Validation of pH Standards and Estimation of the Activity Coefficients

Oct 23, 2018 - We selected and validated the pH values of three standard materials that function in the protic ionic liquid, ethylammonium nitrate (EA...
0 downloads 0 Views 585KB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

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

Validation of pH Standards and Estimation of the Activity Coefficients of Hydrogen and Chloride Ions in an Ionic Liquid, Ethylammonium Nitrate Ryo Kanzaki, Hikaru Daiba, Hitoshi Kodamatani, and Takashi Tomiyasu J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.8b08870 • Publication Date (Web): 23 Oct 2018 Downloaded from http://pubs.acs.org on October 30, 2018

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

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

Page 1 of 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

Validation of pH Standards and Estimation of the Activity Coefficients of Hydrogen and Chloride Ions in an Ionic Liquid, Ethylammonium Nitrate Ryo Kanzaki,* Hikaru Daiba, Hitoshi Kodamatani, Takashi Tomiyasu Department of Earth and Environmental Sciences, Graduate School of Science and Engineering, Kagoshima University, Korimoto, Kagoshima, 890-0065, Japan

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

ABSTRACT

We selected and validated the pH values of three standard materials that function in the protic ionic liquid, ethylammonium nitrate (EAN). The pH values of 0.05 mol kg‒1 phthalate, oxalate acid, and phosphate buffers were 4.93 (0.04), 2.12 (0.04), and 7.13 (0.06), respectively (the values in the parentheses denote the standard deviation). Since the pH of EAN ranges from 0 to 10, with a neutral pH of 5, these materials are usable as acidic, basic, or neutral standards. The standard electrode potential of silver-silver chloride in EAN was 127.2 (1.7) mV. The activity coefficients of hydrogen and chloride ions remain equal to unity in EAN of wide concentration range, which indicates that the effective ionic strength is independent of the solute ion concentration. In addition, the estimated value of the transfer activity coefficient of chloride ion suggests a weaker solvation in EAN compared with water in spite of a ubiquitous cation (C2H5NH3+). These behaviors of ions in EAN can be explained by the unique solvation in the ionic liquid through direct ion-ion electrostatic interactions.

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

Introduction Ionic liquids are electrolytes that have melting temperature below 100 C or around ambient temperature, and provide a special reaction field that is solely filled with ions. Its subclass of onium salt, namely protic ionic liquid (PIL),1-5 is primarily characterized by the cation’s dissociable hydrogen, which makes PILs highly applicable to acid catalyzing reactions,69

fuel cells,10,11 and so on. Since it is linked to the acid-base properties of PILs, a number of

studies concerned with the hydrogen behavior have been performed. In particular, a parameter pKa, based on the difference in the pKa values between the constituent anion and cation in water, exhibits empirically a possible relationship with the ionic character of PIL.1,12-16 However, ions in neat PIL experience serious perturbations in their acidity and basicity. In addition, the pKa values of the conjugate acids of the anions are often unreliable because they tend to behave as strong acids in water. Therefore, a current ongoing debate discusses the actual definition of pKa and what exactly it expresses.17 On the other hand, PILs made from a very weak acid were found to have a surprising ionic property, where hydrogen ions are expected to rarely dissociate according to their pKa values.18-23 From this point of view, the direct observation based on solvatochromism,24-31 NMR,32,33 titrimetric pKa determination,34-41 and so on, of the actual hydrogen ion activity is a significant target of current research. Recently, scaling of pH in ionic liquids using a commonly applicable pH scale is addressed.42-49 Analogous investigations were performed previously for a number of conventional non-aqueous solvents,50-56 however, these types of studies on ionic liquids are few and provide inadequate information. According to IUPAC, pH is a notational definition on the basis of the activity of hydrogen ion (aH), and only the practical procedure to determine pH is provided by

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

potentiometric measurements. That is, the pH value of the sample solution is determined from the potential gap of the pH indicator electrode, often using a glass electrode, from in the pHreference solution.57 For use as the pH-reference, the lists of pH values of several standard aqueous solutions57 and non-aqueous solutions of potassium hydrogen phthalate58 are provided. On the other hand, pH (or aH) can be expressed by the specific molality and the activity coefficient of hydrogen ion. These two pH values, so-called electrochemically determined and concentration-based, are confirmed to be consistent.59-62 However, appropriate pH standards have not been established yet in ionic liquids. Therefore, selection of pH standard materials, validation of their pH value, and comparison with the concentration-based value must first be achieved in ionic liquids. Ethylammonium nitrate (EAN) is the firstly described liquid salt in 1914,63 and nowadays is the most archetypical PIL. Its characteristics, such as macroscopic physicochemistry,64-72 liquid structure,73-79 dynamics,80,81 and so on, are the most widely known out of all PILs. In our previous study, we have demonstrated that EAN is capable of maintaining ordinary acid-base equilibria where the ionization profile is completely described as a function of pH using the dissociation constants.43 In addition, we have also demonstrated that the ion-sensitive field-effect transistor (ISFET) electrode acts as a pH indicator in EAN, in which a glass electrode does not work.71 Therefore, to develop the pH standard materials will promote the investigation and application of acid-base reactions in EAN. In this study, we selected pH standard materials and validated their pH values in accordance with the orthodox procedure that was applied to the conventional molecular solvents. During potentiometric measurements, we observed that the silver-silver chloride electrode acts as a stable reference electrode, and we observed unique behaviors for the activity coefficients of ions in EAN.

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

Experimental Section Reagents: EAN was synthesized following previously reported procedures43,69,71 from aqueous solutions of nitric acid and ethylamine (both 70 %; Kanto Kagaku, Japan). The water content in the synthesized EAN, determined by Karl-Fisher method, was 80-200 ppm. Since a completely neutral EAN is not possible, the excess amount of HNO3 or C2H5NH2 of the synthesized EAN was quantified by separate neutralization titrations, then the same amount of distilled C3H7NH2 or TfOH was added to neutralize. Again, the acid-base balance of this stock EAN was determined by the second neutralization titrations. The excess concentration was less than 1 mM (M = mol dm–3), which was taken into account during the subsequent measurements and analyses. Molarity and molality of EAN solutions were mutually converted using the specific density (1.212), which was determined by a pycnometer. TfOH (>99 %, provided in ampoule; Kanto Kagaku), C2H5NH3Cl (>98 %; Merck, Germany), sodium chloride (99.98 %; Kanto Kagaku), potassium hydrogen phthalate (>99.95 %; Kanto Kagaku), sodium dihydrogen phosphate (99 %; Kanto Kagaku), disodium hydrogen phosphate (>99 %; Nacalai Tesque, Japan), oxalic acid anhydrate (99 %; Merck), and disodium oxalate (99.5 %; Kanto Kagaku) were used without further purification. Potentiometric titration: Potentiometric titrations were performed at 298.15 K in a glass vessel equipped with a thermostatic water jacket. The electrolyzed silver on a platinum wire was partially oxidized to silver chloride for use as a silver-silver chloride electrode. We confirmed in separate experiments that the silver-silver chloride electrode performed as expected in aqueous solution. After titrant was added to the solution with an autoburette, equilibrium was achieved

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

within 5-10 minutes, then electromotive force (emf) was recorded. Hydrogen gas was bubbled in the sample to fill the vessel throughout the titration, and discharged into the atmosphere. The atmospheric pressure was given by an aneroid barometer, and assumed to be equivalent to the partial pressure of hydrogen, 𝑝H2, in the vessel.

Results and Discussion Standard Potential of the Silver Chloride: The silver-silver chloride electrode is one of the best choices as a reference electrode in water. Its standard electrode potential in water is determined in a hydrochloric acid solution with using a Harned cell.57,82 HCl dissociates completely in water to form equal amounts of lyonium H3O+ and solvated Cl–. Instead of HCl,82,83 we used the EAN solution containing equimolar TfOH and C2H5NH3Cl because TfOH is supposed to dissociate completely in EAN to form equal amount of HNO371 and C2H5NH3Cl is the solvate of Cl–. Consequently, this solution is considered to be equivalent to HCl solution during electrochemical measurements. The following represents the Harned cell that we used in these experiments: Pt (H2) | EAN (H+, Cl‒) | Ag/AgCl

(1)

Figure 1 shows the emf of the cell, E, while titration with TfOH + C2H5NH3Cl solution, which corresponds to 0.1 mol kg‒1 or 0.01 mol kg‒1 HCl solution of EAN. The Nernst equation of this cell is expressed as follows:

𝐸 = 𝐸AgCl°(EAN) ―

𝑅𝑇 ( ) 𝐹 ln 𝑎H𝑎Cl

𝑅𝑇

― 2𝐹ln

𝑝H2 𝑝°

(2)

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

where EAgCl(EAN) is the standard silver chloride redox potential in EAN, R = 8.314 J/molK, T = 298.15 K, F = 96485 C/mol, and the standard pressure, p, is 1013 hPa. The effect of the atmospheric pressure on E is negligible, and is hereafter ignored. Using the molalities of H+ and Cl‒ (mH and mCl, respectively), E is expressed by the following equation: 𝐸 = 𝐸AgCl°(EAN) ―

𝑚H𝑚Cl 𝑅𝑇ln 10 𝐹 log 𝑚°2



2𝑅𝑇 𝐹 ln 𝛾 ±

(3)

where  is the mean activity coefficient of H+ and Cl‒, given by 𝛾 ± = 𝛾H𝛾Cl, and m = 1 mol kg‒1. Unlike the HCl aqueous solution, mH and mCl are not strictly identical due to two experimental reasons, i.e., (1) imbalance of HNO3 and C2H5NH2 in synthesized EAN, and (2) not completely equimolar TfOH and C2H5NH3Cl in the titrant. As shown in Figure 1, E of log (mHmCl/m2) falls along a straight line with a slope of 59.73 (0.60) mV (the values in parentheses denote the standard deviation). This value, which is consistent with RT ln 10/F = 59.16 mV at 298.15 K, indicates that there is an ideal Nernst response of the Harned cell in EAN. In addition, excellent linearity suggests that the  value remains constant throughout the molality range. Therefore, E vs. log (mHmCl/m2) was reanalyzed with the ideal slope (59.16 mV), as shown by the solid line in Figure 1. Finally, we obtained EAgCl(EAN) of 127.2 (1.7) mV from the intercept. The experimental points shown in Figure 1 span a region that is significantly diluted. In particular, several data points at the right of the graph plot along the line only when autoprotolysis of EAN is taken into account in mH. One can consider that this is an infinitely dilute condition, where  should be unity. Consequently,  = 1 is retained throughout the concentration range observed. We discuss the activity coefficients for each of the respective ions later.

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry

600 500 400

2

E + RT/2F ln (pH / p°) / mV

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

300 200 100 0

0

1

2 3 4 5 2 –log(mH mCl/m° )

6

7

Figure 1. Electromotive force of the Harned cell. Neat EAN was titrated with 0.1 mol kg‒1 (filled symbols) or 0.01 mol kg‒1 (blank symbols) TfOH + C2H5NH3Cl solution of EAN. A least squares regression provides (127.2  1.7) mV ‒ 59.16 log (mHmCl/m2).

Validation of the pH of standards: According to the IUPAC recommendations,58 0.05 mol kg‒1 of potassium hydrogen phthalate was chosen as the primary pH standard for several non-aqueous solutions. We thus adopted the identical condition for the first pH standard. The emf of the Harned cell (1) was recorded in an EAN solution of 0.05 mol kg‒1 potassium hydrogen phthalate with changing the chloride ion molality by titration. The titrant solution (0.1 mol kg‒1 C2H5NH3Cl along with 0.05 mol kg‒1 potassium hydrogen phthalate) was prepared from the same solution of the titrand in order to keep the analytical concentration of potassium hydrogen phthalate constant throughout the titrations. The emf value, E, in this solution is described using the chloride molality mCl by the following equation.

ACS Paragon Plus Environment

8

Page 9 of 31

𝐸 = 𝐸AgCl°(EAN) +

𝑅𝑇ln 10 𝐹 pH



𝑚Cl 𝑅𝑇ln 10 𝐹 log 𝑚°



𝑅𝑇 𝐹 ln 𝛾Cl

(4)

As shown by the filled circles in Figure 2, E as a function of ‒log mCl/m yields a straight line with a slope of 57.63 (0.38) mV. Although the difference from the ideal Nernst response (59.16 mV) was slightly larger than the standard deviation, we could not find out a thermodynamically proper reason to interpret simultaneously the smaller slope and the linearity. Therefore, we assume the ideal Nernst response in this solution. Besides, 𝛾Cl remained constant throughout the concentration range observed. As a result, similar to the way for obtaining EAgCl(EAN), we reanalyzed Figure 2 with the ideal slope, as shown by the solid line. Finally, we obtain an intercept at log (mCl/m) = 0 of 418.9 mV, which yields a pH of 4.931 (0.043). Note that the neutral pH of EAN is around 5 according to the autoprotolysis constant of EAN.64,71 Therefore, potassium hydrogen phthalate is useful as a pH standard for neutral EAN.

800 700

E / mV

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

600 500 400 300 0

1

2

3

4

–log (mCl/m°)

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

Figure 2. Electromotive force of the Harned cell. EAN solutions of 0.05 mol kg‒1 phthalate (circles), phosphate (squares), and oxalate (triangles) buffers were titrated with a C2H5NH3Cl + 0.05 mol kg‒1 buffer solution of EAN.

We expect that a pair of NaH2PO4 and Na2HPO4 will produce a basic pH buffer according to the pKa value (pKa,2 = 7.6643). Figure 2 filled squares show E for cell (1) depending on Cl‒ molality in the EAN solution containing 0.05 mol kg‒1 of 1:1 NaH2PO4 and Na2HPO4. The cell yields a linear relationship between E and ‒log mCl/m with a slope of 56.16 mV. Following the same procedure mentioned above, we observed a pH of 7.135 (0.056). For an acidic buffer, although its pKa was not known at this stage, we selected oxalic acid. Figure 2 triangles display E for cell (1) depending on Cl‒ molality in the EAN solution containing 0.05 mol kg‒1 of 3:1 H2(ox) and Na2(ox) (ox2‒ = oxalate). The slope was 57.17, and a pH of 2.122 (0.038) was given. These pH values of three buffer solutions are summarized in Table 1. They cover the acidic, neutral, and basic region of EAN. We here compare these pH values with those based on the hydrogen ion molality. Taking into account the acid dissociation constants of phthalic acid in EAN, i.e., pKa,1 = 3.73 and pKa,2 = 5.88,43 a 0.05 mol kg‒1 potassium hydrogen phthalate solution yields an mH of 1.1910‒5 mol/kg. Since acid dissociation constants are molarity-based values, the molality and molarity of all compounds involved in this calculation are simply converted using the specific density of EAN. Finally, this yields a pH = ‒log HmH/m of 4.92, where we apply H = 1 as mentioned later. This value agrees with the pH determined from the electrochemical procedure mentioned above within the standard deviation. With regard to the phosphoric acid solution, using pKas,43 we

ACS Paragon Plus Environment

10

Page 11 of 31

found that ‒log HmH/m is equal to 7.13, which agrees with the pH determined electrochemically. With regard to oxalic acid, the acid dissociation constants in EAN were determined in this work by the separate titrations. The detailed procedure has been described elsewhere.43 Typically, 5 cm3 EAN solution of 10 mM sodium oxalate was titrated by EAN solution of 0.25 M TfOH. In Figure 3, three series of titration data are plotted. By means of least squares fitting, we obtained pKa,1 and pKa,2 of 1.96 (0.06) and 5.28 (0.02), respectively (for only pKas, values in parentheses denote three standard deviations). The calculation curve using the obtained pKa values is drawn by the solid line in Figure 3, which reproduces excellently the experimental points. Using these pKas, ‒log HmH/m is equal to 2.22. This value satisfactorily agrees with the electrochemically determined pH, while the difference is larger than the standard deviation. Overall, these pH buffer solutions associate well with the hydrogen ion concentration.

7 6

–3

–log ([H]/mol dm )

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

5 4 3 2 1

0

1 2 3 TfOH : Na2(ox)

4

Figure 3. A potentiometric titration curve of oxalic acid in EAN. The Na2(ox) solution of EAN is titrated with the TfOH solution of EAN. H means the hydrogen ion in EAN (exists as HNO3).

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

The solid line represents the theoretical curve calculated from finally obtained pKa,1 and pKa,2 values.

The activity coefficients of H+ and Cl‒ in EAN: As previously mentioned, the mean activity coefficient of H+ and Cl‒, , remains constant at unity in EAN, whereas, Cl does not depend on mCl, as is shown in Figure 2. Therefore, H = 1 and Cl = 1 are supposed in wide concentration range in EAN. This is a strange behavior because, in conventional molecular liquids, the activity coefficient of ions depends strongly on their concentration. However, it can be reasonably explained by the nature of the ionic liquids. In molecular solvents, the electrostatic potential generated by an ion decays gradually by the ionic atmosphere of the coexisting ions. On the other hand, in ionic liquids, oppositely charged ions exist ubiquitously around the ion, so that the electrostatic potential is rapidly screened and possibly disappeared at the nearest neighbor by one counterion at a minimum. In other words, the effective ionic strength in the conventional meaning is saturated in ionic liquid and remains constant in spite of the addition of ions. This situation has also been found in C2mimTf2N.84 Hence, it may be a fundamental aspect of ionic liquids. In order to consider the solvation state of H+ and Cl‒ in EAN, their transfer activity coefficients from water to EAN are estimated from emf of the following cell equipped with the double junction bridge: Pt(H2) | sample EAN solution || EANint || NaClaq | Ag/AgCl

(5)

The double bars denote liquid junctions with a 4G glass filter. NaClaq denotes an aqueous solution of 0.1 M NaCl, in which silver-silver chloride is immersed, and EANint is the electrode

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

internal solution, which prevents water from contaminating the EAN sample solution. As previously reported,43 the practical standard potential of the cell is ‒234 mV. Calculating the activity coefficient of Cl‒ in NaClaq using a simple assumption (log 𝛾 ± = ―0.5 0.1), we obtained the redox potential of AgCl to be 165 mV against the standard hydrogen electrode (SHE) in EAN (SHEEAN). If the liquid junction potential between the EAN internal solution and 0.1 M NaClaq is ignored this potential is considered at the same level as the redox potential of AgCl in water. Finally, as indicated in Figure 4, this hypothesis results in 57 mV higher SHEEAN than the SHE in water (SHEW), which yields the transfer activity coefficient of H+ from water to + EAN, log γW→EAN tr,H + , of 0.96. Therefore, HNO3 in EAN is a stronger acid than H3O in water.

Although a rough estimation, this value is excellently consistent with previous results suggesting that EAN is an acidic solvent that is 1 pH unit more acidic than water.43 For Cl‒, the redox potential of AgCl in EAN is 38 mV lower than that in water, resulting in a log γW→EAN of 0.64. tr,Cl ― ‒ The positive log γW→EAN (or γW→EAN tr,Cl ― tr,Cl ― > 1) indicates that Cl is destabilized in EAN, while the

W→EAN transfer Gibbs energy, Δ𝐺W→EAN tr,Cl ― = 𝑅𝑇ln γtr,Cl ― , is only less than 4 kJ/mol. The positive transfer

Gibbs energy was also observed in the transfer from water to most of organic solvents,55,85 however, the it is significantly smaller than the typical values (13.2 kJ/mol for methanol to 40.3 kJ/mol for dimethylsulfoxide). The relatively small Δ𝐺W→EAN value might be surprising if Cl‒ in tr,Cl ― EAN was assumed to be fully solvated by the oppositely charged ions, C2H5NH3+, through electrostatic interactions. Even though the smaller dielectric constant of EAN80 than water should be taken into account, presumably, coexisting NO3‒ predominantly destabilize Cl‒ in EAN due to the repulsive interactions. This competitive solvation contributes negatively to the net solvation Gibbs energy. Such a significant effect of like ion is a distinguishable feature in ionic liquids

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

because, in conventional solvents, like ions rarely make contact with each other and their electrostatic contribution has been almost ignored so far.

Figure 4. A potential diagram in water and EAN. Values with an asterisk denote that AgCl was immersed in water.

Conclusions In this study, we propose three pH standard buffers that function in EAN, and validated their pH values. These values cover the pH range of EAN from acidic to basic. In addition, we confirmed that these pH values are in consistent with the concentration of hydrogen ion. Hence, this set of pH standards are applicable for pH measurements in EAN solutions. Especially, it facilitates using an ISFET electrode because ISFET is a more convenient pH probe than the hydrogen

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

electrode although its standard potential is always not readily available. Since EAN is an easily available ionic liquid, the result of this study may promote the understanding and application of PILs. For thermodynamics, we observed that the activity coefficients of H+ and Cl‒ remain constant at unity in EAN. This suggests that the Debye-Hückel treatment is no longer relevant for ionic liquids. In addition, the transfer activity coefficients of single ions, H+ and Cl–, were estimated from the redox potential of silver chloride in EAN. These estimations reveal that EAN destabilizes Cl‒ to some extent in comparison with water, in spite of the ubiquitously existing C2H5NH3+. Although these trends seem strange in comparison with conventional solvents, they can be explained by the fact that ionic liquids are filled with ions. It is worth noting that there is possibly a negative contribution to the solvation energy caused by the contact of like ions, which can occur exceptionally in ionic liquids. The balance of these contributions should be recognized for considering ion-solvation in ionic liquids. Further knowledge of ion’s behavior in ionic liquids is required.

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

Page 16 of 31

Table 1. Comparison between electrochemically determined and concentration-based pH and. ‒log HmH/m

Buffer

composition (in mol kg‒1)

emf of Harned cell pH (in mV)

Oxalate

H2(ox) 0.03836

252.7 (1.5)

2.12 (0.04) 2.22

418.9 (1.9)

4.93 (0.04) 4.92

548.7 (3.3)

7.12 (0.06) 7.13

Na2(ox) 0.01257 (ox2‒ = oxalate) Phthalate

KHPh 0.05062 (Ph2‒ = phthalate)

Phosphate

NaH2PO4 0.02476 Na2HPO4 0.02477

values in parentheses denote the standard deviation.

AUTHOR INFORMATION Corresponding Author *Ryo KANZAKI, [email protected], +81-99-285-8106 Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding Sources ACKNOWLEDGMENT This work was supported by JSPS KAKENHI Grant Numbers JP26410157 and JP18H03392

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

REFERENCES 1. Yoshizawa, M.; Xu, W.; Angell, C. A. Ionic Liquids by Proton Transfer: Vapor Pressure, Conductivity, and the Relevance of pKa from Aqueous Solutions J. Am. Chem. Soc. 2003, 125, 15411-15419. 2. Angell, C. A.; Ansari, Y.; Zhao, Z. Ionic Liquids: Past, Present and Future Faraday Discuss. 2012, 154, 9-27. 3. Mihichuk, L. M.; Driver, G. W.; Johnson, K. E. Brønsted Acidity and the Medium: Fundamentals with a Focus on Ionic Liquids Chem. Phys. Chem. 2011, 12, 1622-1632. 4. Belieres, J.-P.; Angell, C. A. Protic Ionic Liquids: Preparation, Characterization, and Proton Free Energy Level Representation J. Phys. Chem. B 2007, 111, 4926-4937. 5. Greaves, T. L.; Drummond, C. J. Protic Ionic Liquids: Evolving Structure–Property Relationships and Expanding Applications Chemical Reviews 2015, 115, 11379-11448. 6. Olivier-Bourbigou, H.; Magna, L. Ionic Liquids: Perspectives for Organic and Catalytic Reactions J. Mol. Catal. A Chem. 2012, 182, 419-437. 7. Skoda-Foldes, R. The Use of Supported Acidic Ionic Liquids in Organic Synthesis Molecules 2014, 19, 8840-8884. 8. Amarasekara, A. S. Acidic Ionic Liquids Chem. Rev. 2016, 116, 6133-6183. 9. Hajipour, A. R.; Rafiee, F. Acidic Brønsted Ionic Liquids Org. Prep. Proc. Int. 2010, 42, 285-362.

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

Page 18 of 31

10. Lee, S.-Y.; Ogawa, A.; Kanno, M.; Nakamoto, H.; Yasuda, T.; Watanabe, M. Nonhumidified Intermediate Temperature Fuel Cells Using Protic Ionic Liquids J. Am. Chem. Soc. 2010, 132, 9764-9773. 11. Miran, M. S.; Yasuda, T.; Susan, M. A. B. H.; Dokko, K.; Watanabe, M. Binary Protic Ionic Liquid Mixtures as a Proton Conductor: High Fuel Cell Reaction Activity and Facile Proton Transport J. Phys. Chem. C 2014, 118, 27631-27639. 12. Miran, M. S.; Yasuda, T.; Susan, M. A. B. H.; Dokko, K.; Watanabe, M. Physicochemical Properties Determined by pKa for Protic Ionic Liquids based on an Organic Super-strong Base with Various Brønsted Acids Phys. Chem. Chem. Phys. 2012, 14, 5178-5186. 13. Stoimenovski, J.; Izgorodina, E. I.; MacFarlane, D. R. Ionicity and Proton Transfer in Protic Ionic Liquids Phys. Chem. Chem. Phys. 0341, 12, 1. 14. Angell, C. A.; Byrne, N.; Belieres, J.-P. Parallel Developments in Aprotic and Protic Ionic Liquids: Physical Chemistry and Applications Accounts. Chem. Res. 2007, 40, 1228-1236. 15. Chen, K.; Wang, Y.; Yao, J.; Li, H. Equilibrium in Protic Ionic Liquids: The Degree of Proton Transfer and Thermodynamic Properties J. Phys. Chem. B 2018, 122, 309-315. 16. Shen, M.; Zhang, Y.; Chen, K.; Che, S.; Yao, J.; Li, H. Ionicity of Protic Ionic Liquid: Quantitative Measurement by Spectroscopic Methods J. Phys. Chem. B 2017, 121, 13721376.

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

17. Reid, J. E. S. J.; Filipe, A.; Bernardes, C. E. S.; Martins, F.; Walker, A. J.; Shimizu, S.; Piedade, M. E. M. Structure-Property Relationships in Protic Ionic Liquids: A Thermochemical Study Phys. Chem. Chem. Phys. 2017, 19, 19928-19936. 18. Doi, H.; Song, S.; Minofar, B.; Kanzaki, R.; Takamuku, T.; Umebayashi, Y. A New Proton Conductive Liquid with No Ions: Pseudo-Protic Ionic Liquids Chem. Eur. J. 2013, 19, 11522-11526. 19. Watanabe, H.; Doi, H.; Saito, S.; Sadakane, K.; Fujii, K.; Kanzaki, R.; Kameda, Y.; Umebayashi, Y. Raman Spectroscopic Speciation Analyses and Liquid Structures by HighEnergy X-ray Total Scattering and Molecular Dynamics Simulations for Nmethylimidazolium-Based Protic Ionic Liquids Bull. Chem. Soc. Jpn. 2016, 89, 965-972. 20. Watanabe, H.; Doi, H.; Saito, S.; Matsugami, M.; Fujii, K.; Kanzaki, R.; Kameda, Y.; Umebayashi, Y. Hydrogen bond in Imidazolium Based Protic and Aprotic Ionic Liquids J. Mol. Liquids 2016, 217, 35-42. 21. Orzechowski, K.; Pajdowska, M.; Czarnecki, M.; Kaatze, U. Complexation and Proton Transfer in the Binary System Propionic Acid-Triethylamine Evidence from the Composition Dependencies of Mixture Properties J. Mol. Liquid 2007, 133, 11-16. 22. Kanzaki, R.; Doi, H.; Song, X.; Hara, S.; Ishiguro, S.; Umebayashi, Y. Acid-Base Property of N-Methylimidazolium-Based Protic Ionic Liquids Depending on Anion J. Phys. Chem. B 2012, 116, 14146-14152. 23. Vitorino, J.; Bernardes, C. E. S.; Piedade, M. E. M. A General Strategy for the Experimental Study of the Thermochemistry of Protic Ionic Liquids: Enthalpy of Formation and

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

Vaporization of 1-Methylimidazolium Ethanoate Phys. Chem. Chem. Phys. 2012, 14, 44404446. 24. Thomazeau, C.; Olivier-Bourbigou, H.; Magna, L.; Luts, S.; Gilbert, B. Determination of an Acidic Scale in Room Temperature Ionic Liquids J. Am. Chem. Soc. 2003, 125, 5264-5265. 25. Robert, T.; Magna, L.; Olivier-Bourbigou, H.; Gilbert, B. A Comparison of the Acidity Levels in Room-Temperature Ionic Liquids J. Electrochem. Soc. 2009, 156, F115-F121. 26. Shukla, S. K.; Kumar, A. Do Protic Ionic Liquids and Water Display Analogous Behavior in terms of Hammett Acidity Function? Chem. Phys. Lett. 2013, 566, 12-16. 27. Reid, J. E. S. J.; Bernardes, C. E. S.; Agapito, F.; Martins, F.; Shimizu, S.; Piedade, M. E. M.; Walker, A. J. Structure-Property Relationships in Protic Ionic Liquids: A Study of Solvent–Solvent and Solvent–Solute Interactions Phys. Chem. Chem. Phys. 2017, 19, 2813328138. 28. Valle, J. C.; Blanco, F. G.; Catalan, J. Empirical Parameters for Solvent Acidity, Basicity, Dipolarity, and Polarizability of the Ionic Liquids [BMIM][BF4] and [BMIM][PF6] J. Phys. Chem. B 2015, 119, 4683-4692. 29. Horowitz, A. I.; Arias, P.; Panzer, M. J. Spectroscopic Determination of Relative Bronsted Acidity as a Predictor of Reactivity in Aprotic Ionic Liquids Chem. Commun. 2015, 51, 6651-6654.

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

30. Schmode, S.; Ludwig, R. Utilization of the Dye N-methyl-6-oxyquinolone as an Optical Acidometer in Molecular Solvents and Protic Ionic Liquids Chem. Commun. 2017, 53, 10761-10764. 31. Hasani, M.; Yarger, J. L.; Angell, C. A. On the Use of a Protic Ionic Liquid with a Novel Cation To Study Anion Basicity Chem. Eur. J. 2016, 22, 13312-13319. 32. Zheng, A.; Liu, S.-B., Deng, F. P-31 NMR Chemical Shifts of Phosphorus Probes as Reliable and Practical Acidity Scales for Solid and Liquid Catalysts Chem. Rev. 2017, 117, 1247512531. 33. Zhang, S.; Zhang, T.; Tang, S. Determination of the Hammett Acidity Functions of Triflic Acid/Ionic Liquid Binary Mixtures by the 13C NMR-Probe Method J. Chem. Eng. Data 2016, 61, 2088-2097. 34. Millan, D.; Rojas, M.; Santos, J. G.; Morales, J.; Isaacs, M.; Diaz, C.; Pavez, P. Toward a pKa Scale of N-base Amines in Ionic Liquids J. Phys. Chem. B 2014, 118, 4412-4418. 35. Silvester, D. S.; Aldous, L.; Hardacre, C.; Compton, R. G. An Electrochemical Study of the Oxidation of Hydrogen at Platinum Electrodes in Several Room Temperature Ionic Liquids J. Phys. Chem. B 2007, 111, 5000-5007. 36. D’Anna, F.; Noto, R. Amine Basicity: Measurements of Ion Pair Stability in Ionic Liquid Media Tetrahedron 2007, 63, 11681-11685. 37. D’Anna, F.; Marca, S. L.; Noto, R. p-Nitrophenolate: a Probe for Determining Acid Strength in Ionic Liquids, J. Org. Chem. 2009, 74, 1952-1956

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

38. Barhdadi, R.; Troupel, M.; Comminges, C.; Laurent, M.; Doherty, A. Electrochemical Determination of pK(a) of N-Bases in Ionic Liquid Media J. Phys. Chem. B 2012, 116, 277282. 39. Hashimoto, K.; Fujii, K.; Shibayama, M. Acid-base Property of Protic Ionic Liquid, 1Alkylimidazolium Bis(trifluoromethanesulfonyl)amide Studied by Potentiometric Titration J. Mol. Liquids 2013, 188, 143-147. 40. Fujii, K.; Hashimoto, K.; Sakai, T.; Umebayashi, Y.; Shibayama, M. Brønsted Basicity of Solute Butylamine in an Aprotic Ionic Liquid Investigated by Potentiometric Titration Chem. Lett. 2013, 42, 1250-1251. 41. Bautista-Martinez, J. A.; Tang, L.; Belieres, J.-P.; Zeller, R.; Angell, C. A.; Friesen, C. Hydrogen Redox in Protic Ionic Liquids and a Direct Measurement of Proton Thermodynamics J. Phys. Chem. C 2009, 113, 12586-12593. 42. Himmel, D.; Goll, S. K.; Scholz, F.; Radtke, V.; Leito, I.; Krossing, I. Absolute Brønsted Acidities and pH Scales in Ionic Liquids Chem. Phys. Chem. 2015, 16, 1428-1439. 43. Kanzaki, R.; Kodamatani, H.; Tomiyasu, T.; Watanabe, H.; Umebayashi, Y. A pH Scale for the Protic Ionic Liquid Ethylammonium Nitrate Angew. Chem. Int. Ed. 2016, 55, 6266-6269. 44. Mao, C.; Wang, Z.; Wang, Z.; Ji, P.; Cheng, J.-P. Weakly Polar Aprotic Ionic Liquids Acting as Strong Dissociating Solvent: A Typical “Ionic Liquid Effect” Revealed by Accurate Measurement of Absolute pKa of Ylide Precursor Salts J. Am. Chem. Soc. 2016, 138, 55235526.

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

45. Wang, Z.; Li, X.; Ji, P.; Cheng, J.-P. Absolute pK(a)s of Sulfonamides in Ionic Liquids: Comparisons to Molecular Solvents J. Org. Chem. 2016, 81, 11195-11200. 46. Deng, H.; Li, X.; Chu, Y.; He, J. Q.; Cheng, J.-P. Standard pKa Scales of Carbon-Centered Indicator Acids in Ionic Liquids: Effect of Media and Structural Implication J. Org. Chem. 2012, 77, 7291-7298. 47. Wang, Z.; Ji, P.-J.; Li, X.; Cheng, J.-P. Double-Line Hammett Relationship Revealed through Precise Acidity Measurement of Benzenethiols in Neat Ionic Media: A Typical “Ionic Liquid Effect”? Org. Lett. 2014, 16, 5744-5747. 48. Wang, Z.; Deng, H.; Li, X.; Ji, P.-J.; Cheng, J.-P. Standard and Absolute pKa Scales of Substituted Benzoic Acids in Room Temperature Ionic Liquids J. Org. Chem. 2013, 78, 12487-12493. 49. Mao, C.; Wang, Z.; Ji, P.-J.; Cheng, J. P. Is Amine a Stronger Base in Ionic Liquid Than in Common Molecular Solvent? An Accurate Basicity Scale of Amines J. Org. Chem. 2015, 80, 8384-8389. 50. Himmel, D.; Goll, S. K.; Leito, I.; Krossing, I. A Unified pH Scale for All Phases Angew. Chem. Int. Ed. 2010, 49, 6885-6888. 51. Himmel, D.; Goll, S. K.; Leito, I.; Krossing, I. Anchor Points for the Unified Bronsted Acidity Scale: The rCCC Model for the Calculation of Standard Gibbs Energies of Proton Solvation in Eleven Representative Liquid Media Chem. Eur. J. 2011, 17, 5808-5826.

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

52. Radtke, V.; Himmel, D.; Putz, K.; Goll, S. K.; Krossing, I. The Protoelectric Potential Map (PPM): An Absolute Two-Dimensional Chemical Potential Scale for a Global Understanding of Chemistry Chem. Eur. J. 2014, 20, 4194-4211. 53. Himmel, D.; Radtke, V.; Butschke, B.; Krossing, I. Basic Remarks on Acidity Angew. Chem. Int. Ed. 2018, 57, 4386-4411. 54. The Protoelectric Potential Map, University of Freiburg, https://www.ppm.unifreiburg.de/pH_abs/DeltaG_trans_solv_proton (accessed October 18, 2018) 55. Marcus, Y. Thermodynamic Functions of Transfer of Single Ions from Water to Nonaqueous and Mixed Solvents: Part I - Gibbs Free Energies of Transfer to Nonaqueous Solvents Pure Appl. Chem. 1983, 55, 977-1021. 56. Izutsu, K. Electrochemistry in Nonaqueous Solutions; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2002; pp 78-82. 57. Buck, R. P.; Rondinini, S.; Covington, A. K.; Baucke, F. G. K.; Brett, C. M. A.; Camoes, M. F.; Milton, M. J. T.; Mussini, T.; Naumann, R.; Pratt, K. W.; Spitzer, P.; Wilson, G. S. Measurement of pH Definition, Standards, and Procedures (IUPAC Recommendations 2002) Pure Appl. Chem. 2002, 74, 2169-2200. 58. Rondinini, S.; Mussini, P. R.; Mussini, T. Reference Value Standards and Primary Standards for pH Measurements in Organic Solvents and Water + Organic Solvent Mixtures of Moderate to High Permittivities Pure Appl. Chem. 1987, 59, 1549-1560.

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

The Journal of Physical Chemistry

59. Camoes, M. F.; Lito, M. J. G.; Ferra, M. I. A.; Covington, A. K. Consistency of pH Standard Values with the Corresponding Thermodynamic Acid Dissociation Constants (Technical Report) Pure Appl. Chem. 1997, 69, 1325-1333. 60. Chan, C.-Y.; Eng, Y.-W.; Eu, K.-S. Pitzer Single-Ion Activity Coefficients and pH for Aqueous Solutions of Potassium Hydrogen Phthalate in Mixtures with KCl and with NaCl at 298.15 K J. Chem. Eng. Data 1995, 40, 685-691. 61. Partanen, J. I.; Covington, A. K. Re-evaluation of the First and Second Stoichiometric Dissociation Constants of Phthalic Acid at Temperatures from (0 to 60) °C in Aqueous Phthalate Buffer Solutions with or without Potassium Chloride. 1. Estimation of the Parameters for the Hückel Model Activity Coefficient Equations for Calculation of the Second Dissociation Constant J. Chem. Eng. Data 2006, 51, 777-784. 62. Partanen, J. I.; Covington, A. K. Re-evaluation of the First and Second Stoichiometric Dissociation Constants of Phthalic Acid at Temperatures from (0 to 60) °C in Aqueous Phthalate Buffer Solutions with or without Potassium Chloride. 2. Estimation of Parameters for the Model for the First Dissociation Constant and Tests and Use of the Resulting Activity Coefficient Equations J. Chem. Eng. Data 2006, 51, 2065-2073. 63. Walden, V. P., Ueber die Molelulargrösse und Elektrische Leitfähigkeit Einiger Geschmolzenen Salze Bull. Acad. Imp. Sci. (St. Petersburg) 1914, 8, 405-422. 64. Benlhima, N.; Lemordant, D.; Letellier P. Propriétés Physicochimiques des Mélanges EauNitrate D’éthylammonium Fondu, à 298 K Échelles D’acidité-Solubilités J. Chim. Phys. 1989, 86, 1919-1939

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

65. Evans, D. F.; Chen, S.-H. Thermodynamics of Solution of Nonpolar Gases in a Fused Salt. “Hydrophobic Bonding” Behavior in a Nonaqueous System J. Am. Chem. Soc. 1981, 103, 481-482. 66. Perron, G.; Hardy, A.; Justice, J.-C.; Desnoyers, J. E. Model System for Concentrated Electrolyte Solutions: Thermodynamic and Transport Properties of Ethylammonium Nitrate in Acetonitrile and in Water J. Solution Chem. 1993, 22, 1159-1177. 67. Mirejovsky, D.; Arnett, E. M. Heat Capacities of Solution for Alcohols in Polar Solvents and the New View of Hydrophobic Effects J. Am. Chem. Soc. 1983, 105, 1112-1117. 68. Arancibia, E. L.; Castells, R. C.; Nardillo, A. M. Thermodynamic Study of the Behavior of Two Molten Organic Salts as Stationary Phases in Gas Ghromatography J. Chromatogr. 1987, 298, 21-29. 69. Kanzaki, R.; Song, X.; Umebayashi, Y.; Ishiguro, S. Thermodynamic Study of the Solvation States of Acid and Base in a Protic Ionic Liquid, Ethylammonium Nitrate, and Its Aqueous Mixtures Chem. Lett. 2010, 39, 578-579. 70. Kanzaki, R.; Uchida, K.; Song, X.; Umebayashi, Y.; Ishiguro, S. Acidity and Basicity of Aqueous Mixtures of a Protic Ionic Liquid, Ethylammonium Nitrate Anal. Sci. 2008, 24, 1347-1349. 71. Kanzaki, R.; Uchida, K.; Hara, S.; Umebayashi, Y.; Ishiguro, S.; Nomura, S. Acid-Base Property of Ethylammonium Nitrate Ionic Liquid Directly Obtained Using Ion-selective Field Effect Transistor Electrode Chem. Lett. 2007, 36, 684-685.

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

72. Amman, M.; Caprio, D. D.; Gaillon, L. Interfacial Properties of Mercury/Ethylammonium Nitrate Ionic Liquid + Water System: Electrocapillarity, Surface Charge and Differential Capacitance Electrochim. Acta 2012, 61, 207-215. 73. Hayes, R.; Imberti, S.; Warr, G. G.; Atkin, R. Amphiphilicity Determines Nanostructure in Protic Ionic Liquids Phys. Chem. Chem. Phys. 2011, 13, 3237-3247. 74. Greaves, T. L.; Kennedy, D. F.; Mudie, S. T.; Drummond, C. J. Diversity Observed in the Nanostructure of Protic Ionic Liquids J. Phys. Chem. B 2010, 114, 10022-10031. 75. Umebayashi, Y.; Chung, W.-L.; Mitsugi, T.; Fukuda, S.; Takeuchi, M.; Fujii, K.; Takamuku, T.; Kanzaki, R.; Ishiguro, S. Liquid Structure and the Ion-Ion Interactions of Ethylammonium Nitrate Ionic Liquid Studied by Large Angle X-Ray Scattering and Molecular Dynamics Simulations J. Comp. Chem. Jpn. 2008, 7, 125-134. 76. Song, X.; Hamano, H.; Minofar, B.; Kanzaki, R.; Fujii, K.; Kameda, Y.; Kohara, S.; Watanabe, M.; Ishiguro, S.; Umebayashi, Y. Structural Heterogeneity and Unique Distorted Hydrogen Bonding in Primary Ammonium Nitrate Ionic Liquids Studied by High-Energy Xray Diffraction Experiments and MD Simulations J. Phys. Chem. B 2012, 116, 2801-2803. 77. Atkin, R.; Warr, G. G. The Smallest Amphiphiles:  Nanostructure in Protic RoomTemperature Ionic Liquids with Short Alkyl Groups J. Phys. Chem. B 2008, 112, 4164-4166. 78. Fumino, K.; Wulf, A.; Ludwig, R. Hydrogen Bonding in Protic Ionic Liquids: Reminiscent of Water Angew. Chem. Int. Ed. 2009, 48, 3184-3186.

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

79. Hayes, R.; Silvia, I.; Warr, G. G.; Atkin, R. The Nature of Hydrogen Bonding in Protic Ionic Liquids Angew. Chem. Int. Ed. 2013, 52, 4623-4627. 80. Weingartner, H.; Knocks, A.; Schrader, W.; Kaatze, U. Dielectric Spectroscopy of the Room Temperature Molten Salt Ethylammonium Nitrate J. Phys. Chem. A 2001, 105, 8646-8650. 81. Halder, M.; Headley, L. S.; Mukherjee, P.; Song, X.; Petrich, J.W. Experimental and Theoretical Investigations of Solvation Dynamics of Ionic Fluids: Appropriateness of Dielectric Theory and the Role of DC Conductivity J. Phys. Chem. A 2006, 110, 8623-8626. 82. Harned, H. S.; Ehlers, R. W. The Thermodynamics of Aqueous Hydrochloric Acid Solutions from Electromotive Force Measurements J. Am. Chem. Soc. 1933, 55, 2179-2193. 83. Bates, R. G.; Robinson, R. A. Standardization of silver-silver chloride electrodes from 0 to 60°C J. Solution Chem. 1980, 9, 455-456. 84. Matsubara, Y., Grills, D. C.; Koide, Y. Experimental Insight into the Thermodynamics of the Dissolution of Electrolytes in Room-Temperature Ionic Liquids: From the Mass Action Law to the Absolute Standard Chemical Potential of a Proton ACS Omega 2016, 1, 1393-1411. 85. Gritzner, G. Single-ion Transfer Properties: A Measure of Ion-Solvation in Solvents and Solvent Mixtures Electrochim. Acta 1998, 44, 73-83.

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

SYNOPSIS (Table of Contents Image)

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

Page 30 of 31

Figure 4

AgCl 222

234*

AgCl in 0.1 M NaClaq

165*

AgCl

38 mV 127 AgCl

0 57 mV

in EAN

SHEW 0 in water

ACS Paragon Plus Environment

SHEEAN

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

The Journal of Physical Chemistry

TOC Graphic

AgCl 222

38 mV

57 mV SHEW

0 in water

234*

AgCl in 0.1 M NaClaq

165*

AgCl 127 AgCl

0 in EAN

Cl‒ in water

SHEEAN

ACS Paragon Plus Environment

Cl‒ in EAN