Effect of Surfactant Shape on Solvophobicity and Surface Activity in Alcohol-Water Systems

Size: px
Start display at page:

Download "Effect of Surfactant Shape on Solvophobicity and Surface Activity in Alcohol-Water Systems"

Transcription

1 Effect of Surfactant Shape on Solvophobicity and Surface Activity in Alcohol-Water Systems Phwey S. Gil, Daniel J. Lacks Department of Chemical and Biomolecular Engineering Case Western Reserve University Cleveland, OH Abstract Here we study the relationship between a surfactant s molecular shape and its tendency to partition to the interface in ethanol-water mixtures. In general, finding surfactants that are effective in alcohol-water mixtures is more challenging than finding ones that are effective in pure water. This is because the solvophobic effect that partitions surfactants from bulk solution to the interface becomes weaker as ethanol concentration increases. We use experiments and molecular dynamics to observe the effects of increasing surfactant tail length or width. The results show that increasing surfactant tail length causes the surfactant to partition to the surface better in low ethanol concentrations, but not at high ethanol concentrations. In comparison, increasing surfactant tail width causes the surfactant to partition to the surface better at higher concentrations of ethanol. We examine the liquid structure to elucidate the mechanisms that weaken the partitioning effect as ethanol concentration increases. Ethanol-water mixtures are nanoscopically heterogeneous with protic and aprotic regions in the bulk solution. We see that the surfactant tail is most likely to be solvated in the aprotic regions where it perturbs fewer hydrogen bonds. Introduction Surfactants are molecules that decrease the surface tension between two distinct phases (e.g. water and air or water and oil). They are amphiphilic molecules that have a polar head group and a non-polar tail group. 1

2 The standard picture of how surfactants decrease surface tension is that water attracts the hydrophilic head group while repelling the hydrophobic tail group, partitioning surfactants to the interface between distinct phases. Then the surfactants decrease the surface tension by disrupting attractive forces at the interface. In water, surfactant surface activity is directly related to the hydrophobic effect. Studying surfactants in alcohol-water presents an interesting challenge. Most surfactants are less surface active when there is a high concentration of alcohol in solution. From a practical standpoint, we can deduce that the surfactant tail is less repelled by alcohol molecules compared to water molecules, hence the reduced surfactant surface activity. Certain classes of surfactants maintain their effectiveness in both water and alcohol-water systems. For example, fluorosurfactants and silicone surfactants are more effective at lowering surface tension of alcohol-water than hydrocarbon surfactants. 1 3 Therefore, it is our interest to discover molecular-level characteristics of these surfactants that enable them to be effective in alcohol-water systems. In particular, we are interested in the relationship between solute geometry and solute surface activity. We concluded in our previous work that geometric factors are accountable for the augmented surface activity of fluorosurfactants over hydrocarbon surfactants in alcohol-water environments. 2 We investigated perfluorooctanoic acid (PFOA) and octanoic acid (OA). The two surfactants have analogous structures, and differ only in that the hydrogen atoms along the carbon chain in OA are replaced by fluorine atoms in PFOA. Experiments show that PFOA can reduce surface tension at high ethanol concentrations where OA is unable to reduce the surface tension. Molecular dynamics simulations elucidate these experimental results, and show that PFOA has greater surface activity due to the carbonfluorine bond being longer than the carbon-hydrogen bond. Similar conclusions are also reported by Dalvi et al, 4 who compared the free energy of hydration of alkanes and fluoroalkanes and found that fluoroalkanes have higher free energy of hydration than alkanes due to the larger width of molecules. It is simply this size effect, rather than effects associated with atomic charges or other interatomic forces, that leads to the differences in surfactant surface activity. 2

3 Zygmunt et al. also studied the hydration free energy of perfluoroalcohols and hydrocarbon analogs in water. 5 They found that fluorinating methylene groups adjacent to the alcohol functional group increased the hydration free energy more than fluorinating methylene groups further down the carbon tail. They concluded that the fluorination sterically affects the alcohol group and decreases the average number of hydrogen bonds that the alcohol group can form. The novelty of the present work is the isolation of the separate effects on surfactant activity coming from molecule length and width. To address the effects of molecular length, we examine analogous surfactants that differ only in the length of the hydrocarbon chain. The molecules studied in the simulations are shown in Fig. 1a: butyric acid (which we denote C4), octanoic acid (C8), dodecanoic acid (C12) and hexadodecanoic acid (C16). In our experiments, we study C4 and C8. To address the effects of molecular width, we exploit the capability of molecular simulations to address model molecules in which we can vary one parameter at will to determine the consequences of this one parameter. Note that the isolation of the effect of molecular width is not possible in experiments, as a modification to make the molecule wider, such as methylation or fluorination, will also lead to changes in the intermolecular van der Waals and Coulombic forces that could also have consequences on surface activity. We study modifications of the C8 molecule shown in Fig. 1b, where the carbon-hydrogen bond length has been decreased to 0.9Å (C8 ), increased to 1.332Å (C8 + ), and increased to 1.5 Å (C8 ++ ); in addition we study a modification of C8 molecule where the hydrogen atoms have been replaced with methyl groups (C8m). We also consider how the solvent structure influences the surfactant surface activity. The structure of alcohol-water mixtures is nontrivial. Although ethanol-water solutions are fully miscible, it is known that the excess entropy is largely negative. 6 Experiments using NMR, fluorescence probe, and Raman spectroscopy techniques have explained the negative excess entropy by demonstrating that alcohols selfassociate in aqueous solutions. 7 9 A combination of neutron diffraction and molecular dynamics studies on concentrated methanol-water mixtures revealed that most water molecules hydrogen bond with each other in clusters and strings. 10 More recently, molecular dynamics studies have shown that hydrogen 3

4 bonded networks in ethanol-water mixtures form rings, and can be fully percolating at low ethanol concentrations but non-percolating at high ethanol concentrations. 14,15 Our goal in this work is to relate surfactant surface activity in ethanol-water mixtures to these complex solvent structural effects. Methods Molecular dynamics is a computer simulation method for studying the physical motions of atoms and molecules in a system. The trajectory of the particles is numerically solved for following Newtonian laws of motion. Molecular dynamics force fields define the potential energies of the particles and the forces between the particles. The trajectory of the system is analyzed to determine microscopic and macroscopic properties of the system. Here we use molecular dynamics to simulate linear surfactants in ethanol-water solutions at the liquidvapor interface. To create the liquid-vapor interface, we model the ethanol-water liquid phase in a slab geometry. Periodic boundary conditions apply in all directions. In the z-direction, the lattice parameter is set to a large value so that the molecules cannot fill up the entire cell. Instead, the molecules form a slab that is infinite in the x- and y-dimensions, but finite in the z-direction. This slab creates top and bottom surfaces oriented perpendicular to the z-direction. The large value of the lattice parameter in the z- direction makes interactions between periodic images of the slab negligible. The simulations contain approximately 23,500 atoms inside a simulation cell of size 6 x 6 x 15 nm 3 (the precise number of atoms depends on the composition of the system). This system results in liquid slabs that are about 6 nm thick, which is twice the length of the longest surfactant molecule studied. The simulations are carried out with an NVT ensemble. Temperature is maintained at 300 K using a velocity-rescaling thermostat with time constant of 0.1 ps. 16 The simulations are run for 20 million steps with a time-step of 2 fs for a total of 40 ns. The OPLS-AA force field is used for the ethanol and the surfactant molecules, except the C8m molecule in which the methyl group is modeled with the OPLS-UA 4

5 force field. 17 The SPC/E force field is used for water molecules; 18 previous work has shown that the SPC/E force field better models the surface tension of water better than the TIP3P, TIP4P, and SPC force fields. 19 The Particle Mesh-Ewald method is used to evaluate the Coulombic interactions, with a cutoff distance of 1 nm for the real-space sum. 20 Non-Coulombic interactions are summed to a cutoff distance of 2 nm. Bond constraints are used to keep bond lengths constant, as our previous work showed that the use of bond constraints did not affect results for the surface activity of surfactants. 2 Gromacs software is used for the molecular dynamics simulations, and Avogadro and Visual Molecular Dynamics software are used for visualization We obtain the free energy profile of surfactants with respect to the surface of liquids, at the infinite dilution limit. A single surfactant molecule is inserted to the slab of ethanol-water. The molecular dynamics trajectory is used to determine the probability that the surfactant is at position, denoted, where the position of the surfactant molecule is defined as the position of the carboxylic acid carbon. The center of the liquid slab is taken to be 0, and there are surfaces at both positive and negative due to the symmetry of the system. We average the results from positive and negative to obtain better statistics. The free energy profile, denoted, can be determined, in principle, from, (1) where is Boltzmann s constant, is temperature, and is a constant. In practice, Eq. 1 can be used directly only in cases where the surfactant naturally samples all relevant values of during the trajectory; we find this case occurs at high ethanol concentrations. When the surfactant does not sample all values of z during the simulation, the direct use of Eq. 1 is not possible. This situation occurs at low ethanol concentrations, where the surfactant remains at the surface and does not sample states within the bulk. In this case we use umbrella sampling 24,25 to introduce a bias to force the molecule into the bulk, and the effects of the bias are removed in the analysis stage. For each 5

6 state point five simulations are carried out where the system is biased to different positions with respect to the liquid surface, and a force constant of 100 kj/mol nm 2 is used for the biasing potential. The umbrella sampling methodology is described more fully in our previous works. 2,26 In addition to the simulations described above, we carry out bulk simulations (i.e. without surfaces) to determine the radial distribution functions (RDF) of the surfactants in solution. The NPT ensemble is used for these simulations. Berendsen isotropic pressure coupling is used with a reference pressure of 1 atm and a time constant of 2 ps. 27 The number of atoms in these simulations is approximately 23500, corresponding to simulation cell sizes of approximately 6 x 6 x 6 nm 3. We also carry out experiments to measure the surface tension of water-ethanol solutions with C4, C8 and PFOA surfactants. The Wilhelmy plate method is utilized, which has the advantage that the measurement is independent of solution density. 28,29 The Kruss K100 Tensiometer is used to take the measurements. The surface tension as a function of surfactant concentration is obtained as follows. First, water and ethanol are mixed to a desired concentration. Then, a large amount of surfactant (2-4 g) is added to 5 g of the solution. After mixing thoroughly, the surface tension is measured at room temperature (20-23 C). The solution is diluted with more solvent (ethanol and water) and another measurement is made. A series of dilution steps is repeated to obtain the surface tension of the mixture over the desired range of surfactant concentrations. The critical micelle concentration is found as the surfactant concentration above which there is little change in the surface tension with increasing surfactant concentration. Results Experimental results We carried out surface tension experiments to determine the effectiveness of C4, C8 and PFOA surfactants in reducing the surface tension of ethanol-water mixtures as a function of ethanol concentration. Fig. 2a shows results for the surface tension for water-ethanol mixtures in the absence of 6

7 surfactant, as well as with surfactant at the critical micelle concentration. Fig. 2b shows the reduction in surface tension due to the addition of surfactant, as a function of ethanol concentration. We address chain length effects by comparing results for C4 and C8. The results demonstrate that there is no significant difference in the surfactant effectiveness between C4 and C8. In contrast, our previous experiments showed that PFOA is more effective at reducing the surface tension at high ethanol concentrations, and our analysis showed this effect is a geometric effect due to the width of PFOA being greater than that of C8. 2 Simulation results surfactant free energy We use molecular dynamics to calculate the free energy difference between the bulk and the interface. This quantity allows us to compare the surface activity of the surfactants using simulations. First, we show in Fig. 3 examples of results for the free energy of the surfactant as a function of its position with respect to the liquid surface. The results in 0 wt% ethanol required umbrella sampling, but the results in 70 wt% ethanol were obtained from unbiased molecular dynamics simulations. The minimum free energy of the surfactant is at the interface. The free energy profiles are normalized to the minimum free energy at the surface, around 2.5 nm. We define the free energy difference between the bulk and the interface as G. Then, / describes the ratio of the magnitude of the free energy driving force pushing the surfactant to the interface relative to the thermal energy. When the thermal energy is greater than the free energy driving force (i.e. / 1), the free energy driving force becomes insignificant. In this case the molecule lacks substantial affinity for the interface, and thus the surfactant would not act to reduce the surface tension. A comparison of Fig. 3a and Fig. 3b shows that / depends strongly on ethanol concentration. We calculated free energy profiles for the surfactants in various ethanol concentrations. Fig. 4 shows / as a function of ethanol concentration for each surfactant. Fig. 4a focuses on the effects of surfactant length, while Fig. 4b focuses on the effects of surfactant width. 7

8 We first examine the effects of chain length, by comparing the behavior of C4, C8, C12 and C16. In pure water, increasing the length of the surfactant from C4 to C16 increases the magnitude of / rather significantly (see Fig. 4a). The nonlinearity with increasing chain length could be due to statistical uncertainty in the simulation. As ethanol concentration increases, the value of / for all of these surfactants decreases, and at 70 wt% ethanol where / <1 regardless of chain length. The nonlinear behavior with increasing ethanol concentration is likely due to the changes in the liquid structure (addressed below). At 70 wt% ethanol the value of / is smallest for C4, and it increases with chain length to C8; however, further increases in chain length do not have a significant effect on /. We conducted a two-sample t-test for C4, C8, C12, and C16 surfactants and found statistically significant difference between C4 and other surfactants (95% confidence interval of [-1.867, ]). The differences in / between C8, C12, and C16 were not found to be statistically significant. Thus, our results show that increasing the chain length is not a viable strategy to design more effective surfactants at high ethanol concentrations. On the other hand, increasing surfactant width enhances the surface activity at high ethanol concentrations. As described above, we examine three artificial analogs of C8, where the C-H bonds have been artificially shortened or lengthened (C8, C8 +, C8 ++ ). Our results (Fig. 4b) show that the magnitude of / is a strong function of molecule width at high ethanol concentration. Although C8 + and C8 ++ have smaller molecular volume and surface area than C16, they nonetheless have greater values of / at high ethanol concentrations, as it is only the width and not the length that is relevant to surfactant activity. To study the effect of molecular width without artificially changing a molecular parameter, we examine an analog of C8 where the surfactant tail H atoms have been replaced by methyl groups (C8m). / for C8m is greater than that of C8, as expected because of its greater width. We note that the value of / for C8m is less than or equal to that for C8 + and C8 ++, even though C8m is significantly wider. C8m is wider than C8, but it also has stronger non-bonded attraction; i.e., the Lennard-Jones ε 8

9 parameter is larger for a methyl group than for a hydrogen atom. We have previously shown that increased non-bonded attraction leads to decreased / at high ethanol content. Thus this energetic effect cancels some of the width effect in C8m, causing the value of / for C8m to be slightly less than or equal to that of C8 + and C8 ++. Simulation results liquid structure We analyze the liquid structure in order to better understand the surfactant free energy results presented above. First, we address the structure of water-ethanol mixtures in the absence of surfactant. Although water and alcohol are fully miscible, in mixtures water and alcohol molecules have the tendency to cluster at the molecular scale, as supported by thermodynamic, experimental, and simulation evidence. 6,7,9,10,30 Fig. 5a shows radial distribution functions (RDF) for water molecules in 38 wt% and 70 wt% ethanol. The first peak is larger in 70 wt% ethanol than in 38 wt% ethanol. Similar differences with concentration are observed in the ethanol-ethanol and ethanol-water RDFs, also shown in Fig. 5a. These results show that increasing the ethanol concentration increases the short range order in the system. We plot the ratio of the water-water and ethanol-water RDFs in Fig. 5b. For distances less than approximately 0.8 nm, the water-water RDF is greater than the water-ethanol RDF, which implies that there are waterrich clusters with diameters of approximately 1.5 nm. Ethanol molecules are found to be generally oriented such that the oxygen end faces the water-rich regions. This result is evident in Fig. 5a (right-most plots), where the oxygen head of ethanol is closer to the water regions than the carbon tail. This result shows the strong preference of polar moieties to interact with each other. Thus, we demonstrate that in ethanol-water mixture there exist protic and aprotic regions, where the protic region includes hydrogen bonds and the aprotic region does not include hydrogen bonds. The protic regions are composed of water molecules, with heads of ethanol molecules at the boundaries. 9

10 To illustrate, we show a snapshot of the molecules in a bulk ethanol-water mixture in Fig. 6a. The protic and aprotic regions discussed above are evident i.e., there are water-rich clusters and the oxygen atoms of the ethanol molecules are found at the borders of these regions. Fig. 6b shows that hydrogen bonds are heterogeneously distributed through the system, as the network of hydrogen bonds extends only through the protic regions. In comparison, as shown in a snapshot of pure water in Fig. 6b, pure water has homogeneously distributed hydrogen bonds. The heterogeneous solvent structure will affect the configuration of the surfactant in the solution. Fig. 7a shows the RDFs between the surfactant tail atoms and the solvent molecules. The first peak (r=0.5 nm) corresponds to interactions with ethanol carbon atoms, and the second peak (r=0.6 nm) to interactions with ethanol oxygens; note that interactions with water molecules show no peak (only depletion). This data implies that the surfactant tail is surrounded by ethanol molecules oriented such that their oxygen end points away from the surfactant tail. Fig. 7b shows the RDFs between the surfactant head group and the solvent molecules. At r=0.25 nm there is a tall peak for interactions with water oxygens and a short peak for interactions with ethanol oxygens. This result shows that the surfactant head group forms hydrogen bonds with the polar components in the solvent, with a stronger preference to hydrogen bond to water than to ethanol. This stronger preference occurs even at high ethanol concentration. The depletion of water molecules is due to the ethanol molecules that surround the surfactant tail (discussed above). From the RDF results we conclude that the surfactant tail is surrounded mainly by ethanol molecules oriented with their carbon tail towards the surfactant, and that the surfactant head is surrounded mainly by water molecules but also by some ethanol molecules oriented with their oxygen head towards the surfactant. Fig. 8 shows snapshots from the simulation, in which this structure is evident. While this structure is in some ways reminiscent of the suggestion long ago of clathrate or iceberg structures surrounding hydrophobic solutes in water, 31 we emphasize that the structure we identify is non-rigid and transient, 32 and all types of atoms are significantly sampled in the region adjacent to the surfactant (see Fig. 7). 10

11 Discussion The hydrophobic effect in pure water has been widely studied, and our results can be analyzed in this context. The hydrophobic effect can have entropic or enthalpic origins, depending on the situation. The entropic mechanism is identified when the free energy of hydration, ΔG f, increases with increasing temperature; the enthalpic mechanism is identified when ΔG f decreases with increasing temperature. For macroscopic interfaces between hydrophobic and water phases, and for large hydrophobic solutes (> 1 nm) in water, the enthalpic mechanism is observed. For small solutes (< 1 nm), the entropic mechanism occurs at lower temperatures and the enthalpic mechanism occurs at higher temperatures. The molecular level origins of both the enthalpic and entropic effects involve hydrogen bonding in the water. 33 For macroscopic interfaces and around large solutes, the water molecules near the interface cannot maintain their full complement of two hydrogen bonds per molecule, and the dangling hydrogen bonds which create an enthalpic penalty for the hydration. For small solutes, in contrast, the water molecules can maintain their full complement of hydrogen bonds while wrapping around the solute. In fact, Raman spectroscopy shows that hydrogen bonds around a small solute are stronger than hydrogen bonds in bulk water, 34 and molecular dynamics shows that water molecules around a small solute are more highly ordered (though still dynamic) than in the bulk. 35 While the hydrogen bonds are stronger and more ordered near the small solute, the presence of the solute reduces the number of possible configurations of water, which creates an entropic penalty for the hydration. For non-spherical molecules, such as the surfactants addressed here, the smallest length scale is the most relevant. Single-molecule force-spectroscopy studies of long polymers, which varied the polymer width by changing the monomer pendant groups, showed that increased molecular polymer causes a decrease in the temperature at which the hydrophobic mechanism changes from entropic to enthalpic. 36 Molecular dynamics simulations showed that the structural order of the water molecules surrounding a solute is essentially independent of the length of the solute. 35 The width, rather than the length, matters most 11

12 because the water molecules can most readily maintain its hydrogen bonded network around the width of the molecule. Our results in alcohol-water systems concur with these ideas for the hydrophobic effect in pure water. As discussed above, we find that molecular width, rather than molecular length, is the important factor for making a surfactant effective in alcohol-water systems. This observation is understandable in terms of these previous ideas, in that it is only the solute width that controls whether the solvent molecules can maintain a complete hydrogen bonded network around the solute. The presence of alcohol changes the liquid structure, which in turn affects the magnitude of the hydrophobic effect; these changes can also be understood in terms of the above ideas derived in the context of water. In pure water, hydrogen bonds have hydrogen bond networks that percolate throughout the system. 37 But adding ethanol creates nanometer-sized clusters of water and ethanol molecules. The non-polar regions, which are devoid of hydrogen bonds, do not occur in pure water see Fig. 6b in comparison to Fig. 6d. Previous works show that high ethanol concentrations cause the network of hydrogen bonded water molecules to no longer percolate throughout the system, and leads to discrete water clusters that get smaller with more ethanol in the system. 14,15 We have shown above that the surfactant tail is surrounded mostly by the hydrocarbon tails of ethanol molecules. Thus the surfactant tail will reside in the non-polar regions of the liquid, where it does not perturb the hydrogen bonding network of this liquid. Since the hydrophobic effect is due to solute perturbation of the hydrogen bonding of the solvent, the fact that the surfactant can reside in the liquid without significantly perturbing the hydrogen bonded structure makes the magnitude of hydrophobic effect smaller as ethanol concentration increases. 12

13 Conclusion We investigated the effect of surfactant tail geometry on the molecule s solvophobicity in ethanol-water systems. We calculated the free energy difference between the surfactant molecule at the surface and in the bulk, which determines the propensity of the surfactant molecule to reside at the surface. Longer tail groups increase the free energy difference significantly at low ethanol concentrations, but not significantly at high ethanol concentrations. Wider tail groups increase the free energy difference at all ethanol concentrations. By analyzing radial distribution functions, we determined that ethanol causes the solvent to organize into protic and aprotic regions, which enables surfactant molecules to perturb fewer hydrogen bonds while residing in the bulk liquid, which in turn explains why the surface activity of surfactants decreases as ethanol concentration increases. Acknowledgements This work was supported by the National Science Foundation Grant No. CBET , and the simulations were carried out using the computational resources of the Ohio Supercomputing Center. 13

14 A B Figure 1. Molecular structures of the surfactants addressed in this work. (A) From top to bottom: C4, C8, C12, and C16 surfactants. (B) From left to right: C8, C8, C8 +, C8 ++, where the C-H bond length is arbitrarily varied to determine the effect of this particular parameter on surfactant behavior; at far right, C8m surfactant in which the pendant H atoms are replaced by methyl groups. 14

15 A B Figure 2. Experimental surface tension values. (A) surface tension at the CMC vs ethanol concentration. (B) normalized surface tension. Ethanol-water with no surfactant (Black ); C4 (Pink ); C8 (Blue ); PFOA (Red ). The error bars show 95% confidence intervals. 15

16 A B Figure 3. Free energy profiles of octanoic acid in (A) pure water and (B) 70 wt% ethanol solution. 16

17 A B Figure 4. / versus ethanol concentration for the different surfactants. (A) Surfactants with varying molecular length. (B) Surfactants with varying molecular width. The inset in B shows / versus C-H bond length at 70 wt% ethanol. C4 (Pink ); C8 (Blue ); C12 (Teal ); C16 (Orange ); C8 + (Red ); C8m (Purple. The red dotted line (--) denotes / =1. 17

18 A B C Figure 5. Radial distribution functions for water oxygen atoms (O w ) and ethanol oxygen atoms (O E ). Top row is at 38 wt% ethanol, and bottom row is at 70 wt% ethanol. (A) From left to right: RDF for O w -O w ; O E -O E ; O E -O w. The red dotted line is for ethanol methyl carbon to water oxygen interactions, (CH 3 ) E -O w. (B) Ratio of RDFs for O w -O w to O E -O w. (C) Ratio of RDFs for O E -O E to O W -O E 18

19 A B C D Figure 6. Snapshot of molecule arrangements. (A, B) bulk solution of 70 wt% ethanol. (C, D) bulk solution of pure water. Blue atoms are water, gray atoms are ethanol tail groups, and red atoms are ethanol head groups. Water-water hydrogen bonds are blue, water-alcohol hydrogen bonds are black, and alcohol-alcohol hydrogen bonds are red. 19

20 A B Figure 7. Radial distribution functions for the C8 surfactant molecule, in bulk liquid. Top row is at 38 wt% ethanol, and bottom row is at 70 wt% ethanol. (A) Results for the surfactant tail. RDF from surfactant carbon atoms to water s oxygen atoms (Black Line); RDF from surfactant carbon atoms to ethanol carbon atoms (Red Line); RDF from surfactant carbon atoms to ethanol oxygen atoms (Blue line). (B) Results for the surfactant head. RDF from surfactant oxygen atoms to water s oxygen atoms (Black line); RDF from surfactant oxygen atoms to ethanol carbon atoms (Red Line); RDF from surfactant oxygen atoms to ethanol oxygen atoms (Blue Line). 20

21 A B Figure 8. Snapshots from the trajectory. (A) Molecules around the surfactant tail; the tail is highlighted green. (B) Molecules around the surfactant head; the surfactant tail is highlighted in white. The carboxylic acid carbon is highlighted green, the oxygen atoms orange, and the hydrogen atoms white. The ethanol molecules are shown in blue, and water molecules in red. The oxygen atoms of ethanol are highlighted purple to show the orientation of these molecules. The hydrogen bond is shown with a red dotted line. 21

22 References 1 A. Wang, L. Jiang, G. Mao, and Y. Liu, J. Colloid Interface Sci. 242, 337 (2001). 2 P.S. Gil and D.J. Lacks, J. Phys. Chem. C 119, (2015). 3 I.J.E. Briscoe, P.C. Harris, and G.S. Penny, (1984). 4 V.H. Dalvi and P.J. Rossky, Proc. Natl. Acad. Sci. U.S.A. 107, (2010). 5 W. Zygmunt and J.J. Potoff, Fluid Phase Equilib. 407, 314 (2015). 6 F. and Ives, Q. Rev. 20, 1 (1966). 7 J. Bodet, H.T. Davis, L.E. Scriven, and W.G. Miller, 1, 455 (1988). 8 R. Zana and B. Michels, 2643 (1989). 9 K. Egashira and N. Nishi, J. Phys. Chem. B 102, 4054 (1998). 10 S. Dixit, J. Crain, W.C.K. Poon, J.L. Finney, and a K. Soper, Nature 416, 829 (2002). 11 O. Gereben, J. Mol. Liq. 211, 812 (2015). 12 O. Gereben, Phys. Rev. E 92, (2015). 13 O. Gereben and L. Pusztai, J. Phys. Chem. B 119, 3070 (2015). 14 S.Y. Noskov, G. Lamoureux, and B. Roux, J. Phys. Chem. B. 109, 6705 (2005). 15 A. Ghoufi, F. Artzner, and P. Malfreyt, J. Phys. Chem. B 120, 793 (2016). 16 G. Bussi and M. Parrinello, Comput. Phys. Commun. 179, 26 (2008). 17 W.L. Jorgensen, D.S. Maxwell, and J. Tirado-Rives, J. Am. Chem. Soc. 118, (1996). 18 H.J.C. Berendsen, J.R. Grigera, and T.P. Straatsma, J. Phys. Chem. 91, 6269 (1987). 19 C. Vega and E. De Miguel, J. Chem. Phys. 126, (2007). 20 T. Darden, L. Perera, L. Li, and P. Lee, Structure 7, 55 (1999). 22

23 21 B. Hess, C. Kutzner, D. Van Der Spoel, and E. Lindahl, J. Chem. Theory Comput. 4, 435 (2008). 22 M.D. Hanwell, D.E. Curtis, D.C. Lonie, T. Vandermeerschd, E. Zurek, and G.R. Hutchison, J. Cheminform. 4, 1 (2012). 23 W. Humphrey, A. Dalke, and K. Schulten, J. Mol. Graph. 14, 33 (1996). 24 G.M. Torrie and J.P. Valleau, J. Comput. Phys. 23, 187 (1977). 25 A.M. Ferrenberg, 6, 1195 (1989). 26 A.N. Htet, P.S. Gil, and D.J. Lacks, J. Chem. Phys. 142, (2015). 27 H.J.C. Berendsen, J.P.M. Postma, W.F. van Gunsteren, a DiNola, and J.R. Haak, J. Chem. Phys. 81, 3684 (1984). 28 L.I. Rolo, A.I. Caço, A.J. Queimada, I.M. Marrucho, and J.A.P. Coutinho, J. Chem. Eng. Data 47, 1442 (2002). 29 Rusanov, A. I.; Prokhorov, V. A. Interfacial Tensiometry; Elsevier: Amsterdam, Netherlands, R. Zana, I.C.S.C.R. M, and R. Boussingault, Adv. Colloid Interface Sci. 57, (1995). 31 H.S. Frank and M.W. Evans, J. Chem. Phys. 13, 507 (1945). 32 B.W. Blokzijl and J.B.F.N. Engberts, Angew. Chem. Int. Ed. Engl. 32, 1545 (1993). 33 D. Chandler, Nature 437, 640 (2005). 34 J.G. Davis, K.P. Gierszal, P. Wang, and D. Ben-Amotz, Nature 491, 582 (2012). 35 V.R. Hande and S. Chakrabarty, J. Phys. Chem. B 119, (2015). 36 I.T.S. Li and G.C. Walker, Proc. Natl. Acad. Sci. U. S. A. 108, (2011). 37 F.H. Stillinger, Science 209, 4455 (1980). 23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

Supporting Information. for. Influence of Cononsolvency on the. Aggregation of Tertiary Butyl Alcohol in. Methanol-Water Mixtures

Supporting Information. for. Influence of Cononsolvency on the. Aggregation of Tertiary Butyl Alcohol in. Methanol-Water Mixtures Supporting Information for Influence of Cononsolvency on the Aggregation of Tertiary Butyl Alcohol in Methanol-Water Mixtures Kenji Mochizuki,, Shannon R. Pattenaude, and Dor Ben-Amotz Research Institute

More information

Supplementary Information. Surface Microstructure Engenders Unusual Hydrophobicity in. Phyllosilicates

Supplementary Information. Surface Microstructure Engenders Unusual Hydrophobicity in. Phyllosilicates Electronic Supplementary Material (ESI) for ChemComm. This journal is The Royal Society of Chemistry 2018 Supplementary Information Surface Microstructure Engenders Unusual Hydrophobicity in Phyllosilicates

More information

6 Hydrophobic interactions

6 Hydrophobic interactions The Physics and Chemistry of Water 6 Hydrophobic interactions A non-polar molecule in water disrupts the H- bond structure by forcing some water molecules to give up their hydrogen bonds. As a result,

More information

Title Super- and subcritical hydration of Thermodynamics of hydration Author(s) Matubayasi, N; Nakahara, M Citation JOURNAL OF CHEMICAL PHYSICS (2000), 8109 Issue Date 2000-05-08 URL http://hdl.handle.net/2433/50350

More information

Supporting Information

Supporting Information Supporting Information Structure and Dynamics of Uranyl(VI) and Plutonyl(VI) Cations in Ionic Liquid/Water Mixtures via Molecular Dynamics Simulations Katie A. Maerzke, George S. Goff, Wolfgang H. Runde,

More information

Cluster Formation and Percolation in Ethanol-Water mixtures

Cluster Formation and Percolation in Ethanol-Water mixtures Cluster Formation and Percolation in Ethanol-Water mixtures Orsolya Gereben 1, László Pusztai 1,2,a) 1 Institute for Solid State Physics and Optics, Wigner Research Centre for Physics, Hungarian Academy

More information

Universal Repulsive Contribution to the. Solvent-Induced Interaction Between Sizable, Curved Hydrophobes: Supporting Information

Universal Repulsive Contribution to the. Solvent-Induced Interaction Between Sizable, Curved Hydrophobes: Supporting Information Universal Repulsive Contribution to the Solvent-Induced Interaction Between Sizable, Curved Hydrophobes: Supporting Information B. Shadrack Jabes, Dusan Bratko, and Alenka Luzar Department of Chemistry,

More information

Molecular Dynamics Simulation of Methanol-Water Mixture

Molecular Dynamics Simulation of Methanol-Water Mixture Molecular Dynamics Simulation of Methanol-Water Mixture Palazzo Mancini, Mara Cantoni University of Urbino Carlo Bo Abstract In this study some properties of the methanol-water mixture such as diffusivity,

More information

Determination of Kamlet-Taft parameters for selected solvate ionic liquids.

Determination of Kamlet-Taft parameters for selected solvate ionic liquids. Electronic Supplementary Material (ESI) for Physical Chemistry Chemical Physics. This journal is the Owner Societies 2016 Determination of Kamlet-Taft parameters for selected solvate ionic liquids. Daniel

More information

On the calculation of solvation free energy from Kirkwood- Buff integrals: A large scale molecular dynamics study

On the calculation of solvation free energy from Kirkwood- Buff integrals: A large scale molecular dynamics study On the calculation of solvation free energy from Kirkwood- Buff integrals: A large scale molecular dynamics study Wynand Dednam and André E. Botha Department of Physics, University of South Africa, P.O.

More information

Improved association in a classical density functional theory for water. Abstract

Improved association in a classical density functional theory for water. Abstract Improved association in a classical density functional theory for water Eric J. Krebs, Jeff B. Schulte, and David Roundy Department of Physics, Oregon State University, Corvallis, OR 97331, USA Abstract

More information

Chapter 4. Glutamic Acid in Solution - Correlations

Chapter 4. Glutamic Acid in Solution - Correlations Chapter 4 Glutamic Acid in Solution - Correlations 4. Introduction Glutamic acid crystallises from aqueous solution, therefore the study of these molecules in an aqueous environment is necessary to understand

More information

Some properties of water

Some properties of water Some properties of water Hydrogen bond network Solvation under the microscope 1 Water solutions Oil and water does not mix at equilibrium essentially due to entropy Substances that does not mix with water

More information

Water structure near single and multi-layer nanoscopic hydrophobic plates of varying separation and interaction potentials

Water structure near single and multi-layer nanoscopic hydrophobic plates of varying separation and interaction potentials Bull. Mater. Sci., Vol. 31, No. 3, June 2008, pp. 525 532. Indian Academy of Sciences. Water structure near single and multi-layer nanoscopic hydrophobic plates of varying separation and interaction potentials

More information

Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer

Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer Nuclear Science and Techniques 23 (2012) 252 256 Promoting effect of ethanol on dewetting transition in the confined region of melittin tetramer REN Xiuping 1,2 ZHOU Bo 1,2 WANG Chunlei 1 1 Shanghai Institute

More information

MOLECULAR DYNAMICS SIMULATION OF THE STRUCTURE OF C6 ALKANES INTRODUCTION. A. V. Anikeenko, A. V. Kim, and N. N. Medvedev UDC 544.2: 544.

MOLECULAR DYNAMICS SIMULATION OF THE STRUCTURE OF C6 ALKANES INTRODUCTION. A. V. Anikeenko, A. V. Kim, and N. N. Medvedev UDC 544.2: 544. Journal of Structural Chemistry. Vol. 51, No. 6, pp. 1090-1096, 2010 Original Russian Text Copyright 2010 by A. V. Anikeenko, A. V. Kim, and N. N. Medvedev MOLECULAR DYNAMICS SIMULATION OF THE STRUCTURE

More information

New Perspective on structure and bonding in water using XAS and XRS

New Perspective on structure and bonding in water using XAS and XRS New Perspective on structure and bonding in water using XAS and XRS Anders Nilsson Stanford Synchrotron Radiation Laboratory (SSRL) and Stockholm University, Sweden R. Ludwig Angew. Chem. 40, 1808 (2001)

More information

Supporting Information for. Hydrogen Bonding Structure at Zwitterionic. Lipid/Water Interface

Supporting Information for. Hydrogen Bonding Structure at Zwitterionic. Lipid/Water Interface Supporting Information for Hydrogen Bonding Structure at Zwitterionic Lipid/Water Interface Tatsuya Ishiyama,, Daichi Terada, and Akihiro Morita,, Department of Applied Chemistry, Graduate School of Science

More information

University of Groningen. Characterization of oil/water interfaces van Buuren, Aldert Roelf

University of Groningen. Characterization of oil/water interfaces van Buuren, Aldert Roelf University of Groningen Characterization of oil/water interfaces van Buuren, Aldert Roelf IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it.

More information

On the accurate calculation of the dielectric constant and the diffusion coefficient from molecular dynamics simulations: the case of SPC/E water

On the accurate calculation of the dielectric constant and the diffusion coefficient from molecular dynamics simulations: the case of SPC/E water On the accurate calculation of the dielectric constant and the diffusion coefficient from molecular dynamics simulations: the case of SPC/E water Orsolya Gereben and László Pusztai Research Institute for

More information

Segue Between Favorable and Unfavorable Solvation

Segue Between Favorable and Unfavorable Solvation Segue Between Favorable and Unfavorable Solvation Lutz Maibaum and David Chandler Department of Chemistry, University of California, Berkeley, California 94720 (Dated: October 30, 2018) arxiv:cond-mat/0703686v1

More information

Supporting Information for Solid-liquid Thermal Transport and its Relationship with Wettability and the Interfacial Liquid Structure

Supporting Information for Solid-liquid Thermal Transport and its Relationship with Wettability and the Interfacial Liquid Structure Supporting Information for Solid-liquid Thermal Transport and its Relationship with Wettability and the Interfacial Liquid Structure Bladimir Ramos-Alvarado, Satish Kumar, and G. P. Peterson The George

More information

Diffusion of Water and Diatomic Oxygen in Poly(3-hexylthiophene) Melt: A Molecular Dynamics Simulation Study

Diffusion of Water and Diatomic Oxygen in Poly(3-hexylthiophene) Melt: A Molecular Dynamics Simulation Study Diffusion of Water and Diatomic Oxygen in Poly(3-hexylthiophene) Melt: A Molecular Dynamics Simulation Study Julia Deitz, Yeneneh Yimer, and Mesfin Tsige Department of Polymer Science University of Akron

More information

Hydrogen Bond Kinetics in the Solvation Shell of a Polypeptide

Hydrogen Bond Kinetics in the Solvation Shell of a Polypeptide Hydrogen Bond Kinetics in the Solvation Shell of a Polypeptide Huafeng Xu and B.J. Berne Department of Chemistry and Center for Biomolecular Simulation, Columbia University, 3 Broadway, New York, New York

More information

Supporting Information for: Physics Behind the Water Transport through. Nanoporous Graphene and Boron Nitride

Supporting Information for: Physics Behind the Water Transport through. Nanoporous Graphene and Boron Nitride Supporting Information for: Physics Behind the Water Transport through Nanoporous Graphene and Boron Nitride Ludovic Garnier, Anthony Szymczyk, Patrice Malfreyt, and Aziz Ghoufi, Institut de Physique de

More information

Interfaces and the driving force of hydrophobic assembly David Chandler 1

Interfaces and the driving force of hydrophobic assembly David Chandler 1 INSIGHT REVIEW NATURE Vol 437 29 September 2005 doi:10.1038/nature04162 Interfaces and the driving force of hydrophobic assembly David Chandler 1 The hydrophobic effect the tendency for oil and water to

More information

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled Supporting Information Specific ion effects on the interaction of hydrophobic and hydrophilic self assembled monolayers T. Rios-Carvajal*, N. R. Pedersen, N. Bovet, S.L.S. Stipp, T. Hassenkam. Nano-Science

More information

Critical Micellization Concentration Determination using Surface Tension Phenomenon

Critical Micellization Concentration Determination using Surface Tension Phenomenon Critical Micellization Concentration Determination using Phenomenon 1. Introduction Surface-active agents (surfactants) were already known in ancient times, when their properties were used in everyday

More information

IMPROVED METHOD FOR CALCULATING SURFACE TENSION AND APPLICATION TO WATER

IMPROVED METHOD FOR CALCULATING SURFACE TENSION AND APPLICATION TO WATER IMPROVED METHOD FOR CALCULATING SURFACE TENSION AND APPLICATION TO WATER ABSTRACT Hong Peng 1, Anh V Nguyen 2 and Greg R Birkett* School of Chemical Engineering, The University of Queensland Brisbane,

More information

Unit Cell-Level Thickness Control of Single-Crystalline Zinc Oxide Nanosheets Enabled by Electrical Double Layer Confinement

Unit Cell-Level Thickness Control of Single-Crystalline Zinc Oxide Nanosheets Enabled by Electrical Double Layer Confinement Unit Cell-Level Thickness Control of Single-Crystalline Zinc Oxide Nanosheets Enabled by Electrical Double Layer Confinement Xin Yin, Yeqi Shi, Yanbing Wei, Yongho Joo, Padma Gopalan, Izabela Szlufarska,

More information

Sunyia Hussain 06/15/2012 ChE210D final project. Hydration Dynamics at a Hydrophobic Surface. Abstract:

Sunyia Hussain 06/15/2012 ChE210D final project. Hydration Dynamics at a Hydrophobic Surface. Abstract: Hydration Dynamics at a Hydrophobic Surface Sunyia Hussain 6/15/212 ChE21D final project Abstract: Water is the universal solvent of life, crucial to the function of all biomolecules. Proteins, membranes,

More information

Supplemntary Infomation: The nanostructure of. a lithium glyme solvate ionic liquid at electrified. interfaces

Supplemntary Infomation: The nanostructure of. a lithium glyme solvate ionic liquid at electrified. interfaces Electronic Supplementary Material (ESI) for Physical Chemistry Chemical Physics. This journal is the Owner Societies 207 Supplemntary Infomation: The nanostructure of a lithium glyme solvate ionic liquid

More information

Molecular Dynamic Simulation Study of the Volume Transition of PNIPAAm Hydrogels

Molecular Dynamic Simulation Study of the Volume Transition of PNIPAAm Hydrogels Molecular Dynamic Simulation Study of the Volume Transition of PNIPAAm Hydrogels Jonathan Walter 1, Jadran Vrabec 2, Hans Hasse 1 1 Laboratory of Engineering, University of Kaiserslautern, Germany 2 and

More information

Supplementary Figure 1 Digital images of GG experiments, corresponding to the respective stages shown in Fig. 1a. The aqueous phase is lightly dyed

Supplementary Figure 1 Digital images of GG experiments, corresponding to the respective stages shown in Fig. 1a. The aqueous phase is lightly dyed Supplementary Figure 1 Digital images of GG experiments, corresponding to the respective stages shown in Fig. 1a. The aqueous phase is lightly dyed with methylene blue to aid the visualization of the originally

More information

Global Optimisation of Hydrated Sulfate Clusters

Global Optimisation of Hydrated Sulfate Clusters Global Optimisation of Hydrated Sulfate Clusters Lewis Smeeton University of Birmingham School of Chemistry 15 th December, 2014 1 / 31 1 Introduction The Hofmeister Series Hydrated Sulfate Clusters 2

More information

Molecular modeling. A fragment sequence of 24 residues encompassing the region of interest of WT-

Molecular modeling. A fragment sequence of 24 residues encompassing the region of interest of WT- SUPPLEMENTARY DATA Molecular dynamics Molecular modeling. A fragment sequence of 24 residues encompassing the region of interest of WT- KISS1R, i.e. the last intracellular domain (Figure S1a), has been

More information

arxiv: v2 [physics.chem-ph] 23 Aug 2013

arxiv: v2 [physics.chem-ph] 23 Aug 2013 The quest for self-consistency in hydrogen bond definitions Diego Prada-Gracia, Roman Shevchuk, and Francesco Rao Freiburg Institute for Advanced Studies, School of Soft Matter Research, Albertstrasse

More information

Intermolecular forces

Intermolecular forces Intermolecular forces World of Chemistry, 2000 Updated: August 29, 2013 The attractions of molecules to each other are known as intermolecular forces to distinguish them from intramolecular forces, such

More information

Vol. 114 (2008) ACTA PHYSICA POLONICA A No. 6 A

Vol. 114 (2008) ACTA PHYSICA POLONICA A No. 6 A Vol. 114 (2008) ACTA PHYSICA POLONICA A No. 6 A Optical and Acoustical Methods in Science and Technology Effects of Solvation of 2-Methylpyridine and 2,6-Dimethylpyridine in Dilute Solutions in Water and

More information

Roto-translational motion in liquid water and its structural implication

Roto-translational motion in liquid water and its structural implication Volume 215. number 6 CHEMICAL PHYSICS LETTERS 17 December 1993 Roto-translational motion in liquid water and its structural implication I.M. Svishchev and P.G. Kusalik Department of Chemistry, Dalhousie

More information

Introduction to molecular dynamics

Introduction to molecular dynamics 1 Introduction to molecular dynamics Yves Lansac Université François Rabelais, Tours, France Visiting MSE, GIST for the summer Molecular Simulation 2 Molecular simulation is a computational experiment.

More information

COOPERATIVE ORIGIN OF LOW-DENSITY DOMAINS IN LIQUID WATER. Jeffrey R. Errington, Pablo G. Debenedetti *, and Salvatore Torquato

COOPERATIVE ORIGIN OF LOW-DENSITY DOMAINS IN LIQUID WATER. Jeffrey R. Errington, Pablo G. Debenedetti *, and Salvatore Torquato 6/18/02 COOPERATIVE ORIGIN OF LOW-DENSITY DOMAINS IN LIQUID WATER by Jeffrey R. Errington, Pablo G. Debenedetti *, and Salvatore Torquato Department of Chemical Engineering, Princeton University, Princeton,

More information

COSMO-RS Theory. The Basics

COSMO-RS Theory. The Basics Theory The Basics From µ to properties Property µ 1 µ 2 activity coefficient vapor pressure Infinite dilution Gas phase Pure compound Pure bulk compound Partition coefficient Phase 1 Phase 2 Liquid-liquid

More information

Solutions and Non-Covalent Binding Forces

Solutions and Non-Covalent Binding Forces Chapter 3 Solutions and Non-Covalent Binding Forces 3.1 Solvent and solution properties Molecules stick together using the following forces: dipole-dipole, dipole-induced dipole, hydrogen bond, van der

More information

Nucleation rate (m -3 s -1 ) Radius of water nano droplet (Å) 1e+00 1e-64 1e-128 1e-192 1e-256

Nucleation rate (m -3 s -1 ) Radius of water nano droplet (Å) 1e+00 1e-64 1e-128 1e-192 1e-256 Supplementary Figures Nucleation rate (m -3 s -1 ) 1e+00 1e-64 1e-128 1e-192 1e-256 Calculated R in bulk water Calculated R in droplet Modified CNT 20 30 40 50 60 70 Radius of water nano droplet (Å) Supplementary

More information

Interfaces and the Driving Force of Hydrophobic Assembly

Interfaces and the Driving Force of Hydrophobic Assembly (9/29/03) Submitted Nature insight review article Interfaces and the Driving Force of Hydrophobic Assembly David Chandler Department of Chemistry, University of California, Berkeley, Berkeley, CA 94720

More information

Effect of polarizability of halide anions on the ionic salvation in water clusters

Effect of polarizability of halide anions on the ionic salvation in water clusters University of Nebraska - Lincoln DigitalCommons@University of Nebraska - Lincoln Xiao Cheng Zeng Publications Published Research - Department of Chemistry 9-22-2003 Effect of polarizability of halide anions

More information

Crystal nucleation and growth both from melts and

Crystal nucleation and growth both from melts and pubs.acs.org/jpcl How Crystals Nucleate and Grow in Aqueous NaCl Solution Debashree Chakraborty and G. N. Patey* Department of Chemistry, University of British Columbia, Vancouver, British Columbia, Canada

More information

Module 4: "Surface Thermodynamics" Lecture 21: "" The Lecture Contains: Effect of surfactant on interfacial tension. Objectives_template

Module 4: Surface Thermodynamics Lecture 21:  The Lecture Contains: Effect of surfactant on interfacial tension. Objectives_template The Lecture Contains: Effect of surfactant on interfacial tension file:///e /courses/colloid_interface_science/lecture21/21_1.htm[6/16/2012 1:10:36 PM] Surface Thermodynamics: Roles of Surfactants and

More information

Molecular Dynamics Simulations. Dr. Noelia Faginas Lago Dipartimento di Chimica,Biologia e Biotecnologie Università di Perugia

Molecular Dynamics Simulations. Dr. Noelia Faginas Lago Dipartimento di Chimica,Biologia e Biotecnologie Università di Perugia Molecular Dynamics Simulations Dr. Noelia Faginas Lago Dipartimento di Chimica,Biologia e Biotecnologie Università di Perugia 1 An Introduction to Molecular Dynamics Simulations Macroscopic properties

More information

Water-methanol separation with carbon nanotubes and electric fields

Water-methanol separation with carbon nanotubes and electric fields Electronic Supplementary Material (ESI) for Nanoscale. This journal is The Royal Society of Chemistry 215 Supplementary Information: Water-methanol separation with carbon nanotubes and electric fields

More information

2. Derive ideal mixing and the Flory-Huggins models from the van der Waals mixture partition function.

2. Derive ideal mixing and the Flory-Huggins models from the van der Waals mixture partition function. Lecture #5 1 Lecture 5 Objectives: 1. Identify athermal and residual terms from the van der Waals mixture partition function.. Derive ideal mixing and the Flory-Huggins models from the van der Waals mixture

More information

Some properties of water

Some properties of water Some properties of water Hydrogen bond network Solvation under the microscope 1 NB Queste diapositive sono state preparate per il corso di Biofisica tenuto dal Dr. Attilio V. Vargiu presso il Dipartimento

More information

Chemical Physics Letters

Chemical Physics Letters Chemical Physics Letters 542 (2012) 37 41 Contents lists available at SciVerse ScienceDirect Chemical Physics Letters journal homepage: www.elsevier.com/locate/cplett Thermal conductivity, shear viscosity

More information

Supporting Information

Supporting Information Supporting Information ph-responsive self-assembly of polysaccharide through a rugged energy landscape Brian H. Morrow, Gregory F. Payne, and Jana Shen Department of Pharmaceutical Sciences, School of

More information

A MOLECULAR DYNAMICS SIMULATION OF WATER DROPLET IN CONTACT WITH A PLATINUM SURFACE

A MOLECULAR DYNAMICS SIMULATION OF WATER DROPLET IN CONTACT WITH A PLATINUM SURFACE The 6th ASME-JSME Thermal Engineering Joint Conference March 16-2, 23 TED-AJ3-183 A MOLECULAR DYNAMICS SIMULATION OF WATER DROPLET IN CONTACT WITH A PLATINUM SURFACE Tatsuto KIMURA Department of Mechanical

More information

SUPPLEMENTARY INFORMATION. Very Low Temperature Membrane Free Desalination. by Directional Solvent Extraction

SUPPLEMENTARY INFORMATION. Very Low Temperature Membrane Free Desalination. by Directional Solvent Extraction SUPPLEMENTARY INFORMATION Very Low Temperature Membrane Free Desalination by Directional Solvent Extraction Anurag Bajpayee, Tengfei Luo, Andrew J. Muto, and Gang Chen Department of Mechanical Engineering,

More information

Temperature and Pressure Dependence of the Interfacial Free Energy against a Hard. Surface in Contact with Water and Decane. J.

Temperature and Pressure Dependence of the Interfacial Free Energy against a Hard. Surface in Contact with Water and Decane. J. Temperature and Pressure Dependence of the Interfacial Free Energy against a Hard Surface in Contact with Water and Decane Henry S. Ashbaugh *, Natalia da Silva Moura, Hayden Houser, Yang Wang, Amy Goodson,

More information

Peptide folding in non-aqueous environments investigated with molecular dynamics simulations Soto Becerra, Patricia

Peptide folding in non-aqueous environments investigated with molecular dynamics simulations Soto Becerra, Patricia University of Groningen Peptide folding in non-aqueous environments investigated with molecular dynamics simulations Soto Becerra, Patricia IMPORTANT NOTE: You are advised to consult the publisher's version

More information

Aqueous solutions. Solubility of different compounds in water

Aqueous solutions. Solubility of different compounds in water Aqueous solutions Solubility of different compounds in water The dissolution of molecules into water (in any solvent actually) causes a volume change of the solution; the size of this volume change is

More information

Michael W. Mahoney Department of Physics, Yale University, New Haven, Connecticut 06520

Michael W. Mahoney Department of Physics, Yale University, New Haven, Connecticut 06520 JOURNAL OF CHEMICAL PHYSICS VOLUME 115, NUMBER 23 15 DECEMBER 2001 Quantum, intramolecular flexibility, and polarizability effects on the reproduction of the density anomaly of liquid water by simple potential

More information

Interfacial Thermodynamics of Water and Six Other Liquid Solvents

Interfacial Thermodynamics of Water and Six Other Liquid Solvents pubs.acs.org/jpcb Interfacial Thermodynamics of Water and Six Other Liquid Solvents Tod A. Pascal*, and William A. Goddard, III* Materials and Process Simulation Center, California Institute of Technology,

More information

UB association bias algorithm applied to the simulation of hydrogen fluoride

UB association bias algorithm applied to the simulation of hydrogen fluoride Fluid Phase Equilibria 194 197 (2002) 249 256 UB association bias algorithm applied to the simulation of hydrogen fluoride Scott Wierzchowski, David A. Kofke Department of Chemical Engineering, University

More information

VOLUME 34 NUMBER 12 DECEMBER 2001

VOLUME 34 NUMBER 12 DECEMBER 2001 VOLUME 34 NUMBER 12 DECEMBER 2001 Registered in U.S. Patent and Trademark Office; Copyright 2001 by the American Chemical Society Solvent Size vs Cohesive Energy as the THEMIS LAZARIDIS* Department of

More information

Lower critical solution temperature (LCST) phase. behaviour of an ionic liquid and its control by. supramolecular host-guest interactions

Lower critical solution temperature (LCST) phase. behaviour of an ionic liquid and its control by. supramolecular host-guest interactions Electronic Supplementary Material (ESI) for ChemComm. This journal is The Royal Society of Chemistry 2016 Lower critical solution temperature (LCST) phase behaviour of an ionic liquid and its control by

More information

Relevance of hydrogen bond definitions in liquid water

Relevance of hydrogen bond definitions in liquid water THE JOURNAL OF CHEMICAL PHYSICS 126, 054503 2007 Relevance of hydrogen bond definitions in liquid water Masakazu Matsumoto Department of Chemistry, Nagoya University, Furo-cho, Chikusa, Nagoya 464-8602,

More information

Soft Matter Accepted Manuscript

Soft Matter Accepted Manuscript Accepted Manuscript This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication. Accepted Manuscripts are published online

More information

Proceedings of the ASME 2009 International Mechanical Engineering Congress & Exposition

Proceedings of the ASME 2009 International Mechanical Engineering Congress & Exposition Proceedings of the ASME 9 International Mechanical Engineering Congress & Exposition IMECE9 November 3-9, Lake Buena Vista, Florida, USA Proceedings of the ASME International Mechanical Engineering Congress

More information

Effect of surfactant structure on interfacial properties

Effect of surfactant structure on interfacial properties EUROPHYSICS LETTERS 15 September 2003 Europhys. Lett., 63 (6), pp. 902 907 (2003) Effect of surfactant structure on interfacial properties L. Rekvig 1 ( ), M. Kranenburg 2, B. Hafskjold 1 and B. Smit 2

More information

Kirkwood Buff derived force field for mixtures of acetone and water

Kirkwood Buff derived force field for mixtures of acetone and water JOURNAL OF CHEMICAL PHYSICS VOLUME 118, NUMBER 23 15 JUNE 2003 Kirkwood Buff derived force field for mixtures of acetone and water Samantha Weerasinghe a) and Paul E. Smith b) Department of Biochemistry,

More information

Biophysics II. Hydrophobic Bio-molecules. Key points to be covered. Molecular Interactions in Bio-molecular Structures - van der Waals Interaction

Biophysics II. Hydrophobic Bio-molecules. Key points to be covered. Molecular Interactions in Bio-molecular Structures - van der Waals Interaction Biophysics II Key points to be covered By A/Prof. Xiang Yang Liu Biophysics & Micro/nanostructures Lab Department of Physics, NUS 1. van der Waals Interaction 2. Hydrogen bond 3. Hydrophilic vs hydrophobic

More information

Microscopic analysis of protein oxidative damage: effect of. carbonylation on structure, dynamics and aggregability of.

Microscopic analysis of protein oxidative damage: effect of. carbonylation on structure, dynamics and aggregability of. Microscopic analysis of protein oxidative damage: effect of carbonylation on structure, dynamics and aggregability of villin headpiece Drazen etrov 1,2,3 and Bojan Zagrovic 1,2,3,* Supplementary Information

More information

Introduction. Phys. Chem. Chem. Phys., 2010, Advance Article DOI: /C003407J (Paper)

Introduction. Phys. Chem. Chem. Phys., 2010, Advance Article DOI: /C003407J (Paper) 1 of 15 30/07/2010 17:41 Phys. Chem. Chem. Phys., 2010, Advance Article DOI:10.1039/C003407J (Paper) Lorna Dougan* a, Jason Crain bc, John L. Finney d and Alan K. Soper e a School of Physics and Astronomy,

More information

INTERMOLECULAR AND SURFACE FORCES

INTERMOLECULAR AND SURFACE FORCES INTERMOLECULAR AND SURFACE FORCES SECOND EDITION JACOB N. ISRAELACHVILI Department of Chemical & Nuclear Engineering and Materials Department University of California, Santa Barbara California, USA ACADEMIC

More information

An Introduction to Two Phase Molecular Dynamics Simulation

An Introduction to Two Phase Molecular Dynamics Simulation An Introduction to Two Phase Molecular Dynamics Simulation David Keffer Department of Materials Science & Engineering University of Tennessee, Knoxville date begun: April 19, 2016 date last updated: April

More information

State Key Laboratory of Molecular Reaction Dynamics, Dalian Institute of Chemical Physics, Chinese Academic of Sciences, Dalian , P. R. China.

State Key Laboratory of Molecular Reaction Dynamics, Dalian Institute of Chemical Physics, Chinese Academic of Sciences, Dalian , P. R. China. Electronic Supplementary Material (ESI) for ChemComm. This journal is The Royal Society of Chemistry 2015 Supplementary Information for: Theoretical exploration of MgH 2 and graphene nano-flake in cyclohexane:

More information

unprotonated, x unprotonated, y unprotonated, z protonated, x protonated, y protonated, z A [a.u.] ν [cm -1 ] narrow chain

unprotonated, x unprotonated, y unprotonated, z protonated, x protonated, y protonated, z A [a.u.] ν [cm -1 ] narrow chain a A [a.u.] narrow chain unprotonated, x unprotonated, y unprotonated, z protonated, x protonated, y protonated, z 1 2 3 4 5 A [a.u.] wide chain unprotonated, x unprotonated, y unprotonated, z protonated,

More information

Water liquid-vapor equilibrium by molecular dynamics: Alternative equilibrium pressure estimation

Water liquid-vapor equilibrium by molecular dynamics: Alternative equilibrium pressure estimation Water liquid-vapor equilibrium by molecular dynamics: Alternative equilibrium pressure estimation Michal Ilčin, Martin Michalík, Klára Kováčiková, Lenka Káziková, Vladimír Lukeš Department of Chemical

More information

CHAPTER 6 Intermolecular Forces Attractions between Particles

CHAPTER 6 Intermolecular Forces Attractions between Particles CHAPTER 6 Intermolecular Forces Attractions between Particles Scientists are interested in how matter behaves under unusual circumstances. For example, before the space station could be built, fundamental

More information

Lecture Notes 1: Physical Equilibria Vapor Pressure

Lecture Notes 1: Physical Equilibria Vapor Pressure Lecture Notes 1: Physical Equilibria Vapor Pressure Our first exploration of equilibria will examine physical equilibria (no chemical changes) in which the only changes occurring are matter changes phases.

More information

Structuring of hydrophobic and hydrophilic polymers at interfaces Stephen Donaldson ChE 210D Final Project Abstract

Structuring of hydrophobic and hydrophilic polymers at interfaces Stephen Donaldson ChE 210D Final Project Abstract Structuring of hydrophobic and hydrophilic polymers at interfaces Stephen Donaldson ChE 210D Final Project Abstract In this work, a simplified Lennard-Jones (LJ) sphere model is used to simulate the aggregation,

More information

STRUCTURE OF IONS AND WATER AROUND A POLYELECTROLYTE IN A POLARIZABLE NANOPORE

STRUCTURE OF IONS AND WATER AROUND A POLYELECTROLYTE IN A POLARIZABLE NANOPORE International Journal of Modern Physics C Vol. 2, No. 9 (29) 1485 1492 c World Scientific Publishing Company STRUCTURE OF IONS AND WATER AROUND A POLYELECTROLYTE IN A POLARIZABLE NANOPORE LEI GUO and ERIK

More information

Molecular dynamics simulation of limiting conductances for LiCl, NaBr, and CsBr in supercritical water

Molecular dynamics simulation of limiting conductances for LiCl, NaBr, and CsBr in supercritical water JOURNAL OF CHEMICAL PHYSICS VOLUME 112, NUMBER 2 8 JANUARY 2000 Molecular dynamics simulation of limiting conductances for LiCl, NaBr, and CsBr in supercritical water S. H. Lee Department of Chemistry,

More information

Comparative study on methodology in molecular dynamics simulation of nucleation

Comparative study on methodology in molecular dynamics simulation of nucleation THE JOURNAL OF CHEMICAL PHYSICS 126, 224517 2007 Comparative study on methodology in molecular dynamics simulation of nucleation Jan Julin, Ismo Napari, and Hanna Vehkamäki Department of Physical Sciences,

More information

MODELING THE SURFACE TENSION OF PURE ALKANES AND PERFLUOROALKANES USING THE

MODELING THE SURFACE TENSION OF PURE ALKANES AND PERFLUOROALKANES USING THE MODELING THE SURFACE TENSION OF PURE ALKANES AND PERFLUOROALKANES USING THE soft-saft EoS COUPLED WITH THE GRADIENT THEORY EXTENDED TO THE CRITICAL REGION A. M. A. Dias 1, F. Llovell 2, J. A. P. Coutinho

More information

MODIFIED PROXIMITY CRITERIA FOR THE ANALYSIS OF THE SOLVATION OF A POLYFUNCTIONAL SOLUTE.

MODIFIED PROXIMITY CRITERIA FOR THE ANALYSIS OF THE SOLVATION OF A POLYFUNCTIONAL SOLUTE. Molecular Simulation 1988, Vol. 1 pp.327-332 c 1988 Gordon and Breach Science Publishers S.A. MODIFIED PROXIMITY CRITERIA FOR THE ANALYSIS OF THE SOLVATION OF A POLYFUNCTIONAL SOLUTE. MIHALY MEZEI Department

More information

Molecular Origin of Hydration Heat Capacity Changes of Hydrophobic Solutes: Perturbation of Water Structure around Alkanes

Molecular Origin of Hydration Heat Capacity Changes of Hydrophobic Solutes: Perturbation of Water Structure around Alkanes J. Phys. Chem. B 1997, 101, 11237-11242 11237 Molecular Origin of Hydration Heat Capacity Changes of Hydrophobic Solutes: Perturbation of Water Structure around Alkanes Bhupinder Madan and Kim Sharp* The

More information

An Extended van der Waals Equation of State Based on Molecular Dynamics Simulation

An Extended van der Waals Equation of State Based on Molecular Dynamics Simulation J. Comput. Chem. Jpn., Vol. 8, o. 3, pp. 97 14 (9) c 9 Society of Computer Chemistry, Japan An Extended van der Waals Equation of State Based on Molecular Dynamics Simulation Yosuke KATAOKA* and Yuri YAMADA

More information

Water and Aqueous Solutions. 2. Solvation and Hydrophobicity. Solvation

Water and Aqueous Solutions. 2. Solvation and Hydrophobicity. Solvation Water and Aqueous Solutions. Solvation and Hydrophobicity Solvation Solvation describes the intermolecular interactions of a molecule or ion in solution with the surrounding solvent, which for our purposes

More information

arxiv: v2 [cond-mat.stat-mech] 23 Sep 2009

arxiv: v2 [cond-mat.stat-mech] 23 Sep 2009 arxiv:0909.4097v2 [cond-mat.stat-mech] 23 Sep 2009 Fluctuations of water near extended hydrophobic and hydrophilic surfaces Amish J. Patel and David Chandler Department of Chemistry, University of California,

More information

Force Field for Water Based on Neural Network

Force Field for Water Based on Neural Network Force Field for Water Based on Neural Network Hao Wang Department of Chemistry, Duke University, Durham, NC 27708, USA Weitao Yang* Department of Chemistry, Duke University, Durham, NC 27708, USA Department

More information

Heat capacity of water: a signature of nuclear quantum effects. Abstract

Heat capacity of water: a signature of nuclear quantum effects. Abstract Heat capacity of water: a signature of nuclear quantum effects C. Vega a, M. M. Conde a, C. McBride a, J. L. F. Abascal a, E. G. Noya b, R. Ramirez c and L. M. Sesé d a Departamento de Química Física,

More information

MD simulation of methane in nanochannels

MD simulation of methane in nanochannels MD simulation of methane in nanochannels COCIM, Arica, Chile M. Horsch, M. Heitzig, and J. Vrabec University of Stuttgart November 6, 2008 Scope and structure Molecular model for graphite and the fluid-wall

More information

Water. 2.1 Weak Interactions in Aqueous Sy stems Ionization of Water, Weak Acids, and Weak Bases 58

Water. 2.1 Weak Interactions in Aqueous Sy stems Ionization of Water, Weak Acids, and Weak Bases 58 Home http://www.macmillanhighered.com/launchpad/lehninger6e... 1 of 1 1/6/2016 3:07 PM 2 Printed Page 47 Water 2.1 Weak Interactions in Aqueous Sy stems 47 2.2 Ionization of Water, Weak Acids, and Weak

More information

Molecular Driving Forces

Molecular Driving Forces Molecular Driving Forces Statistical Thermodynamics in Chemistry and Biology SUBGfittingen 7 At 216 513 073 / / Ken A. Dill Sarina Bromberg With the assistance of Dirk Stigter on the Electrostatics chapters

More information

Biochemistry,530:,, Introduc5on,to,Structural,Biology, Autumn,Quarter,2015,

Biochemistry,530:,, Introduc5on,to,Structural,Biology, Autumn,Quarter,2015, Biochemistry,530:,, Introduc5on,to,Structural,Biology, Autumn,Quarter,2015, Course,Informa5on, BIOC%530% GraduateAlevel,discussion,of,the,structure,,func5on,,and,chemistry,of,proteins,and, nucleic,acids,,control,of,enzyma5c,reac5ons.,please,see,the,course,syllabus,and,

More information

Supporting Information

Supporting Information Supporting Information Interface-Induced Affinity Sieving in Nanoporous Graphenes for Liquid-Phase Mixtures Yanan Hou, Zhijun Xu, Xiaoning Yang * State Key Laboratory of Material-Orientated Chemical Engineering,

More information

Free energy, electrostatics, and the hydrophobic effect

Free energy, electrostatics, and the hydrophobic effect Protein Physics 2016 Lecture 3, January 26 Free energy, electrostatics, and the hydrophobic effect Magnus Andersson magnus.andersson@scilifelab.se Theoretical & Computational Biophysics Recap Protein structure

More information

Water models in classical simulations

Water models in classical simulations Water models in classical simulations Maria Fyta Institut für Computerphysik, Universität Stuttgart Stuttgart, Germany Water transparent, odorless, tasteless and ubiquitous really simple: two H atoms attached

More information