Thermodynamics of Mixtures Containing Amines. X. Systems with

Jul 12, 2011 - Ricardo Páramo , Juan Antonio González , Isaías García de la Fuente , Carlos Casanova , José Carlos Cobos. Thermochimica Acta 2015 600,...
0 downloads 0 Views 1MB Size
ARTICLE pubs.acs.org/IECR

Thermodynamics of Mixtures Containing Amines. X. Systems with Cyclic Amines or Morpholine Juan Antonio Gonzalez* G.E.T.E.F., Departamento de Física Aplicada, Facultad de Ciencias, Universidad de Valladolid, 47071 Valladolid, Spain ABSTRACT: Cyclic amine or morpholine + organic solvent mixtures have been investigated in terms of DISQUAC and of the KirkwoodBuff formalism. The amines considered are cyclohexylamine; (c-CH2)uNH (u = 27) and (c-CH2)uNCH3 (u = 4,5). The organic solvents are alkanes and methanol or ethanol. The DISQUAC interaction parameters are reported. The model describes correctly a whole set of thermodynamic properties: vaporliquid equilibria (VLE); excess Gibbs energies, GEm; excess enthalpies, HEm; excess heat capacities at constant pressure, CEP,m; and partial excess enthalpies at infinite dilution, HE,∞ 1 , as well as the main features of the KirkwoodBuff integrals. In (c-CH2)uNH + C6H12 mixtures, amineamine interactions become weaker with and of the effective dipole moment. Interactions increased u values. This is supported by the corresponding decrease of HEm and HE,∞ 1 between amine molecules are also weakened when passing from a primary cyclic amine to an isomeric tertiary cyclic amine. The existence of an aromatic ring in the amine, as in aniline or pyridine, leads to stronger amineamine interactions. Dipolar interactions between morpholine molecules are stronger than those between piperidine molecules, and reveal the existence of proximity effects between the two groups of morpholine. In systems with alkanes, interactions between amine molecules are preferred to those between unlike molecules. In piperidine + methanol and pyrrolidine + ethanol systems, the mixture structure is close to random mixing.

1. INTRODUCTION Amines and oxaalkanes are very important molecules from a theoretical point of view. Primary and secondary amines are weakly self-associated.18 Their Trouton’s constants are close to the value of nonassociated species9 (92.05 J 3 mol1 3 K1). For example, these constants are 90.02 and 92.83 J 3 mol1 3 K1 for hexylamine and aniline, respectively.10 In the case of 1-alkanols,9 the Trouton’s constant is 110.88 J 3 mol1 3 K1. Ether molecules are formally obtained by replacing one or several CH2 groups in an alkane by O atoms (e.g., cyclohexane, oxane, 1,3-dioxane, 1,4dioxane, 1,3,5-trioxane). A large variety of homomorphic molecular species can be so obtained which differ in the number and relative positions of the functional group. Similarly, many different amine molecules may be constructed. Thus, the study of mixtures with oxaalkanes or amines makes possible the examination of the influence of some interesting effects on their thermodynamic properties. Linear amines, CH3(CH2)u1NH2 or CH3(CH2)u1NH(CH2)v1CH3), or linear oxaalkanes, CH3(CH2)u1O(CH2)v1CH3, allow the study of the size and steric effects produced by alkylgroups attached to the amine or to the etheric groups; N,N,N-trialkylamines, the effect of a globular shape; aromatic amines or ethers, the effect of polarizability; cyclic amines or oxaalkanes, the ring strain. Proximity effects between two O groups may be investigated through solutions including linear acetals, CH3(CH2)u1O CH2O(CH2)v1CH3, while mixtures with linear polyoxaalkanes CH3O(CH2CH2O)u1CH3 are useful in the investigation of the effect of increasing the number of oxyethylene groups. From a practical point of view, amines and ethers are very important compounds. Thus, aniline, its derivatives, or piperidine are used in the manufacturing of dyes and of rubber vulcanization accelerators, or in the fabrication of pharmaceuticals and pesticides. The treatment of pyridine systems is a first step for a r 2011 American Chemical Society

better understanding of the pyrrole ring, specially important for modeling typical binding sites on proteins.7 Mixtures containing ethers are also important because they are increasingly used as additives to gasoline owing to their octane-enhancing and pollution-reducing properties.11,12 Cyclic polyethers have attracted interest as model substances for biosystems, particularly in connection with precise molecular recognition, which is essential to living systems,13 separation techniques, or chemical analysis. On the other hand, morpholine is a cyclic molecule which contains both the amine and the ether group and is particularly suitable for extraction, extractive distillation, and solvent application in the petrochemical industry. The aim of this work is to gain insight into the interactions and structure of systems formed by cyclic amines or morpholine and organic solvents, through the application of the DISQUAC model,14,15 and of the KirkwoodBuff formalism.1618 The former is a purely physical model based on the rigid lattice theory developed by Guggenheim.19 The latter is concerned with the study of fluctuations in the number of molecules of each component, and of the cross fluctuations. The amines considered in this work are cyclohexylamine, aziridine, azetidine, pyrrolidine, piperidine, hexamethyleneimine, heptamethyleneimine, methylpyrrolidine, methyl-piperidine, and morpholine. The solvents are cycloalkanes, n-alkanes, and 1-alkanols. No interaction parameters for mixtures with cyclic amines are available in terms of UNIFAC (Dortmund);2023 however, some DISQUAC interaction parameters for cyclic secondary amine + cycloalkane systems have been previously reported.24

Received: May 24, 2011 Accepted: July 12, 2011 Revised: June 23, 2011 Published: July 12, 2011 9810

dx.doi.org/10.1021/ie201120h | Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

ARTICLE

QUAC Table 1. Dispersive (DIS) and Quasichemical (QUAC) Interchange Coefficients, CDIS , for (s,t) Contacts in Cyclic st,l and Cst,l Amine or Morpholine + Organic Solvents Mixtures (l = 1, Gibbs Energy; l = 2, Enthalpy; l = 3, Heat Capacity)

amine

contacta (s,t)

CDIS st,1

CDIS st,2

CDIS st,3

CQUAC st,1

CQUAC st,2

CQUAC st,3

cyclohexylamine

(c,n)

2.7

4

4.5

5.5

(c-CH2)2NH

(a,n) (c,n)

2.05 9.5

3.7 9.1

4.5

4.5 8.5

5.5 19

5

(c-CH2)3NH or (c-CH2)4NH

(c,n)

8.75

8.

4.5

4.5

14

5

(c-CH2)5NH

(c,n)

52b

8

4.5

4.5

10

5

(a,n)

5.4

7.8

2.5c

4.5

10

5

(c-CH2)6NH

(c,n)

4.15

8

4.5

4.5

9.10

5

(c-CH2)7NH

(c,n)

4.15

8

4.5

4.5

4.5

5

(c-CH2)4NCH3

(c,n)d

8

17.2

(c-CH2)5NCH3

(c,n)d (a,n)e

3.7 5

11.9 16.3

Morpholine

(e,n)f

4.3

9

1

1

(e,n)g

10.5

23

1

1

(c-CH2)4NH

(h,n)

20.5

46

1.5

0.5

(c-CH2)5NH

(h,n)

16.8

46

1.5

0.5

a

Abbreviations: a, aliphatic in n-alkane, methyl-pyrrolidine, methyl-piperidine, or 1-alkanols; c, c-CH2 in cycloalkanes, cyclic amines, or morpholine; c e, O in morpholine; n, amine group in cyclic amines or morpholine; h, OH in 1-alkanols. b For cyclopentane, CDIS an,1 = 3.8. For tetradecane, DIS DIS d e CDIS CDIS = 2.1. Interaction parameters determined assuming that in mixtures with cyclohexane, C = C = 0. (octane) =16.6; CDIS an,3 an,1 an,2 an,2(decane) an,2 f g = 17.3. In mixtures with cycloalkanes. In mixtures with n-alkanes.

2. THEORIES 2.1. DISQUAC. In the framework of DISQUAC, cyclic amine or morpholine +organic solvent mixtures are regarded as possessing the following types of surface: (i) type c, (c-CH2 or c-CH in cyclic molecules; (ii) type n, (amine group in cyclic amines or morpholine; (iii) type e, (O group in morpholine); (iv) type s, (s = a, aliphatic (CH3, CH2, in n-alkanes, 1-alkanols, or methyl derivatives of cyclic amines; s = h, hydroxyl (OH in 1-alkanols)). 2.1.1. General Equations. The main features of DISQUAC are as follows: (i) The total molecular volumes, ri, surfaces, qi, and the molecular surface fractions, Ri, of the compounds present in the mixture are calculated additively on the basis of the group volumes RG and surfaces QG recommended by Bondi.25 As volume and surface units, the volume RCH4 and surface QCH4 of methane are taken arbitrarily.26 The geometrical parameters for the groups referred to in this work are given elsewhere.24,26,27 (ii) The partition function is factorized into two terms, in such a way that the excess functions are calculated as the sum of two contributions: a dispersive (DIS) term which represents the contribution from the dispersive forces; and a quasichemical (QUAC) term which arises from the anisotropy of the field forces created by the solution molecules. In the case of the Gibbs , represented by the energy, GEm, a combinatorial term, GE,COMB m FloryHuggins equation26,28 must be considered. Thus COMB DIS QUAC þ GE, þ GE, GEm ¼ GE, m m m

ð1Þ

HmE ¼ HmE, DIS þ HmE, QUAC

ð2Þ

The equations used to calculate the DIS and QUAC contributions to GEm and HEm in the framework of DISQUAC are given elsewhere.24,27 The temperature dependence of the interaction parameters is expressed in terms of the DIS and QUAC interQUAC where s 6¼ t are two conchange coefficients,24,27 CDIS st,l ;Cst,l tact surfaces present in the mixture and l = 1 (Gibbs energy; CDIS/QUAC = gDIS/QUAC (To)/RTo); l = 2 (enthalpy, CDIS/QUAC = st,1 st st,2 DIS/QUAC DIS/ hst (To)/RTo)); l = 3 (heat capacity,CDIS/QUAC = c st,3 pst QUAC (To)/R)). To = 298.15 K is the scaling temperature, and R is the gas constant. 2.2. KirkwoodBuff Formalism. The theory1618 describes thermodynamic properties of solutions in an exact manner in the whole concentration range using the KirkwoodBuff integrals: Z ∞ ðgij  1Þ4πr 2 dr ð3Þ Gij ¼ o

The radial distribution function, gij, denotes the probability of finding a molecule of species i in a volume element at the distance r of the center of a molecule of species j. So, this function provides information about the solution structure on the microscopic level. The Gij values can be interpreted as follows: Gij > 0 represents the excess of molecules of the i type in the space around a given molecule of species j. This means attractive interactions between molecules of i and j. Gij < 0 means that interactions of ii and jj are preferred to mutual interactions.29 The KirkwoodBuff integrals are derived from experimental data of thermodynamic properties as chemical potential; partial molar volumes and isothermal compressibility factor. The resulting equations are17,30

(iii) The interaction parameters are assumed to be dependent on the molecular structure; (iv) The value z = 4 for the coordination number is used for all the polar contacts. This represents one of the more important shortcomings of the model and is partially removed via the hypothesis of considering structure-dependent interaction parameters.

Gii ¼ RTkT þ

Gij ¼ RTkT  9811

xj V̅ 2j xi VD

V̅ i V̅ j VD



V xi

ð4Þ

ði 6¼ jÞ

ð5Þ

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

ARTICLE

where xi and V i are the mole fraction and the partial molar volume of component i, respectively (i = 1, 2); V is the molar volume of the solution; and kT, the isothermal compressibility of the mixture. D is defined as ! x1 x2 ∂2 GEm ð6Þ D¼1 þ RT ∂x21

Table 2. Molar Excess Gibbs Energies, GEm, at Equimolar Composition and Temperature T for Cyclic Amine or Morpholine + Alkane, or +1-Alkanol Mixtures. Comparison of Experimental Results (expt) with DISQUAC (DQ) Calculations Using the Interchange Coefficients from Table 1 GEm (J mol1)

σr (P)a

P, T

T (K) N

system

Using the Gij quantities, it is possible to estimate the so-called linear coefficients of preferential solvation:3032

expt

DQ

expt

DQ ref

C6H13N + n-C8

333.15 11

479

479 0.004 0.007 84

(c-CH2)2NH + C6H12

363.15 11 298.15 18

708 1110

321 0.003 0.10 84 1102 0.004 0.013 85

(c-CH2)3NH + C6H12

298.15 11

620

625 0.004 0.021 85

(c-CH2)4NH + C6H12

298.15 11

533

536 0.004 0.005 85

313.15 13

515

504 0.003 0.007 86

333.15 18

478

463 0.003 0.008 86

(c-CH2)5NH + C5H10

298.15 14

302

303 0.014 0.017 87

3. ESTIMATION OF THE MODEL PARAMETERS The general procedure applied in the estimation of the DISQUAC interaction parameters has been explained in detail elsewhere.34 Final values of the fitted parameters in this work are collected in Table 1. Some important remarks are given as follows.

(c-CH2)5NH + C6H12

298.15 15

366

369 0.008 0.008 85

(c-CH2)5NH + C7H14 (c-CH2)5NH + C8H16

298.15 11 298.15 13

381 372

370 0.004 0.006 87 356 0.004 0.010 87

(c-CH2)6NH + C6H12

298.15 11

305

302 0.008 0.010 85

(c-CH2)5NH + n-C7

298.42

6

512

524 0.005 0.007 88

(c-CH2)4NCH3+ C6H12 298.15 10

114

114 0.002 0.004 85

Cyclohexylamine or Secondary Cyclic Amine + Cycloalkane. These systems are built by one contact (c,n). The corre-

(c-CH2)5NCH3+ C6H12 298.15 12

41

42 0.002 0.004 85

(c-CH2)5NCH3 + n-C7 298.42

5

53

54 0.003 0.003 88

morpholine + C8H16

363.15 16

912

916 0.003 0.007 89

morpholine + n-C7

393.15 20 303.15

839 1206

840 0.004 0.010 89 1171 88

morpholine + n-C8

353.35 14

1060

1048 0.003 0.016 89

383.35 15

952

967 0.002 0.011 89

δii ¼ xi xj ðGii  Gij Þ δij ¼ xi xj ðGij  Gjj Þ

ð7Þ ði 6¼ jÞ

which are useful quantities to determine the local mole fractions, xij, of the i species around the central j molecule.30,33

sponding interaction parameters were obtained from available data on VLE, HEm, and CEpm. Cyclohexylamine or Piperidine + n-Alkane. These mixtures are characterized by three contacts: (a,c), (a,n), and (c,n). The (a,c) contacts are described by DIS parameters only which are known from VLE, HEm and CEpm measurements of cyclohexane + n-alkane mixtures.28 As the interaction parameters for the (c,n) contacts have been obtained previously, only those for the (a,n) contacts must be now determined. Calculations were carried out = CQUAC . This general rule is valid for assuming that CQUAC an,l cn,l many other solutions as, linear oxaalkane,35 linear alkanone,36 linear organic carbonate,37 alkanoic acid anhydride,38 haloalkane,39,40 linear monocarboxylic acids,41 alkanol,4246 alkoxyethanol,47 sulfolane,48 amide,4951 or amine10,52 + n-alkane, or + cyclohexane. Tertiary Cyclic Amine + Alkane. These solutions are built by three contacts: (a,c), (a,n), and (c,n). The CDIS ac,l coefficients have been neglected as in previous DISQUAC applications methylcyclohexane has been treated as a n-alkane.10,37 Note the low HEm of the methylcyclohexane + heptane system (36 J 3 mol1 at 298.15 K and equimolar composition53). The interaction parameters for the (a,n) and (c,n) contacts are assumed to be merely dispersive, as in N,N,N-trialkylamine + alkane systems.54 In the case of cyclohexane mixtures, for simplicity, the interaction parameters for (a,n) contacts have been also neglected. Morpholine + Alkane. The mixtures with cycloalkanes are characterized by three contacts (c,e), (c,n), and (e,n). The CDIS ce,l and CQUAC coefficients are known35 from the study of cyclic ce,l ether + cycloalkane mixtures, and those for the (c,n) contacts have been obtained from data for piperidine systems (see above). Thus, only the interaction parameters for the (e,n) contacts must be fitted. There are six contacts in morpholine + n-alkane systems: (a,c), (a,e), (a,n), (c,e), (c,n), and (e,n). Interaction parameters for the contacts (a,c), (a,e), (a,n), (c,e), (c,n) are known, as the CDIS ae,l and interchange coefficients were obtained in a previous CQUAC ae,l work35 from cyclic ether + n-alkane data. However, it was

(c-CH2)4NH + ethanol

a

313.15 15 1063 1070 0.006 0.021 86 333.15 15

905

(c-CH2)5NH + methanol 298.15 11

831

928 0.004 0.019 86 825 0.008 0.025 72

318.15 11

693

673 0.005 0.021 72

Standard relative deviation, eq 8.

assumed that the different molecular environment (cycloalkane, n-alkane) modify the CDIS en,l coefficients. A similar procedure was applied for systems of alkoxyethanols, and n-alkane, cyclohexane, 1-alkanol or ether.47,55 Pyrrolidine, or Ppiperidine +1-Alkanol. The contacts (a,c), (a,n), (a,h), (c,n), (c,h), and (h,n) are present in these solutions. and CDIS,QUAC are The interchange coefficients CDIS,QUAC ah,l ch,l known from the treatment of 1-alkanol + n-alkane,27 or + cycloalkane42,44 mixtures, respectively. Thus, only the interaction parameters for (h,n) contacts must be fitted.

4. RESULTS 4.1. DISQUAC Results. Results from the DISQUAC model are compared with experimental data for VLE, GEm, HEm, CEPm, and KirkwoodBuff integrals or linear coefficients of HE,∞ 1 preferential solvation in Tables 27. Comparisons for selected mixtures are plotted in Figures 17. For the sake of clarity, relative deviations for the pressure (P) and HEm are defined as

σ r ðPÞ ¼ 9812

8 azetidine > pyrrolidine > piperidine > ... (Tables 4 and 6). This reveals that interactions between amine molecules become weaker when the size of these compounds are increased. The corresponding decrease of the effective dipole moment of cyclic amines (Table 8), a useful magnitude to examine the impact of polarity on bulk properties,34,5962 may explain the mentioned variation 1 E of HEm and HE,∞ 1 . The large positive values of TSm/(J 3 mol ) E E (= Hm  Gm) of these cyclohexane mixtures are remarkable: 510 (aziridine) < 802 (azetidine) > 664 (pyrrolidine) > 461 (piperidine)> 422 (hexamethyleneimine) (all values calculated using experimental data listed in Tables 2 and 4) and reveal that amineamine interactions are mainly of dipolar type. Note that for the typical associated solution 1-pentanol + hexane, GE63 m = 1 and TSEm = 566 J 3 mol1. 1041 J 3 mol1 ;HE64 m = 475 J 3 mol = For the system diethylamine + hexane at 303.15 K, HE,65 m 672 J 3 mol1; GEm = 222 J 3 mol1 (DISQUAC value) and TSEm = 450 J 3 mol1, which is lower than the value for the pyrrolidine + cyclohexane mixture. That is, dipolar interactions are less important in systems with linear secondary amines. On the other hand, in cyclohexane solutions, HEm and HE,∞ 1 changes in the order: cyclohexylamine > hexamethyleneimine > Table 6. Partial Molar Excess Enthalpies at Infinite Dilution and 298.15 K for Cyclic Amine(1) or Morpholine(1) + Cyclohexane(2) Mixtures. Comparison of Experimental Valuesa (expt) with DISQUAC (DQ) Results Obtained Using Interaction Parameters Listed in Table 1

amine cylohexylamine

HE,∞ (kJ 3 mol1) 1

HE,∞ (kJ 3 mol1) 2

expt

DQ

expt

DQ

ref 90

5.46

5.74

2.74

2.65

12.03

14.85

6.09

6.46

85

azetidine pyrrolidine

6.90 6.95

7.49 6.85

5.08 3.95

5.24 4.13

85 85

piperidine

4.68

4.87

2.98

2.79

85

hexamethyleneimine

3.85

4.32

2.44

2.35

85

heptamethyleneimine

2.15

2.50

1.22

1.47

85

N-methylpyrrolidine

1.22

1.00

0.68

0.96

85.

N-methylpiperidine

0.54

0.65

0.33

0.56

85

Morpholineb

9.65

10.09

6.14

6.83

95

aziridine

Obtained from the fitting of the experimental data to a RedlichKister equation. b Value at 308.15 K.

a

N-methylpiperidine (Tables 4 and 6). Therefore, interactions between amine molecules are weakened when passing from a primary amine to an isomeric tertiary amine. The same trend is observed in mixtures containing linear amines. Thus, for systems involving heptane, HEm/(J 3 mol1) = 1064 (hexylamine66)> 454 (dipropylamine67) > 112 (triethylamine68). It is noteworthy that for the cyclohexylamine solution, TSEm = 235 J 3 mol1, value calculated using GEm = 614 J 3 mol1 obtained from DISQUAC, which indicates that amine self-association is in this case more important than in piperidine (see above). HEm results for systems with the aromatic amines aniline, or pyridine and cyclohexane are much higher than those involving cyclohexyamine, piperidine, or N-methylpiperidine. Thus, HEm/(J 3 mol1) = 1880 (aniline;69 T = 308.15 K); 1438 (pyridine70). In addition, aniline is not miscible with alkanes at room temperature and equimolar compostion. The upper critical solution temperature (UCST) of the aniline + heptane system is 342.7 K.71 Therefore, it may be concluded then that the presence of the aromatic ring leads to an enhancement of the amineamine interactions. The HEm/(J 3 mol1) of mixtures including methanol changes in the sequence: 3150 (piperidine)72 < 711 (pyridine)73 < 170 (aniline),74 which may be ascribed in some extent to the fact that the positive contribution to HEm from the disruption of amineamine interactions upon mixing increases in the order aniline > pyridine > piperidine. HEm of morpholine + alkane systems is much higher than that of the corresponding piperidine solutions (Table 4). Moreover, E,∞ (piperidine + C6H12) HE,∞ 1 (morpholine + C6H12) > H1 (Table 6). It may be then concluded that interactions between morpholine molecules are stronger than those between piperidine molecules. These large differences between HEm and HE,∞ 1 values of morpholine and piperidine solutions cannot be merely explained by the larger effective dipole moment of morpholine (Table 8). Proximity effects related to the presence of the N and O atoms in the same molecule should be also taken into consideration, as such effects lead to an enhancement of dipolar interactions between morpholine molecules. This is supported by the higher TSEm values of morpholine mixtures. For the system with cyclohexane, at 308.15 K, GEm = 1029 J 3 mol1 (DISQUAC value) and TSEm = 823 J 3 mol1; higher values than that previously reported for the corresponding mixture with piperidine. It is remarkable that systems with the ether and amine groups placed in different molecules are characterized by lower HEm values (in J 3 mol1): 627 (diethylamine +1,4-dioxane75), 712 (diproylamine +1,4-dioxane75), 523 (butylamine +1,4-dioxane76).

Table 7. KirkwoodBuff Integrals, Gij, and Linear Coefficients of Preferential Solvation, δij, at Equimolar Compositions and Temperature T for Cyclic Amine(1) or Morpholine(1) + Organic Solvent(2) Mixtures. DISQUAC Values are Given in Parentheses system

T (K)

G11 (J cm3)

G22 (J cm3)

G12 (J cm3)

δ11 (J cm3)

δ12 (J cm3)

0.6 (13.9) 26.5 (23.6)

145.6 (138.2) 62.9 (64.6)

197 (207) 162.8 (160.6)

49.4 (55.2) 47.3 (46)

12.8 (17.2) 25 (24)

cyclohexylamine + n-C8a (c-CH2)4NH + C6H12b

333.15 298.15

(c-CH2)5NH + C6H12c

298.15

53 (46.4)

77.7 (72.1)

137.3 (143.4)

21.1 (24.3)

14.9 (17.8)

(c-CH2)5NH + n-C7d

298.42

62.5 (53)

105.1 (108.5)

203.3 (197.5)

66.4 (62.6)

24.5 (22.2)

(c-CH2)5NH + n-C14e

298.15

(180)

(281.2)

(204)

(96)

(19.3)

morpholine + n-C8f

353.35

447.1 (407.7)

63.1 (74.1)

369.8 (348.9)

204.2 (189.2)

76.7 (68.79)

methanol + (c-CH2)5NHg

298.15

52.6 (46.3)

121.7 (120.2)

30 (33.2)

5.6 (3.3)

22.9 (21.8)

ethanol + (c-CH2)4NHh

313.15

90.2 (86)

116.1 (114)

33.2 (36.1)

14.2 (12.5)

20.7 (19.5)

a

VLE, ref 84; VEm, ref 96. b VLE, ref 85; VEm = 0. c VLE, ref 85; VEm, ref 61. d VLE, ref 88; VEm, ref 62. e VEm, ref 62. f VLE, ref 89; VEm = 0. g VLE and VEm, ref 72. h VLE, ref 86; VEm = 0. 9814

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

Figure 1. VLE for cyclic amine(1) + alkane(2) mixtures at temperature T. Points are the experimental results: (b) pyrrolidine(1) + cyclohexane(2) at T = 313.15 K;85 (9) cyclohexylamine(1) + octane(2) at T = 333.15 K.84 Solid lines are the DISQUAC calculations.

Figure 2. VLE for morpholine(1) + cyclooctane(2) mixtures at temperature T. Points are the experimental results:89 (b) T = 353.15 K; (9) T = 383.15 K. Solid lines are the DISQUAC calculations.

Supermolecule calculations show the existence of proximity effects in morpholine.77 Thus, the different hydrogen bonding between water and the amine and ether groups in morpholine and in molecules that contain only one of these groups has been ascribed to the mentioned proximity effects.77 Similar trends are encountered in systems with 1-alkanols or isomeric 2-alkoxyethanols and alkane.47 The effective dipole moments of 2-alkoxyethanols are higher than those of the isomeric 1-alkanols47 (e.g., μ̅ (2-methoxyethanol) = 0.879; μ ̅ (1butanol) = 0.664). The larger μ ̅ values of 2-alkoxyethanols together with the presence of the O and OH groups in the same molecule lead to an enhancement of the dipolar interactions between 2-alkoxyethanol molecules, and to a weakening of the effects related to the intermolecular self-association

ARTICLE

Figure 3. HEm of cyclic amine(1) + cyclohexane(2) mixtures at 298.15 K. Points are the experimental results: (b) cyclohexylamine;90 (9) pyrrolidine;85 (2) aziridine.85 Solid lines are the DISQUAC calculations.

Figure 4. HEm of cyclic amine(1) + alkane(2) mixtures at temperature T. Points are the experimental results: (1) piperidine(1) + cyclohexane(2)85 (T = 298.15 K); (b) piperidine(1) + octane(2)92 (T = 303.15 K); (2) morpholine(1) + cyclohexane(2)95 (T = 308.15 K); (9) morpholine(1) + octane(2)92 (T = 303.15). Solid lines are the DISQUAC calculations.

of these compounds.47 As consequence, the upper critical solution temperatures of the 2-alkoxyethanol + heptane mixtures (319.74 K for the 2-methoxyethanol mixture78) are much higher than those of the systems containing the homomorphic 1-alkanols. In addition, the TSEm curves of the 2-alkoxyethanol mixtures are s-shaped as GEm and HEm are of the same order of magnitude,47 while TSEm values are large and negative for mixtures with 1-alkanols due to GEm ≈ 2HEm (see above). 5.1. Kirkwood-Buff Integrals. The positive δ11 and negative δ12 values encountered for cyclic amine + alkane systems indicate that interactions between amine molecules are preferred to those 9815

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

Figure 5. Molar excess functions at 298.15 K for the methanol(1) + piperidine(2) system. Points are the experimental results:72 (9) F = G, Gibbs energy; (b) F = H, enthalpy. Solid lines are the DISQUAC calculations.

ARTICLE

Figure 7. Linear coefficients of preferential solvation, δij, of cyclic amine(1) + organic solvent(2) mixtures at temperature T. Dashed lines are the experimental results (see Table 7): (a) δ11; (b) δ12 of morpholine(1) + octane(2) at 353.15 K; (c) δ11; (d) δ12 of ethanol(1) + pyrrolidine(2) at 313.15 K. Solid lines are the DISQUAC calculations.

Table 8. Physical Constants of Some Cyclic Amines and Morpholine Va (cm3 3 mol1) Pcb (bar) Tcc (K) μd (D)

Amine

115.02f

cyclohexylamine

j

426g g

̅μ

e

614.9h

1.33i

0.475

g

2.19i

1.162

aziridine

52

67

534.6

pyrrolidine

83.24k

56.1h

568.6h

1.58i

0.663

piperidine

99.37l

46.5h

594.1h

1.19i

0.457

N-methylpiperidine

122.30m

40g

585.7g

0.80i

0.277

morpholine

87.52n

54.7o

618o

1.56i

0.638

a

Molar volume at 298.15 K. b Critical pressure. c Critical temperature. d 2 1/2 Dipole moment. e Effective dipole moment μ ̅ = [μ NA/(4πεoVkBT)] where NA is the Avogadro’s number; εo is the permittivity of the vacuum; kB is the Boltzmann constant. f Reference 97. g Calculated according to Joback’s method, ref 98. h Reference 99. i Reference 100. j Reference 101. k Reference 102. l Reference 61. m Reference 103. n Reference 104. o Reference 98. Figure 6. CEP,m of piperidine(1) + n-alkane(2) systems at 298.15 K. Points are the experimental results:62 (2) heptane; (b) decane; (9) dodecane. Solid lines are the DISQUAC calculations.

between unlike molecules. The rather large positive δ11 values of the morpholine + cyclooctane mixture at 353.15 K are noticeable (Table 7, Figure 7), as reveal that strong interactions between morpholine molecules exist, even at high temperatures. On the other hand, interactions between amine molecules are increasingly preferred when the chain length of the alkane is increased in solutions with piperidine (Table 7). The mixtures examined including 1-alkanols are characterized by rather low |δij| values (Table 7, Figure 7), which are similar to those encountered for 1-alkanol + amide,32,56 or + pyridine79 systems. This has been interpreted assuming that molecules of the same kind have no tendency to form aggregates in the solution.32,56 Thus, despite the strong interactions between

unlike molecules (see the large and negative HEm of the methanol + piperidine system, Table 4, Figure 5), it may be concluded that the distribution of the molecules in the solution is nearly random. 5.2. The DISQUAC Interaction Parameters. Some comments on the dependence with the molecular structure of the interaction parameters are necessary. It is well-known that the segmentation into groups, for example, of cycloalkanes into c-CH2 groups is not strictly justified, and it seems more appropriate to treat each cyclic molecule as an independent entity.24 In the (c-CH2)uNH + cyclohexane series, this may explain in some extent the quite different interactions parameters for aziridine, and in piperidine + cycloalkane systems, the different CDIS cn,1 coefficient for the cyclopentane solution. Despite this, it is (l = 1,3) noteworthy that, for the former mixtures, the CQUAC cn,l coefficients are independent of the amine for u > 2; a behavior rather usual in the framework DISQUAC, as is encountered 9816

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

ARTICLE

Table A1. Physical Constants of Pure Amines Needed for KirkwoodBuff Calculationsa RPb amine

3

(10

cyclohexylamine pyrrolidine

1.05f 1.22i

piperidine

0.96l

morpholine

3 K)

n,o

0.955

kSc 1

(TPa )

kTe

CPd 1

(J 3 mol

1

3K )

(TPa1)

575.1g 626j

200h 160.2k

764 857

965j

177.4m

1061

o

164.8p

620

496

a

For cyclohexane, heptane, octane, methanol and ethanol data are taken from ref 99. b Isobaric thermal expansion ecoefficient. c Adiabatic compressibility. d Isobaric heat capacity. e Isothermal compressibility, kT = kS + (TRp2V)/(Cp,m). f Reference 97. g Reference 105. h Estimated according to the ChuehSwanson method in ref 98. i Reference 102. j Reference 106. k Reference 107. l Reference 108. m Reference 61. n Reference 109. o Reference 110. p Reference 111.

when investigating alcoholic solutions such as alkanol + alkane;27,4246 1-alkanol + cyclic ether,80 + amide,4951 or in alkoxyethanol + alkane mixtures.47 On the other hand, the CDIS cn,l (l = 2,3) coefficients remain also constant for u > 2. Thus, the different behavior of (c-CH2)uNH + cyclohexane systems is QUAC coefficients which decrease merely described by CDIS cn,1 and Ccn,2 with increasing u values. In the case of the enthalpic parameter, the observed variation may be ascribed to an increasing steric effect, which weakens dipolar interactions between amine molecules.81 The CDIS cn,1 variation is somewhat different to that encountered in many other applications, where the increase of the dispersive interaction parameters is typically ascribed to increasing inductive effects.81 For N-methylpiperidine + n-alkane mixtures, CDIS an,2 coefficients depend on the chain length of the n-alkane (Table 1). This reveals the existence of the so-called Patterson effect in such systems, which is ascribed to the order destruction of longer alkanes during the mixing process with nonpolar or weakly polar molecules, of globular or plate-like shapes28,54,82,83 (e.g., benzene, cyclohexane, triethylamine). In the case of piperidine solutions, the mentioned effect are of minor importance due to the large contribution to HEm from the breaking of the polar amineamine interactions. However, a different CDIS an,3 coefficient is needed for an accurate description of the CEP,m of the piperidine + tetradecane system. This merely reveals the close relation between CEP,m and the third interchange coefficients with the molecular structure of the considered solutions. The same trend has been encountered, e.g., when investigating 1-alkanol27 or pyridine10 + n-alkane systems in terms of DISQUAC.

6. CONCLUSIONS DISQUAC describes correctly a whole set of thermodynamics properties: VLE, GEm, HEm, CEP,m, HE,∞ 1 , and the main features of the KirkwoodBuff integrals for the investigated solutions. In (c-CH2)uNH + C6H12 mixtures, amineamine interactions become weaker when the size of the amines is increased. This is supported by the decrease of HEm, HE,∞ 1 , and of the effective dipole moment. Interactions between amine molecules are weakened when passing from a primary cyclic amine to an isomeric tertiary cyclic amine. The existence of an aromatic ring in the amine, as in aniline or pyridine, leads to stronger amineamine interactions. Proximity effects between the two groups of morpholine leads to stronger interactions between these molecules than those between piperidine molecules. In

systems with alkanes, interactions between amine molecules are preferred to those between unlike molecules. In piperidine + methanol and pyrrolidine + ethanol systems, the mixture structure is close to random mixing.

’ APPENDIX See Table A1 for the physical constants of the pure amines needed for the KirkwoodBuff calculations. ’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Fax: +34-983-423136. Tel: +34-983423757.

’ ACKNOWLEDGMENT The author gratefully acknowledges the financial support received from the Consejería de Educacion y Cultura of Junta de Castilla y Leon, under Project VA052A09 and from the Ministerio de Ciencia e Innovacion, under Project FIS201016957. ’ REFERENCES (1) Wolff, H.; Gamer, G. Hydrogen-bonding and complex formation of dimethylamine—Infrared investigations on NH stretching vibration bands. J. Phys. Chem. 1972, 76, 871. (2) Landeck, H.; Wolff, H.; G€otz, H. Two-constant model to describe amine and alcohol association from vapor pressure measurements. J. Phys. Chem. 1977, 81, 718. (3) Cibulka, I.; Tamura, K.; Nagata, I. Thermodynamics of associating component + saturated hydrocarbon mixtures at low pressures. IV. Correlation of vapour pressures and volumetric properties of some aliphatic amines and their mixtures with n-alkanes in terms of association. Fluid Phase Equilib. 1988, 39, 39. (4) Funke, H.; Wetzel, M.; Heintz, A. New applications of the ERAS model. Thermodynamics of amine + alkane and alcohol + amine mixtures. Pure Appl. Chem. 1989, 61, 1429. (5) Wolff, H.; Landeck, H.; Frerichs, H.-P.; Wolff, E. The association of normal and tertiary butylamine in mixtures with n-hexane according to isothermal vapour pressure measurements. Fluid Phase Equilib. 1995, 109, 245. (6) Kehiaian, H. V.; Sosnkwoska-Kehiaian, K. Thermodynamic properties of some binary mixtures of specifically interacting nonelectrolytes I. Aniline + toluene. Trans. Faraday Soc. 1966, 62, 838. (7) Mu~ noz, M. A.; Galan, M.; G omez, L.; Carmona, C.; Guardado, P.; Bal on, M. The pyrrole ring as hydrogen-bonding π-donor base: An experimental and theoretical study of the interactions of N-methylpyrrole with alcohols. Chem. Phys. 2003, 290, 69. (8) Megiel, E.; Kasprzycka-Guttman, T.; Jagielska, A.; Wroblewska, L. A theoretical and experimental study of association of pyridine. J. Mol. Struct. 2001, 569, 111. (9) Vinogradov, S. V.; Linnell, R. H. Hydrogen Bonding; Van Nostrand, Reinhold: New York, 1971. (10) Gonzalez, J. A.; Mozo, I.; García de la Fuente, I.; Cobos, J. C. Thermodynamics of organic mixtures containing amines. V. Systems with pyridine. Thermochim. Acta 2006, 441, 53. (11) Reynolds, R. W.; Smith, J. S.; Steinmetz, T. Methyl ethers as motor fuel components. Abstr. Pap. Am. Chem. Soc. 1974, 11. (12) Csikos, R.; Pallay, J.; Laky, J.; Radchenko, E. D.; Englin, B. A.; Robert, J. A. Low-lead fuel with MTBE and C4 alcohols. Hydrocarbon Process. Int. Ed. 1976, 55, 121. (13) Iida, T.; Endo, T.; Ito, M. M. Evaluation of enthalpies of cyclic alkenebenzene and etherbenzene interactions based on MO calculations. J. Phys. Org. Chem. 2000, 13, 330. 9817

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research (14) Kehiaian, H. V. Group contribution methods for liquid mixtures: A critical review. Fluid Phase Equilib. 1983, 13, 243. (15) Kehiaian, H. V. Thermodynamics of binary liquid organic mixtures. Pure Appl. Chem. 1985, 57, 15. (16) Kirkwood, J. G.; Buff, F. P. The statistical mechanical theory of solutions. 1. J. Chem. Phys. 1951, 19, 774. (17) Ben-Naim, A. Inversion of the KirkwoodBuff theory of solutions: Application to the water-ethanol system. J. Chem. Phys. 1977, 67, 4884. (18) Matteoli, E.; Lepori, L. Solutesolute interactions in water. II. An analysis through the KirkwoodBuff integrals for 14 organic solutes. J. Chem. Phys. 1984, 80, 2856. (19) Guggenheim, E. A. Mixtures; Oxford University Press: Oxford, 1952. (20) Gmehling, J.; Li, J.; Schiller, M. A modified UNIFAC model. 2. Present parameter matrix and results for different thermodynamic properties. Ind. Eng. Chem. Res. 1993, 32, 178. (21) Gmehling, J.; Lohmann, J.; Jakob, A.; Li, J.; Joh, R. A modified UNIFAC (Dortmund) model. 3. Revision and extension. Ind. Eng. Chem. Res. 1998, 37, 4876. (22) Gmehling, J.; Wittig, R.; Lohmann, J.; Joh, R. A modified UNIFAC (Dortmund) model. 4. Revision and extension. Ind. Eng. Chem. Res. 2002, 41, 1678. (23) Jakob, A.; Grensemann, H.; Lohmann, J.; Gmehling, J. Further development of modified UNIFAC (Dortmund): Revision and extension 5. Ind. Eng. Chem. Res. 2006, 45, 7924. (24) Tine, M. R.; Kehiaian, H. V. A comparative study of thermodynamic properties of n-alkane or cycloalkane mixtures with aliphatic, linear or heterocyclic molecules containing the same functional group. Fluid Phase Equilib. 1987, 32, 211. (25) Bondi, A. Physical Properties of Molecular Crystals, Liquids and Glasses; Wiley: New York, 1968. (26) Kehiaian, H. V.; Grolier, J.-P. E.; Benson, G. C. Thermodynamics of organic mixtures. A generalized quasichemical theory in terms of group surface interactions. J. Chim. Phys. 1978, 75, 1031. (27) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C.; Casanova, C. A characterization of the aliphatic/hydroxyl interactions using a group contribution model (DISQUAC). Ber. Bunsen. Phys. Chem. 1991, 95, 1658. (28) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C.; Casanova, C.; Ait-Kaci, A. Application of the zeroth approximation of the DISQUAC model to cyclohexane + n-alkane mixtures using different combinatorial entropy terms. Fluid Phase Equilib. 1995, 112, 63. (29) Marcus, Y. Preferential solvation in mixed solvents. X. Completely miscible aqueous co-solvent binary mixtures at 298.15 K. Monatsh. Chem. 2001, 132, 1387. (30) Zielkiewicz, J. Solvation of DMF in the N,N-dimethylformamide + alcohol + water mixtures investigated by means of the Kirkwood Buff integrals. J. Phys. Chem. 1995, 99, 4787. (31) Zielkiewicz, J. KirkwoodBuff integrals in the binary and ternary mixtures containing heptane and aliphatic alcohol. J. Phys. Chem. 1995, 99, 3357. (32) Zielkiewicz, J. Preferential solvation in (amide + alcohol) binary mixtures. Part 1. The N-methylpyrrolidinone + ethanol mixture at T = 313.15 K: Thermodynamic results and molecular dynamic calculation. Phys. Chem. Chem. Phys. 2003, 5, 1619. (33) Zielkiewicz, J. Solvation of amide group by water and alcohols investigated using the KirkwoodBuff theory of solutions. J. Chem. Soc., Faraday Trans. 1998, 94, 1713. (34) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C. Thermodynamics of mixtures with strongly negative deviations from Raoult’s law. Part 4. Application of the DISQUAC model to mixtures of 1-alkanols with primary or secondary amines. Comparison with Dortmund UNIFAC and ERAS results. Fluid Phase Equilib. 2000, 168, 31. (35) Kehiaian, H. V.; Tine, M. R.; Lepori, L.; Matteoli, E.; Marongiu, B. Thermodynamics of binary mixtures containing oxaalkanes. Part 3. Monoethers, polyethers, acetals, orthoesters and cyclic monoethers + n-alkanes or cyclohexane. Fluid Phase Equilib. 1989, 46, 131.

ARTICLE

(36) Kehiaian, H. V.; Porcedda, S.; Marongiu, B.; Lepori, L.; Matteoli, E. Thermodynamics of binary mixtures containing linear or cyclic alkanones + n-alkanes, or + cycloalkanes. Fluid Phase Equilib. 1991, 63, 231. (37) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C.; Casanova, C.; Kehiaian, H. V. Calorimetric and phase equilibrium data for linear carbonates + hydrocarbons, or + CCl4. Comparison with DISQUAC predictions. Thermochim. Acta 1993, 217, 57. (38) Kehiaian, H. V.; Gonzalez, J. A.; García, I.; Escarda, R.; Cobos, J. C.; Casanova, C. Prediction of liquidliquid equilibria and of enthalpies of mixing in alkanoic acid anhydride + n-alkane mixtures using DISQUAC. Fluid Phase Equilib. 1991, 69, 91. (39) Kehiaian, H. V.; Marongiu, B. A comparative study of thermodynamic properties and molecular interactions in mono- and polychloroalkane + n-alkane or + cyclohexane mixtures. Fluid Phase Equilib. 1988, 40, 23. (40) Soriano, M. J.; Velasco, I.; Otín, S.; Kehiaian, H. V. Thermodynamics of mixtures containing iodoalkanes. II. Excess enthalpies of mixtures of iodoalkane + cyclohexane, + benzene, or + tetrachloromethane. Measurement and analysis in terms of group contributions (DISQUAC). Fluid Phase Equilib. 1989, 45, 205. (41) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C.; Casanova, C. Thermodyanmics of mixtures containing linear monocarboxylic acids. I. DISQUAC predictions on molar excess Gibbs energies, molar excess enthalpies, and solid-liquid equilibria for mixtures of linear monocarboxylic acids with organic solvents. Fluid Phase Equilib. 1994, 99, 19. (42) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C.; Casanova, C. DISQUAC predictions of phase equilibria, molar and standard parcial molar excess quantities for 1-alkanol + cyclohexane mixtures. J. Solution Chem. 1994, 23, 399. (43) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C. Thermodynamics of branched alcohols. I. Extension of DISQUAC to tertalcohols-n-alkanes or tert-alcohols-cyclohexane mixtures. Fluid Phase Equilib. 1996, 119, 81. (44) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C. DISQUAC analysis of binary liquid organic mixtures containing cyclic or linear alkanols and cycloalkanes or n-alkanes. Thermochim. Acta 1996, 278, 57. (45) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C. Estimation of DISQUAC interchange energy parameters for linear secondary alcohols + n-alkanes, or cyclohexane mixtures. Fluid Phase Equilib. 1996, 123, 39. (46) Gonzalez, J. A.; García de la Fuente, I.; Cobos, J. C. Application of a purely physical model (DISQUAC) to binary mixtures of phenol and organic solvents. Ber. Bunsen. Phys. Chem. 1996, 100, 1746. (47) Gonzalez, J. A.; Cobos, J. C.; Carmona, F. J.; García de la Fuente, I.; Bhethanabotla, V. R.; Campbell, S. W. Thermodynamics of mixtures containing alkoxyethanols. Part XV. DISQUAC characterization of systems of alkoxyethanols with n-alkanes or cyclohexane. Phys. Chem. Chem. Phys. 2001, 3, 2856. (48) Gonzalez, J. A.; Domanska, U. Thermodynamics of mixtures containing a very strongly polar compound. Part I. Experimental phase equilibria (solidliquid and liquidliquid) for sulfolane + n-alkan-1-ols systems. Analysis of some mixtures including sulfolane in terms of DISQUAC. Phys. Chem. Chem. Phys. 2001, 3, 1034. (49) Gonzalez, J. A.; Domanska, U.; Lachwa, J. Thermodynamics of mixtures containing a very strongly polar compound. Part 3. DISQUAC characterization of NMP + organic solvent mixtures. Can. J. Chem. 2003, 81, 1451. (50) Gonzalez, J. A.; Cobos, J. C.; García de la Fuente, I. Thermodynamics of binary mixtures containing N-alkylamides. J. Mol. Liq. 2004, 115, 93. (51) Gonzalez, J. A.; Cobos, J. C.; García de la Fuente, I. Thermodynamics of liquid mixtures containing a very strongly polar compound. Part 6. DISQUAC characterization of N,N-dialkylamides. Fluid Phase Equilib. 2004, 224, 169. (52) Gonzalez, J. A.; Mozo, I.; García de la Fuente, I.; Cobos, J. C. Thermodynamics of organic mixtures containing amines. IV. Systems with aniline. Can. J. Chem. 2005, 83, 1812. 9818

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research (53) Wilhelm, E.; Inglese, A.; Grolier, J.-P. E. Excess enthalpies of binary mixtures containing either methylcyclopentane or methylcyclohexane. Thermochim. Acta 1991, 187, 113. (54) Riesco, N.; Gonzalez, J. A.; Villa, S.; García de la Fuente, I.; Cobos, J. C. Thermodynamics of organic mixtures containing amines. III. Excess molar volumes at 298.15 K for tripropylamine + n-alkane systems. Application of the Flory theory to N,N,N-trialkylamine + n-alkane mixtures. Phys. Chem. Liq. 2003, 41, 309. (55) Gonzalez, J. A.; Cobos, J. C.; Mozo, I.; García de la Fuente Thermodynamics of mixtures containing alkoxyethanols. Part XIX. Systems with linear monoethers or 1-alkanols. J. Mol. Liq. 2006, 126, 99. (56) Zielkiewicz, J. Preferential solvation of N-methylformamide, N, N-dimethylformamide and N-methylacetamide by water and alcohols in the binary and ternary mixtures. Phys. Chem. Chem. Phys. 2000, 2, 2925. (57) Marcus, Y. Preferential solvation in mixed solvents. 14. Mixtures of 1,4-dioxane with organic solvents: KirkwoodBuff integrals and volume-corrected preferential solvation parameters. J. Mol. Liq. 2006, 128, 115. (58) Allen, G.; Chai, Z.; Chong, C. L.; Higgins, J. S.; Tripathi., J. Thermodynamics of oligomeric binary mixtures of polyethylene glycol and polypropylene glycol methylethers. Polymer 1984, 25, 239. (59) Wilhelm, E. Thermodynamics of solutions: Selected aspects. Thermochim. Acta 1990, 162, 43. (60) Rowlinson, J. S.; Swinton, F. L. Liquid and Liquid Mixtures, 3rd ed.; Butterworths: Great Britain, 1982. (61) Laínez, A.; Wilhelm, E.; Grolier, J.-P. E. Thermodynamics of liquid mixtures containing hydrocarbons and strongly polar substances: VE and CEP of {pyridine or piperidine + cyclohexane} at 298.15 K. Monatsh. Chem. 1994, 125, 877. (62) Laínez, A.; Rodrigo, M. M.; Wilhelm, E.; Grolier, J.-P. E. Thermodynamics of liquid mixtures containing n-alkanes and strongly polar components: VE and CEP of mixtures containing either pyridine or piperidine. J. Solution Chem. 1992, 21, 49. (63) Sayegh, S. G.; Ratcliff, G. A. Excess Gibbs energies of binary systems of isopentanol and n-pentanol with hexane isomers at 25 °C. Measurement and prediction by analytical group solution model. J. Chem. Eng. Data 1976, 21, 71. (64) Nguyen, T. H.; Ratcliff, G. A. Heats of mixing of binary-systems of isopentanol and n-pentanol with hexane isomers at 25 °C. Measurement and prediction by analytical group solution model. J. Chem. Eng. Data 1975, 20, 256. (65) Velasco, I.; Otin, S.; Gutierrez Losa, C. Excess enthalpies of some aliphatic amine + n-hexane or benzene mixtures. J. Chim. Phys. 1978, 75, 706. (66) Pfestorf, R.; Kuchenbecker, D.; Quitzsch, K. Kalorische exzesscharakteristik fliissiger nichtelektrolytsysteme. IV. Experimentelle bestimmung und vorausberechnung mittlerer molarer mischungsenthalpien von n-alkylamin/n-alkan-systemen. Z. Phys. Chem. (Leipzig) 1982, 263, 233. (67) Paz-Andrade, M. I.; Sarmiento, F. Excess enthalpies of dipropylamine + normal alkane (C6C10) mixtures. Int. DATA Ser., Sel. Data Mixtures 1984, 110. (68) Bender, M.; Hauser, J.; Heintz, A. Themodynamics of the ternary mixture propan-1-ol + triethylamine + n-heptane. Experimental results and ERAS model calculations of HE and VE. Ber. Bunsen. Phys. Chem. 1991, 95, 801. (69) Nicolaides, G. L.; Eckert, C. A. Experimental heats of mixing of some miscible and partially miscible non-electrolyte systems. J. Chem. Eng. Data 1978, 23, 152. (70) Díez, D.; Ruiz, B.; Royo, F. M.; Gutierrez Losa, C. Excess molar enthalpies at 298.15 K of (cyclohexane + a methylpyridine) and of (benzene + a methylpyridine). J. Chem. Thermodyn. 1975, 17, 371. (71) Ochi, K; Momose, M.; Kojima, K. Determination of mutual solubilities in aniline + n-hexane and furfural + cyclohexane systems by a laser-light scattering technique. Can. J. Chem. Eng. 1993, 71, 982. (72) Nakanishi, K.; Wada, H.; Touhara, H. Thermodynamic excess functions of methanol + piperidine at 298.15 K. J. Chem. Thermodyn. 1975, 7, 1125.

ARTICLE

(73) Touhara, H.; Nakanishi, K. Excess molar enthalpies of methanol + pyridine, + methylpyridine, + dimethylpyridine. J. Chem. Thermodyn. 1985, 17, 909. (74) Nakanishi, K.; Touhara, H. Excess molar enthalpies of (methanol +aniline), (methanol + N-methylaniline), and (methanol + N,N-dimethylaniline). J. Chem. Thermodyn. 1986, 18, 657. (75) Letcher, T. M.; Govender, P. U.; Domanska, U. Excess molar enthalpies and volumes of diethylamine or dipropylamine + an ether at 298.15 K. J. Chem. Eng. Data 1999, 44, 274. (76) Letcher, T. M.; Goldon, A. Excess molar enthalpies of (butylamine + and ether) at 298.15 K. J. Chem. Eng. Data 1996, 41, 629. (77) Wu, H. S.; Sandler, S. I. Use of ab initio quantum mechanics calculations in group contribution methods. 1. Theory and the basis for group identifications. Ind. Eng. Chem. Res. 1991, 30, 881. (78) Carmona, F. J.; Gonzalez, J. A.; García de la Fuente, J. A.; Cobos, J. C. Thermodynamic properties of n-alkoxyethanols + organic solvent mixtures. X. Liquidliquid equilibria of systems containing 2-methoxyethanol, 2-(2-methoxyethoxy)ethanol or 2-(2-ethoxyethoxy)ethanol, and selected alkanes. J. Chem. Eng. Data 1999, 44, 892. (79) Gonzalez, J. A.; García de la Fuente, I.; Mozo, I.; Cobos, J. C. Thermodynamics of organic mixtures containing amines. VII. Study of systems containing pyridines in terms of the KirkwoodBuff formalism. Ind. Eng. Chem. Res. 2008, 47, 1729. (80) Gonzalez, J. A.; Mozo, I.; García de la Fuente, I.; Cobos, J. C.; Durov, V. A. Thermodynamics of 1-alkanol + cyclic ether mixtures. Fluid Phase Equilib. 2006, 245, 168. (81) Velasco, I.; Fernandez, J.; Otín, S.; Kehiaian, H. V. Estimation of DISQUAC interchange energy parameters for n-alkylamine + n-alkane mixtures. Fluid Phase Equilib. 1991, 69, 15. (82) Costas, M.; Patterson, D. Order destruction and order creation in binary mixtures of non-electrolytes. Thermochim. Acta 1991, 120, 161. (83) Lam, V. T.; Picker, P.; Patterson, D.; Tancrede, P. Thermodynamic effects or orientational order in chain-molecules mixtures. Part 1. Heats of mixing of globular and normal alkanes. J. Chem. Soc., Faraday Trans. 2 1974, 70, 1465. (84) Grenner, A.; Klauck, M.; Schmelzer, J. An equipment for dynamic measurements of vapourliquid equilibria and results in binary systems containing cyclohexylamine. Fluid Phase Equilib. 2005, 233, 170. (85) Cabani, S.; Ceccanti, N. Thermodynamic properties of binary mixtures of cyclohexane with cyclic amines or cyclic ethers at 298.15 K. J. Chem. Thermodyn. 1973, 5, 9. (86) Wu, H. S.; Locke, W. E., III; Sandler, S. I. Isothermal vaporliquid equilibrium of binary mixtures containing pyrrolidine. J. Chem. Eng. Data 1990, 35, 169. (87) Cabani, S.; Ceccanti, N.; Gianni, P. Liquidvapor equilibria and excess Gibbs energies of piperidine mixtures with cyclopentane, cycloheptane, or cyclooctane. Int. DATA Ser., Selec. Data Mixtures, Ser. A. 1977, 1, 43. (88) Sahki, D.; Belarabi, B. F.; Ait-Kaci, A.; Jose, J. Static measurements of the total pressure of binary mixtures of morpholine + heptane and piperidine or N-methylpiperidine + heptane or + decane between 273 and 353 K. ELDATA: Int. Electron J. Phys-Chem. Data 1999, 5, 85. (89) Wu, H. S.; Locke, W. E., III; Sandler, S. I. Isothermal vaporliquid equilibrium of binary mixtures containing morpholine. J. Chem. Eng. Data 1991, 36, 127. (90) Kern, M.; Abello, L.; Caceres, D.; Pannetier, G. Studies on NH 3 3 3 bonding. 14. Thermodynamical study of autoassociation of normal hexylamine. Energy of autoassociation bonds of cyclohexylamines and normal hexylamine. Bull. Soc. Chim. Fr. 1970, 11, 3849. (91) Velasco, I.; Otín, S.; Gutierrez Losa, C. Excess enthalpies of some cyclohexylamines + cyclohexane or benzene mixtures. J. Chim. Phys. Phys.-Chim. Biol. 1979, 76, 855. (92) Belaribi, B. F.; Belaribi-Boukais, G.; Ait-Kaci, A. Excess enthalpies of binary liquid mixtures of morpholine or piperidine + normal alkanes (heptane, octane, decane) at 303.15 K. ELDATA: Int. Electron J. Phys.-Chem. Data 1998, 4, 59. 9819

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820

Industrial & Engineering Chemistry Research

ARTICLE

(93) Sifaoui, H.; Ait-Kaci, A.; Rogalski, M. Determination des chaleurs de melange des systemes ternaires. Piperidine et hydrocarbures. J. Therm. Anal. Calorim. 2003, 73, 219. (94) Sahki, D.; Belaribi, B. F.; Ait-Kaci, A. Excess enthalpies of binary liquid mixtures of N-methylpiperidine + normal alkanes (hexane, heptane, octane, decane) at 303.15 K. ELDATA: Int. Electron J. PhysChem. Data 1999, 5, 39. (95) Belaribi, F. B.; Dahmonu, A.; Dahmani, A.; Boukais-Belaribi, G.; Mohammadi, A. H.; Richon, D. Experimental measurements and correlations of excess molar enthalpies for the binary and ternary mixtures of (cyclohexane, tetrahydropyran, and morpholine) or (cyclohexane, 1,4-dioxane, and morpholine) at 308.15 K and atmospheric pressure. J. Chem. Eng. Data 2010, 55, 2833. (96) Grenner, A.; Klauck, M.; Kramer, M.; Schmelzer, J. Activity coefficients at infinite dilution of cyclohexylamine + octane, toluene, ethylbenzene, or aniline and excess molar volumes in binary mixtures of cyclohexylamine + heptane, octane, nonane, decane, undecane, aniline, or water. J. Chem. Eng. Data 2006, 51, 176. (97) Kijevcanin, M. L.; Radovic, I. R.; Serbanovic, S. P.; Tasic, A. Z.; Djordjevic, B. D. Experimental determination and modelling of densities and excess molar volumes of ternary system (1-butanol + cyclohexylamine + n-heptane) and corresponding binaries from 288.15 to 323.15 K. Thermochim. Acta 2009, 496, 71. (98) Reid, R. C.; Prausnitz, J. M.; Poling, B. E. The Properties of Gases and Liquids, 4th ed.; McGraw Hill: N.Y., 1987. (99) Riddick, J. A.; Bunger, W. B.; Sakano, T. K. Organic solvents. In Techniques of Chemistry; Weissberger, A., Ed.; Wiley, N.Y., 1986; Vol. II. (100) McClellan, A. L. Tables of Experimental Dipole Moments; Rahara Enterprises: El Cerrito, US, 1974; Vol. 1,2,3. (101) Cabani, S.; Ceccanti, N.; Tine, M. R. Liquidvapor equilibrium, excess Gibbs energy, and excess enthalpy of some cyclic amine (C2-C7) + cyclohexane. Int. DATA Ser., Selec. Data Mixtures, Ser. A. 1986, 4, 185.  lvarez, E.; Gomez-Díaz, D.; La Rubia, D.; Navaza, J. M. (102) A Densities and viscosities of aqueous solutions of pyrrolidine and piperidine from 20 to 50 °C. J. Chem. Eng. Data 2005, 50, 1829. (103) Coulier, Y.; Ballerat-Busserolles, K.; Rodier, L.; Coxam, J.-Y. Temperatures of liquidliquid separation and excess molar volumes of {N-methylpiperidinewater} and {2-methylpiperidinewater} systems. Fluid Phase Equilib. 2010, 290, 206. (104) Park, S.-J.; Fischer, K.; Gmehling, J. Excess volumes for alkanol + morpholine systems at 298.15 and 308.15 K. J. Chem. Eng. Data 1994, 39, 859. (105) Roy, M. N.; Sinha, B.; Dakua, V. K. Excess molar volumes, viscosity deviations, and ultrasonic speeds of sound of binary mixtures of 2-butanone with some alkoxyethanols and amines at 298.15 K. J. Mol. Liq. 2007, 136, 128. (106) Gomez-Díaz, D.; Navaza, J. M. Speed of sound and isentropic compressibility of aqueous solutions of pyrrolidine and piperidine. J. Chem. Eng. Data 2006, 51, 722. (107) Conti, G.; Gianni, P.; Matteoli, E.; Mengheri, M. Capacita termiche molari di alcuni composti organici mono- e bifunzionali nel liquido puro e in soluzione acquosa a 25 °C (Molar thermal capacity of some mono-and bifunctional organic compounds in the pure liquid and in aqueous solution at 25 C). Chim. Ind. (Milan) 1976, 58, 225. (108) Kul, I.; Lieu, T. Thermodynamic properties of aqueous solutions of pyridine and piperidine. Fluid Phase Equilib. 2010, 290, 95. (109) Park, S.-J.; Fischer, K.; Gmehling, J. Excess volumes for alkanol + morpholine systems at 298.15 and 308.15 K. J. Chem. Eng. Data 1994, 39, 859. (110) Rama Rao, G. V.; Sarma, A. V.; Rambabu, C. Excess volumes, deviation in viscosities and compressibilities of binary mixtures consisting of morpholine, piperidine with methanol and pyridine at different temperatures. Ind. J. Pure. Appl. Phys. 2004, 42, 820. (111) Lyashkov, V. I. Thermal characteristics of morpholine. Izv. Vyssh. Uchebn., Khim. Khim. Tekhnol. 1980, 23, 1085.

9820

dx.doi.org/10.1021/ie201120h |Ind. Eng. Chem. Res. 2011, 50, 9810–9820