Colloids and Surfaces A: Physicochemical and Engineering Aspects

Size: px
Start display at page:

Download "Colloids and Surfaces A: Physicochemical and Engineering Aspects"

Transcription

1 Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) Contents lists available at SciVerse ScienceDirect Colloids and Surfaces A: Physicochemical and Engineering Aspects jo ur nal homep a ge: Ions near the air/water interface: I. Compatibility of zeta potential and surface tension experiments Marian Manciu a,, Eli Ruckenstein b a Physics Department, University of Texas at El Paso, TX, United States b Chemical and Biological Engineering Department, State University of New York at Buffalo, United States a r t i c l e i n f o Article history: Received 27 December 211 Received in revised form 18 February 212 Accepted 25 February 212 Available online 5 March 212 Keywords: Air/water interface Surface tension Ion adsorption Zeta potential a b s t r a c t A simple Structure Making/Structure Breaking (SM/SB) Modified Poisson Boltzmann approach is used to describe the distribution of ions in the vicinity of an air/water interface, which is assumed to be a layer possessing a thickness of a few Å. The interface is assumed to be depleted of SM ions, and the adsorption of SB ions on the interface is taken into account via Langmuir adsorption equations. Outside the interface, the ions are assumed to obey a modified Poisson Boltzmann distribution, which includes the screened image forces. The adsorption constants of SB ions, determined by fitting the existing zeta potential data, are further used to determine the surface excesses of salt ions and the dependence of the surface tension on added salt. The number of sites available for the adsorption of OH and H + ions, considered the water molecules of the interface, is (by more than two orders of magnitude) smaller than the available sites in an ice-like layer. This result suggests a (much) lower density of water at the hydrophobic air/water interface than in the bulk. The model can qualitatively explain the dependence of the surface tension upon addition of salt determined experimentally: a large decrease at low concentrations, due to the formation of a double layer, a large increase at intermediate concentrations, where the screened image forces (being less screened than the double layer forces) dominate, and a slower increase at large concentrations, where both double layer and image forces are screened and the ion hydration forces dominate. However, for large surface potentials, the accumulation of ions in the double layer (either as adsorbed on the interface or in the ion clouds in the vicinity of the interface) leads to large surface excesses of ions and consequently to a surface tension that decreases upon addition of salt up to large concentrations, in contrast to experiment. According to the present model, only the zeta potentials experiments that predict low surface potentials and small adsorption of salt anions on the surface are compatible with the surface tension experiments. 212 Elsevier B.V. All rights reserved. 1. Introduction It has been long emphasized that the concentrations of ions at the water/air interface are markedly different from those in bulk water. The presence of a dielectric discontinuity generates an image force on any approaching charge. Furthermore, the dielectric constant of the interfacial water is expected to be lower than that of bulk water, for two main reasons. Firstly, the rotation of the water molecules is somewhat obstructed at the interface, hence they are less likely to screen the electric field generated by surface charges. Secondly, and even more important, there is evidence [1,2] that the density of water at the hydrophobic air/water interface is much smaller than in the bulk, hence there are fewer permanent dipoles available to screen the electric field. Corresponding author. Tel.: ; fax: addresses: mmanciu@utep.edu (M. Manciu), feaeliru@buffalo.edu (E. Ruckenstein). Langmuir [3] even suggested that the increases in the surface tension of water upon addition of salts can be explained by assuming that the first water layer, with a thickness of about 4 Å, is completely depleted of ions. However, experimental data show that the surface tensions of acid solutions decrease with increasing acid concentration [4]. This indicates a positive net surface excess of ions, hence that at least H +, or the anion, or possibly even both are accumulating at the water/air interface. Furthermore, the observation of Jones and Ray [5], that there is a minimum surface tension around 1 mm for many salts, implies that surface excesses of at least one kind of ions are common at low electrolyte concentrations. Among the experiments most directly related to the general behavior of ions near an air/water interface are the determinations of the zeta potential and surface tension as functions of ph and salt concentration. The former experiments show whether the anions or cations are dominant at the interface (being either more adsorbed or less depleted than the ions of opposite charge), whereas the latter experiments indicate whether there is a net surface excess or a net surface depletion of all the ions added to /$ see front matter 212 Elsevier B.V. All rights reserved. doi:1.116/j.colsurfa

2 28 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) the solution. Whereas such conclusions can be derived straightforwardly from experiment, the results are sometimes controversial. For example, it was assumed that the surface tension increases monotonically upon addition of salts [4], in agreement with the screened image forces theories of Wagner [6] and Onsager and Samaras [7]. However, the Jones and Ray experiments [5] exhibit a decrease in surface tension at low electrolyte concentrations, in a range where one should expect, because of reduced screening, stronger image forces than at high electrolyte concentrations. The dependence of the zeta potential on ph and salt concentration is also a topic of debate. Many experiments show that the zeta potential of water is negative, over almost the entire ph range (with the exception of very low phs possible less than 2 [8]). Recently, it was suggested that an isoelectric point might occur around ph = 4 [9]. Gray-Weale and Beattie [1] measured the zeta potential of a water/oil interface and obtained very low absolute values (albeit negative) at low phs. They suggested that an isoelectric point occurs when the absolute values of their experimentally determined zeta potentials are less than 3 mv (the experimental error), values obtained between 3 < ph < 4 [1]. These results are compatible with those of Ref. [9] for an air/water interface; however, the data of Ref. [8] provide in that ph range much larger absolute values for the zeta potential, around 1 mv, which cannot be attributed to experimental inaccuracies. Zeta potential experiments support the hypothesis that some kinds of ions are strongly adsorbed at the hydrophobic air/water interface. Let us assume that the (negative) surface charge of neat water is the result of the depletion of interfacial layer of H +, but not of OH. The corresponding surface charge density is in this case given by = coh q, where is the thickness of the depletion layer, q the electron charge and c OH the bulk concentration of OH. For neat water (c OH = 1 7 M) and a depletion layer = 4 Å, is C/m 2, and the corresponding surface potential (calculated in the linear approximation, S = /εε, with the Debye Huckel length, ε the dielectric constant of water and ε the vacuum permittivity) is only about V. However, experiments [8 1] indicate that the zeta potential of neat water is about 4 orders of magnitude larger, which in turn suggests a strong adsorption of OH on the hydrophobic interface. To explain these experimental results, the interfacial concentration of OH has to be by orders of magnitude larger than in the bulk. Similarly, the large (positive) zeta potential reported at ph = 3 in Ref. [9] cannot be explained by an interfacial depletion of OH ; the concentration of H + should be larger by at least an order of magnitude than in the bulk. Water cluster calculations [11] and molecular dynamics simulations for water slabs [12] indicated that some simple anions (Cl, Br, I ) have a preference for the interface, proportional to their polarizability [12]. Subsequent molecular dynamics calculations supported the idea that there is an excess of hydrogen and of simple anions (with the exception of F and OH ) on the interface [13]. This result, corroborated with the positive zeta potential values measured at very low ph values [9], stimulated the opinion that the surface of water is acidic [14]. This conclusion, mostly based on simulations, drew a lot of attention. It was immediately pointed out that many experimental results (such as the negative zeta potential of air bubbles or oil droplets in neat water, or the formation of hexadecane emulsions in neat water, which are stabilized via the adsorption of OH ), actually suggest that the neat water surface is basic [15]. Recent experiments further deepened the debate; whereas experiments on the zeta potential of water indicated a more basic interface than the bulk [8 1], other experiments (Second Harmonic Generation [16], Vibrational Sum-Frequency Spectroscopy [17], Phase Sensitive Sum-Frequency Vibrational Spectroscopy [18], Photoelectron Spectroscopy [19]) indicated the opposite, namely that the surface adsorption is dominated by H + in detriment of OH. Recent calculations did not resolve the issue: whereas some predicted a lower [12], other predicted a larger [2] concentration of OH at the interface than in bulk water. The goal of this paper is to investigate the compatibility of a very simple Structure Making/Structure Breaking (SM/SB) model [21] with the existing experiments on both the surface tension [5] and the zeta potential at various salt concentrations and ph values [8 1]. When the SM ions approach the interface at a distance of about 4 Å, they cannot be hydrated as well as in bulk and therefore their free energy changes. Whereas the dependence of ion hydration on the distance to the surface is not known, because the bulk hydration free energy is very large (1 2 kt), only a few percent change in ion hydration energy (which occurs over a very short distance), leads to a large change in their free energy [21]. Therefore, the air/water interface can be considered (up to the distance ) practically depleted of SM ions. Their distribution can be approximated by a modified Poisson Boltzmann model (that includes image forces) for distances larger than from the interface with a vanishing concentration at smaller distances. On the other hand, the SB ions are attracted by the interface via complex ion-hydration forces, which account not only for the free energy changes of the SB ions, but also for the free energy changes of the water molecules surrounding these ions. Such forces are difficult to estimate because they must involve the local changes in water density and dielectric constant, volume exclusion effects, ion dispersion forces, etc. For the sake of simplicity, it will be assumed that the adsorption of all SB ions on the interface of thickness follows the Langmuir isotherms. The parameters of the adsorption isotherms, which account for the forces that act close to the interface, will be evaluated by fitting the experimental data regarding the zeta potential. At distances larger than, it will be assumed that the ions follow modified Poisson Boltzmann distributions (which include image forces). At distances larger than, most interactions of the ions (except the mean field and the image forces) are short ranged and can be neglected. As an example of short-range forces, the energies associated with ion-dispersions interactions (which are of the form B/x 3, with the constant B depending on the nature of ion), are typically smaller than kt at x = 4 Å from the interface (between.4 and 1.17 kt, for B ranging from to Jm 3 [21]), hence their effect on the ion distributions outside the interfacial layer is negligible. The Langmuir adsorption model of the interface might constitute an oversimplification; however, it is difficult to build an accurate model of the air/water interface, since the structure of water in the first few Å from the interface is not well known. For example, it was suggested that the density of water in that region is much smaller than in the bulk [1,2], which is indeed supported by the results obtained in this paper. The adsorption constants will be obtained by fitting the zeta potential data; they will be used to calculate the dependence of surface tension on the electrolyte concentration. It will be shown that the zeta potentials obtained by three groups [8 1] are not entirely compatible with each other, and that some of them are not compatible with the experimental data of Jones and Ray regarding the surface tension [5] either. The main reason for the incompatibility is that large surface potentials generate large surface excesses of ions due to the double layer ion clouds in the vicinity of the surface. Such excesses lead to a decrease in surface tension with increasing electrolyte concentration up to large electrolyte concentrations, much larger than the values around 1 mm obtained experimentally [5]. Even more striking, if the surface charges would be generated by the adsorption of salt anions, their surface excess would become so large, that the surface tension would decrease even for large electrolyte concentrations, in contradiction with experiment, which shows that it should increase [4]. Therefore, the critical concentration at which the minimum surface tension is obtained in the Jones Ray effect provides some insight on the nature of ions that

3 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) produce the negative charging of the surface of neat water observed experimentally [8 1]. For agreement with Jones Ray data, the surface potentials should be lower than typically reported [8,1] and the surface charge should be generated mostly by the adsorption of OH. However, the set of adsorption parameters obtained from fitting the data of Ref. [9] obey these conditions and provides good agreement with the Jones Ray data [5]. 2. The adsorption desorption model for SM/SB ions As in zeta potential experiments [8 1], the added salt is NaCl and the change in ph is achieved by using either HCl or NaOH. It will be assumed that the Structure Breaking (SB) anions, OH and Cl, as well as the small cation H +, are adsorbed on the interface, whereas the Structure Making (SM) cations (such as Na + ) prefer the bulk water. The ions adsorbed on the interface are in equilibrium with the ions (of the same kind) located at a distance from the interface, the concentration c i of which is related to their bulk concentration, c,i, by the Boltzmann factor: ( ) c i = c,i exp q i S + W i () (1) kt In Eq. (1), q i is the charge, S the surface potential, k the Boltzmann constant, T the absolute temperature and W i a free energy change due to additional interactions, not included in the mean field potential (such as the image forces [6,7]), calculated at a distance from the interface. Assuming that the added acid, base and salt are completely dissociated, the bulk concentrations of H + and OH are provided by the ph of the solution, whereas the amount of HCl (or NaOH) employed to change the ph affects the dissociation equilibrium of the water molecules: c,h c,oh = (c HCl c R H )(1 7 c R OH ) = 1 14 (2a) c,h c,oh = (c NaOH c R OH )(1 7 c R H ) = 1 14 (2b) where c,h and c,oh are the bulk concentrations of H + and OH ions, 1 7 is the concentration of H + and OH before adding the acid, c HCl is the concentration of H + due to the added acid, c NaOH is the concentration of OH due to the added base, and c R H = c R OH is the number of H + and OH ions, per unit volume, that recombines into water molecules. Consequently, the bulk concentrations of ions are provided by the expressions: c,h = 1 ph ; c,oh = 1 ph 14 ; c,cl = c HCl + c NaCl ; c,na = c NaOH + c NaCl (3) It will be assumed that the surface potential is equal to the zeta potential. The short range interactions (which are smaller than kt at distances larger than a few Å from the interface) will be neglected in Eq. (1), and only the long ranged image forces will be taken into account. In the Onsager Samaras model, they are provided by the equation [7]: ( ε ) ε exp(ai /) q 2 ( i W i (x) = ε + ε 1 + (a i /) 16ε εx exp 2x ), (x > ) (4) where x is the distance from the dielectric boundary, a i the radius of ion i, ε the vacuum permittivity, ε the dielectric constant of water, ε the dielectric constant of the other medium and = (εε kt/ c i q 2 i ) is the Debye Hückel length. Because i exp(a i /) = 1+ (a i /) + (1/2)(a i /) , the ion specificity effects are negligible until very large electrolyte concentrations ( 1 M), when becomes comparable with a i. The mean field potential is assumed to obey the modified Poisson Boltzmann equation for distances larger than : q i c,i exp( ((q i (x) + W i (x))/kt)) 2 (x) = i with the boundary conditions: ε ε (x) x = and = εε d dx x=, (x > ) where the surface charge density is provided by the number of ions adsorbed per unit area, N i : = q(n H N OH N Cl ) (7) q being the proton charge and i = H +, OH or Cl. We will consider that the water molecules at the interface provide the adsorption sites for H + and OH, which are competing for them, and that Cl is located in the free spaces between the water molecules. The H + and OH ions compete for N S sites (assumed to be the number of molecules of water per unit interfacial area), with the equilibrium constants K H and K OH, respectively, while Cl is adsorbed on (different) N A sites, with the equilibrium constant K Cl. The equilibrium equations have the form: K H H = c H (1 H OH ) (8a) K OH OH = c OH (1 H OH ) (8b) K Cl Cl = c Cl (1 Cl ) (8c) where i is the fraction of sites occupied by ions of kind i and c i their concentration near the interface (per unit volume). The number of ions adsorbed per unit area is therefore provided by: N H = N OH = N S 1 + (K H /c H ) + (K H /K OH )(c OH /c H ) N S 1 + (K OH /c OH ) + (K OH /K H )(c H /c OH ) (5) (6) (9a) (9b) N A N Cl = (9c) 1 + (K Cl /c Cl ) For a given set of parameters (N S, N A, K H, K OH, K Cl ), Eqs. (1 7) and (9) can be solved numerically, and the surface potential S = (x) x=, considered equal to the zeta potential, can be calculated. The set of parameters is then optimized to fit the experimental data for the zeta potential as follows: a penalty function, obtained as the sum of the square differences between the calculated and experimental zeta potentials is minimized using the Nelder Mead simplex direct search algorithm implemented in MATLAB. The dependence of the interfacial tension on the uni-univalent electrolyte concentration c E (which can be either an acid, base, or salt) can be calculated using the Gibbs adsorption equation: d = i d i (1) i=1,2 where i is the electrochemical potential and i is the surface excess of ion i: i (c E ) = (c i (x) c E )dx = (c i (x) c E )dx + (c i (x) c E )dx = c E + c i (x)dx + (c i (x) c E )dx (11) with the water/air interface located between x = (corresponding to air) and x =. The last integral corresponds to the surface adsorption of ions in the Poisson Boltzmann range (x > ), and, whereas

4 3 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) the ion distribution c i (x) is not known between and, its integral is provided by the number of ions adsorbed per unit area on the interface, which can be obtained from Eq. (9). The electrochemical potential, which is independent of the distance from the interface, can be related to the bulk electrolyte concentration c E = c i ( ) via: d i = d(kt ln(c i ( )f i )) (12) where f i is the activity coefficient of the electrolyte in the bulk, which will be assumed to be unity. When salt is added to the system, but the ph remains constant, there is no change in the electrochemical potential for either H + and OH, hence their adsorption/desorption does not contribute to changes in surface tension upon addition of salt. Consequently, only the surface excesses of Cl (a SB ion, that can be adsorbed on the interface) and Na + (a SM ion, that cannot be adsorbed on the interface) have to be taken into account. The change in interfacial tension due to the addition of an electrolyte is given by: 1 = Na + Cl 2 (13) kt c E c E where the ratio = (( Na + Cl )/2c E ) has the dimension of a length, which can be interpreted as the thickness of an interfacial region, totally depleted of ions, that produces the same change in surface tension as the actual distribution of ions. A negative value of represents a surface excess of ions, which decreases the surface tension. The surface tension as a function of electrolyte concentration can be obtained by integrating Eq. (13): ce ce (c E ) () = c dc Na (c) + = kt Cl (c) dc c c= ce = 2kT dc (14) 3. Surface charge and surface tension In what follows, we will examine the zeta potentials determined by Li and Somasundaran [8], Takahashi [9] and Gray-Weale and Beattie [1] (see Fig. 1a). In all cases, NaCl was the added salt and the ph was changed either by adding HCl or NaOH. It should be noted that for phs lower than about 4, the bulk concentrations of H +, OH and Cl ions in Ref. [8] (salt concentration c S = 1 5 mol/l) and Ref. [9] (c S = ) are close; however the zeta potentials are sharply different. Furthermore, there are also sharp differences between the data reported in Refs. [8,1], which can be, however, attributed to the different kinds of interfaces involved in their experiments (air/water and oil/water, respectively). Therefore, the tentative fitting of all the data available (Fig. 1a) is not accurate. In Fig. 1b, the surface charge is plotted for the same data as a function of the concentration of OH in the vicinity of the interface (provided by the Boltzmann factor). In spite of the discrepancies between the reported data [8 1], the surface charge densities in all of them are low, indicating either a very weak adsorption of ions on a large number of sites, or a limited number of sites available for adsorption. The first hypothesis is not plausible, because it should predict a linear increase of the negative surface charge with increasing OH concentration (at least at very low salt concentrations c S and sufficiently high phs, for which Cl adsorption can be neglected). In contrast, Fig. 1b shows that the surface charge remains almost constant in a relatively large range of OH concentrations in the vicinity of the air/water interface. The data reported by Gray-Weale and Beattie [1] involving oil in water emulsions exhibit a stronger Fig. 1. (a) Zeta potentials vs. ph for various salt (NaCl) concentrations, reported by Li and Somasundran [8], Takahashi [9] and Gray-Weale and Beattie [1]. The lines represent an unsuccessful attempt to fit all the data together. (b) The surface charges corresponding to the zeta potentials in (a), as functions of ph, for various salt concentrations. dependence of the surface charge on the OH concentration in the vicinity of the interface, which suggests different adsorption equilibria for OH on the oil/water than on the air/water interface. A strong adsorption of OH in the former case is also supported by the increased stability of the oil in water emulsions with increasing ph [1]. The increase (in absolute value) of the negative surface charge density (Fig. 1b), with increasing Cl concentration (generated by adding either salt or acid) indicates that Cl is adsorbed on the water/air interface. The continuous lines in Fig. 1a and b represent results of calculations based on set 1 of parameters (Table 1) obtained by fitting all zeta potential data of Refs. [8 1]. In spite of the large number (5) of fitting parameters, the agreement between calculations and experimental data is quite poor, probably because of the incompatibility between the experimental results discussed above. Let now focus on the surface tension data. As well known, the addition of a salt at constant ph leads to an almost linear increase in surface tension at large electrolyte concentrations [4]. However, at very low electrolyte concentrations the surface tension first decreases and then increases [5], indicating that the total adsorption of salt ions is positive at low concentrations, but becomes negative at large salt concentrations. In Fig. 2, the derivative of the

5 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) Table 1 Equilibrium adsorption parameters. N S (sites/m 2 ) N A (sites/m 2 ) K OH (M) K H (M) K Cl (M) Set 1 (Refs. [8 1]) Set 2 (Refs. [8]) Set 3 (Ref. [9]) Set 4 (Ref. [1]) Fig. 2. The depletion lengths for salt ions vs. salt concentration, calculated from the derivative of the experimental surface tension data, from Refs. [4,5]. The depletion lengths due to double layer, ion hydration and screened image forces alone, as well as all of them together are plotted. surface tension, which is proportional to the depletion length 2, is plotted against the electrolyte concentration. The data have been obtained via the numerical derivative of the surface tension experimental data from Refs. [4,5], which coincide for sufficiently large salt concentrations. At very low salt concentrations, the depletion length is large in absolute value and negative; at intermediate concentrations, it is large and positive; whereas at large concentrations it is positive and almost constant. This behavior can be attributed to the formation of a double layer at low salt concentrations, which being screened stronger becomes less important than the image forces [6,7] at intermediate salt concentrations. When the salt concentration becomes sufficiently large, both the double layer and the image forces are screened and the ion-hydration forces (which are not screened by ions [21]) become dominant. The depletion induced either by double layer forces alone, by image forces alone, or by SM forces alone, as well as the depletion induced by all three together are plotted in Fig. 2. To understand these results, let us examine the depletion length generated by the double layer alone. The total adsorption of ions can be obtained from Eq. (11): Na + Cl = (c Na (x) + c Cl (x) 2c S )dx + ( c S exp ( q (x) kt ( ) ) q (x) +c S exp 2c S dx (15) kt ) After expanding the exponential, the depletion length (Eq. (13)) becomes: = Na + Cl 1 c 2c S 2c Cl (x)dx S ( 1 2 ( q (x) kt ) 2 ( + 1 q (x) 24 kt ) 4 + ) dx (16) where it was assumed that Na + are SM ions, hence c Na (x) = for x < <. At very large electrolyte concentrations, when the double layer collapses, both integrals are vanishingly small and is a positive constant that provides a linear increase in surface tension with increasing electrolyte concentration. At low electrolyte concentrations, the integrals cannot be neglected, and they will be examined separately in what follows. The distribution of Cl between and is not known, but its integral provides the total number of Cl adsorbed per unit surface. The corresponding depletion length is: 1 c 2c Cl (x)dx = ϕ S (17a) S 2c S q where ϕ represents the fraction of the surface charge S generated by the adsorption of Cl ions, while (1 ϕ) S is due to the adsorption of OH and H +. The depletion length generated by the repulsive forces (screened image and ion-hydration forces), at c S = 1 mm, is about 2 3 Å (see Fig. 2). Assuming ϕ = 1 and a surface charge S of the order of 1 2 C/m 2, as suggested by Fig. 1b, the depletion length provided by Eq. (17a) is of the order of 1 3 Å. To obtain a vanishing depletion length at 1 mm, as obtained in the Jones Ray experiment [5], either the surface charge should be much smaller than those calculated from the zeta potentials determined experimentally [8 1], or the Cl adsorption should be very small (ϕ ); as mentioned above, H + and OH surface excesses do not contribute to the changes in surface tension upon addition of salt. Note that the zeta potential experiments from Ref. [9] were performed at c S =, hence the surface charge at c S = 1 mm cannot be inferred directly from these experiments. The second integral in Eq. (16) contains a series of only positive terms. Since the potential behaves in the bulk roughly as (x) Sexp( (x/)), at low electrolyte concentrations (hence large Debye Huckel lengths ), the integral is much larger than and is large in absolute value and negative. By assuming an exponential dependence of the potential (x), the second integral in Eq. (16) can be evaluated: ( ( ) 2 ( ) 4 1 q (x) + 1 q (x) + ) dx 2 kt 24 kt ( ( ) 2 ( ) 4 1 q () = + 1 q () + ) (17b) 4 kt 48 kt Since (q S/kT) is usually of the order of unity ( S = ()), the depletion length due to the ion clouds near the surface

6 32 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) Fig. 3. The depletion length due to image forces, calculated by assuming that the image forces are not screened over a distance l from the interface, vs. l, at various salt concentrations. is of the order of the Debye Huckel length. At c S = 1 mm, 1 Å, much larger than the depletion length due to the Wagner Onsager Samaras image forces (See Fig. 2). Consequently, only for small surface potentials the total depletion length will vanish at c S = 1 mm, as obtained by Jones and Ray [5]. Can a large in absolute value and negative depletion length be compensated by additional repulsive forces acting between the salt ions and the interface (that have been neglected in the present model)? One possibility is that the ions in the vicinity of the interface are not completely free to move and hence the image force is not fully screened in the vicinity of the interface. The non-screened image forces provide divergently large depletion lengths [6]. However, the divergence is logarithmic and increases very slowly with the non-screening distance. To evaluate this effect, let assume that the image forces are provided by: W i (x) = q 2 i 16ε εx, < x < l (18a) W i (x) ( ε ) ε exp(ai /) q 2 ( i = ε + ε 1 + (a i /) 16ε εx exp 2(x ) l), l < x < (18b) where l represents the distance from the surface up to which the image force is not screened. The depletion length is plotted in Fig. 3 vs l for various c S and shows that even for unreasonably large l values, the depletion length is still smaller than 5 Å, therefore cannot compensate the (negative) depletion length generated at 1 mm by the double layer, for large S values. The volume exclusion effects increase slightly the depletion length, but for c S = 1 mm and surface charge S = 1 2 C/m 2 (see Fig. 1b), neither the concentration of Na + in the vicinity of the interface (x > ), which is less than 1 2 M, nor the concentration of Cl ion on the surface (less than 1 ion/1 3 Å 2 ) can provide a long-ranged volume-exclusion repulsion. The strongest ion-dispersion forces lead to a maximum depletion length of less than 1 Å [21], hence they cannot compensate a strong double layer, either. The surface potential compatible with Jones Ray data can be determined by calculating the critical salt concentration at which the minimum in surface tension occurs (when the depletion length vanishes). The calculation implies the existence of a double layer and accounts for the ion hydration and screened image forces, but Fig. 4. (a) The surface potentials vs. the critical salt concentration, calculated for the two extreme cases, ϕ = and ϕ = 1. (b) The surface charges vs. the critical salt concentration, calculated for the two extreme cases, ϕ = and ϕ = 1. is independent of the model for surface charging. Two extreme cases are considered, ϕ = 1 (the surface charge is completely generated by the adsorption of Cl ) and ϕ = (no Cl is adsorbed on the interface); these extremes provide a range of surface potentials compatible with the Jones Ray experiment. In Fig. 4a, the surface potentials are plotted vs. the critical salt concentrations for which a minimum in surface tension has been obtained. To obtain a minimum in surface tension at 1 mm, the surface potential (which is assumed equal to the zeta potential) should be located between 2 and 16 mv, corresponding to a surface charge generated by the adsorption of only Cl (ϕ = 1) and only OH (ϕ = ), respectively. The surface potentials required for the Jones Ray minimum at 1 mm are smaller than the experimental values usually reported in literature. It is possible that the surface charge is generated mostly by the adsorption of the salt anion, as suggested recently [22], but this will lead at even smaller zeta potentials than when the adsorption of OH generates most of the surface charge. The surface charge required for a minimum at 1 mm (Fig. 4b), which is between 1 4 and 1 3 C/m 2, is also much smaller than those corresponding to the zeta potentials determined experimentally (see Fig. 1b). It should be emphasized that the results of Eqs. (17a) and (17b) are general characteristics of the double layer and do not depend

7 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) Fig. 6. (a) The surface tension vs. salt concentration, with the adsorption parameters determined by fitting the data of Refs. [8,1] are not compatible with Jones Ray data. (b) The surface tension vs. salt concentration, with the adsorption parameters determined by fitting the data of Ref. [9] is compatible with Jones Ray data. upon the model selected for the charging of the air/water interface (the Langmuir adsorption isotherm in this paper). A large zeta potential, hence a large surface potential, and a corresponding large surface charge generates large surface excesses of salt ions via two mechanisms: the contribution of the salt ions to the surface charge, and the net accumulation of ions in the ion cloud due to the double layer. While the former can be small if the fraction of the surface charge, ϕ, generated by the salt anions is sufficiently small, the second is small only when the surface potential is sufficiently small. 4. Compatibility between surface tension and zeta potential data Fig. 5. (a) Zeta potentials vs. ph for various salt (NaCl) concentrations, reported by Li and Somasundran [8]. The lines represent the prediction of the present model using fitted adsorption parameters. (b) Zeta potentials vs. ph, reported by Takahashi [9]. The line represents the prediction of the present model using fitted adsorption parameters. (c) Zeta potentials vs. ph, reported by Gray-Weale and Beattie [1]. The line represents the prediction of the present model using fitted adsorption parameters. Whereas the present model cannot describe accurately all the zeta potential data via only one set of parameters, let us try to apply the model to each individual data set. Firstly, the adsorption parameters have been adjusted to fit the zeta potential data reported by Li and Somasundaran [8] (see Table 1, set 2), which cover a large range of ph values for three salt concentrations. In Fig. 5a, the zeta potentials calculated based on these parameter values (continuous lines) are compared with the experimental data. The fitting procedure has been repeated for the data of Refs. [9] (set 3) and [1] (set 4). The zeta potentials calculated with these new sets of parameters (sets 3 and 4 in Table 1) are presented in Fig. 5b and c, respectively.

8 34 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) for adsorption. At very low salt concentrations, the surface charge is generated mostly by the adsorption of OH, and for neat water, the interfacial OH concentration is about 4 orders of magnitude larger than in bulk. This corresponds to a free energy of OH at the interface by about ln(1 4 )kt = 1 kt lower than in bulk. However, at large ph values, the interfacial concentration of OH becomes lower than in the bulk. This surface charge saturation cannot be attributed to interactions between ions at large ph values; the volume excluded effects are very small, because the surface density of OH is less than 1 ion/1 3 Å 2 while the electrostatic energy is also much lower than 1 kt ( S being less than 4 kt, and smaller at large ph values). Therefore, the saturation of the OH adsorption is likely to be caused by the existence of a limited number of adsorption sites on the interface, perhaps individual water molecules. The large zeta potential for neat water predicted by the set 2 and 4 of adsorption parameters, as well as the large contribution of Cl to the surface charge (particularly for set 2), lead to large negative depletion lengths for all electrolyte concentrations, hence to a monotonous decrease of surface tension upon addition of salt (Fig. 6a), in contrast to experiment. It should be emphasized again that set 4 of the parameters has been derived from zeta potential data for an oil/water interface, while the Jones Ray experiment has been performed for the air/water interface. However, using set 3 of parameters (based on the data of Ref. [9]), good agreement has been obtained with the Jones Ray data (Fig. 6b). In this case, the compatibility between the zeta potential and surface tension experiments is a result of low surface potentials and small ϕ values (the fraction of the surface charge generated by the adsorption of Cl ), plotted in Fig. 7a and b, respectively. 5. Conclusions Fig. 7. (a) The zeta potentials vs. salt concentration for the three set of parameters, determined by fitting the experimental data from Refs. [8 1], respectively. (b) The ratio ϕ between the surface charge generated by the adsorption of Cl ions and the total surface charge, vs. salt concentration for the three set of parameters, determined by fitting the experimental data from Refs. [8 1], respectively. The most significant differences in the adsorption parameters resulted from the fitting of the different data sets are the number of adsorption sited for Cl, which is of the same order of magnitude as the number of adsorption sited for H + and OH for set 2, about 1 times smaller for set 4, and about 1 times smaller for set 3. The number of adsorption sites for H + and OH resulted from the fit is in all cases of the order of 1 16 /m 2, which is much smaller than the number of water molecules that would be present at the interface, when assuming that the interfacial water has the same density as the bulk water. Indeed, in bulk each water molecule occupies about 3 Å 3, hence in the first 4 Å from the interface there should be about water molecules/m 2. The number of adsorption sites determined via fitting is by more than two orders of magnitude smaller, and the number of adsorption sited for Cl is even smaller. The much lower density of adsorption sites in the interfacial water than in an ice-like interfacial layer points toward the existence of a region of low density of water near a hydrophobic interface, as demonstrated theoretically earlier [1,2]. The reason for the suggestion that the adsorption sites are interfacial water molecules is the observation that the surface charge does not change much when the ph increases by adding NaOH (see Fig. 1b), which indicate the existence of a saturation mechanism A simple SM/SB Poisson Boltzmann model has been employed to describe simultaneously the experimental data for both the zeta potential and the surface tension. The interface was defined as the volume between and = 4 Å; for distances larger than, a modified Poisson Boltzmann distribution (that accounts also for screened image forces) has been assumed. In addition, all structure breaking ions have been considered to obey Langmuir adsorption isotherms. The H + and OH ions are assumed to compete for the same kind of interfacial sites, provided by the water molecules of the interface. However, the Cl ions have a different kind of adsorption sites on the interface. The values of the adsorption constants have been obtained by fitting the zeta potential data reported by different groups [8 1]. The behavior of the surface tension upon addition of salt can be explained qualitatively very well in terms of positive surface excesses due to the formation of a double layer, and negative surface excesses due to the screened image forces and ion hydration forces (the latter preventing the SM ions to approach the interface). The double layer dominates at very low concentrations, where the depletion length is large in absolute value and negative, hence the surface tension decreases upon addition of salt. At large salt concentrations, both the double layer and the image forces are strongly screened, and the ion hydration forces generate a depletion length independent of electrolyte concentration, leading to a surface tension which increases linearly upon addition of salt. Being screened less than the double layer forces, the image forces dominate at intermediate salt concentrations, where they provide the largest depletion length. In this range, the surface tension as a function of salt concentration has the highest slope. The quantitative description of the behavior of the surface tension upon addition of salt constitutes a more difficult task. For large surface potentials, the accumulation of ions in the double layer is so large, that it cannot be compensated at 1 mm by screened image

9 M. Manciu, E. Ruckenstein / Colloids and Surfaces A: Physicochem. Eng. Aspects 4 (212) forces or ion hydration repulsions. The surface excesses of salt ions are even higher when Cl is strongly adsorbed on the interface. In this case, the surface tension decreases upon addition of salt up to very high concentrations, in contradiction with well accepted experimental data [4]. Whereas our model cannot fit well all the zeta potential data available using only one set of parameters, the data reported by each of the groups can be well described by a single set of parameters. The number of sites for the adsorption of H + and OH (considered as the number of water molecules on the interface) is about the same in all cases and by more than two orders of magnitude smaller than the number of water molecules available in an ice-like interfacial layer. This suggest that the density of water near the hydrophobic air/water interface is much smaller than in the bulk water, as already predicted theoretically [1,2]. The number of sites for Cl adsorption obtained by fitting differs strongly in each of the cases; in the cases when the interfacial adsorption of Cl is large, the surface tension decreases upon addition of salt up to large electrolyte concentrations (as expected from Eq. (16)) but in contrast to experiment. However, the fit of the zeta potential data of Ref. [9] leads to a set of parameters that correspond to small zeta potentials and low Cl adsorption, which are compatible with the surface tension data. References [1] J.H. Stillinger, Structure in aqueous solutions of nonpolar solutes from the standpoint of scaled-particle theory, J. Solution Chem. 2 (1973) [2] E. Ruckenstein, Y.S. Djikaev, Effect of hydrogen bonding between water molecules on their density distribution near a hydrophobic surface, Phys. Chem. Lett. 2 (211) [3] I. Langmuir, The role of attractive and repulsive forces in the formation of tactoids, thixotropic gels, protein crystals and coacervates, J. Chem. Phys. 6 (1938) [4] P.K. Weissenborn, R.J. Pugh, Surface tension of aqueous solution of electrolytes: relationship with ion hydration, oxygen solubility, and bubble coalescence, J. Colloid Interface Sci. 184 (1996) [5] G. Jones, W.A. Ray, The surface tension of solutions of electrolytes as a function of the concentration. I. A differential method for measuring relative surface tension, J. Am. Chem. Soc. 59 (1937) [6] C. Wagner, Phys. Zeit. 25 (1924) 474. [7] L. Onsager, N.N.T. Samaras, The surface tension of Debye Hückel electrolytes, J. Chem. Phys. 2 (1934) [8] C. Li, P. Somasundaran, Reversal of bubble charge in multivalent inorganic salt solutions effect of magnesium, J. Colloid Interface Sci. 146 (1991) [9] M. Takahashi, Zeta potential of microbubbles in aqueous solutions: electrical properties of the gas water interface, J. Phys. Chem. B 19 (25) [1] A. Gray-Weale, J.K. Beattie, An explanation for the charge on water s surface, Phys. Chem. Chem. Phys. 11 (29) [11] L. Perera, M.L. Berkowitz, Many-body effects in molecular dynamics simulations of Na + (H 2O) n and Cl (H 2O) n clusters, J. Chem. Phys. 95 (1991) [12] P. Jungwirth, D.J. Tobias, Ions at the air/water interface, J. Phys. Chem. B 16 (22) [13] M. Mucha, T. Frigato, L.M. Levering, H.C. Allen, D.J. Tobias, L.X. Dang, P. Jungwirth, Unified molecular picture of the surfaces of aqueous, acid, base, and salt solutions, J. Phys. Chem. 19 (25) [14] V. Buch, A. Millet, R. Vacha, P. Jungwirth, J.P. Devlin, Water surface is acidic, Proc. Natl. Acad. Sci. 14 (27) [15] J.K. Beattie, A.M. Djerddjiev, G.G. Warr, The surface of neat water is basic, Faraday Disc. 141 (29) [16] P.B. Peterson, R.J. Saykally, Is the liquid water surface basic or acidic? Macroscopic vs. molecular-scale investigations, Chem. Phys. Lett. 458 (28) [17] T.L. Tarbuck, S.T. Ota, G.L. Richmond, Spectroscopic studies of solvated hydrogen and hydroxide ions at aqueous surfaces, J. Am. Chem. Soc. 128 (26) [18] C. Tian, N. Ji, G.A. Waychunas, Y.R. Shen, Interfacial structures of acidic and basic solutions, J. Am. Chem. Soc. 13 (28) [19] B. Winter, M. Faubel, R. Vacha, P. Jungwirth, Behaviour of hydroxide at the water/vapor interface, Chem. Phys. Lett. 474 (29) [2] C.J. Mundy, I.-F.W. Kuo, M.E. Tuckerman, H.S. Lee, D.J. Tobias, Hydroxide anion at the air water interface, Chem. Phys. Lett. 481 (29) 2 8. [21] M. Manciu, E. Ruckenstein, Specific ion effects via ion hydration: I. Surface tension, Adv. Colloid Interface Sci. 15 (23) [22] P.B. Peterson, R.J. Saykally, Adsorption of ions to the surface of dilute electrolyte solutions: the Jones Ray effect revisited, J. Am. Chem. Soc. 127 (25)

The effect of surface dipoles and of the field generated by a polarization gradient on the repulsive force

The effect of surface dipoles and of the field generated by a polarization gradient on the repulsive force Journal of Colloid and Interface Science 263 (2003) 156 161 www.elsevier.com/locate/jcis The effect of surface dipoles and of the field generated by a polarization gradient on the repulsive force Haohao

More information

On the Chemical Free Energy of the Electrical Double Layer

On the Chemical Free Energy of the Electrical Double Layer 1114 Langmuir 23, 19, 1114-112 On the Chemical Free Energy of the Electrical Double Layer Marian Manciu and Eli Ruckenstein* Department of Chemical Engineering, State University of New York at Buffalo,

More information

Where do ions solvate?

Where do ions solvate? PRAMANA c Indian Academy of Sciences Vol. 64, No. 6 journal of June 25 physics pp. 957 961 YAN LEVIN Instituto de Física, Universidade Federal do Rio Grande do Sul, Caixa Postal 1551, CEP 9151-97, Porto

More information

The change in free energy on transferring an ion from a medium of low dielectric constantε1 to one of high dielectric constant ε2:

The change in free energy on transferring an ion from a medium of low dielectric constantε1 to one of high dielectric constant ε2: The Born Energy of an Ion The free energy density of an electric field E arising from a charge is ½(ε 0 ε E 2 ) per unit volume Integrating the energy density of an ion over all of space = Born energy:

More information

Module 8: "Stability of Colloids" Lecture 38: "" The Lecture Contains: Calculation for CCC (n c )

Module 8: Stability of Colloids Lecture 38:  The Lecture Contains: Calculation for CCC (n c ) The Lecture Contains: Calculation for CCC (n c ) Relation between surface charge and electrostatic potential Extensions to DLVO theory file:///e /courses/colloid_interface_science/lecture38/38_1.htm[6/16/2012

More information

Lecture 3 Charged interfaces

Lecture 3 Charged interfaces Lecture 3 Charged interfaces rigin of Surface Charge Immersion of some materials in an electrolyte solution. Two mechanisms can operate. (1) Dissociation of surface sites. H H H H H M M M +H () Adsorption

More information

957 Lecture #13 of 18

957 Lecture #13 of 18 Lecture #13 of 18 957 958 Q: What was in this set of lectures? A: B&F Chapter 2 main concepts: Section 2.1 : Section 2.3: Salt; Activity; Underpotential deposition Transference numbers; Liquid junction

More information

Water Structuring and Hydroxide Ion Binding at the Interface. between Water and Hydrophobic Walls of Varying Rigidity and. van der Waals Interactions

Water Structuring and Hydroxide Ion Binding at the Interface. between Water and Hydrophobic Walls of Varying Rigidity and. van der Waals Interactions Water Structuring and Hydroxide Ion Binding at the Interface between Water and Hydrophobic Walls of Varying Rigidity and van der Waals Interactions Robert Vácha, 1 Ronen Zangi, 2 Jan B. F. N. Engberts,

More information

16 years ago TODAY (9/11) at 8:46, the first tower was hit at 9:03, the second tower was hit. Lecture 2 (9/11/17)

16 years ago TODAY (9/11) at 8:46, the first tower was hit at 9:03, the second tower was hit. Lecture 2 (9/11/17) 16 years ago TODAY (9/11) at 8:46, the first tower was hit at 9:03, the second tower was hit By Anthony Quintano - https://www.flickr.com/photos/quintanomedia/15071865580, CC BY 2.0, https://commons.wikimedia.org/w/index.php?curid=38538291

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

Ali Eftekhari-Bafrooei and Eric Borguet. Department of Chemistry, Temple University, Philadelphia PA Supporting information

Ali Eftekhari-Bafrooei and Eric Borguet. Department of Chemistry, Temple University, Philadelphia PA Supporting information The Effect of Electric Fields on the Ultrafast Vibrational Relaxation of Water at a Charged Solid-Liquid Interface as Probed by Vibrational Sum Frequency Generation Ali Eftekhari-Bafrooei and Eric Borguet

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

Soft Matter - Theoretical and Industrial Challenges Celebrating the Pioneering Work of Sir Sam Edwards

Soft Matter - Theoretical and Industrial Challenges Celebrating the Pioneering Work of Sir Sam Edwards Soft Matter - Theoretical and Industrial Challenges Celebrating the Pioneering Work of Sir Sam Edwards One Hundred Years of Electrified Interfaces: The Poisson-Boltzmann theory and some recent developments

More information

Contents. 2. Fluids. 1. Introduction

Contents. 2. Fluids. 1. Introduction Contents 1. Introduction 2. Fluids 3. Physics of Microfluidic Systems 4. Microfabrication Technologies 5. Flow Control 6. Micropumps 7. Sensors 8. Ink-Jet Technology 9. Liquid Handling 10.Microarrays 11.Microreactors

More information

2 Structure. 2.1 Coulomb interactions

2 Structure. 2.1 Coulomb interactions 2 Structure 2.1 Coulomb interactions While the information needed for reproduction of living systems is chiefly maintained in the sequence of macromolecules, any practical use of this information must

More information

Surface tensions and surface potentials of acid solutions

Surface tensions and surface potentials of acid solutions Surface tensions and surface potentials of acid solutions Alexandre P. dos Santos 1 and Yan Levin 1 Instituto de Física, Universidade Federal do Rio Grande do Sul, Caixa Postal 15051, CEP 91501-970, Porto

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

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

Properties of Aqueous Solutions

Properties of Aqueous Solutions Properties of Aqueous Solutions Definitions A solution is a homogeneous mixture of two or more substances. The substance present in smaller amount is called the solute. The substance present in larger

More information

Surface interactions part 1: Van der Waals Forces

Surface interactions part 1: Van der Waals Forces CHEM-E150 Interfacial Phenomena in Biobased Systems Surface interactions part 1: Van der Waals Forces Monika Österberg Spring 018 Content Colloidal stability van der Waals Forces Surface Forces and their

More information

1044 Lecture #14 of 18

1044 Lecture #14 of 18 Lecture #14 of 18 1044 1045 Q: What s in this set of lectures? A: B&F Chapter 13 main concepts: Section 1.2.3: Diffuse double layer structure Sections 13.1 & 13.2: Gibbs adsorption isotherm; Electrocapillary

More information

Chapter 12 Intermolecular Forces and Liquids

Chapter 12 Intermolecular Forces and Liquids Chapter 12 Intermolecular Forces and Liquids Jeffrey Mack California State University, Sacramento Why? Why is water usually a liquid and not a gas? Why does liquid water boil at such a high temperature

More information

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai Surfaces and Interfaces Defining of interfacial region Types of interfaces: surface vs interface Surface

More information

An Overview of the Concept, Measurement, Use and Application of Zeta Potential. David Fairhurst, Ph.D. Colloid Consultants, Ltd

An Overview of the Concept, Measurement, Use and Application of Zeta Potential. David Fairhurst, Ph.D. Colloid Consultants, Ltd An Overview of the Concept, Measurement, Use and Application of Zeta Potential David Fairhurst, Ph.D. Colloid Consultants, Ltd Fundamental Parameters that control the Nature and Behavior of all Particulate

More information

Electrical double layer

Electrical double layer Electrical double layer Márta Berka és István Bányai, University of Debrecen Dept of Colloid and Environmental Chemistry http://dragon.unideb.hu/~kolloid/ 7. lecture Adsorption of strong electrolytes from

More information

Chapter 6 Chemistry of Water; Chemistry in Water

Chapter 6 Chemistry of Water; Chemistry in Water Chapter 6 Chemistry of Water; Chemistry in Water Water is one of the most remarkable and important of all chemical species. We, and all living things, are mostly water about 80% of our brain; 65% of our

More information

9.1 Water. Chapter 9 Solutions. Water. Water in Foods

9.1 Water. Chapter 9 Solutions. Water. Water in Foods Chapter 9 s 9.1 Water 9.1 Properties of Water 9.2 s 9.3 Electrolytes and Nonelectrolytes 9.6 Percent Concentration 9.7 Molarity Water is the most common solvent. The water molecule is polar. Hydrogen bonds

More information

Lecture Presentation. Chapter 11. Liquids and Intermolecular Forces. John D. Bookstaver St. Charles Community College Cottleville, MO

Lecture Presentation. Chapter 11. Liquids and Intermolecular Forces. John D. Bookstaver St. Charles Community College Cottleville, MO Lecture Presentation Chapter 11 Liquids and Intermolecular Forces John D. Bookstaver St. Charles Community College Cottleville, MO Properties of Gases, Liquids, and Solids State Volume Shape of State Density

More information

Chapter-2 (Page 22-37) Physical and Chemical Properties of Water

Chapter-2 (Page 22-37) Physical and Chemical Properties of Water Chapter-2 (Page 22-37) Physical and Chemical Properties of Water Introduction About 70% of the mass of the human body is water. Water is central to biochemistry for the following reasons: 1- Biological

More information

Lecture 5: Electrostatic Interactions & Screening

Lecture 5: Electrostatic Interactions & Screening Lecture 5: Electrostatic Interactions & Screening Lecturer: Prof. Brigita Urbanc (brigita@drexel.edu) PHYS 461 & 561, Fall 2009-2010 1 A charged particle (q=+1) in water, at the interface between water

More information

Polarization of Water near Dipolar Surfaces: A Simple Model for Anomalous Dielectric Behavior

Polarization of Water near Dipolar Surfaces: A Simple Model for Anomalous Dielectric Behavior Langmuir 2005, 2, 749-756 749 Polarization of Water near Dipolar Surfaces: A Simple Model for Anomalous Dielectric Behavior Marian Manciu* Department of Physics, University of Texas at El Paso, El Paso,

More information

Multimedia : Boundary Lubrication Podcast, Briscoe, et al. Nature , ( )

Multimedia : Boundary Lubrication Podcast, Briscoe, et al. Nature , ( ) 3.05 Nanomechanics of Materials and Biomaterials Thursday 04/05/07 Prof. C. Ortiz, MITDMSE I LECTURE 14: TE ELECTRICAL DOUBLE LAYER (EDL) Outline : REVIEW LECTURE #11 : INTRODUCTION TO TE ELECTRICAL DOUBLE

More information

Monolayers. Factors affecting the adsorption from solution. Adsorption of amphiphilic molecules on solid support

Monolayers. Factors affecting the adsorption from solution. Adsorption of amphiphilic molecules on solid support Monolayers Adsorption as process Adsorption of gases on solids Adsorption of solutions on solids Factors affecting the adsorption from solution Adsorption of amphiphilic molecules on solid support Adsorption

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

Solids, Liquids and Gases We have already covered these phases of matter. See online section 5.2

Solids, Liquids and Gases We have already covered these phases of matter. See online section 5.2 Chapter 10 This chapter begins to answer the questions: So now that I now what atoms and molecules look like, how do these structures translate into what I see in the world around me. Reading Assignment:

More information

Atomic and molecular interaction forces in biology

Atomic and molecular interaction forces in biology Atomic and molecular interaction forces in biology 1 Outline Types of interactions relevant to biology Van der Waals interactions H-bond interactions Some properties of water Hydrophobic effect 2 Types

More information

Rama Abbady. Zina Smadi. Diala Abu-Hassan

Rama Abbady. Zina Smadi. Diala Abu-Hassan 1 Rama Abbady Zina Smadi Diala Abu-Hassan (00:00) (10:00) Types of Molecules in the Cell 1. Water Molecules: a large portion of the cell mass is water (70% of total cell mass). 2. Organic molecules (carbon

More information

Solutions. Experiment 11. Various Types of Solutions. Solution: A homogenous mixture consisting of ions or molecules

Solutions. Experiment 11. Various Types of Solutions. Solution: A homogenous mixture consisting of ions or molecules Solutions Solution: A homogenous mixture consisting of ions or molecules -Assignment: Ch 15 Questions & Problems : 5, (15b,d), (17a, c), 19, 21, 23, 27, (33b,c), 39, (43c,d),45b, 47, (49b,d), (55a,b),

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

Intermolecular and Intramolecular Forces. Introduction

Intermolecular and Intramolecular Forces. Introduction Intermolecular and Intramolecular Forces Introduction Atoms can form stable units called molecules by sharing electrons. The formation of molecules is the result of intramolecular bonding (within the molecule)

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

*blood and bones contain colloids. *milk is a good example of a colloidal dispersion.

*blood and bones contain colloids. *milk is a good example of a colloidal dispersion. Chap. 3. Colloids 3.1. Introduction - Simple definition of a colloid: a macroscopically heterogeneous system where one component has dimensions in between molecules and macroscopic particles like sand

More information

Chemical Equilibrium

Chemical Equilibrium Chemical Equilibrium Many reactions are reversible, i.e. they can occur in either direction. A + B AB or AB A + B The point reached in a reversible reaction where the rate of the forward reaction (product

More information

Peter Debye and Electrochemistry

Peter Debye and Electrochemistry Peter Debye and Electrochemistry A K Shukla and T Prem Kumar A K Shukla is Professor and Chairman, Solid State and Structural Chemistry Unit, Indian Institute of Science, Bangalore. His research interests

More information

Electrostatic Double Layer Force: Part III

Electrostatic Double Layer Force: Part III NPTEL Chemical Engineering Interfacial Engineering Module 3: Lecture 4 Electrostatic Double Layer Force: Part III Dr. Pallab Ghosh Associate Professor Department of Chemical Engineering IIT Guwahati, Guwahati

More information

A solution is a homogeneous mixture of two or more substances.

A solution is a homogeneous mixture of two or more substances. UNIT (5) SOLUTIONS A solution is a homogeneous mixture of two or more substances. 5.1 Terminology Solute and Solvent A simple solution has two components, a solute, and a solvent. The substance in smaller

More information

Test bank for Chemistry An Introduction to General Organic and Biological Chemistry 12th Edition by Timberlake

Test bank for Chemistry An Introduction to General Organic and Biological Chemistry 12th Edition by Timberlake Test bank for Chemistry An Introduction to General Organic and Biological Chemistry 12th Edition by Timberlake Link download full: http://testbankair.com/download/test-bank-for-chemistry-an-introduction-to-general-organic-and-biological-chemistry-12th-edition-by-timberlak

More information

Chapter 7. Pickering Stabilisation ABSTRACT

Chapter 7. Pickering Stabilisation ABSTRACT Chapter 7 Pickering Stabilisation ABSTRACT In this chapter we investigate the interfacial properties of Pickering emulsions. Based upon findings that indicate these emulsions to be thermodynamically stable,

More information

Thermodynamically Stable Emulsions Using Janus Dumbbells as Colloid Surfactants

Thermodynamically Stable Emulsions Using Janus Dumbbells as Colloid Surfactants Thermodynamically Stable Emulsions Using Janus Dumbbells as Colloid Surfactants Fuquan Tu, Bum Jun Park and Daeyeon Lee*. Description of the term notionally swollen droplets When particles are adsorbed

More information

Discipline Course-I. Semester-II. Paper No: Electricity and Magnetism. Lesson: Polarization and Dielectric Materials

Discipline Course-I. Semester-II. Paper No: Electricity and Magnetism. Lesson: Polarization and Dielectric Materials Discipline CourseI SemesterII Paper No: Electricity and Magnetism Lesson: Polarization and Dielectric Materials Lesson Developer: Dr. Amit Choudhary College/ Department: Physics Department, Deshbandhu

More information

Stability of colloidal systems

Stability of colloidal systems Stability of colloidal systems Colloidal stability DLVO theory Electric double layer in colloidal systems Processes to induce charges at surfaces Key parameters for electric forces (ζ-potential, Debye

More information

Chapter 14 Acids and Bases

Chapter 14 Acids and Bases Properties of Acids and Bases Chapter 14 Acids and Bases Svante Arrhenius (1859-1927) First to develop a theory for acids and bases in aqueous solution Arrhenius Acids Compounds which dissolve (dissociate)

More information

100 C = 100 X = X = 218 g will fit in this solution. 25 C = 100 X = 3640 X = 36.4 g will fit in this solution.

100 C = 100 X = X = 218 g will fit in this solution. 25 C = 100 X = 3640 X = 36.4 g will fit in this solution. 58 Questions for Solutions - You should be able to do ALL of these problems. Use a calculator, write all formulas, watch SF, and find the answers online at Arbuiso.com on the SOLUTIONS page. This is great

More information

Electrostatic Forces & The Electrical Double Layer

Electrostatic Forces & The Electrical Double Layer Electrostatic Forces & The Electrical Double Layer Dry Clay Swollen Clay Repulsive electrostatics control swelling of clays in water LiquidSolid Interface; Colloids Separation techniques such as : column

More information

Atoms can form stable units called molecules by sharing electrons.

Atoms can form stable units called molecules by sharing electrons. Atoms can form stable units called molecules by sharing electrons. The formation of molecules is the result of intramolecular bonding (within the molecule) e.g. ionic, covalent. Forces that cause the aggregation

More information

Chapter 2 Basic Chemistry Outline

Chapter 2 Basic Chemistry Outline Chapter 2 Basic Chemistry Outline 1.0 COMPOSITION OF MATTER 1.1 Atom 1.2 Elements 1.21 Isotopes 1.22 Radioisotopes 1.3 Compounds 1.31 Compounds Formed by Ionic Bonding 1.32 Compounds Formed by Covalent

More information

Water and solutions. Prof. Ramune Morkuniene, Biochemistry Dept., LUHS

Water and solutions. Prof. Ramune Morkuniene, Biochemistry Dept., LUHS Water and solutions Prof. Ramune Morkuniene, Biochemistry Dept., LUHS Characteristics of water molecule Hydrophylic, hydrophobic and amphipatic compounds Types of real solutions Electrolytes and non- electrolytes

More information

Chapter 16. Acid-Base Equilibria

Chapter 16. Acid-Base Equilibria Chapter 16. Acid-Base Equilibria 16.1 Acids and Bases: A Brief Review Acids taste sour and cause certain dyes to change color. Bases taste bitter and feel soapy. Arrhenius concept of acids and bases: An

More information

Chapter 02 The Chemical Basis of Life I: Atoms, Molecules, and Water

Chapter 02 The Chemical Basis of Life I: Atoms, Molecules, and Water Chapter 02 The Chemical Basis of Life I: Atoms, Molecules, and Water Multiple Choice Questions 1. The atomic number of an atom is A. the number of protons in the atom. B. the number of neutrons in the

More information

768 Lecture #11 of 18

768 Lecture #11 of 18 Lecture #11 of 18 768 769 Q: What s in this set of lectures? A: B&F Chapter 2 main concepts: Section 2.1 : Section 2.3: Salt; Activity; Underpotential deposition Transference numbers; Liquid junction potentials

More information

Module 4: "Surface Thermodynamics" Lecture 22: "" The Lecture Contains: Examples on Effect of surfactant on interfacial tension. Objectives_template

Module 4: Surface Thermodynamics Lecture 22:  The Lecture Contains: Examples on Effect of surfactant on interfacial tension. Objectives_template The Lecture Contains: Examples on Effect of surfactant on interfacial tension file:///e /courses/colloid_interface_science/lecture22/22_1.htm[6/16/2012 1:10:07 PM] Example Consider liquid, its vapors and

More information

Chapter 11. Intermolecular Forces and Liquids & Solids

Chapter 11. Intermolecular Forces and Liquids & Solids Chapter 11 Intermolecular Forces and Liquids & Solids The Kinetic Molecular Theory of Liquids & Solids Gases vs. Liquids & Solids difference is distance between molecules Liquids Molecules close together;

More information

Kinetics of surfactant adsorption: the free energy approach

Kinetics of surfactant adsorption: the free energy approach Colloids and Surfaces A: Physicochemical and Engineering Aspects 183 185 (2001) 259 276 www.elsevier.nl/locate/colsurfa Kinetics of surfactant adsorption: the free energy approach Haim Diamant 1, Gil Ariel,

More information

The Effect of Electrostatic Surface Charges on Photoresist Dissolution

The Effect of Electrostatic Surface Charges on Photoresist Dissolution Sean Burns 4/24/00 Dr. Bonnecaze ChE 385M The Effect of Electrostatic Surface Charges on Photoresist Dissolution Introduction/Motivation It would be very useful and economical to have a fundamental model

More information

Brass, a solid solution of Zn and Cu, is used to make musical instruments and many other objects.

Brass, a solid solution of Zn and Cu, is used to make musical instruments and many other objects. Brass, a solid solution of Zn and Cu, is used to make musical instruments and many other objects. 14.1 General Properties of Solutions 14.2 Solubility 14.3 Rate of Dissolving Solids 14.4 Concentration

More information

2. WATER : THE SOLVENT FOR BIOCHEMICAL REACTIONS

2. WATER : THE SOLVENT FOR BIOCHEMICAL REACTIONS 2. WATER : THE SOLVENT FOR BIOCHEMICAL REACTIONS 2.1 Water and Polarity Both geometry and properties of molecule determine polarity Electronegativity - The tendency of an atom to attract electrons to itself

More information

6, Physical Chemistry -II (Statistical Thermodynamics, Chemical Dynamics, Electrochemistry and Macromolecules)

6, Physical Chemistry -II (Statistical Thermodynamics, Chemical Dynamics, Electrochemistry and Macromolecules) Subject Paper No and Title Module No and Title Module Tag 6, Physical -II (Statistical Thermodynamics, Chemical Dynamics, Electrochemistry and Macromolecules) 25, Activity and Mean Activity coefficient

More information

CH 4 AP. Reactions in Aqueous Solutions

CH 4 AP. Reactions in Aqueous Solutions CH 4 AP Reactions in Aqueous Solutions Water Aqueous means dissolved in H 2 O Moderates the Earth s temperature because of high specific heat H-bonds cause strong cohesive and adhesive properties Polar,

More information

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai Zoltán Nagy Department of Physical Chemistry

Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai Zoltán Nagy Department of Physical Chemistry Surface chemistry. Liquid-gas, solid-gas and solid-liquid surfaces. Levente Novák István Bányai Zoltán Nagy Department of Physical Chemistry Surfaces and Interfaces Defining of interfacial region Types

More information

Chapter 4: Types of Chemical Reactions and Solution Stoichiometry

Chapter 4: Types of Chemical Reactions and Solution Stoichiometry Chapter 4: Types of Chemical Reactions and Solution Stoichiometry 4.1 Water, the Common Solvent 4.2 The Nature of Aqueous Solutions: Strong and Weak Electrolytes 4.3 The Composition of Solutions (MOLARITY!)

More information

Water - HW. PSI Chemistry

Water - HW. PSI Chemistry Water - HW PSI Chemistry Name 1) In a single molecule of water, the two hydrogen atoms are bonded to a single oxygen atom by A) hydrogen bonds. B) nonpolar covalent bonds. C) polar covalent bonds. D) ionic

More information

Interfacial Tension of Electrolyte Solutions

Interfacial Tension of Electrolyte Solutions arxiv:cond-mat/992v1 [cond-mat.soft] 6 Sep 2 Interfacial Tension of Electrolyte Solutions Yan Levin Instituto de Física, Universidade Federal do Rio Grande do Sul Caixa Postal 1551, CEP 9151-97, Porto

More information

Lecture 21 Cations, Anions and Hydrolysis in Water:

Lecture 21 Cations, Anions and Hydrolysis in Water: 2P32 Principles of Inorganic Chemistry Dr. M. Pilkington Lecture 21 Cations, Anions and ydrolysis in Water: 1. ydration.energy 2. ydrolysis of metal cations 3. Categories of acidity and observable behavior

More information

Name Chemistry Pre-AP. Notes: Solutions

Name Chemistry Pre-AP. Notes: Solutions Name Chemistry Pre-AP Notes: Solutions Period I. Intermolecular Forces (IMFs) A. Attractions Between Molecules Attractions between molecules are called and are very important in determining the properties

More information

Properties of Solutions

Properties of Solutions Properties of Solutions The States of Matter The state a substance is in at a particular temperature and pressure depends on two antagonistic entities: The kinetic energy of the particles The strength

More information

A More Detailed Look at Chemical Equilibria

A More Detailed Look at Chemical Equilibria A More Detailed Look at Chemical Equilibria effect of ionic strength on solubility calculation of ionic strength activity coefficients effect of ionic strength and size/charge of ions on activity coefficients

More information

Lecture 2 and 3: Review of forces (ctd.) and elementary statistical mechanics. Contributions to protein stability

Lecture 2 and 3: Review of forces (ctd.) and elementary statistical mechanics. Contributions to protein stability Lecture 2 and 3: Review of forces (ctd.) and elementary statistical mechanics. Contributions to protein stability Part I. Review of forces Covalent bonds Non-covalent Interactions: Van der Waals Interactions

More information

Intermolecular forces Liquids and Solids

Intermolecular forces Liquids and Solids Intermolecular forces Liquids and Solids Chapter objectives Understand the three intermolecular forces in pure liquid in relation to molecular structure/polarity Understand the physical properties of liquids

More information

Adsorption of Ions to the Surface of Dilute Electrolyte Solutions: The Jones-Ray Effect Revisited

Adsorption of Ions to the Surface of Dilute Electrolyte Solutions: The Jones-Ray Effect Revisited Published on Web 10/18/2005 Adsorption of Ions to the Surface of Dilute Electrolyte Solutions: The Jones-Ray Effect Revisited Poul B. Petersen and Richard J. Saykally* Contribution from the Department

More information

Surface tension for aqueous electrolyte solutions by the modified mean spherical approximation

Surface tension for aqueous electrolyte solutions by the modified mean spherical approximation Fluid Phase Equilibria 173 (2000) 23 38 Surface tension for aqueous electrolyte solutions by the modified mean spherical approximation Yang-Xin Yu, Guang-Hua Gao, Yi-Gui Li Department of Chemical Engineering,

More information

SBI4U BIOCHEMISTRY. Atoms, Bonding & Molecular Polarity

SBI4U BIOCHEMISTRY. Atoms, Bonding & Molecular Polarity SBI4U BIOCHEMISTRY Atoms, Bonding & Molecular Polarity 6 types of atoms make up 99% of all living organisms Naturally Occurring Elements in the Human Body Element Symbol Atomic # % of human body weight

More information

Chem 321 Lecture 11 - Chemical Activities 10/3/13

Chem 321 Lecture 11 - Chemical Activities 10/3/13 Student Learning Objectives Chem 321 Lecture 11 - Chemical Activities 10/3/13 One of the assumptions that has been made in equilibrium calculations thus far has been to equate K to a ratio of concentrations.

More information

States of matter Part 1

States of matter Part 1 Physical pharmacy I 1. States of matter (2 Lectures) 2. Thermodynamics (2 Lectures) 3. Solution of non-electrolyte 4. Solution of electrolyte 5. Ionic equilibria 6. Buffered and isotonic solution Physical

More information

A.% by mass (like % composition)

A.% by mass (like % composition) Solutions; Colloids Key Words Solute Solvent Solubility effervescence Miscible saturated Supersaturated (metastable system)- a cooled solution contains more solute than it would at equilibrium, desolvation=

More information

CE 530 Molecular Simulation

CE 530 Molecular Simulation 1 CE 530 Molecular Simulation Lecture 14 Molecular Models David A. Kofke Department of Chemical Engineering SUNY Buffalo kofke@eng.buffalo.edu 2 Review Monte Carlo ensemble averaging, no dynamics easy

More information

Ch 9 Liquids & Solids (IMF) Masterson & Hurley

Ch 9 Liquids & Solids (IMF) Masterson & Hurley Ch 9 Liquids & Solids (IMF) Masterson & Hurley Intra- and Intermolecular AP Questions: 2005 Q. 7, 2005 (Form B) Q. 8, 2006 Q. 6, 2007 Q. 2 (d) and (c), Periodic Trends AP Questions: 2001 Q. 8, 2002 Q.

More information

States of matter Part 1. Lecture 1. University of Kerbala. Hamid Alghurabi Assistant Lecturer in Pharmaceutics. Physical Pharmacy

States of matter Part 1. Lecture 1. University of Kerbala. Hamid Alghurabi Assistant Lecturer in Pharmaceutics. Physical Pharmacy Physical pharmacy I 1. States of matter (2 Lectures) 2. Thermodynamics (2 Lectures) 3. Solution of non-electrolyte 4. Solution of electrolyte 5. Ionic equilibria 6. Buffered and isotonic solution Physical

More information

C deposits (63.5/2) g of copper; the quantity passed is therefore

C deposits (63.5/2) g of copper; the quantity passed is therefore 7. SOLUTIONS OF ELECTROLYTES n Faraday s Laws, Molar Conductivity, and Weak Electrolytes 7.1. 96 500 C deposits (63.5/2) g of copper; the quantity passed is therefore 96 500 0.04 2 63.5 C The current was

More information

Initial position, x p (0)/L

Initial position, x p (0)/L .4 ) xp().2 ) ( 2L 2 xp Dc ( Displacement, /L.2.4.5.5 Initial position, x p ()/L Supplementary Figure Computed displacements of (red) positively- and (blue) negatively-charged particles at several CO 2

More information

Modeling Viscosity of Multicomponent Electrolyte Solutions 1

Modeling Viscosity of Multicomponent Electrolyte Solutions 1 International Journal of Thermophysics, Vol. 19, No. 2, 1998 Modeling Viscosity of Multicomponent Electrolyte Solutions 1 M. M. Lencka, 2 A. Anderko, 2,3 S. J. Sanders, 2 and R. D. Young 2 A comprehensive

More information

Chemistry of Life: Water and Solutions

Chemistry of Life: Water and Solutions Chemistry of Life: Water and Solutions Unit Objective I can describe the role of organic and inorganic chemicals important to living things. During this unit, we will answer the following very important

More information

Module17: Intermolecular Force between Surfaces and Particles. Lecture 23: Intermolecular Force between Surfaces and Particles

Module17: Intermolecular Force between Surfaces and Particles. Lecture 23: Intermolecular Force between Surfaces and Particles Module17: Intermolecular Force between Surfaces and Particles Lecture 23: Intermolecular Force between Surfaces and Particles 1 We now try to understand the nature of spontaneous instability in a confined

More information

ADVANCED PLACEMENT CHEMISTRY ACIDS, BASES, AND AQUEOUS EQUILIBRIA

ADVANCED PLACEMENT CHEMISTRY ACIDS, BASES, AND AQUEOUS EQUILIBRIA ADVANCED PLACEMENT CHEMISTRY ACIDS, BASES, AND AQUEOUS EQUILIBRIA Acids- taste sour Bases(alkali)- taste bitter and feel slippery Arrhenius concept- acids produce hydrogen ions in aqueous solution while

More information

MOLECULAR INTERACTIONS NOTES

MOLECULAR INTERACTIONS NOTES - 1 - MOLECULAR INTERACTIONS NOTES Summary of Fundamental Molecular Interactions Q1Q Ion-Ion U ( r) = 4 r πεε o µ Q cosθ U ( r) = 4πεε o r µ 1µ U ( r) 3 r µ 1 µ U ( r) 6 r µ 1 α U ( r) 6 r Ion-Dipole Dipole-Dipole

More information

Charged Interfaces & electrokinetic

Charged Interfaces & electrokinetic Lecture Note #7 Charged Interfaces & electrokinetic phenomena Reading: Shaw, ch. 7 Origin of the charge at colloidal surfaces 1. Ionization Proteins acquire their charge by ionization of COOH and NH 2

More information

Supporting Information for Lysozyme Adsorption in ph-responsive Hydrogel Thin-Films: The non-trivial Role of Acid-Base Equilibrium

Supporting Information for Lysozyme Adsorption in ph-responsive Hydrogel Thin-Films: The non-trivial Role of Acid-Base Equilibrium Electronic Supplementary Material (ESI) for Soft Matter. This journal is The Royal Society of Chemistry 215 Supporting Information for Lysozyme Adsorption in ph-responsive Hydrogel Thin-Films: The non-trivial

More information

Electrical double layer interactions between dissimilar oxide surfaces with charge regulation and Stern Grahame layers

Electrical double layer interactions between dissimilar oxide surfaces with charge regulation and Stern Grahame layers Journal of Colloid and Interface Science 296 (2006) 150 158 www.elsevier.com/locate/jcis Electrical double layer interactions between dissimilar oxide surfaces with charge regulation and Stern Grahame

More information

Colloid Chemistry. La chimica moderna e la sua comunicazione Silvia Gross.

Colloid Chemistry. La chimica moderna e la sua comunicazione Silvia Gross. Colloid Chemistry La chimica moderna e la sua comunicazione Silvia Gross Istituto Dipartimento di Scienze di e Scienze Tecnologie Chimiche Molecolari ISTM-CNR, Università Università degli Studi degli Studi

More information

Effect of Water Chemistry on Zeta Potential of Air Bubbles

Effect of Water Chemistry on Zeta Potential of Air Bubbles Int. J. Electrochem. Sci., 8 (2013) 5828-5837 International Journal of ELECTROCHEMICAL SCIENCE www.electrochemsci.org Effect of Water Chemistry on Zeta Potential of Air Bubbles Weihong Jia, Sili Ren *,

More information

Intermolecular Forces I

Intermolecular Forces I I How does the arrangement of atoms differ in the 3 phases of matter (solid, liquid, gas)? Why doesn t ice just evaporate into a gas? Why does liquid water exist at all? There must be some force between

More information