Matías
Zúñiga-Bustos
a,
Jeffrey
Comer
*b and
Horacio
Poblete
*cd
aPrograma Institucional de Fomento a la Investigacion, Desarrollo e Innovacion (PIDi), Universidad Tecnologica Metropolitana, Santiago 8940577, Chile
bDepartment of Anatomy and Physiology, Kansas State University, Manhattan, 66506-580, Kansas, USA. E-mail: jeffcomer@ksu.edu
cCenter for Bioinformatics and Molecular Simulation, Facultad de Ingenieria, Universidad de Talca, 2 Norte 685, Talca, Chile. E-mail: hopoblete@utalca.cl
dMillennium Nucleus of Ion Channel-Associated Diseases (MiNICAD), Talca, Chile
First published on 17th May 2023
Nanoscale silver particles have growing applications in biomedical and other technologies due to their unique antibacterial, optical, and electrical properties. The preparation of metal nanoparticles requires the action of a capping agent, such as thiol-containing compounds, to provide colloidal stability, prevent agglomeration, stop uncontrolled growth, and attenuate oxidative damage. However, despite the extensive use of these thiol-based capping agents, the structure of the capping agent layers on the metal surface and the thermodynamics of the formation of these layers remains poorly understood. Here, we leverage molecular dynamics simulations and free energy calculation techniques, to study the behavior of citrate and four thiol-containing capping agents commonly used to protect silver nanoparticles from oxidation. We have studied the single-molecule adsorption of these capping agents to the metal–water interface, their coalescence into clusters, and the formation of complete monolayers covering the metal nanoparticle. At sufficiently high concentrations, we find that allylmercaptan, lipoic acid, and mercaptohexanol spontaneously self-assemble into ordered layers with the thiol group in contact with the metal surface. The high density and ordered structure is presumably responsible for their improved protective characteristics relative to the other compounds studied.
In the context of nanomedicine, AgNPs are immersed in a biological environment. Hence, metal nanoparticles need capping agents to prevent oxidation by radical species. Moreover, it has been found that capping agents can be used to control the final shape and size of the nanoparticles.20,21 Thus, various capping agents and surface modifications (functionalization) have been used to enhance in vivo stability for the duration of their desired function. Citrate is the most commonly used capping agent in the synthesis of silver nanoparticles, due to a its high stabilization effect, as well as the controlled size of the nanoparticles generated.22–24 However, other agents are sometimes used to obtain different surface properties or biological activity. Allylmercaptan is a small naturally occurring organosulfur compound derived from garlic that has been used as a protective agent for silver and gold nanoparticles.25,26 Mercaptohexanol has been used as a capping agent in fluorescent gold nanoparticles for sensing mercury(II) ions.27 The disulfide-containing molecule lipoic acid has been shown to facilitate the covalent protein capping of silver nanoparticles, increasing their oxidative stability in biological environments.28,29 Finally, cysteine has been used as a bio-reducing agent capable of generating silver nanoparticles with homogeneous sizes within a gel matrix.30
Experimentally, Toh et al. demostrated distinct behavior between the two thiol-containing molecules MH and Cys combined with initially citrate-capped nanoparticles. They found that MH formed a dense layer including Ag–S–R and Ag–S–Ag motif, while for Cys, they found merely replacement of the citrate capping agent by Cys.31 Thus, it is necessary investigate the behavior of thiol-containing molecules with different physicochemical properties to understand the link between these properties, the structure of the resulting capping layers, and, the stability and surface chemistry of the capped silver nanoparticles.31
Despite the progress made in development of biomolecule-capped silver nanomaterials, the exact mechanisms involved in the stabilization and the structure of the capping agents at the surface–solvent interface remain poorly understood. It has been shown that thiol-containing molecules, such as mercaptohexanol or cysteine, can replace citrate pre-deposited on the surface of silver nanoparticles, giving different electrochemical and oxidative profiles to the nanoparticle.31 Also, alkanethiols are widely used in silver nanoparticle synthesis.32
The capping molecules are generally covalently linked to the surface (chemisorbed); for example, thiols result in capping moieties including Ag–S covalent bonds. However, the energetic barriers for formation of these covalent linkages are high and, therefore, their formation is quite slow relative to the time scale for physical adsorption. For example, the dissociative adsorption of butanethiol to Ag{111} nanoparticles has been studied.33 The process proceeds by cleavage of the S–H bond (dissociation energy of 0.98 eV) yielding a thiolate, which then forms a covalent bond to the Ag{111} surface. Applying the Arrhenius relation suggests several minutes are required for this chemical reaction at room temperature, (although the presence of solvent and solvent pH affect the barrier) while adsorption and aggregation of the capping molecules occurs on the nanosecond time scale. Therefore, the initial adsorption and organization of the capping molecules are likely dominated by non-covalent interactions and that the final covalently linked structures are similar to those prior to any chemical reactions.
Classical non-reactive molecular dynamics allows us to study the structural arrangements driven by non-covalent interactions of capping agents and their associated thermodynamics on the nano- to microsecond time scale. Molecular simulation yields insight into atomic-scale interactions of surfaces and solutes that are difficult or impossible to access through experiments, yielding insight useful in the rational design of functional nanomaterials. Molecular simulations including polarizable silver surfaces (the AgP model34) and explicit solvent have been used to study the interaction of silver nanosurfaces with proteins,35,36 peptides,34,37–40 collagen-like peptides,41,42 and other small organic molecules. We have previously shown that molecular dynamics simulations, coupled with enhanced sampling methods, can reliably estimate differences in the adsorption affinity of biomolecules.38,43,44
As capping agents for metal nanoparticles, peptides have great promise owing to the ease with which custom sequences can be synthesized and the variety of functional groups available, which at the current time are rivaled by no other class of molecules. Thus, it is important to understand the process of adsorption of small protective agents before designing more complex nanocomplexes using peptides. Rationally modified collagen-like peptides, including the silver-anchor motif CLK, have been demonstrated to form strong complexes with AgNPs and stabilize them in aqueous solution,38,41 while the sequence and length of capping peptides are key factors in controlling production of free radicals.45 Molecular dynamics can be used to elucidate the thermodynamics of recognition of metal nanosurfaces by peptides, and to decompose free energies of adsorption into various contributions, such entropic and enthalpic components,44,46,47 thereby highlighting the roles of different physicochemical effects that control the absorption of small substrates to silver nanostructures.
In the present work, we consider capping of a silver surface by protective agents in three stages (see Fig. 2). We have chosen to focus on the {111} facet of face-centered cubic silver, which is a common exposed crystal plane on Ag nanoparticles.48,49 For the first stage, we perform calculations to determine the thermodynamics of adsorption of a single isolated molecule of the 5 capping agents to the silver–water interface using the enhanced sampling method known as adaptive biasing force (ABF) and decomposing the adsorption free energy into the entropic and enthalpic components. We reveal the thermodynamics of adsorption of citrate (CIT), lipoic acid (LA), allylmercaptan (ALM), mercaptohexanol (MH) and cysteine (CYS), agents which have been previously described experimentally as capping agents on gold and silver nanoparticles.45,50 In the second stage, we study the aggregation of adsorbed molecules at the silver–water interface, performing simulations of many capping molecules at this interface and observe their aggregation and, in some cases, assembly into ordered arrangements. Third, we simulated a complete monolayer of protective agents, calculating the number of molecules per unit area and the type of organization achieved.
The surface model of Hughes et al.34 includes a charged particle rigidly bonded to each silver atom to capture atomic polarizability and include image charge effects. It also includes virtual interstitial sites for Ag{111} facets to ensure correct absorption on top of silver atoms and not between them. The charged particles (atom type AGC) carry a charge of 0.308 proton units, while the silver atom centers (type AGS or AGB) carry the opposite charge. The bulk silver atoms (type AGB) serve both as centers of charge and Lennard-Jones interactions, while surface silver atoms (type AGS) have no Lennard-Jones interactions, which are instead assigned the (electrically neutral) virtual interstitial sites (type AGI). Consistent with the prescription of Hughes et al., the silver atom centers and virtual interaction sites (types AGS, AGB, and AGI) were fixed to their initial positions to maintain the ideal fcc structure, while the charged particles (type AGC) were free to rotate at a distance of 0.7 Å from their parent atom (AGS or AGB). The mass of the charged particles was set to 1.0 Da so NAMD would apply rigid bonds. Special Lennard-Jones parameters were used for interactions between silver atoms (or virtual sites) and other atoms, including oxygen atoms in water and capping agent hydroxyl groups, double-bonded carbon atoms of allylmercaptan, the NH3 nitrogen atom of CYS, carbonyl and carboxylate oxygen atoms, polar or thiol hydrogen atoms, and all sulfur atoms, as prescribed by Hughes et al.34
Structures for the capping agents (CIT, LA, ALM, MH) were extracted from the PubChem database,51 while the geometry of the CYS structure was generated from the CHARMM topology files.52 The protonation state of each molecule was estimated using the pKa calculation tool of Marvin Sketch Software.53 CYS adopted its zwitterionic form (with NH3+ and anionic carboxylate groups), while the acid group of LA was deprotonated to a carboxylate anion. Citrate was assigned a singly deprotonated form (total charge of −1), with a central carboxylate and two terminal carboxylic acid groups. The force field parameters and charges of the small molecules were obtained from the ParamChem server (CHARMM General Force Field version 3.0.1).54,55 Parameters for CYS were obtained from the CHARMM36m force field for proteins.52,56–58 Water molecules were placed above and below the Ag{111} surfaces and were represented with the modified TIP3P water model59 conventionally used with the CHARMM force field.
![]() | (1) |
![]() | (2) |
![]() | (3) |
ΔH(Z,T) ≈ −ΔG(Z,T) + TΔS(Z,T), | (4) |
For MH, molecules were only rarely found outside of the central aggregate after 120 ns; hence, there was poor sampling for ρ(s) at large s and ρ* was poorly defined. Hence, in this case, we performed enhanced sampling calculations using ABF. These calculations were performed with 5 replicates, each using a different tagged molecule for the transition coordinate, for a total of 3 μs of simulated time. The geometric (Jacobian) contribution to the free energy, Ggeo(s) = −kBTln(2πs), was subtracted, which gave results consistent to those calculated above from the areal number density ρ(s).
Capping molecule | Contact area (nm2) | ΔGmin (kcal mol−1) | Z min (Å) |
---|---|---|---|
CIT | 0.5 ± 0.1 | −2.71 ± 0.18 | 4.6 |
ALM | 0.4 ± 0.1 | −1.82 ± 0.03 | 4.1 |
CYS | 0.3 ± 0.1 | −1.83 ± 0.08 | 4.7 |
LA | 0.6 ± 0.1 | −4.81 ± 0.12 | 4.0 |
MH | 0.5 ± 0.2 | −1.86 ± 0.09 | 3.8 |
The thermodynamics of adsorption to the Ag{111} from aqueous solution is strongly controlled by the strength with which water interacts with the silver surface. The density functional theory calculations of the Walsh group34,75 show that water binds much more strongly to silver than graphene and slightly more strongly to silver than gold. Accordingly, in the classical models derived from these calculations, the concentration of water molecules is enhanced at the Ag{111}–water interface, more than four times that in bulk water. As shown in Fig. S1A of the ESI,† the density of the first hydration layer is greater than that of Au{111} or the graphite basal plane. There is also greater orientational order of water at the Ag{111} and Au{111} surfaces than at the graphite surface (Fig. S1B of the ESI†). In several cases, but most prominently for ALM, there is a free energy barrier between the long distance plateau and the global free energy minimum, associated with penetration of the first water layer that covers the silver surface, whose density peaks at Z = 5.8 Å. The free-energy curves in Fig. 3A also exhibit more fine features than similar curves calculated for a graphite surface (Fig. S1C of the ESI†), owing to the complexity of interactions between the polarizable silver surface and polar groups of the molecules.
To reveal thermodynamic contributions to adsorption, we performed free energy calculations for each capping agent at multiple temperatures (280, 300, and 320 K), following the protocol proposed by Singam et al.,44 decomposing the adsorption free energies of the capping agents into their entropic and enthalpic components. Fig. 4 shows the entropic (−TΔS, in red) and enthalpic (ΔH, in black) contributions to the adsorption free energy (ΔG, in orange) at 300 K, calculated from the change in the PMF with the temperature. The compounds evaluated in this investigation showed that the free energy of absorption was mainly driven by the entropy as opposed to enthalpy. In all cases, the enthalpy is unfavorable (ΔH > 0) or only weakly favorable (ΔH > 1 kcal mol−1) near the free energy minimum. This can be understood as follows: the strongest enthalpic interactions with the Ag{111} surface are due to its electrical polarizability and few molecules have a polarizability exceeding that of water (as evidenced by the high dielectric constant of water compared to other materials). Although several of the capping agents have polar groups, they must displace multiple water molecules to occupy the Ag{111} surface and their polar interaction with the surface is invariably weaker than that of the water molecules they displace. This conclusion is supported by the fact that the affinity of water for the Ag{111} surface is stronger (Fig. S1A of the ESI†) than other studied materials such as graphite or Au{111}. However, as we previously observed for graphite,44 releasing water from the interface leads to a considerable gain in entropy since water at the interface has a more restricted spatial and orientational distribution compared to bulk. Indeed, the favorable change in the entropy of the water was the largest contribution to the entropy of adsorption on graphite for the solutes considered and was only partially compensated by the loss of conformational and orientational entropy of the solutes, which are more rigid than the multiple water molecules they replace. As demonstrated by Fig. S1B and C (ESI†), water molecules on the Ag{111} surface are considerably more ordered than those on graphite; hence, there should be an even more favorable gain of entropy when they are released. Hence, it may not be surprising that adsorption to the Ag{111} is entropically driven.
As we previously observed for graphite,44 the minimum values the entropic (−TΔS) and enthalpic (ΔS) contributions to the free energies occur at distinct distances from the surface. For the most flexible molecule, MH, the most favorable entropy occurs at a larger distance from the surface (Z = 4.2 Å) than the most favorable enthalpy (Z = 3.3 Å), which is likely due to steric interactions forcing a more planar geometry, reducing conformational entropy, when the molecule closely approaches the surface (Table 1).
Fig. 5 shows the supramolecular organization of the five protective agents considered in this study. It is possible to observe that ALM, LA, and MH form compact aggregates at the silver–water interface. The molecules were initially placed in the solvent some distance from the interface. After 200 ns of simulation, more than 70% of the molecules of ALM, LA, or MH, have adsorbed to the interface and formed multi-molecule aggregates. Likely owing to their relative hydrophobicity, the ALM and MH systems are dominated by one large compact aggregate. The negative charge of LA prevents its aggregates from growing too large and it instead forms a collection of quasi-two-dimensional micelles (Fig. 5A). On the other hand, only about half of the 100 citrate or cysteine molecules adsorb to the surface and any clusters that form are only temporarily bound (Fig. 5C). ALM, LA, and MH agglomerate with the thiol or disulfide groups in contact with the silver surface, similar to the isolated molecules; however, the orientation of the rest of the molecule can differ due to the inter-solute interactions. Notably, while isolated MH molecules typically lie flat parallel to the surface (Fig. 3F), MH molecules within the core of the aggregates orient perpendicular to the interface with their OH groups pointing away from the surface (Fig. 5E).
The middle hydrocarbon portion of MH is hydrophobic and effectively seeks to minimize its water-exposed area. For an isolated molecule, it can do this by lying flat against the silver surface. However, in the MH aggregate, the molecules can reduce exposure of the hydrocarbon region to both water and the polar silver surface by aligning perpendicular to the surface in the interior of the aggregate. This configuration seems to maximize favorable interactions: the hydroxy group of MH makes contact with water, interior hydrocarbon groups are hidden by condensing together, and the thiol group directly contacts the silver surface, for which it has high affinity. At the edges of the aggregate, some of the MH molecules lie flat as they do when isolated at the silver–water interface.
To understand the thermodynamics of the formation of these aggregates at the silver–water interface, we analyzed the distribution of the molecules after the aggregates reached equilibrium (t > 200 ns). As illustrated in Fig. 6A, we calculated the distribution of molecular positions as a function of distance in the xy-plane from the center of mass of the largest aggregate . From these distributions, we estimated the corresponding free energy profiles as described in Methods. These free energy profiles are shown in Fig. 6B. Due to the fact that the MH aggregate was very cohesive, MH molecules were rarely found outside the aggregate after equilibration and sampling of the position distribution was poor for large distances. Hence, it was necessary to use an enhanced sampling algorithm (ABF) to obtain an accurately anchored free energy curve for MH. While formation of pairs or larger clusters is thermodynamically favorable for all capping agents, free energy plateaus at short distances are visible for ALM and MH, corresponding to the free energy change for transfer of an isolated adsorbed molecule to the aggregate. MH exhibits the greatest tendency to aggregate, with a free energy of −3.2 ± 0.2 kcal mol−1 between the isolated adsorbed phase and aggregate phase.
Furthermore, we have determined the diffusion coefficient of MH molecules under two conditions: when they are in a clustered state and when they are freely diffusing on the silver surface. Fig. S3 of the ESI,† illustrates that MH molecules in a clustered state exhibit reduced diffusion rates due to crowded environment of the cluster, whereas isolated MH molecules at the Ag{111}–water interface exhibit more rapid diffusion. LA exhibits a small free energy barrier near s = 25 Å because this capping agent forms quasi-planar micelles (Fig. 5D) that repel each other electrostatically owing to the negatively charged carboxylates forming the surfaces of these micelles (at neutral or high pH).
Fig. 6C shows the average number of molecules in the largest aggregate at the Ag{111}–water interface for the simulations illustrated in Fig. 5. For MH, nearly all 100 molecules are part of a single disc-like aggregrate (Fig. 5E). Similarly, at the studied concentration, the ALM system contains a dominant large aggregate (Fig. 5B) of about 40 molecules, while the other molecules are mostly adsorbed to the silver surface, but remain independent. Aggregates of LA appear to have a limiting size of about 18 molecules due to electrostatic repulsion of the carboxylate groups (this behavior is likely to be different at low pH). Neither CIT nor CYS form large aggregates at the studied concentrations owing to their relative hydrophilicity and resulting poor cohesion. The evolution of the size of the clusters and number of adsorbed molecules during the simulations is shown in Fig. S2 of the ESI.†
Fig. 6 implies that adsorption of many of the agents is cooperative (especially MH as well as ALM and LA): that is, adsorption of one molecule increases the affinity to adsorb a second. AS we have previously demostrated for organic molecules on graphene,43,44 there likely exist a critical concentration below which there is only a low density of molecules on the surface and above which nearly a complete monolayer is formed. However, predicting this critical concentration (or critical chemical potential) would be quite difficult given the complex structure of the aggregates (for example, the micelle-like aggregates of LA or the orientationally ordered aggregates of MH).
Molecule | Total number | Adsorbed number | Coverage (nm2) | Area/molec. (nm2) |
---|---|---|---|---|
CIT | 40 | 8 | 3.85 ± 0.26 | 0.481 ± 0.030 |
ALM | 40 | 30 | 6.17 ± 0.04 | 0.206 ± 0.001 |
CYS | 40 | 22 | 5.36 ± 0.12 | 0.243 ± 0.005 |
LA | 40 | 13 | 4.87 ± 0.21 | 0.375 ± 0.017 |
MH | 40 | 34 | 6.19 ± 0.02 | 0.182 ± 0.001 |
Fig. 7B shows the mean number of molecules making contact with the Ag{111} surface in equilibrium for simulation systems containing different numbers of molecules. The highest density of molecules occurs for MH (6.19 ± 0.02 nm2), where 34 molecules form a complete ordered monolayer. Fig. 7A shows the associated structure, where the MH molecules are oriented perpendicular to the surface, with the –SH group facing the silver while the –OH group is oriented to the solvent. A hexagonal packing pattern in the xy-plane is evident.
For the case of CIT, the we observe the lowest coverage, achieving a total value of 8 molecules adsorbed on the first layer (3.85 ± 0.26 nm2), and no clear order of the layer is observed. However, as described above, chemisorption not modeled in these simulations may lead to stronger binding and greater coverage.74 On the other hand, this chemisorption should not be too strong, since, after synthesis, citrate is often removed or exchanged with another capping agent.76
On the other hand, the ALM molecule also shows high compactation on top of the surface, showing a coverage area of 6.17 ± 0.04 mol nm−2. However, unlike MH, we do not identify an ordered pattern for ALM on the sulver surface (see Fig. 8A) but rather a mixture of horizontal and vertical molecules. Structurally, the –SH group is always facing the surface, but compared to MH, ALM has a shorter hydrophobic chain and lacks the hydrogen bonding interactions present on the termini of MH.
![]() | ||
Fig. 8 Representative configuration the maximum number of adsorbed molecules on the Ag{111} surface. (A) CIT, (B) ALM, (C) CYS, (D) LA, and (E) MH. |
Cysteine, in simulations at high concentrations, shows a medium coverage (5.36 ± 0.04 mol nm−2) compared to the other molecules studied here. Like CIT, CYS does not entirely cover the silver surface. The thiol group of CYS shows a strong affinity for the silver surface, which is likely increased by chemical conjugation. Additionally, several articles have proposed cysteine as an anchorage point for conjugating peptides to silver and gold nanoparticles.38,41,42,77,78 In this sense, CYS could be a stable, economical, and biocompatible protective agent for efficient protection during synthesis and for biomedical applications.
Finally, LA, despite having a cyclic disulfide group, which is oriented toward the silver surface, only occupies a coverage area of 4.87 ± 0.04 mol nm−2, which is due to electrostatic repulsion between the carboxylate (COO−) groups at neutral or high pH. Simulations at high concentration show that the S–S group effectively remains near the surface while the carboxylate groups point into the water. As low pH, LA likely can pack more densely and perhaps form a more ordered supramolecular structure.
On the other hand, cysteine (zwitterionic form) and citrate (doubly-protonated form with charge of −1) exhibit little tendency to aggregate and form ordered patterns.
The study revealed that silver nanoparticles absorb and assemble differently with various thiols-containing molecules, namely MH, LA, ALM, and cysteine. These results indicate that more hydrophobic molecules self-assemble into dense layers on silver, while more hydrophilic molecules, such as zwitterioinc cysteine, exhibit less dense and less structured capping layers, which is consistent with experimental observations.31 This finding underscores the importance of investigating different classes of thiol-containing molecules individually to understand their behavior and how they regulate the properties of the resulting silver nanoparticles. Further research is required to explore the exchange and substitution of citrate with thiol-containing capping agents to understand how they effectively stabilize silver nanoparticles. These findings could have significant implications for the development of silver nanoparticle-based technologies and applications.
A major limitation of this study is that only physical interactions were considered while the capping agents are expected to covalently conjugate with the silver surface. However, under typical synthesis conditions, these chemical reactions are associated with sizable energetic barriers so that physisorption and supramolecular aggregation at the interface likely occurs (on a timescale of nanoseconds) before chemical bonding. Hence, the structures seen in the simulations are likely similar to the structures that persist after chemical conjugation. However, there may be some differences between the physisorbed and chemisorbed structures, such as in the orientation of the citrate molecules relative to the surface.73 Further research including modeling of chemical reactions is required to fully understand the interplay of physical interactions and chemical reactions on the structure of metal surfaces protected by capping agents.
Footnote |
† Electronic supplementary information (ESI) available: A PDF file containing two figures at https://rsc.org. See DOI: https://doi.org/10.1039/d2cp06002g |
This journal is © the Owner Societies 2023 |