Molecular Simulation Hydrogen-bond dynamics of

0 downloads 0 Views 2MB Size Report
Oct 14, 2014 - Hydrogen-bond dynamics of water in presence of an amphiphile, ... Various experimental and simulation studies have suggested that the ...
This article was downloaded by: [S N Bose National Centre for Basic] On: 06 April 2015, At: 23:15 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Molecular Simulation Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/gmos20

Hydrogen-bond dynamics of water in presence of an amphiphile, tetramethylurea: signature of confinement-induced effects a

Sandipa Indra & Ranjit Biswas

a

a

Department of Chemical, Biological and Macromolecular Sciences, S.N. Bose National Centre for Basic Sciences, JD Block, Sector III, Salt Lake, Kolkata 700 098, India Published online: 14 Oct 2014.

Click for updates To cite this article: Sandipa Indra & Ranjit Biswas (2015) Hydrogen-bond dynamics of water in presence of an amphiphile, tetramethylurea: signature of confinement-induced effects, Molecular Simulation, 41:5-6, 471-482, DOI: 10.1080/08927022.2014.965705 To link to this article: http://dx.doi.org/10.1080/08927022.2014.965705

PLEASE SCROLL DOWN FOR ARTICLE Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content. This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http:// www.tandfonline.com/page/terms-and-conditions

Molecular Simulation, 2015 Vol. 41, Nos. 5 – 6, 471–482, http://dx.doi.org/10.1080/08927022.2014.965705

CONFINED FLUIDS Hydrogen-bond dynamics of water in presence of an amphiphile, tetramethylurea: signature of confinement-induced effects Sandipa Indra and Ranjit Biswas* Department of Chemical, Biological and Macromolecular Sciences, S.N. Bose National Centre for Basic Sciences, JD Block, Sector III, Salt Lake, Kolkata 700 098, India

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

(Received 30 May 2014; final version received 9 September 2014) Various experimental and simulation studies have suggested that the presence of amphiphilic molecules in aqueous solutions substantially perturbs the tetrahedral hydrogen-bond (H-bond) network of neat liquid water. Such structural perturbation is expected to impact H-bond lifetime of liquid water. Tetramethylurea (TMU) is an example of an amphiphile because it possesses both hydrophobic and hydrophilic moieties. Molecular dynamics simulations of (water þ TMU) binary mixtures at various compositions have been performed in order to investigate the microscopic mechanism through which the amphiphiles influence the H-bond dynamics of liquid water at room temperature. Present simulations indicate lengthening of both water – water H-bond lifetime and H-bond structural relaxation time upon addition of TMU in aqueous solution. At the highest TMU mole fraction studied, H-bond lifetime and structural relaxation time are, respectively, , 4 and , 8 times longer than those in neat water. This is comparable with the slowing down of H-bond dynamics for water molecules confined in cyclodextrin cavities. Simulated relaxation profiles are multi-exponential in character at all mixture compositions, and simulated radial distribution functions suggest enhanced water – water and water – TMU interactions upon addition of TMU. No evidence for complete encapsulation of TMU by water H-bond network has been found. Keywords: H-bond dynamics; amphiphiles; aqueous solutions; molecular dynamics simulations

1.

Introduction

Study of confined water is an important area of research because water shows significant changes [1–15] in both structural and dynamical properties upon encapsulation. Fluorescence spectroscopic measurements and simulation studies seem to suggest drastic decrease of static dielectric constant (10 ) of bulk water upon confinement.[16–19] Confined water is found in biology such as in lipid bilayers, ion channels and protein cavities [20–23]; ‘biological water’ at protein surfaces [24]; in chemistry such as in fuel cells,[25] gels and reverse micelles [13,26–30]; in technology such as in carbon nanotubes,[31–33] granular and porous materials, [34] etc. Computer simulations on the effects of confinement on structure and dynamics of binary liquid mixtures composed of asymmetric particles [35] have attempted to provide answers to some basic questions of this area. Behaviour of water molecules in the vicinity of hydrophobic solute molecules is an issue of intense research. It has been found from both experimental and theoretical studies that reorientation dynamics of water molecules in the hydration shell of hydrophobes exhibits substantial retardation.[36–43] Dynamic anisotropy measurements of water around a hydrophobic molecule [32] reveal that water molecules become immobilised in the hydration shell of hydrophobic solute, supporting the controversial iceberg

*Corresponding author. Email: [email protected] q 2014 Taylor & Francis

model.[44] Interestingly, a combined computer simulation and theoretical study [38] has suggested no immobilisation of water molecules in the hydrophobic hydration layer and thus debated the validity of the iceberg model. Soluteexcluded volume effects at the transition state are found to be responsible for the slowing down of water reorientation. Concentrated aqueous solutions are found to behave like confining environments due to the steric effects rendered by the solute molecules.[45,46] Hydrophobic molecules tend to aggregate in solution. At higher solute concentrations, these aggregates distribute throughout the solution. Consequently, water molecules are forced to reside by the side of the surfaces created by these aggregated moieties. This results in significant retardation of reorientational and translational dynamics of interfacial water molecules residing at the surface of the solute aggregates.[47] By this way, water shows similar behaviour in concentrated aqueous solutions as found in explicit confinements. The extensive hydrogen-bond (H-bond) network of water plays a key role in determining the structure and dynamics of aqueous systems. Therefore, experiments, [48 – 52] theory and simulations [53 –60] have been employed to gain insight into the mechanism of both H-bond kinetics and dynamics in water and aqueous solutions. Computer simulation study [53] first showed the

472

S. Indra and R. Biswas O H3C

CH3 N

N H3C

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

Scheme 1.

CH3

Chemical structure of TMU.

non-exponential behaviour of the dynamics of forming and breaking H-bonds in liquid water due to the coupling between H-bond population and diffusion. Later, the jump reorientation mechanism [56] showed that water reorientation mechanism involves large amplitude angular jumps, rather than the commonly accepted sequence of small diffusive steps. Determination of H-bond lifetime is also a very important tool to examine the strength of H-bonds. This is largely affected by the structure of the local environment. It is important to mention here that very few attempts [61,62] have been made to calculate H-bond fluctuation dynamics for binary aqueous mixtures. Tetramethylurea (TMU) (shown in Scheme 1) is an amphiphilic molecule possessing both hydrophobic methyl groups and hydrophilic carbonyl group. TMU is easily soluble in water via H-bond interactions. TMU is an important biologically relevant molecule because it is a stronger denaturing agent than urea itself.[63,64] In this article, we show, through molecular dynamics simulations of TMU in water at various composition mixtures, that the confining effects of TMU are transduced on H-bond fluctuation dynamics of water. We have found that the longer lived water – water and water – TMU Hbonds exist in higher TMU concentrations. We have also observed that TMU induces significant influence on local environments that regulate the H-bond dynamics. The organisation of the rest of the paper is as follows. Section 2 contains the molecular model and force field information. Simulation details are provided in Section 3. Simulation results and discussion are provided in Section 4. The paper ends with conclusions in Section 5.

2.

Model and force field

Here we have considered the SPC/E model of water [65] and AMBER-type force field for TMU. The form of the force field is as follows: 2 X  2 X  U ¼ K bij bij 2 b0ij þ K uijk uijk 2 u0ijk bij

uijk

X K fijkl    1 þ cos nfijkl 2 g þ fsc fijkl ( " )  6 # N X sij 12 sij qi qj 41ij 2 þ þ : r ij r ij r ij i,j

ð1Þ

Here, U is the total potential energy of the system. The parameters b; b 0 ; u; u 0 ; f are actual bond length, equilibrium bond length, actual bond angle, equilibrium bond angle and dihedral angle, respectively. K b ; K u ; K f are the force constants for bond, angle and dihedral angle, respectively. fsc denotes the factor by which torsional barrier is divided. g is the phase shift angle and n the periodicity. The non-bonded interactions are included by the Lennard-Jones (LJ) potential.[66] The LJ parameters 1; s; r are the potential well depth, van der Waals radius and distance between atoms, respectively. The parameter q which represents the partial charge of the atom is taken to account for the Coulomb interaction. LJ interactions between unlike atoms are calculated by Lorentz – Berthelot combining rule,[66] pffiffiffiffiffiffiffiffi where 1ij ¼ 1i 1j and ðsij ¼ si þ sj Þ=2.

3.

Simulation details

Molecular dynamics simulations were performed using DL_POLY_Classic [67] package. We have used DL_FIELD [68] to make the input files in DL_POLY format. In this study, the methyl groups of TMU are treated by the united atom approach to avoid the complication of the simulation. All the force field parameters for TMU are taken from the literature.[69,70,71] A total of 512 molecules in a cubic box with periodic boundary condition have been considered. To obtain a reasonable initial configuration, we have run all the simulations in NPT ensemble at 1 atm pressure and 300 K using Hoover thermostat and barostat with relaxation times of 2 and 0.5 ps, respectively. We have run a total of 2 ns simulation in NPT ensemble to attain the experimental density of each system. Then, we have used the final configuration of NPT ensemble as the initial configuration for running simulations in NVT ensemble at 300 K using the Nose´ –Hoover [72,73] thermostat with relaxation time of 0.5 ps. We have further equilibrated the system for a nanosecond to allow complete equilibration and then collected a 4 ns production run. Equations of motion were subsequently solved by leapfrog–Verlet algorithm [66] with time step of 1 fs. Verlet neighbour list shell width was set to ˚ . Trajectories were saved after each 10 fs. Cut-off radius 1.2 A was set nearly to half of the box length. The Shake algorithm [74] was used to constrain all bonds involving hydrogen atoms. Electrostatic interactions were treated via the conventional Ewald summation technique.[66] In addition, we have used the TRANAL utility of MDynaMix program [75] to calculate the static dielectric constant of the SPC/E water and TMU. 4.

Results and discussion

4.1 Validity check of the force field Simulated densities at various compositions of the (water þ TMU) mixtures at 1 atm pressure and 300 K

Molecular Simulation 5

1.02

4

1.00 ρ (g/cc)

XTMU

O(W)-H(W)

T=300K

Simulations, T = 300 K Experiments, T ~ 298 K

XTMU=0.3

3 0.98 2 0.96

0.94

0.2

0.4

0.6

0.8

0 5

1.0

T=300K

XTMU

O(TMU)-H(W)

XTMU

3 g(r)

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

4

Figure 1. (Colour online) A comparison between simulated and experimental TMU mole fraction (X TMU )-dependent density of (water þ TMU) binary mixture at 300 K and 1 atm pressure. Experimental densities are from Refs [74,75].

2

temperature are compared with those from experiments [76,77] in Figure 1. Good agreement between composition-dependent simulated and experimental densities indicates the validity of the force field employed here. We have also calculated 10 for water and TMU from simulated collective moment fluctuations and compared with the relevant experimental data. The following relation [78] has been employed to calculate 10 : i 4p h 2 ð2Þ kM l 2 kMl2 ; 3kB TV P with the collective moment, M ¼ i m i , being the sum over individual dipole moments, mi , V the volume and kB T Boltzmann constant times the absolute temperature. Our calculated 10 for water at 300 K temperature is , 73 and agrees well with experimental [79,80] and simulation [81] results. In contrast, the calculated 10 for TMU is found to be , 13 which is , 60% of the experimental value.[82] This discrepancy between simulation and experiment may arise partly from the united atom description for methyl groups in TMU and partly from system size. Note system size is an important factor in simulations for determining 10 as the latter is connected to the collective (k ! 0; k being the wavenumber) solvent modes. However, the H-bond dynamics that we intend to investigate is governed largely by the local environment and, therefore, consideration of 512 particles is expected to provide a semi-quantitative description of compositiondependent dynamics of this binary mixture. 10 ¼ 1 þ

4.2

Radial distribution functions

The relative structural arrangements of the atoms involved in H-bond interactions in the mixture are shown in Figure 2, where the simulated radial distribution functions

0.3 0.2 0.1 0.07 0.05 0.03 0.01 0.0

XTMU=0

1

0.0

473

XTMU=0.3

0.3 0.2 0.1 0.07 0.05 0.03 0.01

1 XTMU=0.01

0 1 16

2 T=300K

3

4

Me(TMU)-Me(TMU)

XTMU

12 XTMU=0.01

8

0.01 0.03 0.05 0.07 0.1 0.2 0.3

4

0

XTMU=0.3

3

4

5

6

r (Å)

Figure 2. (Colour online) X TMU -dependent representative sitesite RDFs, gðrÞ, for various atomic pairs as a function of distance, r. Relevant atomic sites are indicated in each of the panels. Inset of each plot shows the atomic sites. Colour code for gðrÞ shown in these panels is as follows: neat water: solid black; 0.01, red dot; 0.03, blue long dash; 0.05, pink short dash; 0.07, dark green long dash; 0.1, black dash-dot; 0.2, red short dash; 0.3, dark green dot.

(RDFs, gðrÞ) at several TMU mole fractions (X TMU ) are shown as a function of distance, r. While the upper panel depicts the simulated gðrÞ for oxygen and hydrogen atoms of water (O(W) –H(W)), the RDFs for water hydrogen and TMU oxygen atoms (O(TMU)–H(W)) are shown in the middle panel. At X TMU ¼ 0 (neat water), gðrÞ for O(W) –H ˚ , a distance nearly the (W) exhibits its first peak at , 1.7 A sum of the radii of the atoms involved. Expectedly, the ˚ ) approximately second peak appears at a distance (, 3.3 A equal to the sum of the diameters of these atoms. Note that the O(W) –H(W) gðrÞ peak intensity increases with the

S. Indra and R. Biswas

increase of X TMU in the aqueous mixture, but the peak position remains nearly unaltered. This behaviour suggests that addition of TMU leads to enhanced interaction between oxygen and hydrogen atoms of water molecules. Interestingly, gðrÞ shown in the middle panel also indicate increased interaction between water hydrogen and TMU oxygen upon addition of TMU in water. Growth of both O(W) – H(W) and O(TMU) – H(W) RDFs with increase of TMU concentration suggests that water and TMU molecules arrange themselves in the binary mixture in such a way that the H-bonding interaction between water molecules and that between water and TMU molecules are facilitated upon addition of this amphiphile. This behaviour was observed earlier also for DMSO – water binary mixtures [44] and attributed to structural percolation. It is interesting to note in these panels that the second peak intensity of O(TMU)– H(W) gðrÞ is much less than that of O(W) – H(W), suggesting more number of water molecules in the second solvation shell of water oxygen than that for TMU oxygen. The presence of methyl groups may be the reason for this relatively dwindled population of water molecules. This fact notwithstanding these RDFs does indicate successive enhancement of water – water and water – TMU interactions upon increase of X TMU in the aqueous mixture. This means weakening of interaction among TMU molecules in the presence of water. This is indeed the scenario as evidenced in the lower panel of Figure 2 which depicts H3C – CH3 gðrÞ for TMU molecules in this mixture. Note in this panel the gðrÞ peak intensity decreases steadily and exhibits a trend opposite to that observed in the other two panels above. Importantly, this also indicates no aggregation among TMU molecules via methyl – methyl interaction even at large X TMU . However, the enhanced interactions between water – water and water – TMU suggest decrease of translational diffusion coefficients of these molecules with X TMU . This is what we have investigated next.

Simulated MSDs at various X TMU are presented in Figure 3 2 with upper panel depicting kjDrðtÞj l for oxygen atom of water (upper panel) and lower panel for oxygen atom of TMU. Here we have presented MSD as a function of time in double logarithmic scale in order to show the t 2 and t 1 2 dependencies of kjDrðtÞj l at short and long times, respectively. These dependencies are indicated in these plots by ‘Slope I’ and ‘Slope II’. Note that there is no evidence of ‘rattling in a cage’ motion [85 –87] although simulated gðrÞ suggested stronger interactions between water and TMU molecules through H-bonding. Therefore, caging of TMU, even at low concentrations, by water molecules is probably non-existent in these aqueous binary mixtures. Mixture composition dependence of viscosity (h) and translational diffusion (D) coefficients for water and TMU are investigated next. In simulations, h has been obtained by using the Green –Kubo relation,[83,84] ð V 1 h¼ kPab ðtÞ·Pab ð0Þldt; ð5Þ kB T 0 where a; b ¼ x; y; z. Pab denotes the off-diagonal terms ða – bÞ of the pressure tensor. Figure 4 presents a 100

T = 300 K

Water

pe

Slo

II

10

XTMU 0.00 0.03 0.07 0.10 0.20 0.30 0.50 0.75

1

pe

0.1 (Å2)

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

474

I

o Sl

0.01 100 T = 300 K

TMU

pe

Slo

II

10

4.3 Mean squared displacements The translational self-diffusion coefficient (D) of a particle moving in a fluid can be calculated from the mean squared displacements (MSDs) of centre-of-mass position vectors r i ðtÞ as follows: 2

kjDrðtÞj l ¼

E 1 DX 2 jr i ðtÞ 2 r i ð0Þj ; N

ð3Þ

XTMU 0.03 0.07 0.10 0.20 0.30 0.50 0.75

1

0.1

0.01 0.01

eI

op

Sl

0.1

1

10

100

t (ps) 2

with kjDrðtÞj l denoting the MSD. Subsequently, D has 2 been calculated from the simulated kjDrðtÞj l at long time as follows [83,84]: 1 2 ðkjDrðtÞj lÞ D¼ : ð4Þ 6t t!1

Figure 3. (Colour online) X TMU -dependent simulated MSDs for water (upper panel) and TMU (lower panel) in (water þ TMU) binary mixture. Colour code denotes the following: neat water, solid black; 0.03, red long dash; 0.07, blue short dash; 0.1, pink dash-dot; 0.2, dark green dash-dot-dot; 0.3, cyan long dash; 0.5, dark red short dash; 0.75, grey long dash.

Molecular Simulation

viscosity,[76,89] a much sharper inverted parabolic dependence for D on composition for this mixture. Slip hydrodynamic predictions for D using both simulated (broken lines) and experimental (solid lines) viscosities are also shown in the same figure to highlight the difference between the simulations and hydrodynamic predictions. The difference between the simulated and predicted values may appear from the non-hydrodynamic character of the particle movement in these binary mixtures. Measurements of D for these mixtures are, therefore, necessary to ascertain whether the X TMU dependence of simulated D originates from partial decupling of diffusion from viscosity [90 – 94] in these mixtures or the observed dependence is simply a reflection of the non-quantitativeness of the model interaction potentials used for accurate representation of the complex interactions in these aqueous binary mixtures.

6 Simulated Experimental

5

η (cP)

4 T=

298

3

K

T=

2

300

K

1 0

0.0

0.2

0.4

0.6

0.8

1.0

T = 300K HO

2 D calc (ηsimu)

D/10–5cm2s–1

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

3

2 HO

2 D calc (ηexp)

4.4

1 TMU

D calc (ηsimu) TMU

D calc (ηexp)

0

0.0

0.2

0.4

0.6

475

0.8

XTMU

Figure 4. (Colour online) Composition dependence of viscosity (upper panel) and diffusion coefficients (lower panel) for water and TMU in (water þ TMU) binary mixtures at 300 K. Simulated viscosity coefficients (h) are shown by circles with error bars. In the lower panel, filled black circles represent simulated diffusion coefficients (D) for water and filled red triangles denote the simulated D for TMU. While the solid lines represent the slip hydrodynamic predictions by using the experimental h, the dashed lines denote the same but using the simulated h. For hydrodynamic calculations, radii of water and TMU molecules used are 1.425 and ˚ , respectively. Star denotes the simulated diffusion 3.5 A coefficient for SPC/E water at 300 K reported in Ref. [79].

comparison between the simulated and experimental [76] h for this mixture at , 300 K. Clearly, the agreement is semi-quantitative, indicating insufficiency of the model interaction potentials to represent quantitatively the interactions present in such aqueous binary mixtures, [81,88] and formation of aqueous complex through H-bond interaction.[89] The lower panel of Figure 4 shows the composition dependence of the simulated D for water and TMU for this mixture at 300 K. Simulated D for SPC/E water at , 300 K by other authors [81,88] is also shown to ensure fidelity of the present simulations. Note that D for both the specie first decreases with X TMU and then increases after a shallow minimum. Interestingly, one expects, from X TMU dependencies of medium density (see Figure 1) and

Dynamic heterogeneity in (water 1 TMU) mixtures

Dynamic heterogeneity (DH) can impart non-hydrodynamic character to the particle movements, and can be understood in terms of non-Gaussian (NG) and new non-Gaussian (NNG) parameters, and single particle displacement distributions.[95 – 97] The translational NG parameter (a2 ) can be obtained from the simulated MSD by using the following relation:

a2 ðtÞ ¼

3kdr 4 ðtÞl 5kdr 2 ðtÞl2

2 1;

ð6Þ

P 2 whereP kdr 2 ðtÞl ¼ kN 21 Ni¼1 jDrðtÞj l and kdr 4 ðtÞl ¼ 4 kN 21 Ni¼1 jDrðtÞj l with dr denoting the single particle displacement. DrðtÞ ¼ r i ðtÞ 2 r i ð0Þ. Note that for homogeneous systems (for example, hot liquids), a2 ðtÞ ¼ 0 for both at t ¼ 0 and t ¼ 1 but a2 ðtÞ ¼ 0:2 at intermediate times.[90] Therefore, systems possessing DH would be characterised by a non-monotonic time dependence of a2 ðtÞ with peak higher than 0.2. This peak signifies a DH timescale which is designated as tNG . Interestingly, another much slower DH timescale may also exist in the system which can be accessed via calculating the NNG parameter (gðtÞ) as follows [95]:

 1 1 gðtÞ ¼ kdr 2 ðtÞl 2 1; 3 dr 2 ðtÞ

ð7Þ

P 2 with k1=dr 2 ðtÞl ¼ k1=N Ni¼1 1=jDrðtÞj l. Clearly, gðtÞ is dominated by those particles which execute smaller displacements and contribute to a DH timescale (tNNG ) slower than that reflected by tNG . For systems possessing no spatially varying relaxation rates (that is, DH) tNG ¼ tNNG , and a2 ðtÞ and gðtÞ curves would overlap.

476

S. Indra and R. Biswas

The DH signature can also be revealed by studying the single-particle displacement distribution as follows [95–97]:

0.30

ð8Þ

0.25

where Gs ðdr; tÞ denotes the self-part of the van Hove correlation function.[83,84] Note for a Gaussian Gs ðdr; tÞ, P½log10 ðdrÞ; t becomes independent of time with a peak height of 2.13. Any deviation from the Gaussian shape and peak height would therefore indicate presence of DH in the system and may be interpreted as the origin for the partial decoupling of diffusion from solution viscosity. Figure 5 shows a2 ðtÞ and gðtÞ simulated for water at four different mixture compositions at 300 K. Note that these compositions (that is XTMU ¼ 0.01, 0.2, 0.3 and 0.5) are chosen to represent dilute and concentrated aqueous solutions of TMU, including the composition that corresponds to the maximum in experimental and simulated h. Simulated a2 ðtÞ and gðtÞ for neat water at 300 K, shown in Figure S1 of Supplementary information, depict a non-monotonic dependence with a peak value of , 0.2 at intermediate times, and the peak times (tNG and tNNG ) nearly overlap on each other. However, as XTMU increases, the peak value increases and tNG deviates from tNNG . The appearance of these two ‘slow’ timescales clearly demonstrates the presence of DH in the system. Therefore, partial decoupling between D and h for water noted in Figure 4 at higher TMU concentration may be attributed to the DH of these aqueous binary mixtures. Figure 6 depicts the simulated P½log10 ðdrÞ; t for water at these representative mixture compositions at tNG and tNNG . Corresponding distributions for a Gaussian Gs ðdr; tÞ are also shown for comparison. P½log10 ðdrÞ; t at these DH timescales for ambient water is provided in Figure S2 (Supplementary information). Simulated distributions shown in these panels clearly demonstrate enhancement of DH for water upon increasing TMU concentration in these mixtures. This explains the partial decoupling of solution viscosity from water diffusion observed in Figure 4. Interestingly, the simulated P½log10 ðdrÞ; t at tNG and tNNG for ambient water shows a slight deviation from the Gaussian character, which probably explains our simulated diffusion coefficient for neat water being , 10% larger than that reported in experiments (, 2.3 £ 1025 cm2 s21).

0.15

4.5

H-bond dynamics

Simulated gðrÞ presented in Figure 2 has already revealed pronounced interaction between water and TMU in aqueous mixture, and this is expected to deeply impact the H-bond dynamics of water in solution. We have employed the following geometric definition of H-bonding between a pair of molecules [57 –60]:

T=300K XTMU=0.01

0.20 0.10 0.05 0.00 0.5 0.4

T=300K XTMU=0.2

0.3 0.2 0.1 α2(t), γ(t)

P½log10 ðdrÞ; t ¼ lnð10Þ4pdr 3 Gs ðdr; tÞ;

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

0.35

0.0 0.5 0.4

T=300K XTMU=0.3

0.3 0.2 0.1 0.0 0.6 0.5

T=300K XTMU=0.5

0.4 0.3 0.2 0.1 0.0 0.1

1

10

100

t (ps)

Figure 5. (Colour online) Simulated a2 ðtÞ and gðtÞ for water in (water þ TMU) binary mixtures at a few representative mixture compositions at 300 K. Black curve denotes a2 ðtÞ and the red curve denotes gðtÞ. Vertical broken lines indicate the peak times,tNG and tNNG , for a2 ðtÞ and gðtÞ, respectively. Horizontal line indicates the value of 0.2.

1. Distance between two oxygen atoms must be less ˚. than 3.5 A 2. Angle between OZO bond vector and OZH bond vector must be less than 308. ˚. 3. Distance between O and H must be less than 2.45 A How long a pair of molecules remains H-bonded can be calculated by two different time-correlation functions. [23 – 26] The relaxation of H-bond lifetime is calculated by the time-correlation function, SHB ðtÞ,

Molecular Simulation

3

1.0

τNG τNNG

T=300K XTMU=0.01

477

T = 300 K XTMU = 0 SHB(t)

2 1

0.5

0 10 8

τNG τNNG

T=300K XTMU=0.2

0.0 0.01

0.1

1

1.0

4

T = 300 K

2

XTMU = 0 CHB(t)

P[log10(δr);t]

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

6

0 10 8

0.5

τNG τNNG

T=300K XTMU=0.3

6 0.0 0.01

4

0.1

2

Figure 7. (Colour online) Relaxation profiles of SHB ðtÞ and CHB ðtÞ for neat water along with multi-exponential fits. Simulations are represented by the circles, and solid lines going through them are fits. Average relaxation times for these correlation functions at various X TMUPhave been obtained from fit parameters as follows: ktHB S lx ¼ i ai ti , ai and ti are the normalised amplitudes and time constants obtained via fits to simulated SHB ðtÞ and CHB ðtÞ.

τNG τNNG

T=300K XTMU=0.5

6 4 2

0 –4

10

t (ps)

0 10 8

1

–3

–2

–1

0

1

The structural relaxation time of H-bond is calculated by the time-correlation function, C HB ðtÞ; defined as [23 – 26]

2

log10(δr/σ)

Figure 6. (Colour online) Simulated single particle displacement distributions, P½log10 ðdrÞ; tÞ, for water at a few representative mixture compositions at 300 K. Distributions both at tNG and tNNG are shown. P½log10 ðdrÞ; tÞ corresponding to a Gaussian Gs ðdr; tÞ at these times is also shown for comparison. Horizontal broken lines in these panels indicate a value of 2.13 expected for P½log10 ðdrÞ; tÞ possessing a Gaussian Gs ðdr; tÞ.

SHB ðtÞ ¼

khð0ÞHðtÞl ; khl

ð9Þ

where HðtÞ ¼ 1 if the tagged pair of molecules, for which hð0Þ is calculated, remains continuously H-bonded till time t, else zero. SHB ðtÞ describes the probability that a tagged pair of molecules remains continuously H-bonded up to time t, and approaches to zero when H-bonding between them breaks down for the first time. The time profile of SHB ðtÞ therefore represents relaxation of H-bond lifetime, and hence, the average H-bond lifetime, ktHB S l, can be obtained via time integration of this normalised time correlation function.

CHB ðtÞ ¼

khð0ÞhðtÞl ; khl

ð10Þ

where CHB ðtÞ exhibits a much slower decay than SHB ðtÞ; as the former is independent of possible breaking of H-bonds between the tagged pair of molecules in the intermediate time and considers the reformation of bond after breakage. CHB ðtÞ is, therefore, associated with H-bond structural relaxation, and thus the time integration of this normalised correlation function provides the average structural relaxation time, ktHB C l. Figure 7 shows the decay of SHB ðtÞ and CHB ðtÞ as a function of time simulated at X TMU ¼ 0 and T ¼ 300 K. Bi-exponential fit to SHB ðtÞ and tri-exponential fit to C HB ðtÞ are also shown in these figures. Note simulated SHB ðtÞ and CHB ðtÞ at other mixture compositions can also be similarly described via bi- and tri-exponentials, respectively. Fit parameters required for these correlation functions at all X TMU values are summarised in Tables S1 –S4 of the Supplementary information. Note that multi-exponential

S. Indra and R. Biswas

nature of SHB ðtÞ and C HB ðtÞ has been found earlier also for a variety of systems, ranging from aqueous micellar solutions [60,98,99] and cyclodextrins [100] to moistHB urised proteins and DNA.[101,102] Note ktHB S l and ktC l for pure water are, respectively, , 0.5 and , 4 ps, which are in good agreement with earlier findings.[59 – 61,98 – 102] The multi-exponential character of the relaxations may have origin in the two different density regions that exist in liquid water – one is low density four-coordinated tetrahedral H-bonding region and the other is two or threecoordinated high density region.[103,104] However, further study is required to determine whether density fluctuations in these differently populated regions are indeed responsible for the multi-exponential relaxations of SHB ðtÞ and C HB ðtÞ in neat water and in water-rich solutions. Figure 8 presents the X TMU -dependent relaxations of SHB ðtÞ for water – water and water – TMU H-bond lifetimes. Note SHB ðtÞ for both cases slows down with each successive addition of TMU in the mixture, resulting in longer H-bond lifetimes in the presence of TMU. Interestingly, the enhancement of water – water H-bond lifetime upon gradual addition of TMU in aqueous

1.0 Water-Water 300 K

=0

U

X TM

0.5

SHB(t)

=0

U

.75

X TM

XTMU 0.00 0.03 0.07 0.20 0.30 0.50 0.75

solution resembles the slowing down of water confined in carbon nanotube [105] and water at the surface of an antifreeze protein.[106] Therefore, it may be stated that the presence of TMU slows down the H-bond dynamics in some ways similar to those operative inside the confinement and at the micelle and protein surfaces. One of the frames from the simulation trajectory at X TMU ¼ 0.2 is shown in Figure 9 which depicts the alignment of water and TMU molecules in the binary mixture. Figure 9 indeed suggests that water molecules near the TMU surface can mimic the effects observed either under confinement or near a micellar surface. It is to be mentioned here that longer H-bond lifetime does not necessarily indicate stronger H-bonds.[107] The average lifetime of H-bonds largely depends on local environments. From the simulated gðrÞ we have noticed growth of O – H interactions between water – water and water – TMU molecules. This structural compactness probably forces the water and TMU molecules to remain H-bonded for longer times. Quantum mechanical calculations of H-bond energies may provide further information on this aspect. H-bond structural relaxation functions, CHB ðtÞ, are shown in Figure 10 for water – water (upper panel) and water – TMU (lower panel) specie. As found for SHB ðtÞ in Figure 8, C HB ðtÞ also slows down upon addition of TMU in the aqueous solution, producing longer average structural relaxation times, ktHB C l. This can also be linked to the changes in gðrÞ upon addition of TMU in the mixture. The

0.0 1.0 Water-TMU 300 K

0.5

.75

X TM

=0

U

X TM

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

478

XTMU

.03

=0

U

0.03 0.07 0.20 0.30 0.50 0.75

0.0 0.01

0.1

1

10

t (ps)

Figure 8. (Colour online) X TMU -dependent relaxations of SHB ðtÞ for water – water hydrogen bonding (upper panel) and water – TMU hydrogen bonding (lower panel). X TMU are colour-coded as follows: neat water, solid black; 0.03, red short dash; 0.07, blue long dash; 0.2, dark pink long dash; 0.3, dark green dash-dot; 0.5, pink dash-dot-dot; 0.75, dark green dot.

Figure 9. (Colour online) Snapshot of relative spatial arrangements of water and TMU molecules in (water þ TMU) binary mixture at X TMU ¼ 0.2. TMU molecules are represented by blue colour, whereas water molecules are denoted by using ball and stick. Water oxygen and hydrogen atoms are shown by red and white colours, respectively.

Molecular Simulation 1.0

XTMU

.75 =0

U

τ sHB

(ps)

0.5

120

T = 300 K

τ cHB

TMU

80

X TM

τ cHB

Water

.75 =0

U

40

U

X TM

0.5

Water

1.0

=0

CHB(t)

X TM

1.5

Water-TMU 300 K

0.03 0.07 0.20 0.30 0.50 0.75

TMU

2.0

0.0 1.0

XTMU

0

.03

=0

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

τ sHB

U

X TM

0.5

0.00 0.03 0.07 0.20 0.30 0.50 0.75

T = 300 K

2.5

Water-Water 300 K

479

0.0

0.2

0.4

0.6

0.8

XTMU 0.0 0.01

0.1

1

10

100

1000

t (ps)

Figure 10. (Colour online) X TMU -dependent relaxations of CHB ðtÞ for water – water hydrogen bonding (upper panel) and water – TMU hydrogen bonding (lower panel). Representations of X TMU are colour-coded and follow the sequence shown in Figure 6.

X TMU -induced slowing down of SHB ðtÞ and CHB ðtÞ relaxations is summarised in Figure 11 where average HB relaxation times, ktHB S l and ktC l, are shown as a function of TMU mole fraction in the mixture. Note that both these relaxation times increase linearly with X TMU , and do not show any non-monotonic composition dependence. This clearly indicates that none of these relaxation processes are HB governed by medium viscosity. Moreover, ktHB S l and ktC l for water – TMU are longer at all X TMU than those for water – water. Relatively lower diffusion of TMU may not be the reason for this as these relaxation times do not couple to solution viscosity. Data in Figure 11 also HB indicate that ktHB S l and ktC l slow down by a factor of , 3 – 10 upon changing the medium from neat water to aqueous TMU solution with X TMU ¼ 0.75. Interestingly, this extent of slowing down is comparable to lengthening of these quantities for water encapsulated in cyclodextrin cavities.[100]

5.

Conclusion

In summary, this simulation study suggests considerable slowing down of H-bond relaxation dynamics in the

Figure 11. (Colour online) Plot of average relaxation times for X TMU -dependent H-bond dynamics and structural dynamics. As discussed in the text, ktHB S l has been obtained analytically from fits of SHB ðtÞ and ktHB C l similarly from C HB ðtÞ. Solid circles in both panels represent simulated data reported in Ref. [59].

presence of amphiphilic molecules in bulk binary mixtures, and the extent of slowing down mimics those found for encapsulated water molecules in cavities and those near micellar and biological surfaces. Simulated RDFs indicate increase of water – water and water –TMU interactions upon addition of TMU in the aqueous solution, but decrease of TMU – TMU interaction. Simulated MSDs do not show any signature of cagerattling and this has been interpreted as no evidence for the formation of TMU – water clathrate-type structure even at low TMU concentration. Simulated diffusion coefficients for both the specie seem to indicate partial decoupling of diffusion from medium viscosity. Simulated DH and single particle displacement distributions reflect deviations of particle movements from hydrodynamic behaviour. In addition, simulated diffusion coefficients do not show as pronounced composition dependence as that shown by viscosity coefficients. Use of polarisable potential and explicit treatment of methyl group in the force field for TMU may improve the agreement between the composition-dependent simulated and experimental viscosity coefficients. Relaxation profiles of H-bond lifetime and structure are multi-exponential in character, and we propose that density inhomogeneity inherent to liquid water may be responsible for this non-exponential behaviour. This aspect should be further investigated. In addition, DH of water molecules, particularly of those

480

S. Indra and R. Biswas

present at high TMU concentration, needs to be explored as the simulated diffusion coefficients appear to suggest decoupling from medium viscosity.

Acknowledgements It is a pleasure to thank Dr S. Daschakraborty and Ms T. Pal for helpful discussion. Work described here is supported, in part, by the computing facilities made available at the Centre via a TUECMS project (SR/NM/NS-29/2011(G)), and a UNANST project.

Funding

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

S.I. thanks the UGC, India for a research fellowship.

Supplementary information Supplementary information for this article can be accessed here.

References [1] Swenson J, Bergman R, Longeville S. A neutron spin-echo study of confined water. J Chem Phys. 2001;115:11299–11305. [2] Antognozzi M, Humphris ADL, Miles MJ. Observation of molecular layering in a confined water film and study of the layers viscoelastic properties. Appl Phys Lett. 2001;78:300– 302. [3] Gallo P, Rovere M, Spohr E. Glass transition and layering effects in confined water: a computer simulation study. J Chem Phys. 2000;113:11324–11335. [4] Bellissent-Funel M-C, Lal J, Bosio L. Structural study of water confined in porous glass by neutron scattering. J Chem Phys. 1993;98:4246 –4252. [5] Meyer M, Stanley HE. Liquid–liquid phase transition in confined water: a Monte Carlo study. J Phys Chem B. 1999;103: 9728–9730. [6] Bruni F, Ricci MA, Soper AK. Water confined in Vycor glass. I. A neutron diffraction study. J Chem Phys. 1998;109:1478–1485. [7] Faraone A, Liu L, Mou C-Y, Yen C-W, Chen S-H. Fragile-tostrong liquid transition in deeply supercooled confined water. J Chem Phys. 2004;121:10843–10846. [8] Mamontov E. Comment on ‘Quasielastic neutron scattering of two-dimensional water in a vermiculite clay’ [J. Chem. Phys. 113, 2873 (2000)] and ‘A neutron spin-echo study of confined water’ [J. Chem. Phys. 115, 11299 (2001)]. J Chem Phys. 2004;121: 9193–9194. [9] Tajkhorshid E, Nollert P, Jensen MO, Miercke L, O’Connell J, Stroud RM, Schulten K. Control of the selectivity of the aquaporin water channel family by global orientational tuning. Science. 2002;296:525 –530. [10] Bhattacharyya K. Nature of biological water: a femtosecond study. Chem Commun. 2008 :2848–2857. [11] Bhattacharyya K. Solvation dynamics and proton transfer in supramolecular assemblies. Acc Chem Res. 2003;36:95–101. [12] Bagchi B, Jana B. Solvation dynamics in dipolar liquids. Chem Soc Rev. 2010;39:1936–1954. [13] Guchhait B, Biswas R, Ghorai PK. Solute and solvent dynamics in confined equal-sized aqueous environments of charged and neutral reverse micelles: a combined dynamic fluorescence and all-atom molecular dynamics simulation study. J Phys Chem B. 2013;117: 3345–3361. [14] Levinger NE. Water in confinement. Science. 2002;298: 1722–1723. [15] Coridan RH, Schmidt NW, Lai GH, Abbamonte P, Wong GCL. Dynamics of confined water reconstructed from inelastic X-ray scattering measurements of bulk response functions. Phys Rev E. 2012;85:031501-1–031501-8.

[16] Biswas R, Rohman N, Pradhan T, Buchner R. Intramolecular charge transfer reaction, polarity, and dielectric relaxation in AOT/ water/heptane reverse micelles: pool size dependence. J Phys Chem B. 2008;112:9379–9388. [17] Sarma N, Borah JM, Mahiuddin S, Gazi HAR, Guchhait B, Biswas R. Influence of chain length of alcohols on Stokes’ shift dynamics in catanionic vesicles. J Phys Chem B. 2011;115:9040– 9049. [18] Biswas R, Das AR, Pradhan T, Touraud D, Kunz W, Mahiuddin S. Spectroscopic studies of catanionic reverse microemulsion: correlation with the superactivity of horseradish peroxidase enzyme in a restricted environment. J Phys Chem B. 2008;112:6620–6628. [19] Senapati S, Chandra A. Dielectric constant of water confined in a nanocavity. J Phys Chem B. 2001;105:5106–5109. [20] Rashin AA, Iofin M, Honig B. Internal cavities and buried waters in globular proteins. Biochemistry. 1986;25:3619–3625. [21] Berne`che S, Roux B. Energetics of ion conduction through the K þ channel. Nature. 2001;414:73–77. [22] Ball P. Water as an active constituent in cell biology. Chem Rev. 2008;108:74– 108. [23] Persson E, Halle B. Nanosecond to microsecond protein dynamics probed by magnetic relaxation dispersion of buried water molecules. J Am Chem Soc. 2008;130:1174–1787. [24] Bagchi B. Water dynamics in the hydration layer around proteins and micelles. Chem Rev. 2005;105:3197–3219. [25] Moilanen DE, Piletic IR, Fayer MD. Water dynamics in nafion fuel cell membranes: the effects of confinement and structural changes on the hydrogen bond network. J Phys Chem C. 2007;111: 8884–8891. [26] Faeder J, Ladanyi BM. Molecular dynamics simulations of the interior of aqueous reverse micelles. J Phys Chem B. 2000; 104:1033–1046. [27] Abel S, Sterpone F, Badyopadhyay S, Marchi M. Molecular modeling and simulations of AOT–water reverse micelles in isooctane: structural and dynamic properties. J Phys Chem B. 2004;108:19458–19466. [28] Dokter A, Wourtersen S, Bakker H. Inhomogeneous dynamics in confined water nanodroplets. J Proc Natl Acad Sci USA. 2006; 103:15355–15358. [29] Moilanen DE, Fenn EE, Wong D, Fayer MD. Water dynamics in large and small reverse micelles: from two ensembles to collective behavior. J Chem Phys. 2009;131:014704-1–014704-9. [30] Levinger NE, Swafford LA. Ultrafast dynamics in reverse micelles. Annu Rev Phys Chem. 2009;60:385– 406. [31] Caillez F, Stirnemann G, Boutin A, Demachy I, Fuchs AH. Does water condense in hydrophobic cavities? A molecular simulation study of hydration in heterogeneous nanopores. J Phys Chem C. 2008;112:10435–10445. [32] Walther JH, Jaffe R, Halicioglu T, Koumoutsakos P. Carbon nanotubes in water: structural characteristics and energetic. J Phys Chem B. 2001;105:9980–9987. [33] Javadian S, Taghavi F, Yari F, Hashemianzadeh SM. Phase transition study of confined water molecules inside carbon nanotubes: hierarchical multiscale method from molecular dynamics simulation to ab initio calculation. J Mol Graph Model. 2012;38:40–49. [34] Lockington DA, Parlange J-Y. Anomalous water absorption in porous materials. J Phys D: Appl Phys. 2003;36:760 –767. [35] Daschakraborty S, Biswas R. Asymmetric binary mixtures under cylindrical confinement: a molecular dynamics simulation study. ISRAPS Bull. 2013;25:84–91. [36] Galamba N. Water tetrahedrons, hydrogen-bond dynamics, and the orientational mobility of water around hydrophobic solutes. J Phys Chem B. 2014;118:4169–4176. [37] Rezus YLA, Bakker HJ. Observation of immobilized water molecules around hydrophobic groups. Phys Rev Lett. 2007;99: 148301-1–148301-4. [38] Laage D, Stirnemann G, Hynes JT. Why water reorientation slows without iceberg formation around hydrophobic solutes. J Phys Chem B. 2009;113:2428–2435. [39] Wei H, Fan Y, Gao YQ. Effects of urea, tetramethyl urea, and trimethylamine N-oxide on aqueous solution structure and

Molecular Simulation

[40]

[41]

[42]

[43]

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

[44]

[45]

[46]

[47]

[48]

[49] [50]

[51]

[52] [53] [54] [55] [56] [57] [58]

[59] [60]

[61]

[62]

[63]

solvation of protein backbones: a molecular dynamics simulation study. J Phys Chem B. 2010;114:557 –568. Tielrooij K-J, Hunger J, Buchner R, Bonn M, Bakker HJ. Influence of concentration and temperature on the dynamics of water in the hydrophobic hydration shell of tetramethylurea. J Am Chem Soc. 2010;132:15671–15678. Pradhan T, Ghoshal P, Biswas R. Excited state intramolecular charge transfer reaction in binary mixtures of water and tertiary butanol (TBA): alcohol mole fraction dependence. J Phys Chem A. 2008;112:915 –924. Pradhan T, Ghoshal P, Biswas R. Structural transition in alcohol– water binary mixtures: a spectroscopic study. J Chem Sci. 2008;120:275 –287. Gazi HAR, Biswas R. Heterogeneity in binary mixtures of (water þ tertiary butanol): temperature dependence across mixture composition. J Phys Chem A. 2011;115:2447–2455. Roy S, Banerjee S, Biyani N, Jana B, Bagchi B. Theoretical and computational analysis of static and dynamic anomalies in water– DMSO binary mixture at low DMSO concentrations. J Phys Chem B. 2011;115:685–692. Frank HS, Evans MW. Free volume and entropy in condensed systems III. Entropy in binary liquid mixtures; partial molal entropy in dilute solutions; structure and thermodynamics in aqueous electrolytes. J Chem Phys. 1945;13:507–532. Pomata MHH, Sonoda MT, Skaf MS, Elola MD. Anomalous dynamics of hydration water in carbohydrate solutions. J Phys Chem B. 2009;113:12999–13006. Stirnemann G, Sterpone F, Laage D. Dynamics of water in concentrated solutions of amphiphiles: key roles of local structure and aggregation. J Phys Chem B. 2011;115:3254–3262. Bakker HJ, Skinner JL. Vibrational spectroscopy as a probe of structure and dynamics in liquid water. Chem Rev. 2010; 110:1498–1517. Rezus YLA, Bakker HJ. On the orientational relaxation of HDO in liquid water. J Chem Phys. 2005;123:114502-1–114502-7. Eaves JD, Loparo JJ, Fecko CJ, Roberts ST, Tokmakoff A, Geissler PL. Hydrogen bonds in liquid water are broken only fleetingly. Proc Natl Acad Sci USA. 2005;102:13019–13022. Fayer MD. Dynamics of liquids, molecules, and proteins measured with ultrafast 2D IR vibrational echo chemical exchange spectroscopy. Ann Rev Phys Chem. 2008;60:21–38. Fayer MD, Levinger NE. Analysis of water in confined geometries and at interfaces. Ann Rev Anal Chem. 2010;3:89–107. Luzar A, Chandler D. Hydrogen-bond kinetics in liquid water. Nature. 1996;379:55– 57. Luzar A, Chandler D. Effect of environment on hydrogen bond dynamics in liquid water. Phys Rev Lett. 1996;76:928–931. Sciortino F, Geiger A, Stanley HE. Effect of defects on molecular mobility in liquid water. Nature. 1991;354:218–221. Laage D, Hynes JT. A molecular jump mechanism of water reorientation. Science. 2006;311:832–835. Luzar A. Water hydrogen-bond dynamics close to hydrophobic and hydrophilic groups. Faraday Discuss. 1996;103:29–40. Berkelbach TC, Lee HS, Tuckerman ME. Concerted hydrogenbond dynamics in the transport mechanism of the hydrated proton: a first-principles molecular dynamics study. Phys Rev Lett. 2009;103:238302-1–238302-4. Chandra A. Effects of ion atmosphere on hydrogen-bond dynamics in aqueous electrolyte solutions. Phys Rev Lett. 2000;85:768 –771. Balasubramanian S, Pal S, Bagchi B. Hydrogen-bond dynamics near a micellar surface: origin of the universal slow relaxation at complex aqueous interfaces. Phys Rev Lett. 2002;89:1155011–115501-4. Luzar A, Chandler D. Structure and hydrogen bond dynamics of water– dimethyl sulfoxide mixtures by computer simulations. 1993;98:8160 –8173. Xia Z, Qiang Z, Dong-Xia Z. Hydrogen bond lifetime definitions and the relaxation mechanism in water solutions. Acta Phys Chim Sin. 2011;27:2547 –2552. Pace CN, Marshall HF. A comparison of the effectiveness of protein denaturants for b-lactoglobulin and ribonuclease. Arch Biochem Biophys. 1980;199:270– 276.

481

[64] Barone G, de lVecchio P, Giancola C, Graziano GG. Excess enthalpies of N-acetylglycineamide and N-acetyl-l-leucineamide in concentrated aqueous solutions of tetramethylurea. Thermochim Acta. 1993;227:67– 73. [65] Berendsen HJC, Grigera JR, Straatsma TP. The missing term in effective pair potentials. J Phys Chem. 1987;91:6269–6271. [66] Allen MP, Tildesley DJ. Computer simulations of liquids. New York: Oxford University Press; 1987. [67] Smith W, Forester TR, Todorov IT. DL_POLY classic. Daresbury: STFC Daresbury Laboratory; 2012. [68] Yong CW. DL_FIELD – a force field and model development tool for DL_POLY. Daresbury: STFC Daresbury Laboratory; 2010. [69] Kuffel A, Zielkiewicz J. The hydrogen bond network structure within the hydration shell around simple osmolytes: urea, tetramethylurea, and trimethylamine-N-oxide, investigated using both a fixed charge and a polarizable water model. J Chem Phys. 2010;133:035102-1–035102-8. [70] Smith LJ, Berendsen HJC, van Gunsteren WF. Computer simulation of urea –water mixtures: a test of force field parameters for use in biomolecular simulation. J Phys Chem B. 2004;108: 1065–1071. [71] Belletato P, Freitas LCG, Areˆas EPG, Santos PS. Computer simulation of liquid tetramethylurea and its aqueous solution. Phys Chem Chem Phys. 1999;1:4769–4776. [72] Nose S. A unified formulation of the constant temperature molecular dynamics methods. J Chem Phys. 1984;81:511 –519. [73] Hoover WG. Canonical dynamics: equilibrium phase-space distributions. Phys Rev A. 1985;31:1695–1697. [74] Ryckaert JP, Ciccotti G, Berendsen HJC. Numerical integration of the cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J Comput Phys. 1977;23: 327–341. [75] Lyubartsev AP, Laaksonen AM. DynaMix – a scalable portable parallel MD simulation package for arbitrary molecular mixtures. Comput Phys Commun. 2000;128:565–589. [76] Okpala C, Guiseppi-Elie A, Maharajh DM. Several properties of 1,1,3,3-tetramethylurea – water systems. J Chem Eng Data. 1960;25:384– 386. [77] Sasaki K, Arakawa K. Ultrasonic and thermodynamic studies on the aqueous solutions of tetramethylurea. Bull Chem Soc Jpn. 1973;46:2738– 2741. [78] De Leeuw SW, Perram JW, Smith ER. Computer simulation of the static dielectric constant of systems with permanent electric dipoles. Ann Rev Phys Chem. 1986;37:245–270. [79] Vidulich GA, Evans DF, Kay RL. The dielectric constant of water and heavy water between 0 and 40.degree. J Phys Chem. 1967;71:656– 662. [80] Ferna´ndez DP, Mulev Y, Goodwin ARH, Levelt Sengers JMH. A database for the static dielectric constant of water and steam. J Phys Chem Ref Data. 1995;24:33– 69. [81] van der Spoel D, van Maaren PJ, Berendsen HJC. A systematic study of water models for molecular simulation: derivation of water models optimized for use with a reaction field. J Chem Phys. 1998;108:10220–10230. [82] Fratiello A, Lee RE, Miller DP, Nishida VM. N.M.R. studies of electrolytes in binary solvent mixtures. Mol Phys. 1967;13: 349–359. [83] Hansen JP, McDonald IR. Theory of simple liquids. New York: Academic Press; 1986. [84] Daschakraborty S, Biswas R. Does polar interaction influence medium viscosity? A computer simulation investigation using model liquids. J Chem Sci. 2012;124:763– 771. [85] Roy D, Patel N, Conte S, Maroncelli M. Dynamics in an idealized ionic liquid model. J Phys Chem B. 2010;114:8410–8424. [86] Pal T, Biswas R. Rank-dependent orientational relaxation in an ionic liquid: an all-atom simulation study. Theor Chem Acc. 2013;132:1348-1– 1348-12. [87] Habasaki J, Ngai KL. Molecular dynamics simulation of ion dynamics in glassy ionic conductors: evidence of the primitive ion hopping process. J Non-Cryst Solids. 2006;352:5170–5177. [88] Mark P, Nilsson L. Structure and dynamics of the TIP3P, SPC, and SPC/E water models at 298 K. J Phys Chem A. 2001;105: 9954–9960.

Downloaded by [S N Bose National Centre for Basic] at 23:15 06 April 2015

482

S. Indra and R. Biswas

[89] Areˆas EPG, Areˆas JAG, Hamburger J, Peticolas WL, Santos PS. On the high viscosity of aqueous solution of lysozyme induced by some organic solvents. J Colloid Interface Sci. 1996;180:578 –589. [90] Pal T, Biswas R. Heterogeneity and viscosity decoupling in (acetamide þ electrolyte) molten mixtures: a model simulation study. Chem Phys Lett. 2011;517:180– 185. [91] Guchhait B, Das S, Daschakraborty S, Biswas R. Interaction and dynamics of (alkylamide þ electrolyte) deep eutectics: dependence on alkyl chain length, temperature, and anion identity. J Chem Phys. 2014;140:104514-1–104514-12. [92] Guchhait B, Daschakraborty S, Biswas R. Medium decoupling of dynamics at temperatures ,100 K above glass-transition temperature: a case study with (acetamide þ lithium bromide/ nitrate) melts. J Chem Phys. 2012;136:174503-1–174503-16. [93] Das A, Das S, Biswas R. Fast fluctuations in deep eutectic melts: multi-probe fluorescence measurements and all-atom molecular dynamics simulation study. Chem Phys Lett. 2013;581:47–51. [94] Guchhait B, Gazi HAR, Kashyap HK, Biswas R. Fluorescence spectroscopic studies of (acetamide þ sodium/potassium thiocyanates) molten mixtures: composition and temperature dependence. J Phys Chem B. 2010;114:5066– 5081. [95] Flenner E, Szamel G. Relaxation in a glassy binary mixture: modecoupling-like power laws, dynamic heterogeneity, and a new nonGaussian parameter. Phys Rev E. 2005;72:011205. [96] Kim K, Saito S. Multi-time density correlation functions in glass forming liquids: probing dynamical heterogeneity and its lifetimes. J Chem Phys. 2010;133:044511. [97] Pal T, Biswas R. Slow solvation in ionic liquids: connections to non-Gaussian moves and multipoint correlations. J Chem Phys. 2014;141:104501-1–104501-12. [98] Pal S, Balasubramanian S, Bagchi B. Dynamics of bound and free water in an aqueous micellar solution: analysis of the lifetime and

[99]

[100]

[101]

[102]

[103]

[104] [105]

[106]

[107]

vibrational frequencies of H-bonds at a complex interface. Phys Rev E. 2003;67:061502-1– 061502-10. Pal S, Bagchi B, Balasubramanian S. Hydration layer of a cationic micelle, C10TAB: structure, rigidity, slow reorientation, hydrogen bond lifetime, and solvation dynamics. J Phys Chem B. 2005;109:12879–12890. Jana M, Bandyapadhyay S. Hydration properties of a-, b-, and gcyclodextrins from molecular dynamics simulations. J Phys Chem B. 2011;115:6347–6357. Pal S, Bandyapadhyay S. Importance of protein conformational motions and electrostatic anchoring sites on the dynamics and hydrogen bond properties of hydration water. Langmuir. 2013;29:1162– 1173. Sinha SK, Bandyapadhyay S. Dynamic properties of water around a protein-DNA complex from molecular dynamics simulations. J Chem Phys. 2011;135:135101-1–135101-15. Stra¨ssle Th, Saitta AM, Godec YLe, Hamel G, Klotz S, Loveday JS, Nelmes RJ. Structure of dense liquid water by neutron scattering to 6.5 GPa and 670 K. Phys Rev Lett. 2006;96:0678011–067801-4. Soper AK, Ricci MA. Structures of high-density and low-density water. Phys Rev Lett. 2000;84:2881– 2884. Hummer G, Rasaiah JC, Noworyta JP. Water conduction through the hydrophobic channel of a carbon nanotube. Nature. 2001;414:188– 190. Xu Y, Gnanasekaran R, Leitner DM. Analysis of water and hydrogen bond dynamics at the surface of an antifreeze protein. J Atom Mol Opt Phys. 2012 :125071-1–125071-6. Laage D, Hynes JT. Do more strongly hydrogen-bonded water molecules reorient more slowly? Chem Phys Lett. 2006;433: 80–85.