Abstract
Protein adsorption at the surface affects the material biocompatibility directly as it is the first reaction that happens when a foreign material comes in contact with blood. In this study, the mechanism of albumin adsorption on hydrophilic and hydrophobic surfaces is investigated. Although it is studied extensively and has been of keen interest for decades, the adsorptive nature of albumin is still not fully understood with contradicting reported studies. This problem results from previous works focusing on mostly qualitative and quantitative adsorption properties of albumin, rather than the specific interaction mechanisms. The variable local surface properties across albumin can significantly impact adsorption and must be explored. In this work, the effect of hydration is found to significantly increase adsorption with minor reductions. The adsorption of albumin on hydrophilic or hydrophobic surfaces is dependent on albumin orientation, which is dictated by local charge effects. Based on these findings, an optimized material surface is proposed to minimize albumin adsorption using functional groups to limit surface availability for hydrophobic interactions while inhibiting excess electrostatic effects at hydrophilic sites. The extent of albumin adsorption and shape change are characterized herein using the heat capacity. Current study identifies interaction mechanisms previously missing in literature, which are responsible for inconsistent adsorption results.
Keywords: Hemocompatibility, Albumin Adsorption, Hydrophobic, Hydrophilic, Heat capacity
Graphical Abstract

1. Introduction
Hemocompatibility, a key factor in the design of implantable and other biomedical devices, is a thrombotic response to a surface in contact with blood [1]. It pertains to a wide range of materials suitable for bioapplication such as needles for vascular graft and heart valves [2], load bearing for spinal disks [3], membranes for blood filtration [4], etc. Generally, the mechanism of host reaction to a biomaterial is very complicated and involves many different processes such as protein adsorption and deformation, platelets adhesion and complement activation, and provisional matrix formation. When a biomaterial is introduced to blood, depending on the protein surface properties, instantaneous protein biofouling may occur and trigger an inherent biological response [1] to a foreign body invasion, which include a series of reactions such as coagulation and platelet adhesion. Surface modification and blending specific compounds with the base material are known methods to improve hemocompatibility and several efforts have been made to utilize them for improving biocompatibility of materials [4–11]. Plasma protein adsorption is the first and fastest reaction that occurs and results in a set of subsequent responses [12–14]. Human serum albumin (HSA) is an important protein in blood with various important functionalities [15–18] and forms 60% of whole blood proteins. Therefore, understanding of its interaction with surfaces provides good insight into the material’s hemocompatibility. Generally, fibrinogen and other proteins mediate the platelet adhesion to surfaces while HSA is considered inert [14] because its adhesion to a surface could prevent platelet adhesion [19].
Hydrophilicity and hydrophobicity of materials are the key factors known to impact protein-protein interaction [20] and protein adsorption on the surface which can be tuned by surface modification [21–24]. Many studies suggest that hydrophilic surfaces and compounds in the base material through blending [25] or addition [26–29] reduce HSA adsorption while hydrophobic surfaces increase it. Jesus et al. [30] studied the adsorption of HSA on surfaces of hydrophobic and hydrophilic Ti6A14V powder and reported lower protein adsorption on hydrophilic surfaces. Oss et al. [31] investigated the interaction between proteins and inorganic oxides and mentioned that silica has a low HSA adsorption due to its hydrophilicity and a negatively charged surface measured by microelectrophoresis [32] that repels HSA. Mücksch and Urbassek [33] studied the effect of surface hydrophobicity on bovine serum albumin (BSA), which has similar surface properties to HSA [34], and indicated that BSA adsorbs to the surface, partially unfolds, and experiences denaturation.
On the contrary, some studies suggest that hydrophobic surfaces have lower amounts of adsorption. Rupert et al. [35] studied the BSA adsorption on different surfaces and showed that at the low BSA concentration, 0.1 mg mL−1, the most hydrophobic surface has the lowest amount of BSA adsorption on the surface. However, at a high BSA concentration of 10 mg mL−1, which is closer to that of HSA in blood i.e. 35–50 mg mL−1, a slightly hydrophilic surface results in minimum adsorption. Supported by simulation, Jeyachandran et al. [36] showed experimentally that the interaction forces between BSA and hydrophilic surfaces are stronger than that of the hydrophobic one and ~95% of the hydrophilic surface is covered by BSA, whereas it is ~50% for the hydrophobic one. Absolom et al. [37] reported that siliconized glass made by treating glass surface with silicon oil according to the Neumann and Renzow method [38] has slightly higher hydrophobicity than Teflon and results in lower protein adsorption. It is hypothesized that it is attributed to the “screening” phenomenon where the adsorbed protein on surface changes conformationally, effectively screening new proteins. Sousa et al. [39] reported the same behavior for hydrophobic TiO2 sp versus hydrophilic TiO2 cp where it shows a lower work of adhesion and easier exchange of adsorbed HSA with solution. Finally, some studies mention that the adsorption happens on both surfaces regardless of the surface type, such as a study by Malgoratza et al. [40].
Despite the extensive studies on the effect of surface hydrophilicity and hydrophobicity on HSA adsorption, the reported results are inconsistent and the underlying mechanism of this process is not clearly known. It is pertinent to mention that HSA is a large protein with various folded hydrophobic sites inside and charged hydrophilic sites on surface [41,42] that interact with a surface, resulting in different adsorption properties depending on the direction in which it approaches the surface. However, the focus of prior studies have been on adsorption properties of HSA regardless of its orientation and local surface properties. In this study, the local surface properties of HSA at the molecular level are considered and it is shown how hydration and surface charge affect its adsorption. Moreover, the effect of different orientations of HSA when approaching various surfaces are evaluated. In each case, the binding energy between the protein and surface is calculated and the heat capacity variation [43] is used to evaluate its shape change and extent of adsorption. This work provides a better understanding of blood protein interaction with various surfaces such as GO biocompatible membranes, which has precise molecular cut-off and 40 times higher permeability relative to high flux dialysis membranes [44].
2. Setup and Methodology
2.1. Simulation box
Simulations are performed by Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) [45]. In each simulation there is one HSA molecule, water molecules, and a graphene or graphene oxide sheet. In each case, a set of simulations are performed in six directions for albumin to evaluate its interaction with different surfaces from all facets it approaches them. These directions are the original position in the data file, +90X, −90X, +180X, −90Y, and +90Y where the sign shows the rotation direction using right hand rule and the letter shows the axis of rotation. Graphene and graphene oxide (GO) sheets are used as hydrophobic and hydrophilic surfaces, respectively, and positioned in the bottom of the box i.e. −z direction. Three different levels of graphene oxide is used to investigate the effect of oxidation on the results. For all simulations, periodic boundaries in x and y-directions, fix boundary in −z, and fix-wall command in the +z-direction are applied. The fixed boundary and fix-wall command in the +z direction allows HSA to interact with the given surface only inside the simulation box and not through the boundary. Length (Lx), width (Ly), and height (Lz) of the simulation box are 10 nm, which is large enough to avoid any self-interaction of HSA through periodic boundaries. Figure 1 shows a simulation box with HSA surrounded by water molecules above a graphene sheet.
Figure 1.

Human serum albumin molecule in original direction (base case), water molecules, and graphene as a hydrophobic surface in simulations (as seen in VMD software).
2.2. Ensembles and Simulation Parameters
Graphene and GO sheets are fixed and NVT ensemble is used for 0.1 ns to equilibrate water molecules and HSA. NPT ensemble is used for 10 ns to run simulations with a time step of 1 fs. Unless otherwise stated, Nose-Hoover thermostat is used to control the temperature at 310 K, which is equal to the temperature of blood or 37 °C. Real units, periodic boundaries, bond style of harmonic, angle and dihedral styles of hybrid consisting of charm and harmonic, and improper style of harmonic are used. Cut-off of 12 Å is applied for both Lenard Jones and columbic terms. To model the non-bonded interatomic interactions, the Lenard Jones (LJ) model of 6–12 is used as follows [46]:
where φij is the LJ potential, εij is the potential well depth, σij is the equilibrium distance between atoms, and rij is distance between atoms.
2.3. Data Files and Models
The SPC/E water model explained by Berendsen et al. [47] with an initial density of 1 g/cm3 is used in the simulations. The graphene and GO data files were made according to the existing structures and models [48–57]. The graphene sheet is a 2D hexagonal lattice of carbon atoms with a bond length of 1.4 Å. Graphene and GO layers and functional groups are fixed in all simulations and no interaction is considered between the atoms in a sheet. The hydrophilic groups on GO sheets are on the edges and defects where carbon has free electrons [58]. Thus, to make the GO sheets from graphene, an epoxy or hydroxyl group is introduced to each carbon atom at defects in the form of a dual or single bond, respectively. Three levels of oxidations are used in simulation by removing every 3, 6, and 9 carbon atoms to make a defect.
The water model and a schematic of G and GO structures with the functional groups on GO are shown in Figure 2. To generate different levels of graphene oxide one carbon atom is removed every three, six, and nine atoms, which corresponds to high, normal, and low oxidation levels respectively, and is replaced by functional groups. Among various interatomic potentials between carbon and water molecules, the one with simulation results that best matched the experimental properties, such as contact angle, was selected for Carbon-Water interactions [59].
Figure 2.

Water model (SPC/E), graphene, and functional groups.
In this model, σC-O = 3.19 Å and εC-O = 0.0937 kcal/mol. The parameters for water SPC/E [47] model are used to account for Water-Water interactions. The interaction parameters between atoms within the functional groups are adopted from a study by Jiao et al. [60]. The bond length and angle coefficients are adopted from Levitt et al. [61]. These parameters are summarized in Table 1. For those interactions that are not explicitly set, the mix arithmetic pair modify function is used for LJ parameters as and [62]. This method was successfully used previously [63]. HSA data and topology files were downloaded from Protein Data Bank (RCSB PDB) [64] and used in VMD [65] to generate a LAMMPS data file. Charmm36 force-field [66,67] along with cmap36 file necessary for proteins [68,69] are adopted.
Table 1.
Interaction parameters. σC-O and εC-O are defined separately.
| Molecule | Atom | q (e) | σ (Å) | ε (kcal/mol) | b0 (Å) | kb (kcal/mol) | θ (°) | Kθ (kcal/mol) |
|---|---|---|---|---|---|---|---|---|
| Graphene | C | 0.0 | 3.83 | 0.0551 | C-C, 1.4 | C-C, 250 | C-C-C, 120 | C-C-C, 50 |
| Epoxy | C | +0.2 | 3.4 | 0.0693 | C-O, 1.41 | C-O, 450 | C-C-O, 90 | C-C-O, 50 |
| O | −0.4 | 2.9 | 0.142 | |||||
| Hydroxyl | C | +0.1966 | 3.4 | 0.0693 | C-O, 1.41 | C-O, 450 | C-C-O, 90 | C-C-O, 50 |
| O | −0.5260 | 3.166 | 0.14 | O-H, 0.945 | O-H, 450 | C-O-H, 110 | C-O-H, 50 | |
| H | +0.3294 | 0.0 | 0.0 | |||||
| Water | O | −0.8476 | 3.166 | 0.1553 | O-H, 1.0 | O-H, 450 | H-O-H, 109.47 | H-O-H, 55 |
| H | +0.4238 | 0.0 | 0.0 |
HSA has a globular protein composed of 585 amino acids with molecular weight of 68000 Da [69]. About 7% of the blood is protein and albumin forms almost 2/3 (about 60%) of the whole blood protein volume [18,70]. HSA has a heart-like shape and is about 80 Å on the side with an average thickness of about 30 Å. Its volume is 88249 Å3 in non-solvated state and increases up to 20% when solvated in water [17].
2.4. Methodology
The binding energy and heat capacity are used to evaluate interactions of the HSA with the surface and its shape change, respectively, as described below.
Binding Energy: To calculate the binding energy, two atom groups of “HSA” for albumin and “Surface” for graphene or graphene oxide are defined. Using group/group command in LAMMPS, the total energy and force interaction between these two groups are calculated.
- Heat Capacity: To calculate the heat capacity, two methods are utilized:
-
Total energy change with temperature:where q and E are heat and total energy of the system, respectively.
-
Fluctuation of potential energy [71]:where 〈 〉, E, and kb are ensemble average, total energy, and Boltzmann constant, respectively.
-
3. RESULTS AND DISCUSSION
First, the effect of HSA hydration on its adsorption to the surface was investigated and it was shown that reducing the hydration level increases the adsorption significantly. Then, the effect of surface hydrophobicity and hydrophilicity with different oxidation levels on adsorption of HSA is examined and the binding energy for each case is calculated. It is shown that the binding energy correlates with surface charge directly. Finally, the heat capacity methods, validated by calculating water heat capacity, are used to investigate the shape change of HSA.
3.1. Effect of Hydration on Albumin Adsorption
HSA adsorption on a hydrophobic surface requires dehydration between HSA and surface. Simulation of HSA on graphene as a hydrophobic surface reveals a weak binding between them, which is expected as the hydrophobic sites are inside HSA and hydrophilic sites are covered with water molecules on the surface. Results show that any reduction in the number of water molecules around HSA result in higher adsorption on the surface. Partial dehydration of HSA can be achieved by either reducing water content in the simulation box or adding ions or any other molecules to water as they result in dehydration by decreasing the number of water molecules available for HSA. This explains why higher concentrations increase adsorption on hydrophobic surfaces [34]. Results show that reducing the HSA hydration by ~10% result in ~2.5 times stronger adsorption on the surface, while the effect on water adsorption and its binding energy at the surface were negligible. Table 2 shows the ratio of binding energy among water, surface, and albumin for partial dehydration over full hydration. In full hydration the empty space inside the simulation box is full of water molecules with bulk density of 1g/cm3, while in reduced conditions the water content is decreased by 20%. The provided data are the averaged values for six different facets of HSA approaching the hydrophobic surface of graphene. The provided error bars show ~21% error for binding energy between HSA and graphene using three different results for each facet, while it is negligible for the other two.
Table 2.
Binding energy ratio of HSA on surface of graphene (partial dehydration over full hydration) and its standard error.
| Binding Surfaces | Binding Energy Ratio, Cal/mol | Standard Error |
|---|---|---|
| HSA-Water | 0.894 | 0.006 |
| HSA-Graphene | 2.501 | 0.213 |
| Graphene-Water | 0.997 | 0.002 |
The same simulations for graphene oxide reveal that the hydrophilic-hydrophilic interaction between different sites on the surface of HSA and functional groups on graphene oxide is strong enough to dehydrate the surface easily. The rest of simulations in this study are performed with reduced water content to allow for HSA adsorption as it occurs in real conditions in the presence of ions and other species in blood. The lower water numbers decrease the computational costs as well.
3.2. HSA adsorption on hydrophilic/hydrophobic surface
Binding energy between the surface and HSA molecules is used as a scale to compare the interaction forces between them in different conditions. Results show that HSA adsorbs on all hydrophobic and hydrophilic surfaces regardless of the local surface charge and oxidation type, however the binding energy magnitude is a function of the above-mentioned parameters. Figure 3 shows the results of binding energy for HSA on the hydrophilic surface of graphene oxide with an average oxidation level in all six directions.
Figure 3.

Binding energy between HSA and hydrophilic graphene oxide with average oxidation level. Different facets of HSA results in different binding energy values.
The same simulations are performed for all surfaces in six directions and repeated three times. The average value for all cases are summarized in Table 3.
Table 3.
Average values of binding energy for HSA and different surfaces
| HAS Direction\Type | G | GOLow | GOAve | GOHigh |
|---|---|---|---|---|
| +90X, kcal/mol | −32 | −4 | −27 | −115 |
| +90Y, kcal/mol | −108 | −229 | −73 | −211 |
| +180X, kcal/mol | −177 | −114 | −99 | −445 |
| Base Case, kcal/mol | −84 | −20 | 53 | 28 |
| −90X, kcal/mol | −67 | −109 | −121 | −319 |
| −90Y, kcal/mol | −96 | −72 | −150 | −78 |
| Average, kcal/mol | −94 | −91 | −69 | −190 |
According to Table 3, the binding energy starts with a moderate value at the hydrophobic surface of graphene, reduces over the low oxidized graphene, and becomes minimum over an average oxidation level. This behavior is reported by Rupert et al. where they have examined the BSA adsorption on different surfaces and have mentioned that at high concentrations, close that of HSA in blood, the adsorption on slightly hydrophilic surfaces is minimum [35]. However, the adsorption increases significantly at high oxidation level, which can be compared to a study by Jeyachandran et al. where they have reported that 95% of their hydrophilic surface was covered by BSA [36]. The strength and nature of interaction is a function of both surface and protein properties. When the surface is hydrophobic, the dehydration of surface and hydrophobic-hydrophobic interaction is the determining factor and binding energy gradually increases over time as dehydration progresses. However, in the presence of surface charges, i.e. hydrophilic surfaces of graphene oxide, electrostatic forces are dominant and the binding energy increases with oxidation rate. Figure 4 shows the net charge of the HSA surface at two extreme situations in terms of adsorption which are base case and the one rotated 180 degree around X i.e. +180X as a function of involved length of HSA starting from its closest atom to GO surface. The positive surface charge of +180X and negative charge of the base case are good indicators of the charge dominance on adsorption processes as the surface of graphene oxide has many oxygen with negatively charged electron clouds that adsorb the positive charges and repel the negative ones.
Figure 4.

Net charge of HSA surface as a function of its involved length starting from its closest atom to the hydrophilic surface of graphene oxide.
The results of binding energy indicate that a surface with a moderate level of oxidation minimizes the HSA adsorption, while both hydrophobic and highly hydrophilic surfaces increase the adsorption. The reason behind this is that the slightly oxidized graphene results in a weaker hydrophilic-hydrophilic interaction compared to highly oxidized one and at the same time it limits the availability for GO hydrophobic domains and avoids hydrophobic-hydrophobic interaction.
3.3. HSA shape change on hydrophilic/hydrophobic surface
During HSA adsorption, protein shape can change and result in its denaturation which may lead to malfunctioning of the protein. In order to evaluate shape change, two methods of heat capacity were used and they have been validated by calculating heat capacity of water prior to apply them to HSA. Figure 5 shows the results of water heat capacity simulations for both methods, which are in good agreement with experimental values.
Figure 5.

Validation of heat capacity method by calculating heat capacity of water. Both methods are in good agreement with experimental water heat capacity of 1 cal/g.K.
These methods are used for solvated HSA in water and they both show close results as shown in Figure 6. While HSA heat capacity is below that of bulk water at 1 J/mol.K, results show that it increases significantly by about 8 times in the presence of both hydrophobic graphene and hydrophilic graphene oxide surfaces. One reason behind this is that the added heat to the HSA will be adsorbed to overcome the interaction between HSA hydrophilic sites and both water and hydrophilic molecules on the surface of graphene oxide leave no energy for the albumin shape change as seen in simulations.
Figure 6.

Heat capacity of HSA in water using total energy method (on the left) and fluctuation of energy around average energy in ensemble (on the right). Results show 10% difference.
Another reason for this behavior is the protein denaturation process under different conditions, such as protein shape changes, which is dominant on the hydrophobic surface of graphene where hydrophobic-hydrophobic interactions unfold the protein. Heat capacity increase during unfolding is reported in different studies previously. Privalov et al. have shown how a protein’s heat capacity increases with temperature [43], Cooper has reported a significant increase in heat capacity of a small globular protein (lysozyme) during unfolding [72], and Myers et al. investigated 45 proteins and reported the increase in heat capacity of these proteins upon unfolding. This sudden increase in heat capacity is the same as the melting process in other materials such as water where the added heat is used to perform the phase change. The calculated heat capacity value for all cases of graphene and graphene oxide with different oxidation levels are almost the same with a maximum of 20% variation. Figure 7 shows the result of heat capacity of HSA on graphene and graphene oxide with an average oxidation level.
Figure 7.

Heat Capacity of HSA on the surface of graphene on the left and graphene oxide with average oxidation level on the right side.
4. CONCLUSION
This study sheds light on the underlying mechanism of human serum albumin adsorption and shape changes on hydrophilic and hydrophobic surfaces using molecular dynamics simulations. To evaluate the adsorption of HSA on a surface the binding energy method was used to show that the adsorption is a function of hydration degree and local surface properties of the HSA facet, such as charge, when it approaches the material surface. The protein surface properties were used to introduce an optimum level of oxidation that minimizes the average adsorption of HSA by limiting the hydrophobic surface availability and minimizing hydrophilic interaction. In addition, the shape change was investigated using heat capacity variation methods validated by calculating water heat capacity. It was shown that the heat capacity of hydrated HSA is lower than that of water, but it increases significantly in the presence of any surfaces which is an indication of high binding energy between the protein and hydrophilic surfaces and its denaturation on hydrophobic surfaces. The current study provides a good insight on the mechanism of HSA adsorption on different surfaces and reveals how surface properties of protein affect the adsorption process and how these properties can be utilized to optimize the adsorption and control the protein deformation.
Slight albumin dehydration increases the interaction energy with surface significantly.
Hydrophilic surface increases binding energy, but doesn’t deform albumin significantly.
Hydrophobic surface deforms the protein and results in protein denaturation.
Binding energy is minimum on a moderately hydrophilic surface.
Net charges of the surface and albumin are dominant factors in adsorption behavior.
ACKNOWLEDGMENTS
This work is supported in part by National Institutes of Health (1R21EB023527-01A1).
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Declaration of interests
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
DATA AVAILABILITY
The data support the findings on this study are available from the corresponding author upon reasonable request.
REFERENCES
- [1].Mulinti P, Brooks JE, Lervick B, Pullan JE, Brooks AE, 10 - Strategies to improve the hemocompatibility of biodegradable biomaterials, in: C.A.B.T.-H. of B. for C.A. Siedlecki (Ed.), Woodhead Publishing, 2018: pp. 253–278. https://doi.org/ 10.1016/B978-0-08-100497-5.00017-3. [DOI] [Google Scholar]
- [2].Schettini N, Jaroszeski MJ, West L, Saddow SE, Chapter 5 – Hemocompatibility Assessment of 3C-SiC for Cardiovascular Applications, in: S.E.B.T.-S.C.B. Saddow (Ed.), Elsevier, Oxford, 2012: pp. 153–208. https://doi.org/ 10.1016/B978-0-12-385906-8.00005-2. [DOI] [Google Scholar]
- [3].Boyan BD, Guldberg RE, Kennedy SJ, Ku DN, Schwartz Z, Load bearing biocompatible device, (2011).
- [4].Irfan M, Idris A, Overview of PES biocompatible/hemodialysis membranes: PES–blood interactions and modification techniques, Mater. Sci. Eng. C 56 (2015) 574–592. [DOI] [PubMed] [Google Scholar]
- [5].Rana D, Matsuura T, Surface modifications for antifouling membranes, Chem. Rev 110 (2010) 2448–2471. [DOI] [PubMed] [Google Scholar]
- [6].Pavlenko D, Van Geffen E, Van Steenbergen MJ, Glorieux G, Vanholder R, Gerritsen KGF, Stamatialis D, New low-flux mixed matrix membranes that offer superior removal of protein-bound toxins from human plasma, Sci. Rep 6 (2016) 34429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Li S-S, Xie Y, Xiang T, Ma L, He C, Sun S, Zhao C-S, Heparin-mimicking polyethersulfone membranes–hemocompatibility, cytocompatibility, antifouling and antibacterial properties, J. Memb. Sci 498 (2016) 135–146. [Google Scholar]
- [8].Ciolkowski M, Petersen JF, Ficker M, Janaszewska A, Christensen JB, Klajnert B, Bryszewska M, Surface modification of PAMAM dendrimer improves its biocompatibility, Nanomedicine Nanotechnology, Biol. Med 8 (2012) 815–817. https://doi.org/ 10.1016/j.nano.2012.03.009. [DOI] [PubMed] [Google Scholar]
- [9].Vuppaladadium SSR, Agarwal T, Kulanthaivel S, Mohanty B, Barik CS, Maiti TK, Pal S, Pal K, Banerjee I, Silanization improves biocompatibility of graphene oxide, Mater. Sci. Eng. C 110 (2020) 110647. https://doi.org/ 10.1016/j.msec.2020.110647. [DOI] [PubMed] [Google Scholar]
- [10].Stan MS, Badea MA, Pircalabioru GG, Chifiriuc MC, Diamandescu L, Dumitrescu I, Trica B, Lambert C, Dinischiotu A, Designing cotton fibers impregnated with photocatalytic graphene oxide/Fe, N-doped TiO2 particles as prospective industrial self-cleaning and biocompatible textiles, Mater. Sci. Eng. C 94 (2019) 318–332. https://doi.org/ 10.1016/j.msec.2018.09.046. [DOI] [PubMed] [Google Scholar]
- [11].Sen S, Konar S, Das B, Pathak A, Dhara S, Dasgupta S, DasGupta S, Inhibition of fibrillation of human serum albumin through interaction with chitosan-based biocompatible silver nanoparticles, RSC Adv. 6 (2016) 43104–43115. [Google Scholar]
- [12].Horbett TA, Principles underlying the role of adsorbed plasma proteins in blood interactions with foreign materials, Cardiovasc. Pathol 2 (1993) 137–148.25990608 [Google Scholar]
- [13].Tang L, Thevenot P, Hu W, Surface chemistry influences implant biocompatibility, Curr. Top. Med. Chem 8 (2008) 270–280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Xu L-C, Bauer JW, Siedlecki CA, Proteins, platelets, and blood coagulation at biomaterial interfaces, Colloids Surf. B. Biointerfaces 124 (2014) 49–68. 10.1016/j.colsurfb.2014.09.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [15].Malleda C, Ahalawat N, Gokara M, Subramanyam R, Molecular dynamics simulation studies of betulinic acid with human serum albumin, J. Mol. Model 18 (2012) 2589–2597. 10.1007/s00894-011-1287-x. [DOI] [PubMed] [Google Scholar]
- [16].Carballal S, Alvarez B, Turell L, Botti H, Freeman BA, Radi R, Sulfenic acid in human serum albumin, Amino Acids. 32 (2007) 543–551. 10.1007/s00726-006-0430-y. [DOI] [PubMed] [Google Scholar]
- [17].Artali R, Bombieri G, Calabi L, Del Pra A, A molecular dynamics study of human serum albumin binding sites, Farm. 60 (2005) 485–495. https://doi.org/ 10.1016/j.farmac.2005.04.010. [DOI] [PubMed] [Google Scholar]
- [18].Ghuman J, Zunszain PA, Petitpas I, Bhattacharya AA, Otagiri M, Curry S, Structural Basis of the Drug-binding Specificity of Human Serum Albumin, J. Mol. Biol 353 (2005) 38–52. https://doi.org/ 10.1016/j.jmb.2005.07.075. [DOI] [PubMed] [Google Scholar]
- [19].Brash JL, Exploiting the current paradigm of blood–material interactions for the rational design of blood-compatible materials, J. Biomater. Sci. Polym. Ed 11 (2000) 1135–1146. 10.1163/156856200744237. [DOI] [PubMed] [Google Scholar]
- [20].Feng M, Kang H, Yang Z, Luan B, Zhou R, Potential disruption of protein-protein interactions by graphene oxide, J. Chem. Phys 144 (2016) 225102. 10.1063/1.4953562. [DOI] [PubMed] [Google Scholar]
- [21].Wang H, Yu T, Zhao C, Du Q, Improvement of hydrophilicity and blood compatibility on polyethersulfone membrane by adding polyvinylpyrrolidone, Fibers Polym. 10 (2009) 1–5. [Google Scholar]
- [22].Xiang T, Xie Y, Wang R, Wu M-B, Sun S-D, Zhao C-S, Facile chemical modification of polysulfone membrane with improved hydrophilicity and blood compatibility, Mater. Lett 137 (2014) 192–195. [Google Scholar]
- [23].Kuo Z, Fang M, Wu T, Yang T, Tseng H, Chen C, Cheng C, Hydrophilic films: How hydrophilicity affects blood compatibility and cellular compatibility, Adv. Polym. Technol 37 (2018) 1635–1642. [Google Scholar]
- [24].Ikada Y, Iwata H, Horii F, Matsunaga T, Taniguchi M, Suzuki M, Taki W, Yamagata S, Yonekawa Y, Handa H, Blood compatibility of hydrophilic polymers, J. Biomed. Mater. Res 15 (1981) 697–718. [DOI] [PubMed] [Google Scholar]
- [25].Safarpour M, Vatanpour V, Khataee A, Preparation and characterization of graphene oxide/TiO2 blended PES nanofiltration membrane with improved antifouling and separation performance, Desalination. 393 (2016) 65–78. [Google Scholar]
- [26].Modi A, Verma SK, Bellare J, Graphene oxide-doping improves the biocompatibility and separation performance of polyethersulfone hollow fiber membranes for bioartificial kidney application, J. Colloid Interface Sci 514 (2018) 750–759. [DOI] [PubMed] [Google Scholar]
- [27].Kaleekkal NJ, Thanigaivelan A, Durga M, Girish R, Rana D, Soundararajan P, Mohan D, Graphene oxide nanocomposite incorporated poly (ether imide) mixed matrix membranes for in vitro evaluation of its efficacy in blood purification applications, Ind. Eng. Chem. Res 54 (2015) 7899–7913. [Google Scholar]
- [28].Wang S-D, Ma Q, Wang K, Chen H-W, Improving antibacterial activity and biocompatibility of bioinspired electrospinning silk fibroin nanofibers modified by graphene oxide, ACS Omega. 3 (2018) 406–413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [29].Ganesh BM, Isloor AM, Ismail AF, Enhanced hydrophilicity and salt rejection study of graphene oxide-polysulfone mixed matrix membrane, Desalination. 313 (2013) 199–207. [Google Scholar]
- [30].Rodríguez-Sánchez J, Gallardo-Moreno AM, Bruque JM, González-Martín ML, Adsorption of human fibrinogen and albumin onto hydrophobic and hydrophilic Ti6Al4V powder, Appl. Surf. Sci 376 (2016) 269–275. https://doi.org/ 10.1016/j.apsusc.2016.03.014. [DOI] [Google Scholar]
- [31].van Oss CJ, Wu W, Giese RF, Naim JO, Interaction between proteins and inorganic oxides — Adsorption of albumin and its desorption with a complexing agent, Colloids Surfaces B Biointerfaces. 4 (1995) 185–189. https://doi.org/ 10.1016/0927-7765(94)01170-A. [DOI] [Google Scholar]
- [32].Van Oss CJ, Fike RM, Good RJ, Reing JM, Cell microelectrophoresis simplified by the reduction and uniformization of the electroosmotic backflow, Anal. Biochem 60 (1974) 242–251. [DOI] [PubMed] [Google Scholar]
- [33].Mücksch C, Urbassek HM, Molecular Dynamics Simulation of Free and Forced BSA Adsorption on a Hydrophobic Graphite Surface, Langmuir. 27 (2011) 12938–12943. 10.1021/la201972f. [DOI] [PubMed] [Google Scholar]
- [34].Lin VJC, Koenig JL, Raman studies of bovine serum albumin, Biopolym. Orig. Res. Biomol 15 (1976) 203–218. [DOI] [PubMed] [Google Scholar]
- [35].Kargl R, Bračič M, Resnik M, Mozetič M, Bauer W, Stana Kleinschek K, Mohan T, Affinity of Serum Albumin and Fibrinogen to Cellulose, Its Hydrophobic Derivatives and Blends, Front. Chem. 7 (2019) 581. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [36].Jeyachandran YL, Mielczarski E, Rai B, Mielczarski JA, Quantitative and Qualitative Evaluation of Adsorption/Desorption of Bovine Serum Albumin on Hydrophilic and Hydrophobic Surfaces, Langmuir. 25 (2009) 11614–11620. 10.1021/la901453a. [DOI] [PubMed] [Google Scholar]
- [37].Absolom DR, Zingg W, Neumann AW, Protein adsorption to polymer particles: Role of surface properties, J. Biomed. Mater. Res 21 (1987) 161–171. 10.1002/jbm.820210202. [DOI] [PubMed] [Google Scholar]
- [38].Li D, Xie M, Neumann AW, Vapour adsorption and contact angles on hydrophobic solid surfaces, Colloid Polym. Sci 271 (1993) 573–580. [Google Scholar]
- [39].Sousa SR, Moradas-Ferreira P, Saramago B, Viseu Melo L, Barbosa MA, Human Serum Albumin Adsorption on TiO2 from Single Protein Solutions and from Plasma, Langmuir. 20 (2004) 9745–9754. 10.1021/la049158d. [DOI] [PubMed] [Google Scholar]
- [40].Ferenc M, Katir N, Milowska K, Bousmina M, Brahmi Y, Felczak A, Lisowska K, Bryszewska M, El Kadib A, Impact of mesoporous silica surface functionalization on human serum albumin interaction, cytotoxicity and antibacterial activity, Microporous Mesoporous Mater. 231 (2016) 47–56. https://doi.org/ 10.1016/j.micromeso.2016.05.012. [DOI] [Google Scholar]
- [41].Gromiha MM, Protein bioinformatics: from sequence to function, academic press, 2010. [Google Scholar]
- [42].Perutz MF, Kendrew JC, Watson HC, Structure and function of haemoglobin: II. Some relations between polypeptide chain configuration and amino acid sequence, J. Mol. Biol 13 (1965) 669–678. [Google Scholar]
- [43].Privalov PL, Makhatadze GI, Heat capacity of proteins: II. Partial molar heat capacity of the unfolded polypeptide chain of proteins: Protein unfolding effects, J. Mol. Biol 213 (1990) 385–391. https://doi.org/ 10.1016/S0022-2836(05)80198-6. [DOI] [PubMed] [Google Scholar]
- [44].Rode RP, Chung HH, Miller HN, Gaborski TR, Moghaddam S, Trilayer Interlinked Graphene Oxide Membrane for Wearable Hemodialyzer, Adv. Mater. Interfaces. n/a (2020) 2001985. https://doi.org/ 10.1002/admi.202001985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Plimpton S, Fast parallel algorithms for short-range molecular dynamics, J. Comput. Phys 117 (1995) 1–19. [Google Scholar]
- [46].Shit SP, Ghosh NK, Pal S, Thermal conductivity of water base nanofluids containing loaded graphene nanosheets using molecular dynamics simulation, AIP Conf. Proc 2220 (2020) 20004. 10.1063/5.0001590. [DOI] [Google Scholar]
- [47].Berendsen HJC, Grigera JR, Straatsma TP, The missing term in effective pair potentials, J. Phys. Chem 91 (1987) 6269–6271. 10.1021/j100308a038. [DOI] [Google Scholar]
- [48].Mkhoyan KA, Contryman AW, Silcox J, Stewart DA, Eda G, Mattevi C, Miller S, Chhowalla M, Atomic and electronic structure of graphene-oxide, Nano Lett. 9 (2009) 1058–1063. [DOI] [PubMed] [Google Scholar]
- [49].Navaee A, Salimi A, Efficient amine functionalization of graphene oxide through the Bucherer reaction: an extraordinary metal-free electrocatalyst for the oxygen reduction reaction, Rsc Adv. 5 (2015) 59874–59880. [Google Scholar]
- [50].Xu K, Ji X, Chen C, Wan H, Miao L, Jiang J, Electrochemical double layer near polar reduced graphene oxide electrode: Insights from molecular dynamic study, Electrochim. Acta 166 (2015) 142–149. [Google Scholar]
- [51].Nakajima T, Matsuo Y, Formation process and structure of graphite oxide, Carbon N. Y 32 (1994) 469–475. [Google Scholar]
- [52].Pendolino F, Armata N, Synthesis, characterization and models of Graphene oxide, in: Graphene Oxide Environ. Remediat. Process, Springer, 2017: pp. 5–21. [Google Scholar]
- [53].Krishnamoorthy K, Veerapandian M, Yun K, Kim S-J, The chemical and structural analysis of graphene oxide with different degrees of oxidation, Carbon N. Y 53 (2013) 38–49. [Google Scholar]
- [54].Stauffer D, Dragneva N, Floriano WB, Mawhinney RC, Fanchini G, French S, Rubel O, An atomic charge model for graphene oxide for exploring its bioadhesive properties in explicit water, J. Chem. Phys 141 (2014) 44705. 10.1063/1.4890503. [DOI] [PubMed] [Google Scholar]
- [55].Dabhi SD, Gupta SD, Jha PK, Structural, electronic, mechanical, and dynamical properties of graphene oxides: A first principles study, J. Appl. Phys 115 (2014) 203517. 10.1063/1.4878938. [DOI] [Google Scholar]
- [56].Huang L, Seredych M, Bandosz TJ, van Duin ACT, Lu X, Gubbins KE, Controllable atomistic graphene oxide model and its application in hydrogen sulfide removal, J. Chem. Phys 139 (2013) 194707. 10.1063/1.4832039. [DOI] [PubMed] [Google Scholar]
- [57].Anees P, Valsakumar MC, Panigrahi BK, High temperature phonon dispersion in graphene using classical molecular dynamics, AIP Conf. Proc 1591 (2014) 1070–1071. 10.1063/1.4872856. [DOI] [Google Scholar]
- [58].Huang H, Song Z, Wei N, Shi L, Mao Y, Ying Y, Sun L, Xu Z, Peng X, Ultrafast viscous water flow through nanostrand-channelled graphene oxide membranes., Nat. Commun 4 (2013) 2979. 10.1038/ncomms3979. [DOI] [PubMed] [Google Scholar]
- [59].Werder T, Walther JH, Jaffe RL, Halicioglu T, Koumoutsakos P, Field M, E. Corporation, W.F.A. V, V. Sunny, On the Water−Carbon Interaction for Use in Molecular Dynamics Simulations of Graphite and Carbon Nanotubes, J. Phys. Chem. B 107 (2003) 1345–1352. 10.1021/jp0268112. [DOI] [Google Scholar]
- [60].Jiao S, Xu Z, Selective Gas Diffusion in Graphene Oxides Membranes: A Molecular Dynamics Simulations Study, ACS Appl. Mater. Interfaces 7 (2015) 9052–9059. 10.1021/am509048k. [DOI] [PubMed] [Google Scholar]
- [61].Levitt M, Hirshberg M, Sharon R, Daggett V, Potential energy function and parameters for simulations of the molecular dynamics of proteins and nucleic acids in solution, Comput. Phys. Commun 91 (1995) 215–231. 10.1016/0010-4655(95)00049-L. [DOI] [Google Scholar]
- [62].Hosseini SB, Moradi M, Raoufi D, Study of Carbon Atoms Deposited on Graphene Layer Using Molecular Dynamics Simulation, AIP Conf. Proc 1400 (2011) 108–113. 10.1063/1.3663095. [DOI] [Google Scholar]
- [63].Moulod M, Hwang G, Water self-diffusivity confined in graphene nanogap using molecular dynamics simulations, J. Appl. Phys 120 (2016) 194302. 10.1063/1.4967797. [DOI] [Google Scholar]
- [64].RCSB PDB, https://www.rcsb.org/..
- [65].Humphrey K, W., Dalke A and Schulten, VMD - Visual Molecular Dynamics, J. Molec. Graph 1996, Vol. 14, Pp. 33–38. (n.d.). [DOI] [PubMed] [Google Scholar]
- [66].Vanommeslaeghe K, Hatcher E, Acharya C, Kundu S, Zhong S, Shim J, Darian E, Guvench O, Lopes P, Vorobyov I, MacKerell AD, CHARMM General Force Field (CGenFF): A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields, J. Comput. Chem 31 (2010) 671–690. 10.1002/jcc.21367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Huang J, MacKerell AD Jr, CHARMM36 all-atom additive protein force field: validation based on comparison to NMR data, J. Comput. Chem 34 (2013) 2135–2145. 10.1002/jcc.23354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [68].Bjelkmar P, Larsson P, Cuendet MA, Hess B, Lindahl E, Implementation of the CHARMM force field in GROMACS: Analysis of protein stability effects from correction maps, virtual interaction sites, and water models, J. Chem. Theory Comput 6 (2010) 459–466. [DOI] [PubMed] [Google Scholar]
- [69].Dockal M, Carter DC, Rüker F, The three recombinant domains of human serum albumin structural characterization and ligand binding properties, J. Biol. Chem 274 (1999) 29303–29310. [DOI] [PubMed] [Google Scholar]
- [70].Kragh-Hansen U, Structure and ligand binding properties of human serum albumin, Dan. Med. Bull 37 (1990) 57–84. [PubMed] [Google Scholar]
- [71].Rajabpour A, Akizi FY, Heyhat MM, Gordiz K, Molecular dynamics simulation of the specific heat capacity of water-Cu nanofluids, Int. Nano Lett 3 (2013) 58. [Google Scholar]
- [72].Cooper A, Protein Heat Capacity: An Anomaly that Maybe Never Was, J. Phys. Chem. Lett 1 (2010) 3298–3304. 10.1021/jz1012142. [DOI] [Google Scholar]
- [73].Myers JK, Nick Pace C, Martin Scholtz J, Denaturant m values and heat capacity changes: Relation to changes in accessible surface areas of protein unfolding, Protein Sci. 4 (1995) 2138–2148. 10.1002/pro.5560041020. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data support the findings on this study are available from the corresponding author upon reasonable request.
