Hydrogen Bond Networks in Binary Mixtures of Water and Organic

May 2, 2019 - Bertie, J. E.; Lan, Z. Liquid Water–Acetonitrile Mixtures at 25 °C: The Hydrogen-Bonded Structure Studied through Infrared Absolute I...
0 downloads 0 Views 899KB Size
Subscriber access provided by UNIV OF LOUISIANA

B: Liquids, Chemical and Dynamical Processes in Solution, Spectroscopy in Solution

Hydrogen Bond Networks in Binary Mixtures of Water and Organic Solvent Simon Stehle, and Andreas Siegfried Braeuer J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.9b02829 • Publication Date (Web): 02 May 2019 Downloaded from http://pubs.acs.org on May 2, 2019

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

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

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

The Journal of Physical Chemistry

1

Hydrogen Bond Networks in Binary Mixtures of

2

Water and Organic Solvent

3

Simon Stehle‡,§ and Andreas Siegfried Braeuer‡,*

4



5

Institute of Thermal-, Environmental- and Resources‘ Process Engineering (ITUN), Technische Universität Bergakademie Freiberg (TUBAF), Leipziger Strasse 28, 09599 Freiberg, Germany

6

§

7

Universitaet Erlangen-Nuernberg (FAU), Paul-Gordan-Straße 6, 91052 Erlangen, Germany

8

*[email protected]

Erlangen Graduate School in Advanced Optical Technologies (SAOT), Friedrich-Alexander-

9

10

ABSTRACT. We here present an approach for the optical in situ characterization of hydrogen

11

bond networks in binary mixtures of water and organic solvent, such as methanol, ethanol and

12

acetonitrile. Hydrogen bond networks are characterized based on (i) the analysis of experimental

13

molar Raman spectra of the mixture, (ii) partial molar Raman spectra of the mixture constituents

14

and (iii) computed ideal molar Raman spectra of the mixture. Especially the consideration of the

15

partial molar Raman spectra provides insights into the development of hydrogen bonds of

16

molecules of one species to their neighbors. The obtained Raman spectra are evaluated with

17

respect to the centroid of the symmetric stretching vibration Raman signal of water and to the

18

hydroxyl stretching vibration of alcohols. We show the influence of composition and

ACS Paragon Plus Environment

1

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

Page 2 of 31

19

temperature on the development of the hydrogen bond network of the mixtures, the hydrogen

20

bond network of water and the hydrogen bond network of the organic solvent molecules.

21

1 Introduction

22

Binary mixtures of water and an organic solvent as acetonitrile, methanol or ethanol are

23

miscible in a single liquid and transparent phase over the entire composition range at moderate

24

temperature and pressure. These binary mixtures have in common that their electric conductivity,

25

their permittivity, their specific volume, enthalpy of mixing or their scattering behavior exhibit a

26

non-monotonic behavior when regarded as a function of the composition of the mixture. A

27

controversial discussion about the origin of this non-monotonic behavior has grown over the last

28

70 years and still theories of the intermolecular interactions in these mixtures are ambiguous. As

29

the “organic solvent” contains a hydrophobic part, the terms “hydrophobic solvation” or

30

“hydrophobic effects” can summarize all the subsequently mentioned theories.

31

The most debated theory is the so called “iceberg theory” from Frank and Evans who

32

discussed in 1945 entropic anomalies in several water diluted binary mixtures.1 They have

33

chosen the term iceberg because they proposed that water molecules surrounding a strong diluted

34

hydrophobic or amphiphile compound create a stronger hydrogen bond network to their water

35

neighbors – like in ice – than those in bulk water. Later, in accordance to the hydration of salt

36

ions dissolved in water, also the term “hydrophobic hydration” was common2–5 in order to

37

describe the development of “hydrate-like” or “ice-like” formations of water molecules.6 This

38

concept was confirmed2,7–11 but also challenged12–16 by experimental and theoretical

39

investigations. However, not only the formation of water molecules in these mixtures is under

40

discussion. For example, other theories about the development of

41

 an incomplete mixing on a molecular scale,13,17,18

ACS Paragon Plus Environment

2

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

The Journal of Physical Chemistry

42

 water diluted mixtures without strengthening the structure of water,19–22

43

 a

stronger

network

between

the

water

molecules

surrounding

a

hydrophobic/amphiphile molecule,4,9,11,14,23–29

44 45

 hydration shells around single or few hydrophobic/amphiphile molecules,30,31

46

 hydrophobic/amphiphile clusters at low concentration of the hydrophobic/amphiphile compound,4,14,19,22–24,30,32,33

47 48

 water clusters at low water concentration,13,22,30,34–36

49

 greater percolating networks of water and the hydrophobic/amphiphile compound around equimolar compositions,15,20,22,27,37–41

50 51

 hydrophilic and hydrophobic/amphiphile clusters over a broad composition range,42– 44

52

and

53

 a complete and homogeneous mixing over a defined composition range28,31,32

54

contribute to the current debate. This incomplete above listed abstract of the debate about

55

liquid binary water/alcohol and water/nitrile mixtures shows the dispute and the challenging task

56

to draw a complete and coherent picture from the comprehensive results stemming from

57

thermodynamic studies, molecular dynamic simulations and a variety of measurement

58

techniques.

59

Common to the systems analyzed in the above listed studies is that their molecular formation

60

and the dominant intermolecular forces are strongly influenced by the development of a

61

hydrogen bond network (HBN). The unique ability of water to form as many hydrogen (H)-

62

bonds as covalent bonds per molecule45 is supposed to be one reason for the abnormalities

63

observable in these mixtures. As the intermolecular H-bonds affect the intramolecular covalent

64

bonds of water and alcohols, spectroscopic analysis of the intramolecular covalent bond

65

vibrations is – beside thermodynamic and theoretical studies – often used to investigate into the

66

intramolecular HBN of the mixture. In this context infrared (IR) absorption is one frequently

67

applied spectroscopic technique. For example, Takamuku et al. investigated water/acetonitrile

ACS Paragon Plus Environment

3

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

Page 4 of 31

68

mixtures over the whole composition range with X-ray diffraction and IR spectroscopy.29 At

69

acetonitrile (Acn) molar fractions of 𝑥𝐴𝑐𝑛 < 0.6 the “X-ray radial distribution functions and IR

70

spectra suggested that the hydrogen bond network of water is enhanced”. Furthermore within the

71

composition range 0.2 < 𝑥𝐴𝑐𝑛 < 0.6 they observed a microheterogeneity with coexisting water

72

and acetonitrile clusters. They supposed that the clustering is an explanation for the maximum of

73

the enthalpy of mixing at 𝑥𝐴𝑐𝑛 = 0.6. Many other examples of the application of IR spectroscopy

74

can be found in the literature.20,21,46–50

75

Raman spectroscopy is an alternative and also frequently applied technique for the

76

investigation into HBN.7,11,31,38,51–53 In 2012, Lin et al. evaluated polarized Raman spectra from

77

different water/alcohol mixtures (methanol, ethanol, 1-propanol and n-propanol) over the whole

78

composition range.31 The stronger the intermolecular interactions are, the slower are the

79

translational motions that can be deducted from the Raman spectra. They observed a sharp

80

increase in the translational relaxation time during the addition of small amounts of alcohol to

81

water and interpreted this with the development of water hydration shells around the alcohol

82

molecules. In the water rich regime, they interpreted this as a confirmation for the “iceberg

83

model”.

84

In contrast to the majority of previous Raman investigations we here do not analyze the Raman

85

spectrum of the mixture as a whole, but separate the shares originating from each species in the

86

mixture. This makes possible the extraction of information about the involvement of each species

87

in the HBN in the mixture. The information about the development of the HBN is deduced from

88

the symmetric stretching vibration of water and of the OH stretching vibration of alcohols.

89

Unfortunately, both Raman signatures completely overlap in the spectral range of Raman shifts

90

between 3000 cm-1 and 3800 cm-1. Therefore, we introduce a method for the efficient separation

ACS Paragon Plus Environment

4

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

The Journal of Physical Chemistry

91

of the two overlapping Raman signatures that is based on the consideration of partial molar

92

spectra. Finally, from the analysis of the partial molar spectra of each compound in the mixture

93

information about the role of this compound in the HBN network can be gathered, while the

94

analysis of the spectrum of the mixture provides information of the overall development of the

95

HBN.

96

2 Experimental

97

2.1 Materials

98

Experiments were carried out with acetonitrile (Acn) (Merck LiChrosolv®, purity > 99.9 %,

99

Germany), methanol (MeOH) (VWR HiPerSolv®, purity > 99.8 %, USA), ethanol (EtOH)

100

(Merck Uvasol®, purity > 99.9 %, Germany) and distilled water (W) (KERNDL, conductivity

101

< 10 μS cm-1, Germany).

102

2.2 Experimental setup

103

The binary mixtures were analyzed via Raman spectroscopy in a microcapillary setup (MCS)

104

that according to Figure 1 consists of the microcapillary itself (1), three syringe pumps (2) and

105

(2-1), a microcapillary heating system (3) and a Raman system (4).

ACS Paragon Plus Environment

5

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

Page 6 of 31

106 107

Figure 1. Schematic of the experimental setup (1-3) and the Raman system (4); L1,2: Convex

108

lenses; O: Microscope objective; D: Dichroic mirror; F1,2: Optical Fibers; R: Long pass filter

109

(535 nm); S: slit.

110 111

The investigated compounds (water and organic solvent) flow through the MCS from a starting

112

point to an end, wherefore the functionality of the MCS and the function of its components will

113

be described along the flow direction. At the starting point, two of the three syringe pumps ((2)

114

in Figure 1) (2x Teledyne Isco 100 DX, USA) one containing water and the other one containing

115

either MeOH, EtOH or Acn reduce their volumes to feed a specific volume flow (+-1 µl min-1)

116

into the MCS. The three-way valves next to the outlets of the syringe pumps serve for the

117

refill/renew or the cleaning of the respective syringe pump. Within the subsequent T-junction,

118

the two fed compounds merge and mix within a stainless steel capillary (ID = 1 mm,

119

length = 0.6 m) from which 0.25 m are located within the heating system area (3), to preheat the

120

mixture before it reaches the fused silica microcapillary. The temperature conditioning of the

ACS Paragon Plus Environment

6

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

The Journal of Physical Chemistry

121

binary mixture is assured by the microcapillary heating system (3). A fan channels air above a

122

heating coil according to the simple principle of a hairdryer whereupon a nozzle guides the

123

heated air laminar along the fused silica microcapillary ((1) in Figure 1). A controller (JUMO,

124

iTRON 32, Switzerland) that receives the temperature actual value from a temperature sensor

125

located 5 mm beside the measuring volume automatically determines the heating power of the

126

heating coil required for a certain temperature set point. The transition from the stainless steel

127

capillary towards the fused silica microcapillary (ID = 700 µm, OD = 850 µm) (Molex,

128

Polymicro TechnologiesTM) is sealed by a PEEK sleeve – located between the stainless steel

129

capillary and the fused silica microcapillary – and two component glue. Further downstream,

130

immediately after the transition from the microcapillary to a short stainless steel capillary, a

131

Coriolis densitometer (Bronkhorst, Mini Cori-Flow ML120, Germany) determines the density

132

and the mass flow of the mixture. The Coriolis densitometer is conditioned to exactly the

133

temperature and the pressure of the fluid mixture in the measuring volume. A further stainless

134

steel capillary connects the Coriolis densitometer with the third syringe pump ((2-1) in Figure 1)

135

(Teledyne Isco 500 D). This pump runs reverse and takes up the binary mixture in a way that the

136

desired pressure within the entire microcapillary path is constant. According to the maximal

137

volume flow of the summation of all delivered compounds of 220 µl min-1 the maximal flow rate

138

is 0.95 cm s-1 at the lowest inner diameter of the MCS. The pressure indications of the two feed

139

pumps at the start point of the MCS and the one reverse running pump at the end point of the

140

MCS do not show a measurable pressure drop along the MCS.

141

The Raman system ((4) in Figure 1) whereof the Raman sensor is sketched, includes also the

142

excitation continuous wave diode pumped solid-state laser (SmabaTM, Cobolt, 531.93 nm,

143

0.250 W, Sweden) and for the signal detection a spectrometer (QE Pro, Ocean Optics, 100 µm

ACS Paragon Plus Environment

7

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

Page 8 of 31

144

entrance slit, USA) including a cooled 1024x58 (pixels) charge coupled device (CCD)-detector.

145

An optical fiber (F1) guides the excitation laser beam towards the Raman sensor where the beam

146

is collimated (L1) and reflected through a high power microscope objective (High Power

147

MicroSpot LMH-20x, ThorLabs, NA = 0.4, USA) (O) that focuses the excitation laser beam into

148

the probe volume in the center of the microcapillary (C). The high NA of the microscope

149

objective ensures that no signals from the microcapillary walls reach the spectrometer and

150

interfere with the desired Raman signal of the fluids contained inside the capillary. An aperture

151

before the objective (S) lets pass only the high intensity part of the laser beam with originally a

152

Gaussian intensity profile. The red-shifted Raman signal scattered in the probe volume passes

153

straight the dichroitic mirror (reflecting wavelengths smaller than 535 nm and transmitting

154

wavelengths larger than 535 nm) (D) whereon a razor edge long-pass filter (R) blocks the

155

remaining elastically scattered light. The second lens (L2: f = 60 mm and Ø = 25.4 mm;

156

NA = 0.21) focuses the Raman signals into the detection fiber (core Ø = 200 µm) (F2). The

157

detection fiber guides the detected signals to the spectrometer.

158 159

2.3 Raman spectra acquisition using MCS

160

At the beginning of each experimental run, we flush the system with pure liquid water at the

161

desired temperature of 308 K, 318 K or 328 K. Therefore, the first water containing syringe

162

pump delivers 200 µL min-1 of water until the pressure within the third syringe pump at the end

163

of the microcapillary reaches 0.4 MPa. Now the third syringe pump starts to run reverse and

164

takes up the fed water in a way that the pressure within the MCS remains constant at 0.4 MPa.

165

We record the first Raman spectra from pure water at the set temperature that corresponds to a

166

mixture composition of the molar fraction 𝑥𝑊 = 1. Afterwards, the second syringe pump

ACS Paragon Plus Environment

8

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

The Journal of Physical Chemistry

167

containing either Acn, EtOH or MeOH starts delivering the compound into the MCS while at the

168

same time the feed flow of the water is reduced accordingly. The two compounds mix at the T-

169

junction and stream afterwards through the stainless steel capillary through the heating system,

170

before the mixture enters the fused silica microcapillary. Via the set flow rates of the two feed

171

syringe pumps the composition of the mixture inside the capillary system can be adjusted. For

172

example, 152 µl min-1 ethanol and 47 µl min-1 water were set to achieve an equimolar binary

173

mixture. The overall volumetric flow rate with respect to the summation of the pure substances

174

water and solvent is kept between 200-220 µL min-1. The small flow velocities (less than 1 cm s-

175

1)

176

system, before they pass the Raman measurement location. Gradually the composition of the

177

binary mixture is varied from pure water 𝑥𝑊 = 1 to pure organic solvent 𝑥𝑊 = 0. When the

178

mixture is varied from one composition to the next one, it takes approximately 20 minutes to

179

reach stationary conditions. Stationary conditions are firstly indicated by the stationary and non-

180

fluctuating mixture density and mass flow (Coriolis) and secondly by a constant Raman

181

spectrum. At the end of the experimental run, the MCS is flushed (200 µL min-1) with the

182

organic solvent in order to record a pure organic solvent spectrum 𝑥𝑊 = 0. During the entire

183

procedure, the mixture density and the mass flow are recorded (Coriolis). The density of the

184

mixture is required to convert the molar compositions (mol mol-1) of the mixture into

185

concentrations (mol L-1).

assure that the fluids inside the capillary adopt the temperature of the air inside the heating

186

For each stationary mixture composition 100 Raman spectra are taken with an exposure time

187

of 3.4 s spectrum-1. Consequently, the acquisition of 100 Raman spectra takes ~340 s per

188

adjusted mixture composition. The Raman spectra are processed and evaluated with respect to

ACS Paragon Plus Environment

9

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

Page 10 of 31

189

the development of hydrogen bonds according to the procedure described in the section that

190

follows.

191

192

3 Data evaluation

193

3.1 Processing of the spectra

194

“OH stretching vibration” from now on refers to the symmetric stretching vibration of water

195

and to the OH stretching vibration of the hydroxyl functional group of alcohols. For every single

196

analyzed mixture composition, a mean spectrum is averaged from the 100 Raman spectra. Each

197

experiment was triplicated, in total on three different days. The standard deviation of the

198

triplication is provided as error bars in the figures that follow.

199

Figure 2 summarizes from the top to the bottom how the mean spectra were treated in order to

200

determine (i) the experimental molar Raman spectra of the mixture and (ii) the partial molar

201

Raman spectra of the mixture constituents. A detailed description of the chain of processing the

202

spectra is provided in the supporting information (SI). Here a short overview is given.

ACS Paragon Plus Environment

10

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

The Journal of Physical Chemistry

203

ACS Paragon Plus Environment

11

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

Page 12 of 31

204

Figure 2. Raman spectra of an example W/MeOH mixture (𝑥𝑊 = 0.84) at 308 K and 0.4 MPa

205

showing the Raman shift region of the CH- and the OH stretching vibration. From top to bottom:

206

Pure spectrum as recorded 𝑆𝑚𝑖𝑥(𝜈𝑆), spectrum after the baseline correction (BLC) 𝑆𝑒𝑥𝑝 𝑚𝑖𝑥(𝜈𝑆),

207

isolated OH Raman spectrum 𝑆𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆) after the subtraction of the CH Raman spectrum (CHS),

208

𝑂𝐻 molar OH Raman spectrum 𝑠𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆) and partial molar OH Raman spectra of water 𝑠𝑊(𝑀𝑒𝑂𝐻)(𝜈𝑆)

209

in MeOH and methanol 𝑠𝑂𝐻 𝑀𝑒𝑂𝐻(𝑊)(𝜈𝑆) in water. The blue shaded rectangles mark the Raman shift

210

region from which the centroid is calculated. A detailed description of the chain of processing

211

the spectra is provided in the supporting information (SI).

212 213

Figure 2 shows at the top the raw mean spectrum 𝑆𝑚𝑖𝑥(𝜈𝑆) acquired from a water/methanol

214

(W/MeOH) mixture (𝑥𝑊 = 0.84) as a function of the Raman shift 𝜈𝑆. The background is fitted by

215

a cubic spline function and subtracted from 𝑆𝑚𝑖𝑥(𝜈𝑆) in order to achieve a background

216

compensated spectrum 𝑆𝑒𝑥𝑝 𝑚𝑖𝑥(𝜈𝑆) that is shown as second spectrum in Figure 2. The Raman

217

spectral features that can be assigned to the CH vibration of the organic solvents and to the OH

218

stretching vibration of water and the alcohols are labelled with “CH” or “OH”, respectively. The

219

CH Raman signatures are subtracted in order to obtain an isolated OH Raman spectrum 𝑆𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆)

220

of the mixture, which is shown as third spectrum in Figure 2. 𝑆𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆) (capital letter “𝑆”) is

221

divided by the concentration (mol L-1) of the mixture in order to obtain a molar Raman spectrum

222

𝑠𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆) (lowercase letter “𝑠”) of the OH vibration of the mixture that is shown as fourth

223

spectrum in Figure 2. The concentration of the mixture is known from the density measurements

224

(Coriolis), the molar masses and the set composition. Finally, the molar Raman OH spectrum of

225

the mixture is deconstructed (for details see the supporting information (SI)) into the shares

ACS Paragon Plus Environment

12

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

The Journal of Physical Chemistry

226

assignable to water and the organic solvent. As the construction method is identical to the

227

determination of partial molar properties from molar mixture properties, the corresponding and

228

resulting spectra shown in Figure 2 at the bottom are referred to as partial molar Raman spectra

229

𝑂𝐻 𝑠𝑂𝐻 𝑖(𝑗)(𝜈𝑆) of the respective compound 𝑖 in a mixture with compound 𝑗. For example 𝑠𝑀𝑒𝑂𝐻(𝑊)(𝜈𝑆)

230

shows the partial molar Raman spectrum of the OH vibration of MeOH in the mixture with

231

water. 𝑠𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻)(𝜈𝑆) shows the partial molar Raman spectrum of the OH vibration of water in the

232

mixture with MeOH. The partial molar Raman spectrum of 𝑖 in 𝑗 represents (exactly as partial

233

molar thermodynamic properties do), how the molar Raman spectrum of the mixture changes if

234

an infinite small amount of 𝑖 is added to the mixture. The molar fraction weighted summation 𝑂𝐻 𝑂𝐻 𝑠𝑂𝐻 𝑊/𝑀𝑒𝑂𝐻(𝜈𝑆) = 𝑥𝑊𝑠𝑊(𝑀𝑒𝑂𝐻)(𝜈𝑆) + (1 ― 𝑥𝑊)𝑠𝑀𝑒𝑂𝐻(𝑊)(𝜈𝑆)

(1)

235

𝑂𝐻 of the partial molar spectra 𝑠𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻)(𝜈𝑆) and 𝑠𝑀𝑒𝑂𝐻(𝑊)(𝜈𝑆) is the molar spectrum of the

236

mixture 𝑠𝑂𝐻 𝑊/𝑀𝑒𝑂𝐻(𝜈𝑆). According to the authors’ knowledge, the deconstruction of mixture

237

spectra into their spectrally overlapping shares based on the computation of partial molar spectra

238

is new.

239 240

Figure 3 shows for the example binary system W/MeOH, how the ideal molar mixture spectrum 𝑂𝐻,0 𝑂𝐻,0 𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑊/𝑀𝑒𝑂𝐻(𝜈𝑆) = 𝑥𝑊𝑠𝑊 (𝜈𝑆) + (1 ― 𝑥𝑊)𝑠𝑀𝑒𝑂𝐻(𝜈𝑆)

(2)

241

is computed from the molar Raman spectra of the pure (upper index 0) compounds 𝑠𝑂𝐻,0 𝑊 (𝜈𝑆) and

242

𝑠𝑂𝐻,0 𝑀𝑒𝑂𝐻(𝜈𝑆) as molar fraction weighted average. Figure 3 also shows that the ideal spectrum

243

(𝜈𝑆) is different from the real molar spectrum of the mixture 𝑠𝑂𝐻 𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑚𝑖𝑥 𝑚𝑖𝑥(𝜈𝑆), exactly as it is the

244

case for many thermodynamic properties, such as entropy or Gibbs energy. As it will be shown

245

in the next section, the difference between the two OH spectra can be quantified based on the

ACS Paragon Plus Environment

13

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

Page 14 of 31

246

Raman shifts of their centroids. This is why also the centroids are indicated in Figure 2 and

247

Figure 3.

248 249

Figure 3. Isolated OH stretching vibration of the molar Raman spectra of pure water and pure

250

methanol at 308 K and 0.4 MPa (upper row). Molar Raman spectra of the real (black) and ideal

251

(blue) W/MeOH-mixture at 𝑥𝑊 = 0.84 (lower row). The blue shaded rectangle marks the Raman

252

shift region from which the centroid is calculated

253

ACS Paragon Plus Environment

14

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

The Journal of Physical Chemistry

254

In Figure 2 and Figure 3 the positive and negative zig-zag peaks between 2800 and 3000 cm-1

255

remain as mathematical residual after the subtraction of the CH Raman signal. They do not have

256

any physical meaning and they are not regarded in the further processing of the data.

257 258

3.2 Determination of the centroid of the OH stretching vibration Raman signal

259 260 261

The Raman shift of the centroid 𝑣𝑂𝐻 (centroid position) of the OH stretching vibration is computed as generally described in Reference54 𝑣𝑂𝐻 =

262





267

(3)

In equation (3) we consider the molar OH Raman spectrum of the mixture 𝑠(𝜈𝑆) = 𝑠𝑂𝐻 𝑚𝑖𝑥

We consider the partial molar spectra of the constituents of the mixture 𝑠(𝜈𝑆) = 𝑠𝑂𝐻 𝑖(𝑗)(𝜈𝑆) 𝑂𝐻

for the computation of their centroid position 𝑣𝑂𝐻 = 𝑣𝑖(𝑗).

265 266

∫3033 𝑐𝑚 ―1𝑠(𝜈𝑆)𝜈𝑆 𝑑𝜈𝑆.

(𝜈𝑆) for the computation of its centroid position 𝑣𝑂𝐻 = 𝑣𝑂𝐻 𝑚𝑖𝑥.

263 264

3798 𝑐𝑚 ―1

1 3798 𝑐𝑚 ―1 ∫3033 𝑐𝑚 ―1𝑠(𝜈𝑆)𝑑𝜈𝑆



𝑖𝑑𝑒𝑎𝑙( ) We consider the ideal molar mixture spectra 𝑠(𝜈𝑆) = 𝑠𝑂𝐻, 𝜈𝑆 for the computation of 𝑚𝑖𝑥 𝑂𝐻,𝑖𝑑𝑒𝑎𝑙

𝑣𝑂𝐻 = 𝑣𝑚𝑖𝑥

.

268

The calculation of the centroid of the OH stretching vibration is a straightforward method to

269

observe and characterize changes in the shape of the OH stretching vibration that is affected by

270

the average development of the HBN of the mixture. The stronger the HBN is developed, the less

271

energy is stored in the OH vibration of alcohols and/or water and therefore the smaller is the

272

centroid position 𝑣𝑂𝐻.

273

ACS Paragon Plus Environment

15

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

274

Page 16 of 31

4 Results and Discussion

275

Figure 4 shows the centroid positions of the molar OH Raman spectra of the three binary

276

mixtures water/methanol (W/MeOH), water/ethanol (W/EtOH) and water/acetonitrile (W/Acn)

277

as a function of the water molar fraction in the mixture 𝑥𝑊. The centroids 𝜈𝑂𝐻 𝑚𝑜𝑙𝑎𝑟 computed from

278

𝑂𝐻 𝑂𝐻 the real mixture spectra 𝑠𝑂𝐻 𝑊/𝑀𝑒𝑂𝐻(𝜈𝑆), 𝑠𝑊/𝐸𝑡𝑂𝐻(𝜈𝑆) and 𝑠𝑊/𝐴𝑐𝑛(𝜈𝑆) are given as solid discs, while

279

𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 the centroids 𝜈𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 computed from the ideal mixture spectra 𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑊/𝑀𝑒𝑂𝐻(𝜈𝑆), 𝑠𝑊/𝐸𝑡𝑂𝐻(𝜈𝑆) and 𝑚𝑖𝑥

280

𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑊/𝐴𝑐𝑛 (𝜈𝑆) are given as black crosses. The very right values (𝑥𝑊 = 1) represent the centroid

281

positions of the OH Raman spectrum of pure water and the very left values (𝑥𝑊 = 0) represent

282

the centroid positions for the OH Raman spectrum of pure organic solvents. While the alcohols

283

MeOH and EtOH feature an OH Raman signal, Acn – simply due to the lack of the hydroxyl

284

group – does not feature an OH Raman signal. Therefore, no data points are given for pure Acn (

285

𝑥𝑊 = 0). The Acn-richest W/Acn mixture, in which the OH Raman spectrum of water is still

286

observable is 𝑥𝑊 = 0.02 (see supporting information). Therefore, only in the case of W/Acn the

287

ideal mixture spectra 𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑊/𝐴𝑐𝑛 (𝜈𝑆) were not computed according to equation (2) but according to 𝑠𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑊/𝐴𝑐𝑛 (𝜈𝑆) =

(𝑥𝑊 ― 0.02) 0.98

𝑠𝑂𝐻,0 𝑤 (𝜈𝑆) +

(1 ― 𝑥𝑊) 𝑂𝐻,𝑥 = 0.02 𝑤 𝑠 (𝜈𝑆) 0.98 𝑊/𝐴𝑐𝑛

(4)

288

from the molar Raman spectrum of pure water 𝑠𝑂𝐻,0 𝑊 (𝜈𝑆) and the molar Raman spectrum

289

𝑤 = 0.02 (𝜈𝑆) of the Acn-rich mixture with 𝑥𝑊 = 0.02. The centroid positions were then 𝑠𝑂𝐻,𝑥 𝑊/𝐴𝑐𝑛

290

computed as always according to equation (3).

ACS Paragon Plus Environment

16

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

The Journal of Physical Chemistry

291 292

Figure 4. Centroid positions 𝜈𝑂𝐻 𝑚𝑖𝑥 of the molar spectra as a function of the mixture composition

293

for the binary mixtures W/MeOH, W/EtOH and W/Acn at 308 K, 318 K and 328 K. The dashed

294

lines are to guide the eyes and the error bars represent the standard deviation of the triplication of

295

each data point (here smaller than the data points).

296 297

The centroid position 𝑣𝑂𝐻 𝑚𝑖𝑥 increases if water is diluted with Acn, but decreases when water is

298

diluted with MeOH or EtOH. The addition of acetonitrile to water weakens the HBN of the

299

mixture in the entire composition range, while the HBN of the mixture is strengthened when

300

MeOH or EtOH are added. The centroid positions 𝑣𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 (black crosses) of the ideal molar OH 𝑚𝑖𝑥

301

𝑖𝑑𝑒𝑎𝑙 Raman spectra 𝑠𝑂𝐻, (𝜈𝑆) are close to the ones (𝑣𝑂𝐻 𝑚𝑖𝑥 𝑚𝑖𝑥) computed from the real molar OH

302

Raman spectra 𝑠𝑂𝐻 𝑚𝑖𝑥(𝜈𝑆) for every investigated mixture. With increasing temperature the HBN

303

gets less developed. Therefore 𝜈𝑂𝐻 𝑚𝑖𝑥 increases with increasing temperature.

304 305

𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 In order to quantify the deviation of the real 𝜈𝑂𝐻 centroid positions, the 𝑚𝑖𝑥 and the ideal 𝜈𝑚𝑖𝑥

relative deviations

ACS Paragon Plus Environment

17

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

∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥

=

𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝜈𝑂𝐻 𝑚𝑖𝑥 ― 𝜈𝑚𝑖𝑥

𝜈𝑂𝐻,𝑖𝑑𝑒𝑎𝑙 𝑚𝑖𝑥

Page 18 of 31

(5)

306

for the three mixtures are provided as a function of the mixture composition for three different

307

temperatures in Figure 5.

308 309

Figure 5. Relative deviations ∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥 of the real and ideal centroid positions as a function of

310

the composition of the binary mixtures W/MeOH, W/EtOH and W/Acn at 308 K, 318 K, 328 K.

311

The dashed lines are to guide the eyes and the error bars represent the standard deviation of the

312

triplication of each data point.

313 314

Here negative values imply that the HBN represented by the mixture real spectrum is more

315

developed than the HBN represented by the mixture ideal spectrum. Vice versa, positive values

316

imply that the HBN shown by the mixture real spectrum is less developed than in the ideal one.

317

It should be underlined here that according to equation (2) the ideal molar OH Raman spectra

318

𝑖𝑑𝑒𝑎𝑙 (𝜈𝑆) are simply computed as molar fraction weighted average of the pure compound OH 𝑠𝑂𝐻, 𝑚𝑖𝑥

319

Raman spectra. Therefore, any decay of ∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥 from zero indicates that the HBN in the real

320

mixture not simply is a weighted average of the HBN of the pure compounds but the result of a

ACS Paragon Plus Environment

18

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

The Journal of Physical Chemistry

321

change in the intermolecular forces and of the orientation of the molecules relative to each other.

322

Figure 5 shows that the negative deviations of ∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥 increase with decreasing temperature

323

and that the positive deviations of ∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥 increase with increasing temperature. Therefore,

324

cold temperatures support the formation of HBN in real mixtures that are more intense than in

325

ideal mixtures. Warm temperatures support the formation of HBN that are less intense than in the

326

ideal mixture case. The HBN deviations from the ideal mixture behavior can explain the

327

anomalies found for the respective mixtures with respect to their density, their conductivity and

328

other properties.27,55–58 ∆𝑖𝑑𝑒𝑎𝑙𝜈𝑂𝐻 𝑚𝑖𝑥 was found to be most temperature sensitive for the system

329

W/Acn. This can be explained by the proximity of the analyzed temperatures to the upper critical

330

solution temperature of this binary system, which at ambient pressure is at 272 K.59

331

𝑂𝐻 𝑂𝐻 Figure 6 shows the centroid positions 𝑣𝑂𝐻 𝑖(𝑗) and 𝑣𝑗(𝑖) of the partial molar Raman spectra 𝑠𝑖(𝑗)

332

(𝜈𝑆) and 𝑠𝑂𝐻 𝑗(𝑖)(𝜈𝑆), respectively, for all three analyzed mixtures in three columns and for all three

333

analyzed temperatures as a function of the binary mixture composition 𝑥𝑤. The upper three

334

diagrams show the centroid position of the partial molar Raman spectrum of water in the mixture

335

with the organic solvent and thus represent the development of the water-HBN in the mixture. In

336

this context, water-HBN refers to the HBN that water molecules develop to their neighbors,

337

which can be other water molecules and/or organic solvent molecules. The lower three diagrams

338

show vice versa the centroid position of the partial molar Raman spectrum of the organic solvent

339

(OS) in the mixture with water and thus represent the development of the OS-HBN of the

340

mixture. In this context, OS-HBN refers to the HBN that organic solvent molecules develop to

341

their neighbors, which can be other organic solvent molecules and/or water molecules.

342

𝑂𝐻 In the following paragraph we first describe in detail the behaviors of 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) and 𝑣𝑀𝑒𝑂𝐻(𝑊).

ACS Paragon Plus Environment

19

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

Page 20 of 31

343

𝑂𝐻 For pure water 𝑥𝑤 = 1 it can be seen that 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) and 𝑣𝑀𝑒𝑂𝐻(𝑊) increase with increasing

344

temperature. Thus, the water-HBN and the OS-HBN get less developed with increasing

345

temperature.

346 347

𝑂𝐻 𝑂𝐻 𝑂𝐻 Figure 6. Centroid positions 𝑣𝑂𝐻 𝑖(𝑗) and 𝑣𝑗(𝑖) of the partial molar Raman spectra 𝑠𝑖(𝑗)(𝜈𝑆) and 𝑠𝑗(𝑖)

348

(𝜈𝑆), for the binary mixtures W/MeOH, W/EtOH and W/Acn and for 308 K, 318 K and 328 K as

349

a function of the binary mixture composition 𝑥𝑤.

350 351

For molar fractions 1 > 𝑥𝑊 > 0.8 (regime 1) 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) decreases to a minimum at 𝑥𝑊 ≈ 0.8,

352

when pure water is diluted with MeOH. This means that in regime 1 the water-HBN develops

353

stronger the more MeOH is added to the mixture until a maximum is reached at 𝑥𝑊 ≈ 0.8. On the

ACS Paragon Plus Environment

20

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

The Journal of Physical Chemistry

354

contrary 𝑣𝑂𝐻 𝑀𝑒𝑂𝐻(𝑊) (lower diagram in Figure 6) increases in the same regime when MeOH is

355

added to pure water to a maximum at 𝑥𝑊 ≈ 0.8. This implies that MeOH at 𝑥𝑊 ≈ 0.8 develops

356

least H-bonds to its neighbors. Obviously the minimum of 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) and the maximum of

357

𝑣𝑂𝐻 𝑀𝑒𝑂𝐻(𝑊) can be found at the same mixture composition 𝑥𝑊 ≈ 0.8, meaning that the maximum

358

development of the water-HBN exists when the minimum development of the OS-HBN exists.

359

These observations are in support of theories that suppose a formation of an intense water-HBN

360

around organic solvent molecules, where the organic solvent molecules participate the less in the

361

HBN, the more intense the water-HBN growth. This development turns around in the

362

composition regime 2 for molar fractions 0.8 > 𝑥𝑊 > 0.2. With the addition of more MeOH

363

𝑂𝐻 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) increases and 𝑣𝑀𝑒𝑂𝐻(𝑊) decreases. Therefore, the water-HBN gets less developed and

364

the OS-HBN develops stronger in regime 2 with the addition of more MeOH. One explanation is

365

that when more than 𝑥𝑊 ≈ 0.8 MeOH molecules are present in the mixture, their dissolution is no

366

more accomplished by surrounding them in a strong water-HBN, but by integrating them into the

367

HBN of the mixture. Therefore, the OS-HBN gets more intense, while the water-HBN gets less

368

intense. At the left border of regime 2 for mixture compositions of 𝑥𝑊 ≈ 0.2 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) shows a

369

local maximum while 𝑣𝑂𝐻 𝑀𝑒𝑂𝐻(𝑊) has already reached approximately the value that we also

370

measured for pure MeOH 𝑥𝑊 = 0. The local maximum of 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) implies that at 𝑥𝑊 ≈ 0.2

371

water molecules develop least H-bonds to their neighbors. Diluting the mixture with even more

372

MeOH in regime 3, which covers the compositions 0.2 > 𝑥𝑊 > 0.0, shows that 𝑣𝑂𝐻 𝑀𝑒𝑂𝐻(𝑊) remains

373

at a constant level, irrespectively of the mixture composition. Thus, the development of the OS-

374

HBN seems not to be influenced by the mixture composition in regime 3. This is different for

375

𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻), whose course in regime 3 very much depends on the mixture composition and on

ACS Paragon Plus Environment

21

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

Page 22 of 31

376

temperature. At 308 K 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) decreases with increasing MeOH content in the mixture. This

377

indicates that the water molecules develop the more H-bonds to their neighbors, the more MeOH

378

is in the mixture. As at the same time the OS-HBN remains unaltered, we speculate that in

379

regime 3 water molecules tend to develop H-bonds preferentially to other water molecules and

380

seem to avoid MeOH molecules. This interpretation is in agreement with the theory of the

381

formation of water-rich and MeOH-rich sub-micro domains (nanostructured fluids) that were

382

found in water/alcohol mixtures.13,22,30 At 328 K 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) shows a local minimum at 𝑥𝑊 ≈ 0.1.

383

Therefore, the accumulation of water molecules in water-rich sub-micro domains, in which the

384

water molecules tend to develop H-bonds mainly to other water-molecules, has a maximum at

385

𝑥𝑊 ≈ 0.1. In mixtures that at 328 K are even richer in MeOH 0.1 > 𝑥𝑊 > 0.0, these nanodomains

386

disintegrate again, which is represented by the increase of 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) towards 𝑥𝑊 = 0.

387

What has been described above for the mixture W/MeOH can be transferred directly to the

388

mixture W/EtOH. Therefore, we skip the description of the mixture W/EtOH. Nonetheless, the

389

mixture compositions at which transitions from regime 1 to regime 2 and from regime 2 to

390

regime 3 occur are slightly different. For both systems W/MeOH and W/EtOH the composition

391

at which the transition from regime 1 to regime 2 occurs (local minima of 𝑣𝑂𝐻 𝑊(𝑀𝑒𝑂𝐻) and

392

𝑣𝑂𝐻 𝑊(𝐸𝑡𝑂𝐻)) coincide with the composition range at which the binary systems feature the extremes

393

of their molar excess enthalpy.

394

In the mixture, W/Acn the course of 𝑣𝑂𝐻 𝑊(𝐴𝑐𝑛) as a function of the mixture composition is very

395

different compared to the mixtures W/MeOH and W/EtOH. When diluting water with Acn

396

𝑣𝑂𝐻 𝑊(𝐴𝑐𝑛) increases monotonically. Therefore the water-HBN is weakened the more, the more Acn

397

is added to the mixture. Putting this observation into the context of the known existence of

ACS Paragon Plus Environment

22

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

The Journal of Physical Chemistry

398

water-rich and acetonitrile-rich clusters for mixture compositions 0.4 < 𝑥𝑊 < 0.8,29,60 we must

399

conclude that the water-HBN in the water-rich clusters are less developed than in pure water. It

400

has been mentioned already above that the partial molar Raman spectrum of Acn contains an OH

401

Raman signal, though the Acn molecule does not feature an OH vibration. This is due to the fact

402

that partial molar properties show how the molar property of a mixture changes when

403

infinitesimal small amounts of Acn are added to the mixture. Thus, the partial molar Raman

404

spectrum of Acn shows how the OH Raman signal of the mixture is influenced when the mixture

405

is diluted with Acn. In this context large values of 𝑣𝑂𝐻 𝐴𝑐𝑛(𝑊) imply that Acn interacts with water

406

molecules that develop a less intense HBN and small values of 𝑣𝑂𝐻 𝐴𝑐𝑛(𝑊) imply that Acn interacts

407

with water molecules that develop an intense HBN.

408

5 Conclusion

409

We here showed some insight into the mixtures of water and organic solvent that are based on

410

the consideration of partial molar Raman spectra. The idea behind partial molar Raman spectra is

411

identical to the concept of partial molar thermodynamic properties such as partial molar volume,

412

enthalpy, entropy, energy and others, but in our case only applied to Raman spectra.

413

Figure 6 and what has been discussed in its context show that information can be extracted

414

from partial molar Raman spectra that is not extractable from mixture Raman spectra. The

415

consideration of partial molar Raman spectra of water and of the organic solvent made possible

416

the analysis of the hydrogen bond network that water molecules develop to their neighbors and

417

that organic solvent molecules develop to their neighbors. Because of this, for the mixtures

418

water/methanol and water/ethanol we found three different regimes, in which the interaction

419

mechanisms between water and alcohol changes.

ACS Paragon Plus Environment

23

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

Page 24 of 31

420

In general, the computation of partial molar Raman spectra is not limited to binary mixtures.

421

However, the more constituents are present in the mixture, the more experimental Raman spectra

422

have to be determined in order to compute the partial molar Raman spectra of each compound.

423

What we showed for the hydrogen bond network is possible for any intermolecular interaction

424

that modulates a signal peak in the Raman spectra. We carried out these experiments at a

425

pressure of 0.4 MPa, as some pressure is required to obtain reproducible operational conditions

426

in the microcapillary setup. Due to the low compressibility of the binary systems analyzed, we

427

speculate that the binary systems at ambient pressure would show very similar results.

ACS Paragon Plus Environment

24

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

428 429

The Journal of Physical Chemistry

ASSOCIATED CONTENT Supporting Information. Detailed description of the chain of processing the spectra and a list

430

of analyzed mixture compositions with their measured densities.

431

AUTHOR INFORMATION

432

Corresponding Author

433

*E-Mail: [email protected]

434

Author Contributions

435

The manuscript was written through contributions of all authors. All authors have given approval

436

to the final version of the manuscript.

437 438

ACKNOWLEDGMENT

439

The project leading to this contribution has received funding from the European Union’s Horizon

440

2020 research and innovation programme under grant agreement No. 637654 (Inhomogeneities).

441

The authors also gratefully acknowledge the funding of the Erlangen Graduate School in

442

Advanced Optical Technologies (SAOT) by the German Research Foundation (DFG) in the

443

framework of the German excellence initiative.

444

445

6 References

446 447 448 449

(1) Frank, H. S.; Evans, M. W. Free Volume and Entropy in Condensed Systems III. Entropy in Binary Liquid Mixtures; Partial Molal Entropy in Dilute Solutions; Structure and Thermodynamics in Aqueous Electrolytes. J. Chem. Phys. 1945, 13, 507–532, DOI: 10.1063/1.1723985.

ACS Paragon Plus Environment

25

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

Page 26 of 31

450 451 452

(2) Titantah, J. T.; Karttunen, M. Long-Time Correlations and Hydrophobe-Modified Hydrogen-Bonding Dynamics in Hydrophobic Hydration. J. Am. Chem. Soc. 2012, 134, 9362–9368, DOI: 10.1021/ja301908a.

453 454 455

(3) Straatsma, T. P.; Berendsen, H. J. C.; Postma, J. P. M. Free Energy of Hydrophobic Hydration: A Molecular Dynamics Study of Noble Gases in Water. J. Chem. Phys. 1986, 85, 6720–6727, DOI: 10.1063/1.451846.

456 457 458

(4) Tan, M.-L.; Miller, B. T.; Te, J.; Cendagorta, J. R.; Brooks, B. R.; Ichiye, T. Hydrophobic Hydration and the Anomalous Partial Molar Volumes in Ethanol-Water Mixtures. The Journal of chemical physics 2015, 142, 64501, DOI: 10.1063/1.4906750.

459 460

(5) Pangali, C.; Rao, M.; Berne, B. J. Hydrophobic Hydration around a Pair of Apolar Species in Water. J Solution Chem 1979, 71, 2982, DOI: 10.1063/1.438702.

461 462

(6) Blokzijl, W.; Engberts, J. B. F. N. Hydrophobic Effects. Opinions and Facts. Angew. Chem. Int. Ed. Engl. 1993, 32, 1545–1579, DOI: 10.1002/anie.199315451.

463 464 465

(7) Dolenko, T.; Burikov, S.; Hojo, M.; Patsaeva, S.; Yuzhakov, V. Hydrogen Bonding in Aqueous Ethanol Solutions Studied by Raman Spectroscopy 2010, DOI: 10.1117/12.871101.

466 467 468

(8) Rezus, Y. L. A.; Bakker, H. J. Observation of Immobilized Water Molecules around Hydrophobic Groups. Phys. Rev. Lett. 2007, 99, 148301, DOI: 10.1103/PhysRevLett.99.148301.

469 470

(9) Galamba, N. Water's Structure around Hydrophobic Solutes and the Iceberg Model. The journal of physical chemistry. B 2013, 117, 2153–2159, DOI: 10.1021/jp310649n.

471 472 473

(10) Grdadolnik, J.; Merzel, F.; Avbelj, F. Origin of Hydrophobicity and Enhanced Water Hydrogen Bond Strength near purely Hydrophobic Solutes. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 322–327, DOI: 10.1073/pnas.1612480114.

474 475 476 477 478

(11) Dolenko, T. A.; Burikov, S. A.; Dolenko, S. A.; Efitorov, A. O.; Plastinin, I. V.; Yuzhakov, V. I.; Patsaeva, S. V. Raman Spectroscopy of Water-Ethanol Solutions: The Estimation of Hydrogen Bonding Energy and the Appearance of Clathrate-like Structures in Solutions. The journal of physical chemistry. A 2015, 119, 10806–10815, DOI: 10.1021/acs.jpca.5b06678.

479 480

(12) Gill, S. J.; Dec, S. F.; Olofsson, G.; Wadsoe, I. Anomalous Heat Capacity of Hydrophobic Solvation. J. Phys. Chem. 1985, 89, 3758–3761, DOI: 10.1021/j100263a034.

481 482 483

(13) Benson, S. P.; Pleiss, J. Incomplete Mixing versus Clathrate-like Structures: A Molecular View on Hydrophobicity in Methanol-Water Mixtures. Journal of molecular modeling 2013, 19, 3427–3436, DOI: 10.1007/s00894-013-1857-1.

484 485

(14) Chandler, D. Interfaces and the Driving Force of Hydrophobic Assembly. Nature 2005, 437, 640–647, DOI: 10.1038/nature04162.

ACS Paragon Plus Environment

26

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

The Journal of Physical Chemistry

486 487 488

(15) Dougan, L.; Hargreaves, R.; Bates, S. P.; Finney, J. L.; Réat, V.; Soper, A. K.; Crain, J. Segregation in Aqueous Methanol Enhanced by Cooling and Compression. The Journal of chemical physics 2005, 122, 174514, DOI: 10.1063/1.1888405.

489 490 491

(16) Laage, D.; Stirnemann, G.; Hynes, J. T. Why Water Reorientation slows without Iceberg Formation around Hydrophobic Solutes. The journal of physical chemistry. B 2009, 113, 2428–2435, DOI: 10.1021/jp809521t.

492 493 494

(17) Dixit, S.; Crain, J.; Poon, W. C. K.; Finney, J. L.; Soper, A. K. Molecular Segregation Observed in a Concentrated Alcohol-Water Solution. Nature 2002, 416, 829–832, DOI: 10.1038/416829a.

495 496 497

(18) Guo, J.-H.; Luo, Y.; Augustsson, A.; Kashtanov, S.; Rubensson, J.-E.; Shuh, D. K.; Agren, H.; Nordgren, J. Molecular Structure of Alcohol-Water Mixtures. Phys. Rev. Lett. 2003, 91, 157401, DOI: 10.1103/PhysRevLett.91.157401.

498 499 500

(19) Moreau, C.; Douhéret, G. Thermodynamic and Physical Behaviour of Water + Acetonitrile Mixtures. Dielectric Properties. The Journal of Chemical Thermodynamics 1976, 8, 403– 410, DOI: 10.1016/0021-9614(76)90060-4.

501 502 503

(20) Jamroz, D.; Stangret, J.; Lindgren, J. An Infrared Spectroscopic Study of the Preferential Solvation in Water-Acetonitrile Mixtures. J. Am. Chem. Soc. 1993, 115, 6165–6168, DOI: 10.1021/ja00067a036.

504 505 506

(21) Bertie, J. E.; Lan, Z. Liquid Water−Acetonitrile Mixtures at 25 °C: The Hydrogen-Bonded Structure Studied through Infrared Absolute Integrated Absorption Intensities. J. Phys. Chem. B 1997, 101, 4111–4119, DOI: 10.1021/jp9639511.

507 508

(22) Chen, J.; Sit, P. H.-L. Ab Initio Study of the Structural Properties of Acetonitrile–Water Mixtures. Chemical Physics 2015, 457, 87–97, DOI: 10.1016/j.chemphys.2015.05.022.

509 510

(23) Bakker, H. J. Physical Chemistry: Water's Response to the Fear of Water. Nature 2012, 491, 533–535, DOI: 10.1038/491533a.

511 512

(24) Davis, J. G.; Gierszal, K. P.; Wang, P.; Ben-Amotz, D. Water Structural Transformation at Molecular Hydrophobic Interfaces. Nature 2012, 491, 582–585, DOI: 10.1038/nature11570.

513 514 515

(25) Idrissi, A.; Longelin, S.; Sokolić, F. Study of Aqueous Acetone Solution at Various Concentrations: Low-Frequency Raman and Molecular Dynamics Simulations. J. Phys. Chem. B 2001, 105, 6004–6009, DOI: 10.1021/jp004217r.

516 517 518

(26) Kovacs, H.; Laaksonen, A. Molecular Dynamics Simulation and NMR Study of WaterAcetonitrile Mixtures. J. Am. Chem. Soc. 1991, 113, 5596–5605, DOI: 10.1021/ja00015a013.

519 520

(27) Marcus, Y.; Migron, Y. Polarity, Hydrogen Bonding, and Structure of Mixtures of Water and Cyanomethane. J. Phys. Chem. 1991, 95, 400–406, DOI: 10.1021/j100154a070.

ACS Paragon Plus Environment

27

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

Page 28 of 31

521 522 523

(28) Byakov, V. M.; Lanshina, L. V.; Stepanova, O. P.; Stepanov, S. V. The Nanoheterogeneous Structure of Aqueous Solutions of n-Propanol. Russ. J. Phys. Chem. 2009, 83, 214–219, DOI: 10.1134/S0036024409020125.

524 525 526

(29) Takamuku, T.; Tabata, M.; Yamaguchi, A.; Nishimoto, J.; Kumamoto, M.; Wakita, H.; Yamaguchi, T. Liquid Structure of Acetonitrile−Water Mixtures by X-ray Diffraction and Infrared Spectroscopy. J. Phys. Chem. B 1998, 102, 8880–8888, DOI: 10.1021/jp9824297.

527 528 529 530

(30) Li, R.; D'Agostino, C.; McGregor, J.; Mantle, M. D.; Zeitler, J. A.; Gladden, L. F. Mesoscopic Structuring and Dynamics of Alcohol/Water Solutions probed by Terahertz Time-domain Spectroscopy and pulsed Field Gradient Nuclear Magnetic Resonance. The journal of physical chemistry. B 2014, 118, 10156–10166, DOI: 10.1021/jp502799x.

531 532

(31) Lin, K.; Hu, N.; Zhou, X.; Liu, S.; Luo, Y. Reorientation Dynamics in Liquid Alcohols from Raman Spectroscopy. J. Raman Spectrosc. 2012, 43, 82–88, DOI: 10.1002/jrs.2997.

533 534 535

(32) Hands, M. D.; Slipchenko, L. V. Intermolecular Interactions in Complex Liquids: Effective Fragment Potential Investigation of Water-tert-Butanol Mixtures. The journal of physical chemistry. B 2012, 116, 2775–2786, DOI: 10.1021/jp2077566.

536 537 538

(33) Subramanian, D.; Anisimov, M. A. Phase Behavior and Mesoscale Solubilization in Aqueous Solutions of Hydrotropes. Fluid Phase Equilibria 2014, 362, 170–176, DOI: 10.1016/j.fluid.2013.09.064.

539 540 541

(34) Venables, D. S.; Schmuttenmaer, C. A. Spectroscopy and Dynamics of Mixtures of Water with Acetone, Acetonitrile, and Methanol. J. Chem. Phys. 2000, 113, 11222–11236, DOI: 10.1063/1.1328072.

542 543

(35) Bergman, D. L.; Laaksonen, A. Topological and Spatial Structure in the Liquid-Water– Acetonitrile Mixture. Phys. Rev. E 1998, 58, 4706–4715, DOI: 10.1103/PhysRevE.58.4706.

544 545

(36) Mountain, R. D. Microstructure and Hydrogen Bonding in Water-Acetonitrile Mixtures. The journal of physical chemistry. B 2010, 114, 16460–16464, DOI: 10.1021/jp105272q.

546 547 548

(37) Dougan, L.; Bates, S. P.; Hargreaves, R.; Fox, J. P.; Crain, J.; Finney, J. L.; Reat, V.; Soper, A. K. Methanol-Water Solutions: A bi-Percolating Liquid Mixture. The Journal of chemical physics 2004, 121, 6456–6462, DOI: 10.1063/1.1789951.

549 550 551

(38) Egashira, K.; Nishi, N. Low-Frequency Raman Spectroscopy of Ethanol−Water Binary Solution: Evidence for Self-Association of Solute and Solvent Molecules. J. Phys. Chem. B 1998, 102, 4054–4057, DOI: 10.1021/jp9806359.

552 553 554

(39) Cringus, D.; Yeremenko, S.; Pshenichnikov, M. S.; Wiersma, D. A. Hydrogen Bonding and Vibrational Energy Relaxation in Water−Acetonitrile Mixtures. J. Phys. Chem. B 2004, 108, 10376–10387, DOI: 10.1021/jp0495141.

555 556

(40) Mountain, R. D. Molecular Dynamics Study of Water−Acetonitrile Mixtures. J. Phys. Chem. A 1999, 103, 10744–10748, DOI: 10.1021/jp992305+.

ACS Paragon Plus Environment

28

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

The Journal of Physical Chemistry

557 558 559

(41) Noskov, S. Y.; Lamoureux, G.; Roux, B. Molecular Dynamics Study of Hydration in Ethanol-Water Mixtures using a Polarizable Force Field. J. Phys. Chem. B 2005, 109, 6705– 6713, DOI: 10.1021/jp045438q.

560 561 562

(42) Wakisaka, A.; Matsuura, K. Microheterogeneity of Ethanol–Water Binary Mixtures Observed at the Cluster Level. Journal of Molecular Liquids 2006, 129, 25–32, DOI: 10.1016/j.molliq.2006.08.010.

563 564 565

(43) Soper, A. K.; Dougan, L.; Crain, J.; Finney, J. L. Excess Entropy in Alcohol-Water Solutions: A Simple Clustering Explanation. J. Phys. Chem. B 2006, 110, 3472–3476, DOI: 10.1021/jp054556q.

566 567 568 569

(44) Bakó, I.; Megyes, T.; Grósz, T.; Pálinkás, G.; Dore, J. Structural Investigation of Water– Acetonitrile Mixtures: Small-Angle and Wide-Angle Neutron Diffraction Study Compared to Molecular Dynamics Simulation. Journal of Molecular Liquids 2005, 125, 174–180, DOI: 10.1016/j.molliq.2005.11.022.

570

(45) Yves Maréchal. The Hydrogen Bond and the Water Molecule; Elsevier, 2007.

571 572 573

(46) Nedić, M.; Wassermann, T. N.; Larsen, R. W.; Suhm, M. A. A Combined Raman- and Infrared Jet Study of Mixed Methanol-Water and Ethanol-Water Clusters. Physical chemistry chemical physics : PCCP 2011, 13, 14050–14063, DOI: 10.1039/c1cp20182d.

574 575 576 577

(47) Koga, Y.; Sebe, F.; Minami, T.; Otake, K.; Saitow, K.-i.; Nishikawa, K. Spectrum of Excess Partial Molar Absorptivity. I. Near Infrared Spectroscopic Study of Aqueous Acetonitrile and Acetone. The journal of physical chemistry. B 2009, 113, 11928–11935, DOI: 10.1021/jp901934c.

578 579 580

(48) Li, Q.; Wang, N.; Zhou, Q.; Sun, S.; Yu, Z. Excess Infrared Absorption Spectroscopy and its Applications in the Studies of Hydrogen Bonds in Alcohol-containing Binary Mixtures. Applied spectroscopy 2008, 62, 166–170, DOI: 10.1366/000370208783575663.

581 582 583 584

(49) Wallace, V. M.; Dhumal, N. R.; Zehentbauer, F. M.; Kim, H. J.; Kiefer, J. Revisiting the Aqueous Solutions of Dimethyl Sulfoxide by Spectroscopy in the Mid- and Near-Infrared: Experiments and Car-Parrinello Simulations. The journal of physical chemistry. B 2015, 119, 14780–14789, DOI: 10.1021/acs.jpcb.5b09196.

585 586 587 588

(50) Takebayashi, Y.; Mashimo, Y.; Koike, D.; Yoda, S.; Furuya, T.; Sagisaka, M.; Otake, K.; Sakai, H.; Abe, M. Fourier Transform Infrared Spectroscopic Study of Water-inSupercritical CO2 Microemulsion as a Function of Water Content. The journal of physical chemistry. B 2008, 112, 8943–8949, DOI: 10.1021/jp802578y.

589 590 591

(51) Burikov, S.; Dolenko, S.; Dolenko, T.; Patsaeva, S.; Yuzhakov, V. Decomposition of Water Raman Stretching Band with a Combination of Optimization Methods. Mol. Phys. 2010, 108, 739–747, DOI: 10.1080/00268970903567288.

592 593 594

(52) Dolenko, T. A.; Burikov, S. A.; Patsaeva, S. V.; Yuzhakov, V. I. Manifestation of Hydrogen Bonds of Aqueous Ethanol Solutions in the Raman Scattering Spectra. Quantum Electron. 2011, 41, 267–272, DOI: 10.1070/QE2011v041n03ABEH014401.

ACS Paragon Plus Environment

29

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

Page 30 of 31

595 596 597

(53) Fukasawa, T.; Amo, Y.; Tominaga, Y. Low-Frequency Raman Scattering Study of tert Butyl Alcohol–Water and Tetrahydrofuran–Water Binary Mixtures. J. Chem. Phys. 2003, 118, 6387–6393, DOI: 10.1063/1.1559672.

598 599 600

(54) Grimaldi, N.; Rojas, P. E.; Stehle, S.; Cordoba, A.; Schweins, R.; Sala, S.; Luelsdorf, S.; Piña, D.; Veciana, J.; Faraudo, J. et al. Pressure-Responsive, Surfactant-Free CO2-Based Nanostructured Fluids. ACS nano 2017, DOI: 10.1021/acsnano.7b02500.

601 602 603

(55) Cunningham, G. P.; Vidulich, G. A.; Kay, R. L. Several Properties of Acetonitrile-Water, Acetonitrile-Methanol, and Ethylene Carbonate-Water Systems. J. Chem. Eng. Data 1967, 12, 336–337, DOI: 10.1021/je60034a013.

604 605 606 607

(56) Armitage, D. A.; Blandamer, M. J.; Foster, M. J.; Hidden, N. J.; Morcom, K. W.; Symons, M. C. R.; Wootten, M. J. Thermodynamic, Ultrasonic, Spectroscopic and Miscibility Studies of Water + Methyl Cyanide Solutions. Trans. Faraday Soc. 1968, 64, 1193, DOI: 10.1039/tf9686401193.

608 609 610

(57) Davis, M. I. Analyses of Excess Molar Volumes of the Acetonitrile-Water System using Segmented Composition Models. Thermochimica Acta 1983, 71, 59–78, DOI: 10.1016/0040-6031(83)80355-4.

611 612 613

(58) Stokes, R.H. Excess Partial Molar Enthalpies for (Acetonitrile + Water) from 278 to 318 K. The Journal of Chemical Thermodynamics 1987, 19, 977–983, DOI: 10.1016/00219614(87)90044-9.

614 615

(59) Schneider, G. Druckeinfluß auf die Entmischung flüssiger Systeme. Zeitschrift für Physikalische Chemie 1964, 41, 327–338, DOI: 10.1524/zpch.1964.41.5_6.327.

616 617 618 619

(60) Takamuku, T.; Yamaguchi, A.; Matsuo, D.; Tabata, M.; Kumamoto, M.; Nishimoto, J.; Yoshida, K.; Yamaguchi, T.; Nagao, M.; Otomo, T. et al. Large-Angle X-ray Scattering and Small-Angle Neutron Scattering Study on Phase Separation of Acetonitrile−Water Mixtures by Addition of NaCl. J. Phys. Chem. B 2001, 105, 6236–6245, DOI: 10.1021/jp003011n.

620

ACS Paragon Plus Environment

30

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

The Journal of Physical Chemistry

621 622

TOC Graphic

ACS Paragon Plus Environment

31