Molecular Dynamics Simulations of Metal Nanoparticles in Deep

Likewise, MNPs synthetic methods in solution are highly convenient and .... constants (Au=4.062 Å, Ag=4.062 Å, Pd=3.876 Å, Pt=3.913 Å and Ni=3.508...
1 downloads 0 Views 937KB Size
Subscriber access provided by Kaohsiung Medical University

C: Physical Processes in Nanomaterials and Nanostructures

Molecular Dynamics Simulations of Metal Nanoparticles in Deep Eutectic Solvents Mert Atilhan, and Santiago Aparicio J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b02582 • Publication Date (Web): 06 Jul 2018 Downloaded from http://pubs.acs.org on July 7, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Molecular Dynamics Simulations of Metal Nanoparticles in Deep Eutectic Solvents Mert Atilhan*a,b and Santiago Aparicio*c

a

Department of Chemical Engineering, Texas A&M University at Qatar, Doha, Qatar b

Gas and Fuels Research Center, Texas A&M University, College Station, TX, USA c

Department of Chemistry, University of Burgos, 09001 Burgos, Spain

*

Corresponding authors: [email protected] (M. A.) and [email protected] (S. A.)

ABSTRACT: A molecular dynamics study on the solvation of metal nanoparticles in deep eutectic solvents is reported in this work. The solvation process was analysed in terms of the type of metal, geometry of the nanoparticles and properties of the deep eutectic solvent. Simulations results in the microsecond range allowed to infer the properties of the solvation shells and the effects of the nanoparticles on the liquid structuring. The possible aggregation of metal nanoparticles in the studied solvents was analysed and discussed in terms of the screening effect of the solvents and the efficient nanoparticle – solvent intermolecular forces. The reported results show deep eutectic solvents acting as metal nanoparticle stabilizers, thus providing a new platform for nanoparticles technologies.

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1. Introduction Metal nanoparticles (MNPs, i.e. particles with diameters lower than 100 nm) have led to a large collection of technological applications1,2 in diverse fields such as physics,3,4 chemistry,5,6 electronics,7,8 optical,9,10,11 materials science12,13,14,15 and biomedical applications.16,17 These applications rise from their outstanding physicochemical properties,18 which are largely correlated with MNPs size and shape19 leading to a way for tailoring MNPs properties via morphology control.20,21,22,23 The successful application of MNPs for catalytic purposes is well known,24,25 and it has led to great advances in petrochemicals manufacturing.26 Physical phenomena such as surface plasmon on MNPs have also great relevance,27,28 and they offer many opportunities for application in several technologies via observing nanometer scale alterations in thickness, density oscillations or molecular absorption on metal surfaces.29,30 For medical applications, the use of MNPs for cancer therapy,31 for diagnostic and imaging,32 as drug delivery agents,33 or as antifungal – antimicrobial agents34,35,36 have been widely studied. The development of new methods for MNPs synthesis and surface modification has allowed to produce MNPs with tailored sizes, shapes and composition,37,38,39 i.e. tuning the MNPs by their morphology.40,41 The preparation and properties of stable MNPs colloidal dispersions have been studied in different solvents,42,43,44,45 which is of great relevance in applications such as catalysis.46,47 Likewise, MNPs synthetic methods in solution are highly convenient and versatile,48,49,50 and thus, MNPs fabrication in polar and nonpolar solvents has been considered. The use of water as solvent for MNPs synthesis has been widely reported because of environmental, economic and physicochemical properties,51,50 but other solvents such as supercritical fluids,52 carbonates,53 and most recently ionic liquids (ILs)50,52,54,55 have also been considered as new solvents for the same purpose. For the case of MNPs in ILs has attracted great attention because of the possibility of tailoring IL properties to fulfil requirements for MNPs synthesis and colloidal stabilization. It is wellknown that MNPs induce large restructuring in surrounding solvents56 due to the solvent – 2

ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

MNP interface and the involved intermolecular forces. This phenomenon is also relevant especially for ILs due to the complex intermolecular forces.57,58 Due to such ionic character of the solvent it leads to a remarkable structuring at the interfaces. Despite the promising properties of ILs for developing MNPs, some concerns59,60 about their toxicity and biodegradability,61 economic viability,

62

large scale application63,64 or unsuitable

physicochemical properties65 have been reported, which have shifted, at least partially, the interest toward other related relevant solvent alternatives. Deep eutectic solvents (DESs)66 are a type of fluids, which are produced by a combination of a hydrogen bond acceptor (HBA) and a hydrogen bond donor (HBD), leading to a mixture with low melting temperature. DESs have been proposed as sustainable media for the development of nanoscale materials67 because of their to low toxicity,68 suitable biodegradability,69,70 and low production costs, which lead to agreement on DESs as suitable platforms for developing MNPs within a sustainable chemistry framework.71,72,73 Nevertheless, studies on MNPs at DESs are in their infancy and the mechanisms of MNPs growth, solvation and stabilization in DESs remain largely unexplored. Lee74 recently reviewed the use of DESs for the synthesis of MNPs showing great potential for developing MNPs applications. Moreover, Wang et al.75 studied the electrochemical preparation of copper MNPs in choline chloride – urea DES and showed how DESs can successfully replace traditional methods for preparing copper nanoparticles. Several studies have also reported successful preparation of different types of metallic nanoparticles,76,77,78 and thus, confirming the suitability of these solvents for preparing MNPs. The large number of possible HBA:HBD combinations leading to DESs is advantageous on one side since it allows to tailor DESs properties for the selected process, but on the other side it requires a deep knowledge of the properties of these solvents as a function of the involved HBA:HBD and their interaction with MNPs. Therefore, the analysis of solvation and stability of MNPs in selected solvents require a know-how on the intermolecular forces at microscopic level and structuring at particle – solvent (DESs) interfaces. Theoretical studies using molecular dynamics simulation (MD) have proven to be useful for characterizing MNPs in different solutions,79,80,81 but MD studies for DESs-

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

MNPs are absent in the literature. MD studies on DESs-MNPs would allow to obtain molecular level information on the relationships between the involved HBA:HBD and its ability to solvate and interact with MNPs, thus contributing to the development of structure – property relationships for screening purposes of DESs regarding MNPs. Therefore, we report herein a MD study on the solvation and aggregation of gold, silver, palladium, platinum and nickel nanoparticles in choline chloride (ChCl) – based DES with urea (URE), ethylene glycol (EG), glycerol (GLY), malonic acid (MAL), levulinic acid (LEV) and phenyl acetic acid (PAA) as HBDs, which were selected to consider different molecular functionalities to analyse their effect on the interaction with the selected MNPs. ChCl is selected as HBA since it is a non-toxic and biodegradable compound and it can be obtained at very low costs.82,83 MNPs size and shape effects on the mechanism of solvation were also studied. The objective of the study is to analyse for the first time the mechanisms of MNPs solvation and aggregation in DESs as a function of the type of metal, size and shape of the nanoparticles, and involved HBDs.

Methods MD simulations were carried out using the ACEMD84 program running on GPUs. Force field parameterizations for the considered DESs were previously reported.85 Metal atoms (Au, Ag, Pd, Pt and Ni) were described as non-charged 12-6 Lennard–Jones sites according to the parameterization reported by Heinz et al.86 Icosahedral and spherical MNPs were built using the nanoparticle builders available in openMD software.87 Icosahedral MNPs with 4 shells (Au4, Ag4, Pd4, Pt4 and Ni4) were built with the corresponding experimental lattice constants (Au=4.062 Å, Ag=4.062 Å, Pd=3.876 Å, Pt=3.913 Å and Ni=3.508 Å),88 and thus leading to nanoparticles with diameters Au4=22.0 Å, Ag4=18.8 Å, Pd4=17.8 Å, Pt4=17.9 Å and Ni4=16.1 Å. Gold icosahedral nanoparticles with larger sizes were also built to analyse the MNPs size effect on properties, and thus Au icosahedral nanoparticles with 6 (Au6, diameter=28 Å) and 8 (Au8, diameter=44 Å) shells were also built. Gold spherical nanoparticles with 14 Å diameter (Au_sph) were also built. The number of atoms for each studied MNP are reported in Table S1 (Supporting Information). Systems and conditions 4

ACS Paragon Plus Environment

Page 4 of 37

Page 5 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

for simulating DES + MNPs were built as defined in Table S2 (Supporting Information) with the objectives of analyse the effects of HBD type, metal type, and MNPs size and shape effect. Initial simulation boxes with the characteristics reported in Table S2 (Supporting Information) were built with the Packmol software.89 In the case of systems containing a single MNP, boxes were built with the corresponding MNP placed in the center of a cubic simulation box and surrounded by the corresponding molecules forming the DES. For systems containing five MNPs (for the study of MNP aggregation), these nanoparticles were initially distributed in the cubic simulation boxes as reported in Figure S1 and surrounded by the corresponding DES molecules. MD simulations were carried in two steps for all the systems reported in Table S2 (Supporting Information): i) equilibration runs using 20 ns long simulations in the NVT ensemble with fixed MNPs (for equilibrating the surrounding DES solvent) at 300 K, and ii) production runs using 1 μs long simulations in the NPT ensemble at 300 K and 1 bar. Temperature was controlled with a Langevin type thermostat,90 with 0.1 ps-1 damping constant, and pressure with a Berendsen barostat,91 with 4 ps relaxation time. Nonbonded cross interaction terms were calculated using Lorentz-Berthelot mixing rule. Particle mesh Ewald method was used for handling electrostatic interactions.92 All of the studied systems were treated by considering periodic boundary conditions.

Results and discussion The addition of MNPs to the studied DESs should lead to changes in the liquid structure because of changes in intermolecular forces to produce efficient solvation around MNPs, and thus, changes in the intermolecular forces between the molecules forming the DESs should be expected with largest changes in solvation shells around the MNPs. The properties of DESs stand on the hydrogen bonding between the salt (ChCl in this work) and the considered HBDs. Therefore, for the first DESs studied in this work, ChCl:URE, the Ch to Cl and Cl to URE hydrogen bonding93,94 should experience changes upon adding MNPs to the fluid. Radial pair distribution functions, RDFs, for relevant molecular pairs of 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ChCl:URE DES are reported in presence and absence of one MNP (Au4, Ag4, Pd4, Pt4 and Ni4), Fig. 1. The characteristic narrow RDF peaks, showing hydrogen bonding is maintained upon addition of the MNPs, although minor shifting and certain rearrangement of structuring at distances larger than 7 Å are inferred. Therefore, the ChCl:URE DES is able to solvate efficiently the MNPs maintaining most of the features of their liquid structuring. Likewise, there are no differences in the type of considered MNPs, with all of them leading to the same minor changes in the DES. Nevertheless, slightly visible minor changes as reported in Fig. 1 are indicative of molecular rearrangements in the solvation shells around the MNPs. The structuring of the molecules around the MNPs is analysed from the RDFs between the centers-of-mass of the MNPs and those for cation, anion and URE (for ChCl:URE DES), Fig. 2. The results reported in Figs. 2a to 2c show the formation of a first solvation layer around the MNPs with additional disrupting features produced by the presence of the MNPs on DES liquid structure beyond this first layer. The analysis of the properties of this first solvation layer shows that they are largely dependent on the type of metal forming the MNP and they are richer in URE (HBD) than in salt ions, Figs. 2d to 2f. Two regions may be defined around the MNPs, the first one with r < 12 Å (where r stands for the distance to the MNPs center-of-mass), corresponding to the limit of the first solvation layer, and the second one with r in the 12 Å to 15 Å range, which can be assigned roughly to the second solvation shell. The quantification of the molecules in these solvation shells is reported in Fig. 3. Results in Fig. 3a shows a large excess of URE molecules in the first solvation shell, especially for Au MNPs, which almost vanishes in the second solvation shell. This excess of URE molecules is almost negligible for Ni MNPs. Likewise, the first solvation shell is characterized by a deficit of Ch cations in comparison with Cl anions, especially for Au and Ag MNPs, which almost vanishes on going to the second solvation spheres. Therefore, the solvation of the studied MNPs is characterized by a first strong and well-defined solvation sphere, rich in URE molecules, followed by a second weak solvation shell with properties close to the bulk liquid DES, and with very minor perturbations of the liquid structure beyond these two solvation spheres, Fig. S2 (Supporting Information). Regarding the orientation of molecules around the MNPs, using

6

ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

vectors as defined in Figs. 4 and S3 (Supporting Information). Results in Fig. 4a for URE molecules show that in the first layer they tend to be parallel to the surface, whereas URE molecules are skewed in the second solvation shell. In the case of cholinium cations, Fig. 4b, ions are skewed in the first shell, whereas except for Ni MNPs they lay parallel to the MNPs in the second solvation shell. The spatial structure in the first solvation shell around Au MNP is reported in Fig. 5a (similar arrangements are inferred for the remaining MNPs) showing how most of the MNP is covered by URE molecules with isolated spots of chlorine and cholinium ions. A visual picture of the parallel arrangement of URE molecules on top of Au MNP is reported in Fig. 5b. The adsorption of DES molecules, especially of URE, should lead to changes in the molecular mobility for those molecules in the first solvation shell in comparison with those in the bulk liquid phase. These effects are quantified by the self-diffusion coefficients reported in Fig. 6 (obtained from the mean square displacements and Einstein’s equation) and compared with the values in the bulk fluid (well-beyond the solvation spheres). These results show an increase in molecular mobility upon adsorption on MNPs for all the considered metals, which can be justified considering that the first solvation shell is largely composed of URE molecules (Fig. 3a and 5a). The strength of MNP-DES interactions is quantified in Fig. 7, following the ordering URE > Ch> Cl, in agreement with the composition of the first solvation shell around MNP, therefore the stabilization of MNPs in DES is produced mainly by the interaction with the hydrogen bond donor (e.g. URE). Nevertheless, the rearrangement of DES molecules leads to local ordering around the MNPs by the preferential solvation by URE molecules but the disruptive effect of the MNPs is limited to the liquid region close to the MNPs (Figs. 1 and S2, Supporting Information), with minor changes in the structuring of the bulk liquid phase as showed by the extension of hydrogen bonding reported in Fig. 8. MNPs solvation and stabilization is characterized by the MNP-HBD interaction for all the studied MNPs, the effect of the type of involved HBD was also analysed in this work. RDFs reported in Figs. 9a to 9c show that the first solvation shell around MNPs is always characterized by the presence of an excess of HBD molecules. Nevertheless, the

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

type of HBD leads to subtle differences for the properties of the solvation shells, especially for the first solvation shell because of the size and shape of HBD molecules. The first peak in RDFs corresponding to MNP-HBD pairs, Fig. 9c, provided for all the studied HBDs and the structure of the peak changes depending on the considered HBD. Likewise, the characteristics of the distribution of Ch and Cl ions around the MNPs is also dependent on the type of HBD. Nevertheless, first solvation shells are always rich in HBD for all the considered DES (Figs. 9d to 9f), but these heterogeneities vanish beyond the second solvation shells. These heterogeneities are quantified by the molecular ratios reported in Fig. 10 with a large excess of HBDs in the first solvation shells but almost vanishing on going to the second shell, and thus, confirming that the disruptive effect of MNPs is very local and the bulk liquid structure in DES is mostly maintained for the studied MNPs concentrations. The HBD rich solvation shells around Au MNPs are reported in Fig. 11 showing that PAA leads to a sphere almost covered by the HBD and MAL leads to a lower presence of the HBD in the solvation shell. Nevertheless, strong MNP-HBD interactions are reported for all the considered HBDs, Fig. 12, with the weakest interactions obtained for MAL in agreement with results in Figs. 10a and 11c. Regarding the molecular orientation of HBDs in the first solvation shell around Au MNP, results in Fig. 13 show parallel arrangement of HBDs on top of MNP for HBDs in the first solvation shell (with the exception of slight skewing for MAL and EG), whereas molecules are skewed in the second shell. Therefore, HBDs lay parallel to MNPs surface for all the studied DES, increasing Lennard-Jones interactions and thus leading to the large interaction energies reported in Fig. 12. Nevertheless, the MNPs solvation is almost limited to the first sphere and the hydrogen bonding structuring of DES is maintained for all the considered HBDs, Fig. 14. The effect of the MNP size is analysed and provided in Fig. 15. The prevailing presence of URE (HBD) in the first solvation shell is maintained for the studied MNP sizes. Nevertheless, RDFs peaks corresponding both to Ch and Cl ions in the first solvation shells around the MNPs increase with increasing MNP size, thus decreasing the heterogeneities in the first solvation shells. Therefore, the increase in the size of MNPs allows the

8

ACS Paragon Plus Environment

Page 8 of 37

Page 9 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

presence of a large number of URE molecules in the first solvation shell, but on the other side a remarkable presence of ions is also possible, which was not present in small Au8 MNPs. Nevertheless, the disrupting effect of the MNPs is almost limited to the DES region in close contact with the MNP, even for the large Au8 MNPs. The effect of the MNP shape is analysed in Fig. 16 by comparison of the solvation of Au MNP with icosahedral and spherical geometries, both with similar size. RDFs results in Figs. 16a to 16c show that for both geometries similar mechanisms of solvation are inferred, with preferential presence of URE molecules in the first solvation shell. The main difference between the solvation of spherical and icosahedral MNPs stands on a minor decrease of the position of the maxima at which RDFs corresponding to the first solvation shell appears (∼ 1 Å shift), showing that spherical MNPs allows a closer approach of URE molecules (and of ions) to the surface of the particle leading to a larger number of solvating molecules, especially for URE, close to the surface, Figs. 16d to 16f. Therefore, this seems to be a purely geometrical factor controlling also the geometry, and thus, availability of space of the first solvation shells around MNPs. In the last stage of this work, the solvation of a larger number of MNPs (i.e. larger MNP concentration) and possible MNP aggregation was analysed using ChCl:URE as model DES. Raghuwanshi et al.78 studied the formation and self-assembly of Au MNP in ChCl:URE using an experimental approach showing that the formation of Au MNP with a regular size of 5 nm on average can be obtained in this DES. Likewise, these authors showed that MNP tend to self-aggregate leading to short-range ordering with MNP to MNP separation as close as 0.5 Å. Nevertheless, these experimental results showed that self-aggregation is largely dependent on MNP concentration, i.e. occurring for concentrated solutions, and requires several hours to form the MNP aggregates. Likewise, the authors showed the templating effect of DES for solvating MNPs. The behaviour of a group of MNPs, initially arranged in a tetrahedral formation (reported in Fig. S1 of Supporting Information) was studied in the considered model DES. The initial distance between the centers-of-mass of nanoparticles was roughly 34 Å, which leads to 12 Å surface-to-surface distance, which is almost the double to that experimentally reported by Raghuwanshi et al.78 for the

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

aggregates obtained experimentally (5 Å). The evolution of the interparticle distances was studied for 1 μs, showing a slight decrease, mainly in the initial stages of the simulation, the nanoparticles stay at distances larger than those reported by Raghuwanshi et al.78, i.e. no aggregation is inferred by MD within the studied simulation timeframe, Fig. 17. The formation of stable solvation shells around MNPs hinders the MNP aggregation, maintaining MNPs beyond the aggregation limit; the large MNP-DES interaction energies reported in Fig. 7 stabilizes the MNP upon solution in DES, and leads to a screening effect delaying particle aggregation. Nevertheless, this effect seems to be mainly kinetic because experimental results (Raghuwanshi et al.78) showed MNP aggregation after time which is beyond the limits of atomistic MD simulations. MD results confirm that efficient MNP solvation is produced even for highly concentrated MNP solutions in DES as studied in this work. The efficient MNPs solvation is inferred for all the studied metals, although some metals such as Pd leads to closer MNP inter-distances in the studied simulation time, Fig. 18a. Likewise, the analysis of HBD effect on these highly concentrated MNP solutions shows that some HBDs such as MAL lead to lower inter-particle distances, in agreement with the weaker MNP-DES intermolecular interactions reported in Fig. 12, but for all the systems inter-particle distances are larger than those experimentally obtained for aggregates. Therefore, MD results confirm that solutions with large MNP concentrations can be efficiently solvated by DES molecules, delaying MNP aggregation with DES screening MNP-MNP intermolecular forces.

Conclusions The solvation of metal nanoparticles in deep eutectic solvents was studied using molecular dynamics simulations. The effects of type of metal, type of deep eutectic solvent, particle size and shape were analysed. Efficient solvation was inferred for all the considered cases, with a solvation structure characterized by a first solvation shell very rich in molecules of the corresponding hydrogen bond donor, followed by a second shell with properties close to the bulk liquid phases. The disruptive effect of the nanoparticles is almost limited to these two solvation layers, with the remaining fluid maintaining its 10

ACS Paragon Plus Environment

Page 10 of 37

Page 11 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hydrogen-bonding network. The formation of this strong first solvation layer stands on the large nanoparticle – hydrogen bond donor Lennard-Jones interactions, and in minor extensions intermolecular forces with available anions and cations.

The effective

solvation of the nanoparticles is maintained for all the studied metals (Au, Ag, Pd, Pt and Ni) and for the considered hydrogen bond donors, although in some cases such as Ni nanoparticles or malonic acid the solvation is less effective. The increase of nanoparticles size decreases the heterogeneity in the composition of the first solvation shell, allowing both the preferential presence of hydrogen bond donor molecules but also of hydrogen bond acceptors. Likewise, the analysis of shape effect of nanoparticles on solvation shows that spherical nanoparticles leads a slightly more efficient solvation in comparison with icosahedral ones. Finally, the molecular dynamics study of the solvation of concentrated nanoparticles solution in deep eutectic solvents showed that even for these concentrated solutions, deep eutectic solvents can solvate the metal nanoparticles screening the interaction between nanoparticles and hindering nanoparticle aggregation in the studied microsecond simulation range. Therefore, the molecules of deep eutectic solvents can efficiently solvate different types of metal nanoparticles, acting as nanoparticle stabilizers, and delaying nanoparticles aggregation, and thus, they can be considered as suitable matrix for developing stable solution of metal nanoparticles with controlled particle size.

Supporting Information Table S1 (MNPs used in this work); Table S2 (systems used for MD simulations); Figure S1 (initial arrangement of MNPs); Figure S2 (MNP-DES radial distribution functions); Figure S3 (definition of orientation angles).

Acknowledgements The authors acknowledge SCAYLE (Supercomputación Castilla y León, Spain) for providing supercomputing facilities. The statements made herein are solely the responsibility of the authors. 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) Daniel, M. C.; Astruc, D. Gold Nanoparticles: Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications Toward Biology, Catalysis, and Nanotechnology. Chem. Rev. 2004, 104, 293– 346. (2) Louis, C.; Pluchery (Eds.) Gold Nanoparticles for Physics, Chemistry and Biology. Ed. World Scientific, London, 2017. (3) Reguera, J.; Langer, J.; Jiménez de Aberasturi, D.; Liz-Marzán, L. M. Anisotropic Metal Nanoparticles for Surface Enhanced Raman Scattering. Chem. Soc. Rev. 2017, 46, 3866-3885. (4) Lermé, J.; Bonnet, C.; Lebeault, M. A.; Pellarin, M.; Cottancin, E. Surface Plasmon Resonance Damping in Spheroidal Metal Particles: Quantum Confinement, Shape, and Polarization Dependences. J. Phys. Chem. C 2017, 121, 5693-5708. (5) Alonso, F.; Moglie, Y.; Radivoy, G. Copper Nanoparticles in Click Chemistry. Acc. Chem. Res. 2015, 48, 2516-2528. (6) Liu, P.; Qin, R.; Fu, G.; Zheng, N. Surface Coordination Chemistry of Metal Nanomaterials. J. Am. Chem. Soc. 2017, 139, 2122-2131. (7) Kamyshny, A.; Magdassi, S. Conductive Nanomaterilas for Printed Electronics. Small 2014, 10, 3515-3535. (8) Gayen, A. L.; Paul, B. K.; Roy, D.; Kar, S.; Bnadyopagyay, P.; Basu, R.; Das, S.; Bhar, D. S.; Manchanda, R. K.; Khurana, A. K.; et al. Enhanced Dielectric Properties and Conductivity of Triturated Copper and Cobalt Nanoparticles-Doped PVDF-HFP Film and Their Possible Use in Electronic Industry. Materials Research Innovations 2017, 21, 166-171. (9) Zhang, Y. X.; Wang, Y. H. Nonlinear Optical Properties of Metal Nanoparticles. RSC Adv. 2017, 7, 4512945144. (10) Hartland, G. V.; Besteiro, L. V.; Johns, P.; Govorov, A. O. What’s So Hot About Electrons in Metal Nanoparticles? ACS Energy Lett. 2017, 2, 1641-1653. (11) Gong, C.; Sousa Dias, M. R.; Wessler, G. C.; Taillon, J. A.; Salamanca-Riba, L. G.; Leite, M. S. Near-Field Optical Properties of Fully Alloyed Noble Metal Nanoparticles. Adv Opt. Mater. 2017, 5, 1600568. (12) Schreiber, R.; Do, J.; Roller, E. M.; Zhang, T.; Schïller, V. J.; Nickels, P. C.; Feldmann, J.; Liedl, T. Hierarchical Assembly Of Metal Nanoparticles, Quantum Dots and Organic Dyes Using DNA Origami Scaffolds. Nat. Nanotechnol. 2014, 9, 74-78. (13) Ahmadi, M.; Mistry, H.; Cuenya, B. R. Tailoring the Catalytic Properties of Metal Nanoparticles via Support Interactions. J. Phys. Chem. Lett. 2016, 7, 3519-3533. (14) Klechikov, A.; Sun, J.; Hu, G.; Zheng, M.; Wagberg, T.; Talyzin, A. V.; Graphene Decorated with Metal Nanoparticles: Hydrogen Sorption and Related Artefacts. Micropor. Mesopor. Mat. 2017, 250, 27-34. (15) Yang, Q.; Liu, W.; Wang, B.; Zhang, W.; Zeng, X.; Zhang, C.; Qin, Y.; Sun, X.; Wu, T.; Liu, J.; et al. Regulating The Spatial Distribution of Metal Nanoparticles Within Metal-Organic Frameworks To Enhance Catalytic Efficiency. Nat. Comm. 2017, 8, 14429. (16) Yang, X.; Yang, M.; Pang, B.; Vara, M.; Xia, Y. Gold Nanoparticles at Work in Biomedicine. Chem. Rev. 2015, 115, 10410-10488. (17) Zhang, X. Gold NanoparticlesÇ Recent Advances in the Biomedical Applications. Cell Biochem. Biophys. 2015, 72, 771-775. (18) Daniel, M. C.; Astruc, D. Gold Nanoparticles: Assembly, Supramolecular Chemistry, Qauntum-SizeRelated Properties, and Applications Toward Bilogy, Catalysis, and Nanotechnology. Chem. Rev. 2004, 104, 293-346.3 (19) Rao, C. N. R.; Kulkarni, G. U.; Thomas, P. J.; Edwards, P. P. Metal Nanoparticles and Their Assemblies. Chem. Soc. Rev. 2000, 29, 27-35. (20) Kelly, K. L.; Coronado, E.; Zhao, L. L.; Schatz, G. C. The Optical Properties of Metal Nanoparticles: The Influence of Size, Shape, and Dielectric Environment. J. Phys. Chem. B 2003, 107, 668-677. 12

ACS Paragon Plus Environment

Page 12 of 37

Page 13 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(21) Liz-Marzán, L. M. Tailoring Surface Plasmons Through the Morphology and Assembly of Metal Nanoparticles. Langmuir 2006, 22, 32-41. (22) Sau, T. K.; Rogach, A. L. Nonspherical Noble Metal Nanoparticles: Colloid – Chemical Syntheis and Morphology Control. Adv. Mater. 2010, 22, 1781-1804. (23) Roldan Cuenya, B.; Behafarid, F. Nanocatalysis: Size- and Shape Dependent Chemisorption and Catalytic Reactivity. Surf. Sci. Rep. 2015, 70, 135−187. (24) Jia, C. J.; Schüth, F. Colloidal Metal Nanoparticles As a Component of Designed Catalyst. Phys. Chem. Chem. Phys. 2011, 13, 2457-2487. (25) Metal Nanoparticles for Catalysis: Advances and Applications, Tao F. (Ed). RSC Catalysis Series No. 17. The Royal Society of Chemistry, Cambridge, 2014. (26) Talapin, D. V.; Shevchenko, E. V. Introduction: Nanoparticle Chemistry. Chem. Rev. 2016, 116, 1034310345. (27) Noguez, C. Surface Plasmons on Metal Nanoparticles: The Influence of Shape and Physical Environment. J. Phys. Chem. C 2007, 111, 3806-3819. (28) Rycenga, M.; Cobley, C. M.; Zeng, J.; Li, W.; Moran, C. H.; Zhang, Q.; Qin, D.; Xia, Y. Controlling the Synthesis and Assembly of Silver Nanostructures for Plasmonic Applications. Chem. Rev. 2011, 111, 36693712. (29) Baffou, G.; Quidant, R. Nanoplasmonic for Chemistry. Chem. Soc. Rev. 2014, 43, 3898-3907. (30) Linic.S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical Transformations on Plasmonic Metal Nanoparticles. Nat. Mater. 2015, 14, 567-576. (31) Zhao, J.; Lee, P.; Wallace, M. J.; Melancon, M. P. Gold Nanoparticles in Cancer Therapy: Efficiency, Biodistribution, and Toxicity. Curr. Pharm. Des. 2015, 21, 4240-4251. (32) Sharma, H.; Mishra, P. K.; Talegaonkar, S.; Vaidya, B. Metal Nanoparticles: A Theranostic Nanotool Against Cancer. Drug Discovery Today 2015, 20, 1143-1151. (33) Jeong, E. H.; Jung, G.; Honh, C. A.; Lee, H. Gold Nanoparticle (AuNP)-based Drug Delivery and Molecular Imaging for Biomedical Applications. Arch. Pharm Res. 2014, 37, 53-59. (34) Paril, P.; Baar, J.; Cermak, P.; Rademacher, P.; Prucek, R.; Sivera, M.; Panacek, A. Antifungal Effects of Copper and Silver Nanoparticles Against White and Brown-Rot Fungi. J. Mater. Sci. 2017, 52, 2720-2729. (35) Dizaj, S. M.; Lotfipour, F.; Barzegar-Jalali, M.; Zarrintan, M. H.; Adibkia, K. Antimicrobial Activity of the Metals and Metal Oxide Panoparticles. Materials Sci. Eng C 2014, 44, 278-284. (36) Singh, R.; Nawale, L. U.; Arkile, M.; Shedbalkar, U. U.; Wadhwani, S. A.; Sarkar, D.; Chopade, B. A. Chemical and Biological Metal Nanoparticles as Antimycobacterial Agents: A Comparative Study. Int. J. Antimicrob. Ag. 2015, 46, 183-188. (37) Sun, Y.; Xia, Y. Shape-Controlled Synthesis of Gold and Silver Nanoparticles. Science 2012, 298, 21762179. (38) Xia, Y.; Xiong, Y.; Lim, B.; Skrabalak, S. E. Shape-Controlled Synthesis of Metal Nanocrystals: Simple Chemistry Meets Complex Physics? Angew. Chem. Int. Ed. 2008, 48, 60-103. (39) Xia, Y.; Xia, X.; Peng, H. C. Shape-Controlled Synthesis of Colloidal Metal Nanocrystals: Thermodynamic versus Kinetic Products. J. Am. Chem. Soc. 2015, 137, 7947-7966. (40) Kosinova, A.; Wang, D.; Baradács, E.; Parditka, B.; Kups, T.; Kilinger, L.; Erdélyi, Z.; Schaaf, P.; Rabkin, E. Tuning the Nanoscale Morphology and Optical Properties of Porous Gold Nanoparticles by Surface Passivation and Annealing. Acta Mater. 2017, 127, 108-116. (41) Duan, Z.; Li, M.; Mwenya, T.; Li, Y.; Song, D. Morphology Optimization of Silver Nanoparticles Used to Improve the Light Absorption in Thin-Film Silicon Solar Cells. Plasmonics 2018, 13, 555-561. (42) Nath, S.; Jana, S.; Pradhan, M.; Pal, T. Ligand-stabilized Metal Nanoparticles in Organic Solvent. J. Colloid. Inter. Sci. 2010, 341, 333-352. (43) Thanh, N. T. K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. 2014, 114, 7610-7630. (44) Lista, M.; Liu, D. Z.; Mulvaney, P. Phase Transfer of Noble Metal Nanoparticles to Organic Solvents. Langmuir 2014, 30, 1932-1938. (45) Polte, J. Fundamental Growth Principles of Colloidal Metal Nanoparticles – A New Perspective. Cryst. Eng. Comm. 2015, 17, 6809-6830. 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(46) Narayanan, R.; El-Sayed, M. Catalysis with Transition Metal Nanoparticles in Colloidal Solution:  Nanoparticle Shape Dependence and Stability. J. Phys. Chem. B 2005, 109, 12663-12676. (47) Kaur, R.; Giordano, C.; Gradzielski, M.; Mehta, S. K. Synthesis of Highly Stable, Water-Dispersible Copper Nanoparticles as Catalysts for Nitrobenzene Reduction. Chem. Asian J. 2014, 9, 189-198. (48) Martin, M. N.; Basham, J. I.; Chando, P.; Eah, S. K. Charged Gold Nanoparticles in Non-Polar Solvents: 10-min Synthesis and 2D Self-Assembly. Langmuir 2010, 26, 7410-7417. (49) Ju-Nam, Y.; Abdussalam-Mohammed, W.; Ojeda, J. J. Highly Stable Noble Metal Nanoparticles Dispersible in Biocompatible Solvents: Synthesis of Cationic Phosphonium Gold Nanoparticles in Water and DMSO. Faraday Discuss. 2016, 186, 77-93. (50) Duan, H.; Wang, D.; Li, Y. Green Chemistry for Nanoparticle Synthesis. Chem. Soc. Rev. 2015, 44, 57785792. (51) Anderson, R.; Buscall, R.; Eldridge, R.; Mulvaney, P.; Scales, P. Concentrated Synthesis of Metal Nanoparticles in Water. RSC Adv. 2014, 4, 31914-31925. (52) Türk, M.; Erkey, C. Synthesis of Supported Nanoparticles in Supercritical Fluids by Supercritical Fluid Reactive Deposition: Current State, Further Perspectives and Needs. J. Supercrit. Fluid. 2018, 134, 176183. (53) Schütte, K.; Meyer, H.; Gemel, C.; Barthel, J.; Fischer, R. A.; Janiak, C. Synthesis of Cu, Zn and Cu/Zn Brass Alloy Nanoparticles from Metal Amidinate Precursors in Ionic Liquids or Propylene Carbonate with Relevance to Methanol Synthesis. Nanoscale 2014, 6, 3116-3126. (54) Antionietti, M.; Kuang, D.; Smarsly, B.; Zhou, Y. Ionic Liquids for the Convenient Synthesis of Functional Nanoparticles and Other Inorganic Nanostructures. Angew. Chem. Int. Ed. 2004, 43, 4988-4992. (55) Charan, K. T.; Pothanaghandi, N.; Vijayakrishna, K.; Sivaramakrishna, A.; Mecerreyes, D., Sreedar, B. Poly(ionic liquids) as “Smart” Stabilizers for Metal Nanoparticles. Eur. Polym. J. 2014, 60, 114-122. (56) Zobel, M.; Neder, R. B.; Kimber, S. A. Universal Solvent Restructuring Induced by Colloidal Nanoparticles. Science 2015, 347, 292-294. (57) Pensado, A. S.; Pádua, A. A. H. Solvation and Stabilization of Metallic Nanoparticles in Ionic Liquids. Angew. Chem. 2011, 123, 8842-8846. (58) He, Z.; Alexandridis, P. Nanoparticles in Ionic Liquids: Interactions and Organization. Phys. Chem. Chem. Phys. 2015, 17, 18238-18261. (59) Cevasco, G.; Chiappe, C. Are Ionic Liquids a Proper Solution to Current Environmental Challenges? Green Chem. 2014, 16, 2375-2385. (60) Egorova, K. S.; Ananikov, V. P. Toxicity of Ionic Liquids: Eco(cyto)activity as Complicated, but Unavoidable Parameter for Task-Specific Optimization. ChemSusChem. 2014, 7, 336-360. (61) Neumann, J.; Steudte, S.; Cho, C. W.; Thöming, J.; Stolte, S. Biodegradability of 27 Pyrrolidinium, Morpholinium, Piperidinium, Imidazolium and Pyridinium Ionic Liquid Cations Under Aerobic Conditions. Green Chem. 2014, 16, 2174-2184. (62) Chen, L.; Sharifzadeh, M.; MacDowell, N.; Welton, T.; Shah, N.; Hallett, J. P. Inexpensive Ionic Liquids: [HSO4]- - Based Solvent Production at Bulk Scale. Green Chem. 2014, 16, 3098-3106. (63) Daillo, A. O.; Len, C.; Morgan, A. B.; Marlair, G. Revisiting Physico-Chemical Hazards of Ionic Liquids. Separ. Purif. Sci. Technol. 2012, 97, 228-234. (64) McCrary, P. D.; Rogers, R. D. 1-Ethyl-3-methylimidazolium Hexafluorophosphate: From Ionic Liquid Prototype to Antitype. Chem. Commun. 2013, 49, 6011-6014. (65) Paduszynski, P.; Domanska, U. Viscosity of Ionic Liquids: An Extensive Database and a New Group Contribution Model Based on a Feed-Forward Artificial Neural Network. J. Chem. Inf. Model. 2014, 54, 1311-1324. (66) Abbott, A. P.; Boothby, D.; Capper, G.; Davies, D. L.; Rasheed, R. K. Deep Eutectic Solvents Formed Between Choline Chloride and Carboxylic Acids: Versatile Alternatives to Ionic Liquids. J. Am. Chem. Soc. 2004, 126, 9142-9147. (67) Wagle, D. V.; Zhao, H.; Baker, G. A. Deep Eutectic Solvents: Sustainable Media for Nanoscale and Functional Materials. Acc. Chem. Res. 2014, 47, 2299-2308. (68) Mbous, Y. P.; Hayyan, M.; Wong, W. F.; Looi, C. Y.; Hashim, M. A. Unraveling the Cytotoxicity and Metabolic Pathways of Binary Natural Deep Eutectic Solvent Systems. Scientific Reports 2017, 7, 41257. 14

ACS Paragon Plus Environment

Page 14 of 37

Page 15 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(69) Paiva, A.; Craveiro, R.; Aroso, I.; Martins, M.; Reis, R. L.; Duarte, A. R. C. Natural Deep Eutectic Solvents – Solvents for the 21st Century. ACS Sustainable Chem. Eng. 2014, 2, 1063-1071. (70) Dai, Y.; van Spronsen, J.; Witkamp, G. J.; Verpoorte, R.; Choi, Y. H. Natural Deep Eutectic Solvents As New Potential Media for Green Technology. Anal. Chim. Acta 2013, 766, 61-68. (71) Pena-Pereira, F.; Namiesnik, J. Ionic Liquids and Deep Eutectic Mixtures: Sustainable Solvents for Extraction Processes. ChemSusChem 2014, 7, 1784-1800. (72) Abo-Hamad, A., Hayyan, M.; AlSaadi, M. A., Hashim, M. A. Potential Applications of Deep Eutectic Solvents in Nanotechnology. Chem. Eng. J. 2015, 273, 551-567. (73) Ge, X.; Gu, C.; Wang, X., Tu, J. Deep Eutectic Solvents (DESs)-Derived Advanced Functional Materials for Energy and Environmental Applications: Challenges, Opportunities, and Future Vision. J. Mater. Chem. A 2017, 5, 8209-8229. (74) Lee, J. S. Deep Eutectic Solvents as Versatile Media for the Synthesis of Noble Metal Nanomaterials. Nanotechnol. Rev. 2017, 6, 271-278. (75) Wang, R.; Hua, Y.; Zhang, Q. Electrochemical Preparation of Copper Nanoparticles in Choline ChlorideUrea Deep Eutectic Solvent. ECS Trans. 2014, 59, 505-511. (76) Söldner, A.; Zach, J.; Iwanow, M.; Gärtner, T.; Schlosser, M.; Pfitzner, A.; König, B. Preparation of Magnesium, Cobalt and Nickel Ferrite Nanoparticles from Metal Oxides using Deep Eutectic Solvents. Chem. Eur. J. 2016, 22, 13108-13113. (77) Huang, Y., Shen, F., La, J.; Luo, G.; Lai, J.; Liu, C.; Chu, G. Synthesis and Characterization of CuCl Nanoparticles in Deep Eutectic Solvents. Particul. Sci. Technol. 2013, 31, 81-84. (78) Raghuwanshi, V. S.; Ochman, M.; Hoell, A.; Polzer, F.; Rademann, K. Deep Eutectic Solvents for the SelfAssembly of Gold Nanoparticles: a SAXS, UV-Vis, and TEM Investigation. Langmuir 2014, 30, 6038-6046. (79) Tian, P. Molecular Dynamics Simulations of Nanoparticles. Annu. Rep. Prog. Chem., Sect. C: Phys. Chem. 2008, 104, 142-164. (80) Milano, G.; Santangelo, G.; Ragone, F.; Cavallo, L.; Di Matteo, A. Gold Nanoparticle/Polymer Interfaces: All Atom Structures from Molecular Dynamics Simulations. J. Phys. Chem. C 2011, 115, 15154-15163. (81) Milek, T.; Zahn, D. Molecular Simulation of Ag Nanoparticle Nucleation from Solution: Redox-Reactions Direct the Evolution of Shape and Structure. Nano Lett. 2014, 14, 4913-4917. (82) Morrison, H. G.; Sun, C. C.; Neervannan, S. Characterization of Thermal Behavior of Deep Eutectic Solvents and Their Potential as Drug Solubilization Vehicles. Int. J. Pharm. 2009, 378, 136-139. (83) Singh, B. S.; Lobo, H. R.; Shankarling, G. S. Choline Chloride Based Eutectic Solvents: Magical Catalytic System for Carbon–Carbon Bond Formation in the Rapid Synthesis of β-Hydroxy Functionalized Derivatives. Catal. Commun. 2012, 24, 70-74. (84) Harvey, M. J.; Giupponi, G.; De Fabritis, G. ACEMD: Accelerating Biomolecular Dynamics in the Microsecond Time Scale. J. Chem. Theory Comput. 2009, 5, 2371-2377. (85) Atilhan, M.; Costa, L. T., Aparicio, S. On the Behaviour of Aqueous Solutions of Deep Eutectic Solvents at Lipid Biomembranes. J. Mol. Liq. 2017, 247, 116-125. (86) Heinz, H.; Vaia, R. A.; Farmer, B. L.;Naik, R. R. Accurate Simulation of Surfaces and Interfaces of FaceCentered Cubic Metals Using 12-6 and 9-6 Lennard – Jones Potentials. J. Phys. Chem. C 2008, 112, 1728117290. (87) Gezelter, J. D.; Michalka, J.; Kuang, S.; Marr, J.; Stocker, K.; Li, C.; Vardeman, C. F.; Lin, T.; Fennell, C. J.; Sun, X.; et al. OpenMD, an Open Source Engine for Molecular Dynamics. Available at http://openmd.org. (accessed April 1, 2018). (88) Haas, P.; Tran, F.; Blaha, P. Calculating of the Lattice Constant of Solids with Semilocal Functionals. Phys. Rev. B 2009, 79, 085104. (89) Martínez, L.; Andrade, R.; Birgin, E. G.; Martínez, J. M. PACKMOL: A Package for Building Initial Configurations for Molecular Dynamics Simulations. J. Comput. Chem. 2009, 30, 2157-2164. (90) Izaguirre, J. A.; Catarello, D. P.; Wozniak, J. M.; Skeel, R. D. Langevin Stabilization of Molecular Dynamics. J. Chem. Phys. 2001, 114, 2090–2098. (91) Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R. Molecular Dynamics with Coupling to an External Bath. J. Chem. Phys. 1984, 81, 3684–3690.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(92) Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald: An N⋅log(N) Method for Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089–10092. (93) Atilhan, M.; Aparicio, S. Behavior of Deep Eutectic Solvents under External Electric Fields: A Molecular Dynamics Approach. J. Phys. Chem. B 2017, 121, 221-232. (94) Hammond, O. S.; Bowron, D. T.; Edler, K. J. Liquid Structure of the Choline Chloride-Urea Deep Eutectic Solvent (Reline) From Neutron Diffraction and Atomistic Modelling. Green Chem. 2016, 18, 2736-2744.

TOC Graphic

16

ACS Paragon Plus Environment

Page 16 of 37

Page 17 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure Captions.

Figure 1. Center-of-mass radial distribution functions, g(r), for the reported molecular pairs for pure ChCl:URE and for the ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 2. (a to c) Center-of-mass radial distribution functions, g(r), and (d to f) the corresponding solvation numbers, N, for the reported pairs for the systems of ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar. In panels (a to c) curves are shifted to improve visibility. r stands for the distance to the center-of-mass of the MNPs.

Figure 3. Relevant ratios of number of molecules, Ni, in solvation spheres around central MNPs as a function of the radius of the sphere, r, defined from the center-of-mass of MNPs. Dashed lines show values corresponding to stoichiometric ratios in bulk DES. The systems are formed by ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 4. Orientation angles, φ and δ, as a function of the distance, r, between MNPs and (a) URE and (b) Ch, for vectors defined in the Figure. The systems are formed by ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 5. (a) Distribution of molecules and (b) snapshot of URE molecules around a central Au4 MNP, for molecules inside MNP solvation sphere with r = 12 Å. The systems are formed by ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 6. Ratio of self-diffusion coefficients, D, for molecules in the first solvation sphere around MNPs (r < 12 Å) and for those in bulk liquid phase. The systems are formed by ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 7. Intermolecular interaction energies, Einter, for MNP-molecule pairs. The systems are formed by ChCl:URE containing one icosahedral MNP. Values at 300 K and 1 bar.

Figure 8. Average number of hydrogen bonds per molecule between the reported molecular pairs in systems containing ChCl:URE and one icosahedral MNP. The criterion for defining hydrogen bond was 3.5 Å and 60⁰ for donor – acceptor separation and angle. Values are calculated per Cl molecule (URE-Cl pairs) or per Ch molecule, for the remaining interactions. Values at 300 K and 1 bar.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9. (a to c) Center-of-mass radial distribution functions, g(r), and (d to f) the corresponding solvation numbers, N, for the reported pairs for the systems of ChCl:HBD (where HBD stands for URE, EG, GLY, MAL, LEV or PAA) containing one icosahedral Au4 MNP. Values at 300 K and 1 bar. In panels (a to c) curves are shifted to improve visibility. r stands for the distance to the center-of-mass of the Au4.

Figure 10. Relevant ratios of number of molecules, Ni, in solvation spheres around central Au4 as a function of the radius of the sphere, r, defined from the center-of-mass of Au4. Dashed lines show values corresponding to stoichiometric ratios in bulk DES (NHBD / NCH = 2 for all the DES except when HBD = MAL). The systems are formed by ChCl: HBD (where HBD stands for URE, EG, GLY, MAL, LEV or PAA) containing one icosahedral Au4 MNP. Values at 300 K and 1 bar.

Figure 11. Distribution of molecules around a central icosahedral Au4 MNP, for molecules inside Au4 solvation sphere with r = 12 Å. The systems are formed by ChCl:HBD (where HBD stands for URE, EG, GLY, MAL, LEV or PAA) containing one icosahedral Au4 MNP. Values at 300 K and 1 bar.

Figure 12. Intermolecular interaction energies, Einter, for Au4-molecule pairs. The systems are formed by ChCl:HBD (where HBD stands for URE, EG, GLY, MAL, LEV or PAA) containing one icosahedral Au4 MNP. Values at 300 K and 1 bar.

Figure 13. Orientation angle, φ, as a function of the distance, r, between Au4 center-of-mass and HBDs. The systems are formed by ChCl:HBD containing one icosahedral Au4 MNP. Values at 300 K and 1 bar. Vectors are defined in Figure S3 (Supporting Information).

Figure 14. Average number of hydrogen bonds per molecule between the reported molecular pairs in systems containing ChCl:HBD and one icosahedral Au4 MNP. The criterion for defining hydrogen bond was 3.5 Å and 60⁰ for donor – acceptor separation and angle. Values are calculated per Cl molecule (URE-Cl pairs) or per Ch molecule, for the remaining interactions. Values at 300 K and 1 bar.

Figure 15. (a to c) Center-of-mass radial distribution functions, g(r), and (d to f) the corresponding solvation numbers, N, for the reported pairs for the systems of ChCl:URE containing one icosahedral Au MNP with different sizes (Au4, Au6 and Au8). Values at 300 K and 1 bar. r stands for the distance to the center-of-mass of the Au MNPs.

Figure 16. (a to c) Center-of-mass radial distribution functions, g(r), and (d to f) the corresponding solvation numbers, N, for the reported pairs for the systems of ChCl:URE containing one icosahedral Au6 or Au MNP with spherical shape, Au_sph. Values at 300 K and 1 bar. r stands for the distance to the center-of-mass of the Au MNPs.

18

ACS Paragon Plus Environment

Page 18 of 37

Page 19 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 17. Time evolution of the reported distances between the centers-of-mass of icosahedral Au4 MNPs in ChCl:URE. Values at 300 K and 1 bar.

Figure 18. Time evolution of the minimum distance between the centers-of-mass of two MNPs (the remaining MNP-MNP distances are larger than those reported in each case). (a) Systems with ChCl:URE and different icosahedral MNPs and (b) systems containing icosahedral Au4 MNPs and different ChCl:HBD. Values at 300 K and 1 bar.

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry

4

pure ChCl:URE Au4 Ag4 Pd4 Pt4 Ni4

(a) Ch-Cl

g(r)

3 2 1 0 6 0

(b) Cl-URE

5

10

15

10

15

10

15

r/Å

g(r)

4

2

0 2.5 0

(c) Ch-URE

5 r/Å

2

g(r)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 37

1.5 1 0.5 0 0

5

r/Å Figure 1.

20

ACS Paragon Plus Environment

Page 21 of 37 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 37

r = 12 Å URE/CH r = 15 Å

12

(a) N URE / N Ch

10 8 6 4 2 0 Au4

Ag4

Pd4 CH/Cl

Pt4

Ni4

1.5

(b) N Ch / N Cl 1.0

0.5

0.0 Au4

Ag4

Pd4

Pt4

Figure 3.

22

ACS Paragon Plus Environment

Ni4

Page 23 of 37

180

O H

2N

(a) MNP - URE

ϕ / deg

φ NH

2

MNP MNP

135

90

45 180 H3C

OH

+

10

(b) MNP - Ch

15

20

15

20

r/Å

N H3C

δ

CH3

MNP MNP

δ / deg

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

135

90

45 10

r/Å Figure 4.

23

ACS Paragon Plus Environment

Au4 Ag4 Pd4 Pt4 Ni4

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5.

24

ACS Paragon Plus Environment

Page 24 of 37

Page 25 of 37

60 URE Ch

100 × (Dr