:1078. doi:10.3390/met11071078
SUPPORTING INFORMATION
Additional supporting information can be found online in the Support-
ing Information section at the end of this article.
How to cite this article: Sánchez-López L, Chico B,
García-Alonso MC, Lozano RM. Macrophage proteomic
analysis of covalent immobilization of hyaluronic acid and
graphene oxide on CoCr alloy in a tribocorrosive environment.
J Biomed Mater Res. 2024;119. doi:10.1002/jbm.a.37751
S´
ANCHEZ-LÓPEZ ET AL.19
15524965, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/jbm.a.37751 by Readcube (Labtiva Inc.), Wiley Online Library on [02/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
In this work, impermeable and ultrathin surface nanomodifications for joint applications based on graphene oxide (GO) are assembled on CoCr surfaces via covalent immobilization between GO nanosheets and silane monolayers. Two silane curing temperatures, 45 • C for 24 h and 75 • C for 30 min, on CoCr surfaces and two incubation times for GO suspension, 12 h and 24 h, on silanized CoCr surfaces are prepared. Electrochemical characterization is performed using electrochemical impedance spectroscopy (EIS) in a 3 g/L hyaluronic acid solution. Results show that GO nanosheets immobilized with silane covalent bonding confer impermeability of sp2 networks on GO and strong interfacial adhesion of GO sheets anchored to silanized CoCr via organosilane chemistry, which prevents the permeation of oxidant species at the metal interface. At short GO incubation times (12 h), the Rs values decrease with the immersion time, indicating that small species, such as metal ions, are able to diffuse through the interlayer gaps of nanolayers. Longer GO incubation times (24 h) favor the formation of bonds between the GO and the silane, thus slowing downdiffusion and metal ion release into the medium. EIS data confirm the impermeability of GO nanocoatings with lengthening GO incubation time for medical application of metallic implants.
Article
Full-text available
In recent years, the development of hyaluronic acid or hyaluronan (HA) based scaffolds, medical devices, bioconjugate systems have expanded into a broad range of research and clinical applications. Research findings over the last two decades suggest that the abundance of HA in most mammalian tissues with distinctive biological roles and chemical simplicity for modifications have made it an attractive material with a rapidly growing global market. Besides its use as native forms, HA has received much interest on so‐called “HA‐bioconjugates” and “modified‐HA systems”. In this review, the importance of chemical modifications of HA, underlying rationale approaches, and various advancements of bioconjugate derivatives with their potential physicochemical, and pharmacological advantages are summarized. This review also highlights the current and emerging HA‐based conjugates of small molecules, macromolecules, crosslinked systems, and surface coating strategies with their biological implications, including their potentials and key challenges discussed in detail.
Article
Full-text available
The presence of a worn surface in the implanted material, as in the case of a replacement of a damaged osteoarticular joint, is the normal condition after implantation. This manuscript focuses precisely on the comparative study of the cellular behavior on worn CoCr surfaces, analyzing the effect of different surface modifications on macrophages’ responses. CoCr surfaces were modified by the deposition of electrochemically reduced graphene oxide (CoCrErGO), followed by additional surface functionalization with hyaluronic acid (CoCrErGOHA). After the wear corrosion processes, the macrophage response was studied. In addition, macrophage supernatants exposed to the surfaces, before and after wear, were also evaluated for osteoblast response through the analysis of the metabolic activity, plasma membrane damage, and phosphatase alkaline activity (ALP). The proteomic analysis and the quantitative TNF-α/IL-10 ratios of the J774A.1 macrophages exposed to the surfaces under study showed a polarization shift from M0 (basal state) to M1, associated with the pro-inflammatory response of all surfaces. A lower M1 polarization was observed upon exposure to the surface modification with ErGO, whereas posterior HA functionalization attenuated, even more, the M1 polarization. The wear corrosion process contributed to inflammation and exacerbated the M1 polarization response on macrophages to CoCr, which was diminished for the ErGO and attenuated the most for the ErGOHA surfaces. Comparative proteomics showed that the pathways related to M1 polarization were downregulated on the surfaces of CoCrErGOHA, which suggests mechanisms for the observed attenuation of M1 polarization. The suitable immuno-modulatory potential induced by the ErGOHA surface, with and without wear, together with the stimulation of ALP activity in osteoblasts induced by macrophage supernatants, promotes the mineralization processes necessary for bone repair. This makes it feasible to consider the adsorption of ErGOHA on CoCr as a recommended surface treatment for the use of biomaterials in osseous joint applications.
Article
Full-text available
Periprosthetic osteolysis remains a leading complication of total hip and knee arthroplasty, often resulting in aseptic loosening of the implant and necessitating revision surgery. Wear-induced particulate debris is the main cause initiating this destructive process. The purpose of this article is to review recent advances in understanding of how wear debris causes osteolysis, and emergent strategies for the avoidance and treatment of this disease. A strong activator of the peri-implant innate immune this debris-induced inflammatory cascade is dictated by macrophage secretion of TNF-α, IL-1, IL-6, and IL-8, and PGE2, leading to peri-implant bone resorption through activation of osteoclasts and inhibition of osteoblasts through several mechanisms, including the RANK/RANKL/OPG pathway. Therapeutic agents against proinflammatory mediators, such as those targeting tumor necrosis factor (TNF), osteoclasts, and sclerostin, have shown promise in reducing peri-implant osteolysis in vitro and in vivo; however, radiographic changes and clinical diagnosis often lag considerably behind the initiation of osteolysis, making timely treatment difficult. Considerable efforts are underway to develop such diagnostic tools, therapies, and identify novel targets for therapeutic intervention.
Article
Full-text available
The chemical reduction efficiencies of graphene oxide (GO) are critically important in achieving graphene-like properties in reduced graphene oxide (rGO). In this study, we assessed GO lateral size and its degree of oxidation effect on its chemical reduction efficiency in both suspension and film and the electrical conductivity of the corresponding rGO films. We show that while GO-reduction efficiency increases with the GO size of lower oxidation in suspension, the trend is opposite for film. FESEM, XRD, and Raman analyses reveal that the GO reduction efficiency in film is affected not only by GO size and degree of oxidation but also by its interlayer spacing (restacking) and the efficiency is tunable based on the use of mixed GO. Moreover, we show that the electrical conductivity of rGO films depends linearly on the C/O and Raman ID/IG ratio of rGO and not the lateral size of GO. In this study, an optimal chemical reduction was achieved using premixed large and small GO (L/SGO) at a ratio of 3:1 (w/w). Consequently, the highest electrical conductivity of 85,283 S/m was achieved out of all rGO films reported so far. We hope that our findings may help to pave the way for a simple and scalable method to fabricate tunable, electrically conductive rGO films for electronic applications.
Article
Full-text available
Enzyme immobilization has been developing since the 1960s and although many industrial biocatalytic processes use the technology to improve enzyme performance, still today we are far from full exploitation of the field. One clear reason is that many evaluate immobilization based on only a few experiments that are not always well-designed. In contrast to many other reviews on the subject, here we highlight the pitfalls of using incorrectly designed immobilization protocols and explain why in many cases sub-optimal results are obtained. We also describe solutions to overcome these challenges and come to the conclusion that recent developments in material science, bioprocess engineering and protein science continue to open new opportunities for the future. In this way, enzyme immobilization, far from being a mature discipline, remains as a subject of high interest and where intense research is still necessary to take full advantage of the possibilities.
Article
Full-text available
Both cell–cell and cell–matrix adhesions are regulated by mechanical signals, but the mechanobiological processes that mediate the cross talk between these structures are poorly understood. Here we show that α-catenin, a mechanosensitive protein that is classically linked with cadherin-based adhesions, associates with and regulates integrin adhesions. α-Catenin is recruited to the edges of mesenchymal cells, where it interacts with F-actin. This is followed by mutual retrograde flow of α-catenin and F-actin from the cell edge, during which α-catenin interacts with vinculin within integrin adhesions. This interaction affects adhesion maturation, stress-fiber assembly, and force transmission to the matrix. In epithelial cells, α-catenin is present in cell–cell adhesions and absent from cell–matrix adhesions. However, when these cells undergo epithelial-to-mesenchymal transition, α-catenin transitions to the cell edge, where it facilitates proper mechanosensing. This is highlighted by the ability of α-catenin–depleted cells to grow on soft matrices. These results suggest a dual role of α-catenin in mechanosensing, through both cell–cell and cell–matrix adhesions.
Article
Full-text available
Improvements in durable lubrication together with minimized wear are essential for obtaining long-term, functioning metallic joint prostheses. To achieve this objective, CoCr surface was functionalized with Graphene Oxide (GO) and characterized by FTIR and XPS. CoCr alloy was subjected to alkalinization in order to generate high hydroxyl content on the surface. FTIR and XPS revealed reactive OH groups, enabling intermediate coupling via (3-aminopropyl) triethoxysilane (APTES), which was cured at 45 °C and 75 °C for 24 h and 30 min, respectively. FTIR revealed cross-linked films (Si–O–Si), inferring condensation and self-assembly of silane layers, while XPS revealed the presence of NH2, enabling chemical binding of GO. Silane-coated CoCr disks were immersed in GO solution at 60 °C for 12h and 24h, respectively. FTIR displayed CC band confirming the assembly of GO on silane-coated CoCr surfaces. XPS revealed three possible surface mechanisms: (1) reaction between primary amines of APTES and epoxy groups of GO; (2) free –OH groups in APTES and carboxyl groups in GO; and (3) reaction between APTES primary amines and –OH from carboxyl groups of GO. Overall, the multilayer system CoCr–OH–Si45-GO24h showed covalent functionalization of metal substrate with GO to a large extension of surface area among all the multilayer systems studied.
Article
Surface modification techniques are crucial in biomedical applications as they determine the direct response to the biological environment against the interfacing materials. Physical and chemical modification techniques have been extensively studied, with the latter being more commonly used due to the need for molecularly thin layers. Achieving uniform and conformal surface coverage on target geometries is critical to optimizing sensor and biomedical applications. However, achieving molecularly thin layers is practically challenging, and thick layers can alter the original properties of the bulk material. Furthermore, delamination of coated layers in humid or aqueous environments is also a concern, which can be prevented by covalent bonding of the functional groups on the substrate or incorporating appropriate functional groups or charges for solid adhesion. In this review, we provide an overview of the consolidated techniques for surface modification of materials for biomedical applications, including protein immobilization, chemical grafting, thin film coating, and plasma treatments. It also discusses the most frequently used surface modification techniques and their applications in the field. Overall, optimizing surface engineering for each case is crucial, even if the method is the same, to achieve a uniform and conformal surface coverage on target geometries for various biomedical devices, sensors, and implants.