Volumetric Properties of Binary Mixtures of 1-Butyl-1

Mar 19, 2014 - Pannuru Kiran Kumar , Varadhi Govinda , Karlapudi Sreenuvasulu , Pannuru Venkatesu , Indra Bahadur , Eno E. Ebenso. Journal of Molecula...
0 downloads 0 Views 438KB Size
Article pubs.acs.org/jced

Volumetric Properties of Binary Mixtures of 1‑Butyl-1Methylpyrrolidinium Tris(pentafluoroethyl)trifluorophosphate with N‑Methylformamide, N‑Ethylformamide, N,N‑Dimethylformamide, N,N‑Dibutylformamide, and N,N‑Dimethylacetamide from (293.15 to 323.15) K Slobodan Gadžurić, Aleksandar Tot, Nebojša Zec, Snežana Papović, and Milan Vraneš* Faculty of Science, Department of Chemistry, Biochemistry and Environmental Protection, University of Novi Sad, Trg Dositeja Obradovića 3, 21000 Novi Sad, Serbia S Supporting Information *

ABSTRACT: Experimental densities of binary mixtures of ionic liquid 1-butyl-1methylpyrrolidinium tris(pentafluoroethyl)trifluorophosphate ([BMPYR][FAP]) with two protic amidesN-methylformamide (NMF) and N-ethylformamide (NEF)and three aproticN,N-dimethylformamide (DMF), N,N-dibutylformamide (DBF), and N,N-dimethylacetamide (DMA)are reported in the temperature range from (293.15 to 323.15) K and at atmospheric pressure (0.1 MPa) over the whole composition range. This ionic liquid was selected due to its hydrophobic properties and hydrolytic stability. The excess molar volumes derived from the experimental densities appear to be positive for all of the studied mixtures in the whole composition range at all temperatures. The systems with aprotic amides (DMF, DBF, DMA) show higher values of excess molar volume compared to those with protic amides (NMF, NEF), while the number of the methyl groups at the carbon atom does not affect the composition corresponding to the maximum of the excess molar volume.



INTRODUCTION It is well-known that ionic liquids are the salts with melting points below 373.15 K.1 They consist of large, asymmetric organic cations and organic or inorganic anions,2 and it is the largest class of the compounds in chemistry. Because of its physical and chemical properties, such as negligible vapor pressure,3 nonflammability,4 high ionic conductivity,5 thermal and electrochemical stability,6,7 and the ability of dissolution and extraction of inorganic and organic components,8 ionic liquids nowadays are gaining attention of the scientific community. Due to these features together with tunable properties of ionic liquids9 they are considered to be perfect “designed” and “green” solvents.10 Understanding the physical and chemical properties of ionic liquids is essential for the enormous potential application of ionic liquids in organic synthesis,11,12 catalysis,13 electrochemical,14 and separation processes.15,16 Ionic liquids with a highly fluorinated anion as tris(pentafluoroethyl)trifluorophosphate, [FAP], have properties that are attracting growing attention of many researchers.17−19 The synthesis of ionic liquids containing the [FAP] anion was described by Ignat’ev et al.17 These ionic liquids are highly hydrophobic2,18 and have a high heat capacity,20 a wide electrochemical window,21 high thermal stability,22 and very low vapor pressure.23 They are ideal candidates for direct © 2014 American Chemical Society

immersion extraction from an aqueous matrix due to their hydrophobic properties and hydrolytic stability.18 Also, [FAP] based ionic liquids were tested as a lubricant additives for different coatings,24 synthesis of polymers25 by electrowetting,26 and electropolymerization.27 Duffy and Bond28 used them as supporting electrolyte for the electrochemical techniques if the solvents are aromatic compounds. It is well-known that hydrophobic ionic liquids, including those with the [FAP] anion, encourage self-association of the amphiphiles including lower amides.29,30 Lower amides such as NMF and NEF are the nonaqueous organic solvents with the capability of the self-association,31 unlike of DMF, DBF, and DMA. Thus, the volumetric properties of [BMPYR][FAP] binary mixtures with two protic (NMF and NEF) and three aprotic (DMF, DBF, and DMA) amides were examined and compared in the temperature range from (293.15 to 323.15) K and at atmospheric pressure.



EXPERIMENTAL SECTION Chemicals. The purity of purchased chemicals is listed in Table 1. Prior to their use, all amides were distilled, and Received: September 6, 2013 Accepted: March 10, 2014 Published: March 19, 2014 1225

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

Table 2. Experimental and Literature Values of Densities, d, of Pure Liquids at the Specified Temperatures and at Atmospheric Pressurea

Table 1. Provenance and Purity of the Samples chemical name 1-butyl-1-methylpyrrolidinium tris(pentafluoroethyl)trifluorophosphate N-methylformamide N,N-dimethylformamide N-ethylformamide N,N-dibutylformamide N,N-dimethylacetamide

provenance

mass fraction purity

Merck

≥ 0.99

Sigma Aldrich J.T. Baker Sigma Aldrich Sigma Aldrich Merck

≥ 0.99

d/(g·cm−3) component [BMPYR] [FAP]

≥ 0.998 ≥ 0.99 ≥ 0.99 ≥ 0.99

NMF

collected middle fractions were dried for 15 h under vacuum. Dried substances were stored in the sealed dark bottles over the 4 Å molecular sieves. Traces of water in commercial [BMPYR][FAP] ionic liquid were removed by drying under reduced pressure for several days. The amount of water was checked by Karl Fisher titration using a 831 Karl Fischer coulometer, and it was found to be less than 1·10−4 mass fraction. Binary mixtures containing [BMPYR][FAP] and selected amides were prepared by measuring an appropriate amount of the components on a Denver analytical balance. The combined experimental uncertainty (k = 2) of mass fraction was less than 5·10−5. Binary mixtures covering the whole composition range were prepared. Density. For measuring density, a Rudolph Research Analytical DDM 2911 vibrating tube densimeter was used. The densimeter is equipped with devices for automatic viscosity correction and Peltier-type maintenance of temperature within ± 0.01 K. The instrument was calibrated at the atmospheric pressure using ambient air and bidistilled ultra pure water in the temperature range (293.15 to 323.15) K before each series of measurements. The repeatability of the densimeter was ± 0.00001 g·cm−3. Each experimental density value used in our further calculations is the mean value of at least three measurements at the selected temperature. The experimental measurements showed repeatability less than 0.01 %, and an average value was shown in this paper. The combined experimental uncertainty (k = 2) of determining the density is less than 2.0·10−5 g·cm−3.

DMF

NEF

DBF



RESULTS AND DISCUSSION Densities of pure components and {[BMPYR][FAP] (1) + amides (2)} mixtures were measured as a function of temperature in the temperature range from (293.15 to 323.15) K, and the results are shown in Table 2 and Table S1 in the Supporting Information of this manuscript. As expected, the density of the mixtures decreases with the increasing temperature and increases with the mole fraction of ionic liquid. Obtained experimental density data were fitted as a function of temperature using a linear fit. These parameters are given in Table S2 and graphically represented in Figure S1 in the Supporting Information. From the obtained experimental density of the mixture, d, densities of the pure components, di, the corresponding mole fractions, xi, and molar masses, Mi, the excess molar volume, VE, was calculated using the following equation: ⎛1 ⎛1 1⎞ 1⎞ V E = x1M1⎜ − ⎟ + x 2M 2⎜ − ⎟ d1 ⎠ d2 ⎠ ⎝d ⎝d

DMA

T/K

this work

293.15 298.15 303.15 308.15 313.15 318.15 323.15 293.15 298.15 303.15 308.15 313.15 318.15 323.15 293.15

1.58856 1.58315 1.57767 1.57216 1.56662 1.56106 1.55548 1.00335 0.99891 0.99449 0.99001 0.98547 0.98088 0.97624 0.94872

298.15

0.94386

303.15

0.93900

308.15

0.93408

313.15

0.92913

318.15 323.15 293.15 298.15 303.15 308.15 313.15 318.15 323.15 293.15 298.15 303.15 308.15 313.15 318.15 323.15 293.15 298.15

0.92406 0.91896 0.95159 0.94748 0.94329 0.93904 0.93478 0.93045 0.92606 0.87905 0.87512 0.87117 0.86715 0.86311 0.85901 0.85482 0.94092 0.93630

303.15 308.15 313.15 318.15 323.15

0.93162 0.92689 0.92209 0.91723 0.91223

reference 1.5886;32 1.6037 (293.19 K)33 1.5832;32 1.58261;34 1.5826335 1.5928 (303.18 K)33 1.5720634 1.5819 (313.20 K)33 1.5615734 1.5710 (323.14 K)33 0.99858;36 0.9988;37 0.99929;38 0.9988939 0.994840 0.990540 0.986140 0.981840 0.9491;41 0.948051;42 0.9491;43 0.94858;44 0.94865;45 0.948978;46 0.94868;47 0.94406;38 0.94383;39 0.9446;41 0.942915;42 0.9449;43 0.94389;45 0.944061;46 0.9445;48 0.94457;49 0.9438550 0.9401;41 0.938876;42 0.9402;43 0.93946;44 0.93910;45 0.939376;46 0.9398;48 0.93793;49 0.939451 0.9357;41 0.933964;42 0.9355;43 0.93433;45 0.934402;46 0.9351;48 0.934251 0.9312;41 0.929549;42 0.9307;43 0.93012;44 0.92952;45 0.929860;46 0.929435;47 0.9302;48 0.92653;49 0.930151 0.9267;41 0.924674;42 0.9259;43 0.925351 0.92055;44 0.920151 0.955252 0.944753 0.936453

0.87554

0.94155;55 0.93982456 0.93634;55 0.9364;57 0.93654;58 0.93617;59 0.93615;60 0.936382;61 0.93633762 0.9363163 0.9316655 0.92704;55 0.926297;56 0.92714;61 0.9270863 0.917216;56 0.917863;61 0.9178263 0.91312964

a Standard uncertainties are: u(d) = 2·10−5 g·cm−3, u(x) = 5·10−5, and u(T) = 0.01 K.

where x1, M1, and d1 relate to [BMPYR][FAP], whereas x2, M2, and d2 relate to amides. Calculated values of the excess molar volumes are given in Table S1 and graphically presented in Figure 1 and Figures S2− S6 in the Supporting Information, using a Redlich−Kister-type equation65 where VE represents the excess property, Ai refers to

(1) 1226

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

cm3·mol−1 and nm3 are reported in Table S4 as the Supporting Information. According to Glaser,66 the standard entropy can be calculated on the basis of the calculated molar volumes of the components at 298.15 K using a simple expression: Smo(298.15 K)/(J ·K−1·mol−1) = 1246.5·(Vm/nm 3) + 29.5 (5)

The excess standard entropy at 298.15 K can be calculated using the following expression: S E = Smo − [x1S1o + x 2S2o]

(6)

Soi

where xi and are the mole fraction and standard entropy of the components, respectively. Som is the standard entropy of binary mixtures calculated using eq 5. The values of the excess standard entropy at 298.15 K are given in Table S5 in the Supporting Information. The standard entropy of mixture increases as the function of the ionic liquids mole fraction. The excess standard entropy values for all investigated systems are positive, with the maximum value obtained in the 0.4 ≤ xIL ≤ 0.6 interval. These results are graphically presented in Figure 2, and their agreement with the values obtained for the excess molar volume can be observed.

Figure 1. Excess molar volumes, VE, for ([BMPYR][FAP] + amide) binary mixtures as a function of the [BMPYR][FAP] mole fraction composition, x1, at the temperature T = 298.15 K: ■, NMF; ●, DMF; △, NEF; ▽, DBF; ◀, DMA.The lines represent the Redlich−Kistertype fittings with the parameters indicated in Table S3.

the adjustable parameters, and n is the number of the coefficients in the equation: n

V E = x1x 2 ∑ Ai (2x1 − 1)i i=0

(2)

The standard deviation of the fit can be calculated as: m E E 2 σV E = [∑ (Vexp − Vcalc ) /(m − n)]1/2 1

(3)

where m denotes the number of experimental points. The coefficients Ai and the standard deviations of the fit are tabulated in Table S3 in the Supporting Information. For all examined systems excess molar volumes have positive values in the whole range of concentrations, with maximum values between 0.4 and 0.6 of ionic liquid mole fraction. Positive VE values indicate that the interactions within the pure components are much stronger than the interactions established between the ions and amide after their mixing. It can be seen from Figure 1 that systems with aprotic amides (DMF, DBF, DMA) have higher values of VE compared to systems with protic amides (NMF, NEF). This can be explained on the basis of the steric hindrance of amide carbonyl groups, due to a weak ion−dipole interactions between [BMPYR]+ and the partially negative charge on the oxygen atom of the carbonyl group. By increasing the number of substituted methylene groups on the nitrogen atom, VE maximum will be shifted toward higher ionic liquids mole ratio in the case of both protic (NMF, NEF) and aprotic amides (DMF, DBF). Comparing the values obtained for the mixtures containing DMA and DMF, it may be observed that number of the methyl groups at carbon atom does not change the position of VE maximum. Another volumetric property, molar volume, Vm, can be calculated using the following equation: Vm =

(x1M1 + x 2M 2) Nad

Figure 2. Excess standard entropy, SE, of ([BMPYR][FAP] + amides) binary mixtures as a function of [BMPYR][FAP] mole fraction at T = 298.15 K; ■, NMF; ●, DMF; △, NEF; ▽, DBF; ◀, DMA.

The apparent molar volumes, Vϕ1 and Vϕ2, were obtained using the following expressions: Vϕ1 =

(d 2 − d ) M + 1 m1dd 2 d

(7)

Vϕ2 =

(d1 − d) M + 2 m2dd1 d

(8)

Here, Vϕ1 and Vϕ2 are the apparent molar volumes, and m1 and m2 are molalities of [BMPYR][FAP] and amides, respectively. The apparent molar volumes are reported in Table S1 and Figure 3. Knowing the molar volumes of the pure components, Vo1 ([BMPYR][FAP]) and Vo2 (amides), the partial molar volumes of the components, V1 and V2, can be calculated from the expressions (eq 9) and (eq 10):

(4)

where Na is the Avogadro’s constant. The values of molar volume of pure components and their mixtures expressed in 1227

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

to the possibility of a hydrophobic interaction between the alkyl substituents of amide and carbon structure of the pyrrolidinium cation and FAP anion. In the case when x1 = 0 or x2 = 0, eqs 9 and 10 become: i=n

∑ Ai(−1)i

V1∞ = V1o +

(x1 → 0)

i=0

(11)

i=n

V 2∞ = V 2o +

∑ Ai

(x 2 → 0)

i=0

V∞ 1

where and are the partial molar volumes of the components at infinite dilution. These properties provide useful information about ion−dipole interactions, since ion−ion interactions can be neglected at infinite dilution. Partial excess molar volumes at infinite dilution of the components (VE1 )∞ and (VE2 )∞ can be obtained after the rearrangement of eqs 11 and 12:

Figure 3. Apparent molar volumes, VΦ1, of ([BMPYR][FAP] + amides) binary mixtures as a function of [BMPYR][FAP] mole fraction at T = 298.15 K; ■, NMF; ●, DMF; △, NEF; ▽, DBF; ◀, DMA.

i=n

i=n

V1 =

V1o

2

+ (1 − x1)

(12)

V∞ 2

i

∑ Ai(1 − 2x1)

(V1E)∞

2

− 2x1(1 − x1) ·

=

∑ Ai(−1)i i=0

i=0

(13)

i=n

∑ Ai(i)(1 − 2x1)i− 1 i=0

i=n

(V2E)∞ =

(9)

∑ Ai i=0

(14)

i=n

The partial molar volumes at infinite dilution and partial molar excess volumes at infinite dilution are listed in Table S6. Positive values of (VEi )∞ for all of the systems indicate weaker ion−dipole interactions between the ionic liquid and amides, compared with interactions in the pure amides and ionic liquids. Slightly higher (VEi )∞ values in the mixtures with DMA, DMF, and DBF compared to those with NMF and NEF are the result of higher dielectric constants of NMF and NEF (εNMF = 182.4; εDMF = 36.7; εDMA = 37.80; εNEF = 102.7; εDBF = 18), which is consistent with our previous results.66 On the basis of the volumetric data, the thermal expansion coefficients, αip and αp, for the pure components and mixtures, respectively, can be defined as:

V2 = V 2o + x12 ∑ Ai (1 − 2x1)i + 2x12(1 − x1) · i=0 i=n

∑ Ai(i)(1 − 2x1)i− 1 i=0

(10)

The values for the partial molar volumes are given in Table S1 and Figure 4. In the mixtures of [BMPYR][FAP] with amides

αip =

αp =

o 1 ⎛ ∂V i ⎞ ⎜ ⎟ V io ⎝ ∂T ⎠ P

1 ⎛⎜ ∂V ⎞⎟ V ⎝ ∂T ⎠ P

(15)

(16)

The expression (eq 16) can be transformed to: αp = − Figure 4. Partial molar volumes, V1, of ([BMPYR][FAP] + amides) binary mixtures as a function of [BMPYR][FAP] mole fraction at T = 298.15 K; ■, NMF; ●, DMF; △, NEF; ▽, DBF; ◀, DMA.

1 ⎛ ∂d ⎞ ⎜ ⎟ d ⎝ ∂T ⎠ P

(17)

if the molality of the component is a constant. The values of the corresponding isobaric thermal expansion coefficients calculated at different temperatures using eq 17, αip and αp, are given in Table S7. The excess thermal expansions, αEp , were calculated at T = 293.15 K using eq 18:

substituted with longer alkyl chains (DBF and NEF), it can be seen that the values of the partial and apparent molar volumes of ionic liquids increase at low [BMPYR][FAP] mole fraction, reaching its maximum at the xIL ≈ 0.2, and then decrease. Conversely, in the systems with shorter alkyl substituent on the N-atom (NMF, DMF, and DMA), the partial and apparent molar volumes of ionic liquid monotonically decrease with the increase of the ionic liquid mole fraction. This difference is due

i=2

αpE = αp −

∑ ϕα i ip i=1

(18)

where ϕi is the volume fraction of component i, defined as: 1228

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

i=2

ϕi = xiV io/∑ xiV io i=1

Funding

This work was financially supported by the Ministry of Education and Science of Serbia under project contract ON172012 and The Provincial Secretariat for Science and Technological Development of APV.

(19)

The obtained results are presented in Figure 5. It is clear that systems with NMF and NEF have positive values of excess

Notes

The authors declare no competing financial interest.



Figure 5. Variation of excess thermal expansion coefficients, αEp , of ([BMPYR][FAP] + amides) binary mixtures as a function of [BMPYR][FAP] volume fraction, ϕ1, at T = 298.15 K; ■, NMF; ●, DMF; △, NEF; ▽, DBF; ◀, DMA.

thermal expansion coefficient, which is typical for the systems containing molecules capable to self-associate.68,69 It is also known that NMF and NEF, as protic amides, realize selfassociation through N−H···CO hydrogen bonding.29 On the other hand, the absence of amide protons in DMA and DBF prevents the self-association, resulting in a negative deviation of excess thermal expansion coefficient. The system with DMF shows negative deviations of excess thermal expansion coefficient in both the ionic liquid and the DMF rich region. Comparing our results with those obtained previously in the mixtures DMF and NMF in the 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ionic liquid ([BMPYR][NTf2]),67 we can conclude that the presence of the highly hydrophobic [FAP] anion enhances amphiphile self-association.



ASSOCIATED CONTENT

S Supporting Information *

Density, excess molar volume, apparent molar volume, and partial molar volume at different temperatures and compositions of the studied mixtures; linear fitting coefficients of the density of the ([BMPYR][FAP] + amides) binary mixture; Redlich−Kister fitting coefficients of the VE of the ([BMPYR][FAP] + amides) binary mixtures; molar volumes of the ([BMPYR][FAP] + amides) binary mixtures; real, ideal, and excess standard entropy of the ([BMPYR][FAP] + amides) binary mixtures; partial molar volume at infinite dilution and partial molar excess volume at infinite dilution for the components of the mixtures; thermal expansion coefficients; and supporting figures. This material is available free of charge via the Internet at http://pubs.acs.org.



REFERENCES

(1) Crosthwaite, J. M.; Muldoon, M. J.; Aki, S. N. V. K.; Maginn, E. J.; Brennecke, J. F. Liquid phase behavior of ionic liquids with alcohols: Experimental studies and modeling. J. Phys. Chem. B 2006, 110, 9354− 9361. (2) O’Mahony, A. M.; Silvester, D. S.; Aldous, L.; Hardacre, C.; Compton, R. G. Effect of water on the electrochemical window and potential limits of room-temperature ionic liquids. J. Chem. Eng. Data 2008, 53, 2884−2891. (3) Earle, M. J.; Esperança, J.; Gilea, M. A.; Lopes, J. N. C.; Rebelo, L. P. N.; Magee, J. W.; Seddon, K. R.; Widegren, J. A. The distillation and volatility of ionic liquids. Nature 2006, 439, 831−834. (4) Pauliukaite, R.; Doherty, A. P.; Murnaghan, K. D.; Brett, C. M. A. Characterization and application of carbon film electrodes in room temperature ionic liquid media. J. Electroanal. Chem. 2008, 616, 14− 26. (5) Zarrougui, R.; Dhahbi, M.; Lemordant, D. Effect of temperature and composition on the transport and thermodynamic properties of binary mixtures of ionic liquid N-butyl-N-methylpyrrolidinium bis(Trifluoromethanesulfonyl)imide and propylene carbonate. J. Solution Chem. 2010, 39, 921−942. (6) Ge, R.; Hardacre, C.; Nancarrow, P.; Rooney, D. W. Thermal conductivities of ionic liquids over the temperature range from 293 to 353 K. J. Chem. Eng. Data 2007, 52, 1819−1823. (7) Doyle, K. P.; Lang, C. M.; Kim, K.; Kohl, P. A. Dentrite-free electrochemical deposition of Li-Na alloys from an ionic liquid electrolyte. J. Electrochem. Soc. 2006, 153, 1353−1357. (8) Chun, S.; Dzyuba, S. V.; Bartsch, R. A. Influence of structural variations in room-temperature ionic liquids on the selectivity and efficiency of competitive alkali metal salt extraction by a crown ether. Anal. Chem. 2001, 73, 3737−3741. (9) Visser, A. E.; Rogers, R. D. Room-temperature ionic liquids: new solvents for f-element separations and associated solution chemistry. J. Solid State Chem. 2003, 171, 109−113. (10) Rilo, E.; Pico, J.; Garabal, S. G.; Varela, L. M.; Cabeza, O. Density and surface tension in binary mixtures of CnMIM-BF4 ionic liquids with water and ethanol. Fluid Phase Equilib. 2009, 285, 83−89. (11) Singh, T.; Kumar, A. Thermodynamics of dilute aqueous solutions of imidazolium based ionic liquids. J. Chem. Thermodyn. 2011, 43, 958−965. (12) Pârvulescu, V. I.; Hardacre, C. Catalysis in ionic liquids. Chem. Rev. 2007, 107, 2615−2665. (13) Paul, C. E.; Gotor-Fernández, V.; Lavandera, I.; MontejoBernardo, J.; García-Granda, S.; Gotor, V. Chemoenzymatic preparation of optically active 3-(1H-imidazol-1-yl) cyclohexanolbased ionic liquids: application in organocatalysis and toxicity studies. RSC Adv. 2012, 2, 6455−6463. (14) Rogers, E. I.; Šljukić, B.; Hardacre, C.; Compton, R. G. Electrochemistry in room-temperature ionic liquids: Potential windows at mercury electrodes. J. Chem. Eng. Data 2009, 54, 2049− 2053. (15) Poole, C. F. Chromatographic and spectroscopic methods for the determination of solvent properties of room temperature ionic liquids. J. Chromatogr., A 2004, 1037, 49−82. (16) Anderson, J. L.; Armstrong, D. W. High-stability ionic liquids. A new class of stationary phases for gas chromatography. Anal. Chem. 2003, 75, 4851−4858. (17) Ignat’ev, N. V.; Welz-Biermann, U.; Kucheryna, A.; Bissky, G.; Willner, H. New ionic liquids with tris(perfluoroalkyl)-

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. 1229

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

trifluorophosphate (FAP) anions. J. Fluorine Chem. 2005, 126, 1150− 1159. (18) Yao, C.; Pitner, W. R.; Anderson, J. L. Ionic liquids containing the tris(pentafluoroethyl) trifluorophosphate anion: a new class of highly selective and ultra hydrophobic solvents for the extraction of polycyclic aromatic hydrocarbons using single drop microextraction. Anal. Chem. 2009, 81, 5054−5063. (19) Yan, P. F.; Yang, M.; Liu, X. M.; Liu, Q. S.; Tan, Z. C.; WelzBiermann, U. Activity coefficients at infinite dilution of organic solutes in 1-ethyl-3-methylimidazolium tris(pentafluoroethyl)trifluorophosphate [EMIM][FAP] using gas-liquid chromatography. J. Chem. Eng. Data 2010, 55, 2444−2450. (20) Aparicio, S.; Mert, A.; Ferdi, K. Thermophysical properties of pure ionic liquids: Review of present situation. Ind. Eng. Chem. Res. 2010, 49, 9580−9595. (21) Paulsen, B. D.; Frisbie, C. D. Dependence of conductivity on charge density and electrochemical potential in polymer semiconductors gated with ionic liquids. J. Phys. Chem. C 2012, 116, 3132−3141. (22) Ferreira, A. F.; Simo̅es, P. N.; Ferreira, A. G. M. Quaternary phosphonium-based ionic liquids: Thermal stability and heat capacity of the liquid phase. J. Chem. Thermodyn. 2012, 45, 16−27. (23) Deyko, A.; Lovelock, K. R. J.; Corfield, J. A.; Taylor, A. W.; Gooden, P. N.; Villar-Garcia, I. J.; Licence, P.; Jones, R. G.; Krasovskiy, V. G.; Chernikova, E. A.; Kustov, L. M. Measuring and predicting ΔvapH298 values of ionic liquids. Phys. Chem. Chem. Phys. 2009, 11, 8544−8555. (24) González, R.; Hernández Battez, A.; Blanco, D.; Viesca, J. L.; Fernández-González, A. Lubrication of TiN, CrN and DLC PVD coatings with 1-butyl-1-methylpyrrolidinium tris(pentafluoroethyl)trifluorophosphate. Tribol. Lett. 2010, 40, 269−277. (25) Zoubi, M. A.; Endres, F. Electrochemical synthesis of poly(pphenylene) and poly(p -phenylene)/TiO2 nanowires in an ionic liquid. Electrochim. Acta 2011, 56, 5872−5876. (26) Millefiorini, S.; Tkaczyk, A. H.; Sedev, R.; Efthimiadis, J.; Ralston, J. Electrowetting of ionic liquids. J. Am. Chem. Soc. 2006, 128, 3098−3101. (27) Abedin, S. Z. E.; Borissenko, N.; Endres, F. Electropolymerization of benzene in a room temperature ionic liquid. Electrochem. Commun. 2004, 6, 422−426. (28) Duffy, N. W.; Bond, A. M. Macroelectrode voltammetry in toluene using a phosphonium-phosphate ionic liquid as the supporting electrolyte. Electrochem. Commun. 2006, 8, 892−898. (29) Greaves, T. L.; Drummond, C. J. Solvent nanostructure, the solvophobic effect and amphiphile self-assembly in ionic liquids. Chem. Soc. Rev. 2013, 42, 1096−1120. (30) Smirnova, N. A.; Safonova, E. A. Micellization in solutions of ionic liquids. Colloid J. 2012, 74, 254−265. (31) Dannhauser, W.; Johari, G. P. Intermolecular association and dielectric relaxation in some liquid amide. Can. J. Chem. 1968, 46, 3143−3149. (32) Gaciño, F. M.; Regueira, T.; Lugo, L.; Comuñas, M. J. P.; Fernández, J. Influence of molecular structure on densities and viscosities of several ionic liquids. J. Chem. Eng. Data 2011, 56, 4984− 4999. (33) Stevanović, S.; Costa Gomes, M. F. Solubility of carbon dioxide, nitrous oxide, ethane and nitrogen in 1-butyl-1-methylpyrrolidinium and trihexyl(tetradecyl)phosphonium tris(pentafluoroethyl)trifluorophosphate (eFAP) ionic liquids. J. Chem. Thermodyn. 2013, 59, 65−71. (34) Domańska, U.; Lukoshko, E. V.; Królikowski, M. Measurements of activity coefficients at infinite dilution for organic solutes and water in the ionic liquid 1-butyl-1-methylpyrrolidinium tris(pentafluoroethyl)trifluorophosphate ([BMPYR][FAP]). Chem. Eng. J. 2012, 183, 261−270. (35) Domańska, U.; Lukoshko, E. V.; Królikowski, M. Phase behaviour of ionic liquid 1-butyl-1-methylpyrrolidinium tris(pentafluoroethyl)trifluorophosphate with alcohols, water and aromatic hydrocarbons. Fluid Phase Equilib. 2013, 345, 18−22.

(36) Noh, H. J.; Park, S. J.; In, S. J. Excess molar volumes and deviations of refractive indices at 298.15 K for binary and ternary mixtures with pyridine or aniline or quinoline. J. Ind. Eng. Chem. 2010, 16, 200−206. (37) Han, K. J.; Oh, J. H.; Park, S. J.; Gmehling, J. Excess molar volumes and viscosity deviations for the ternary system N,Ndimethylformamide + N-methylformamide + water and the binary subsystems at 298.15 K. J. Chem. Eng. Data 2005, 50, 1951−1955. (38) García, B.; Alcalde, R.; Leal, J. M.; Matos, J. S. Solute-solvent interactions in amide-water mixed solvents. J. Phys. Chem. B 1997, 101, 7991−7997. (39) Davis, M. I. Determination and analysis of the excess molar volumes of some amide-water systems. Thermochim. Acta 1987, 120, 299−314. (40) Nikolić, A.; Jović, B.; Krstić, V.; Tričković, J. Excess molar volumes of N-methylformamide + tetrahydropyran, + 2-pentanone, + n-propylacetate at the temperatures between 298.15 and 313.15 K. J. Mol. Liq. 2007, 133, 39−42. (41) Nain, A. K. Densities and volumetric properties of (acetonitrile + an amide) binary mixtures at temperatures between 293.15 and 318.15 K. J. Chem. Thermodyn. 2006, 38, 1362−1370. (42) Scharlin, P.; Steinby, K.; Domańska, U. Volumetric properties of binary mixtures of N,N-dimethylformamide with water or water-d2 at temperatures from 277.13 to 318.15 K. J. Chem. Thermodyn. 2002, 34, 927−957. (43) Geng, Y.; Wang, T.; Yu, D.; Peng, C.; Liu, H.; Hu, Y. Densities and viscosities of the ionic liquid [C4mim][PF6] + N,N-dimethylformamide binary mixtures at 293.15 to 318.15 K. Chin. J. Chem. Eng. 2008, 16, 256−262. (44) Peng, S. J.; Hou, H. Y.; Zhou, C. S.; Yang, T. Densities and excess volumes of binary mixtures of N,N-dimethylformamide with aromatic hydrocarbon at different temperature. J. Chem. Thermodyn. 2007, 39, 474−482. (45) Juárez-Camacho, E. P.; Manríquez-Ramírez, M. E.; Reza-San Germán, C. M.; Zúñiga-Moreno, A. Volumetric properties of the binary system trihexyltetradecylphosphonium bromide (CYPHOS IL 102) + N,N-dimethylformamide (DMF) at temperatures from T = 293.15 to 313.15 K at atmospheric pressure. J. Sol. Chem. 2012, 41, 1575−1586. (46) Chu, D. Y.; Chang, Y.; Hu, I. Y.; Liu, R. L. Excess molar volumes of mixtures of N,N-dimethylformamide and water and apparent molal volumes and partial molal volumes of N,N-dimethylformamide in water from 278.15 to 318.15 K. Acta Phys. Chim. Sin. 1990, 6, 203− 208 in Chinese. (47) Volpe, C. D.; Guarino, G.; Sartorio, R.; Vitagliano, V. Diffusion, viscosity and refractivity data on the system dimethylformamide-water at 20 and 40°C. J. Chem. Eng. Data 1986, 31, 37−40. (48) Nikam, P. S.; Kharat, S. J. Densities and viscosities of binary mixtures of N,N-dimethylformamide with benzyl alcohol and acetophenone at (298.15, 303.15, 308.15, and 313.15) K. J. Chem. Eng. Data 2003, 48, 1291−1295. (49) Xu, L.; Lin, G.; Wang, X.; Lin, R. Densities and volumetric properties of 2-chlorethanol with N,N-dimethylformamide and water at different temperatures. J. Mol. Liq. 2006, 123, 130−133. (50) Attri, P.; Venkatesu, P.; Hofman, T. Temperature dependence measurements and structural characterization of trimethyl ammonium ionic liquids with a highly polar solvent. J. Phys. Chem. B 2011, 115, 10086−10097. (51) Bhuiyan, M. M. H.; Uddin, M. H. Excess molar volumes and excess viscosities for mixtures of N,N-dimethylformamide with methanol, ethanol and 2-propanol at different temperatures. J. Mol. Liq. 2008, 138, 139−146. (52) Beilsteins Handbuch der Organischen Chemie, 4th ed.; Verlag von Julius Springer: Berlin, 1922; p 109. (53) Sears, P. G.; O’Brien, W. C. Dielectric constants and viscosities of some mono-N-substituted amides and cyclic esters. J. Chem. Eng. Data 1968, 13, 112−115. (54) Blomendal, M.; Marcus, Y.; Booij, M.; Hofstee, R.; Somsen, G. Enthalpies of solution and solvation of amides in N,N-dimethylforma1230

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231

Journal of Chemical & Engineering Data

Article

mide: Application of the random contact point approach. J. Solution Chem. 1988, 17, 15−19. (55) Papamatthaiakis, D.; Aroni, F.; Havredaki, V. Isentropic compressibilities of (amide + water) mixtures: A comparative study. J. Chem. Thermodyn. 2008, 40, 107−118. (56) Scharlin, P.; Steinby, K. Excess thermodynamic properties of binary mixtures of N,N-dimethylacetamide with water or water-d2 at temperatures from 277.13 to 318.15 K. J. Chem. Thermodyn. 2003, 35, 279−300. (57) Chen, F.; Wu, J.; Wang, Z. Volumetric properties of the binary liquid mixtures of N,N-dimethylacetamide + benzene, + toluene, or + ethylbenzene at different temperatures and atmospheric pressure. J. Mol. Liq. 2008, 140, 6−9. (58) Pal, A.; Kumar, A. Excess molar volumes and kinematic viscosities for binary mixtures of dipropylene glycol monobutyl ether and dipropylene glycol tert-butyl ether with 2-pyrrolidinone, Nmethyl-2-pyrrolidinone, N,N-dimethylformamide, and N,N-dimethylacetamide at 298.15 K. J. Chem. Eng. Data 2005, 50, 856−862. (59) Sekhar, G. C.; Venkatesu, P.; Rao, M. V. P. Excess molar volumes and speeds of sound of N,N-dimethylacetamide with chloroethanes and chloroethenes at 303.15 K. J. Chem. Eng. Data 2001, 46, 377−380. (60) Davis, M. I.; Hernandez, M. E. Excess molar volumes for N,Ndimethylacetamide + water at 25 °C. J. Chem. Eng. Data 1995, 40, 674−678. (61) Ivanov, E. V.; Abrosimov, V. K.; Lebedeva, E. Y. Apparent molar volumes and expansibilities of H2O and D2O in N,N-dimethylformamide and N,N-dimethylacetamide in the range of T = (278.15 to 318.15) K at p = 0.1 MPa: A comparative analysis. J. Chem. Thermodyn. 2012, 53, 131−139. (62) Bǿje, L.; Hvidt, A. Densities of aqueous mixtures of nonelectrolytes. J. Chem. Thermodyn. 1971, 3, 663−673. (63) Iloukhani, H.; Rakhshi, M. Excess molar volumes, viscosities, and refractive indices for binary and ternary mixtures of {cyclohexanone (1) + N,N-dimethylacetamide (2) + N,N-diethylethanolamine (3)} at (298.15, 308.15, and 318.15) K. J. Mol. Liq. 2009, 149, 86−95. (64) Zafarani-Moattar, M. T.; Shekaari, H. Density and speed of sound of lithium bromide with organic solvents: Measurement and correlation. J. Chem. Thermodyn. 2007, 39, 1649−1660. (65) Redlich, O.; Kister, A. T. Algebraic representation of thermodynamic properties and the classification of solutions. Ind. Eng. Chem. 1948, 40, 345−348. (66) Glasser, L. Lattice and phase transition, thermodynamics of ionic liquids. Thermochim. Acta 2004, 421, 87−93. (67) Vraneš, M. B.; Dožic, S.; Đeric, V.; Gadžuric, S. B. Volumetric properties of binary mixtures of 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide with N-methylformamide and N,Ndimethylformamide from (293.15 to 323.15) K. J. Chem. Eng. Data 2013, 58, 1092−1102. (68) Tamura, K.; Nakamura, M.; Murakami, S. Excess volumes of water + acetonitrile and water + dimethylsulfoxide at 30°C and the effect of the excess thermal expansivity coefficients on derivated thermodynamic properties. J. Solution Chem. 1997, 26, 1199−1207. (69) Pires, J.; Timperman, L.; Jacquemin, J.; Balducci, A.; Anouti, M. Density, conductivity, viscosity, and excess properties of (pyrrolidinium nitrate-based protic ionic liquid + propylene carbonate) binary mixture. J. Chem. Thermodyn. 2013, 59, 10−19.

1231

dx.doi.org/10.1021/je400803f | J. Chem. Eng. Data 2014, 59, 1225−1231