You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/225449601

An Equation for the Correlation of Viscosities of Binary Mixtures

Article  in  Journal of Solution Chemistry · February 2008


DOI: 10.1007/s10953-007-9226-2

CITATIONS READS

71 506

4 authors:

J. V. Herráez R. Belda
University of Valencia University of Valencia
42 PUBLICATIONS   798 CITATIONS    17 PUBLICATIONS   572 CITATIONS   

SEE PROFILE SEE PROFILE

OCTAVIO DIEZ SALES M. Herraez


University of Valencia University of Valencia
100 PUBLICATIONS   2,031 CITATIONS    51 PUBLICATIONS   756 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nanovesicles View project

Alternative methods View project

All content following this page was uploaded by J. V. Herráez on 01 June 2014.

The user has requested enhancement of the downloaded file.


J Solution Chem (2012) 41:1334–1351
DOI 10.1007/s10953-012-9878-4

The Relative Reduced Redlich–Kister and Herráez


Equations for Correlating Excess Properties
of N,N-Dimethylacetamide + Formamide Binary
Mixtures at Temperatures from 298.15 K to 318.15 K

D. Das · N. Ouerfelli

Received: 29 April 2011 / Accepted: 4 November 2011 / Published online: 22 August 2012
© Springer Science+Business Media, LLC 2012

Abstract Excess molar volumes and viscosity deviations in binary mixtures of N,N-di-
methylacetamide with formamide at (298.15, 308.15 and 318.15) K were calculated from
experimental density and viscosity data presented in previous work. The density and viscos-
ity data as well as their corresponding derived functions were used to test the applicability
of two correlative equations: the reduced Redlich–Kister equation and the recently proposed
Herráez equation. Their correlation abilities at different temperatures, and the use of differ-
ent numbers of parameters, are discussed for the case of limiting experimental data. These
relative functions are important to reduce the effect of temperature and, consequently, to
reveal the effects of different types of interactions. Correlation between the two Arrhenius
parameters of viscosity over the composition domains shows the existence of two distinct
behaviors.

Keywords Density and viscosity correlation · Binary liquid mixture ·


N,N-dimethylacetamide · Formamide · Reduced Redlich–Kister equation ·
Herráez equation

1 Introduction

Among the physico-chemical properties of liquid mixtures, density and viscosity are valu-
able tools for developing new theoretical models to understand the liquid state. On the other
hand, excess thermodynamic functions and deviations of analogous non-thermodynamic
functions of binary liquid mixtures are fundamental for understanding different types of
intermolecular interactions in these mixtures.

D. Das
Department of Chemistry, Dinhata College, Dinhata - 736135, Cooch-Behar, West Bengal, India
e-mail: debu_nbu@rediffmail.com

N. Ouerfelli ()
Laboratoire de Chimie Analytique et Electrochimie, Département de Chimie, Faculté des Sciences
de Tunis, Campus Universitaire, 2092 El Manar, Tunisia
e-mail: nouerfelli@yahoo.fr
J Solution Chem (2012) 41:1334–1351 1335

This paper is a continuation of our earlier work that includes the study of the binary liquid
mixtures of N,N-dimethylacetamide (DMA) + formamide (FA) [1]. In this previous paper
the measured density and viscosity values were used to calculate the excess properties and
then fitted to the Redlich–Kister equation. Here, density and viscosity deviations are fitted
to the Herráez and the reduced Redlich–Kister (R-K) equations [2–4], through Legendre
polynomials, and are interpreted in terms of molecular interactions and structural effects.
Comparison between the two models at different temperatures, and the effect of using dif-
ferent numbers of parameters, will be discussed. Note that similar comparisons were made
in previous papers for isobutyric acid + water mixtures both near to, and far away from,
the critical temperatures, and for the 1,4-dioxane + water system for different temperatures
[5–8]. In addition, the reduced R-K function relative to the viscosity Arrhenius activation
energy, which is equivalent to an apparent molar property, is more sensitive than the excess
property to interactions and gives more information. We have noted that, at infinite dilution
of one component the Herráez exponential polynomials’ values converge to the universal
exponent 0.5 independent of temperature, which was justified [7, 8] and considered to be an
indicator of predominantly solute–solute interactions at infinite dilution.

2 Experimental Details

2.1 Purification of the Chemicals

N,N-dimethylacetamide (DMA, SRL, India) was shaken with CaO (A.R., B.D.H.) for 1–2 h,
then decanted and distilled twice; the middle fraction was collected and used. Its density
(0.93644 g·cm−3 ) and viscosity (0.93700 mPa·s) at 298.15 K compared well with the lit-
erature values, which are 0.9366 g·cm−3 and 0.919 mPa·s, respectively [9]. Formamide
(E. Merck, >99%) was dried over freshly ignited quicklime for several hours and then dis-
tilled over reduced pressure, and the purified solvent had a density of 1.1254 g·cm−3 at
298.15 K and viscosity of 2.9527 mPa·s also at 298.15 K. All these values are in good
agreement with the literature values [9].

2.2 Apparatus and Procedure

Densities were measured with an Ostwald–Sprengel type pycnometer having a bulb volume
of 25 cm3 and an internal capillary diameter of about 1 mm. The pycnometer was calibrated
at 298.15, 308.15 and 318.15 K with doubly-distilled water. The kinematic viscosities were
measured by means of a suspended-level Ubbelohde viscometer with a flow time of about
539 s for distilled water at 25 °C. The time of efflux was measured with a stopwatch capable
of recording to 0.1 s. The methods and techniques have been described in detail in our
earlier papers [1, 10]. The excess molar volume, viscosity deviation, enthalpies of activation
of viscous flow and Arrhenius parameters have been investigated from density and shear
viscosity measurements of DMA–FA mixtures over the entire range of mole fractions with
20 composition points at the temperatures 298.15, 308.15 and 318.15 K.

3 Empirical Equations

The excess molar volume V E and viscosity deviation values η were correlated with the
mole fraction using two procedures. First, the R-K [2] expression, as described below, was
1336 J Solution Chem (2012) 41:1334–1351

fitted with least-squares polynomials. Secondly, the recently proposed Herráez [3] correla-
tion equation that uses an exponential polynomial as described below, and which is a func-
tion of the mole fraction x1 of DMA, was fitted at the three different temperatures (298.15,
308.15 and 318.15) K. Note that these models are used for nonelectrolytes solution mixtures
having nonideal behavior. For all approaches, the goodness of fit is discussed as a function
of the number of the adjustable parameters for the different temperatures.

3.1 Redlich–Kister Equation

3.1.1 The Reduced Redlich–Kister Excess Function

In an earlier paper [1] the measured density and viscosity values of DMA and FA binary
mixtures were used to calculate the excess molar volumes and viscosity deviations of this
system and then these values were fitted to the Redlich–Kister equation [2]:

p=M
YjE = x1 (1 − x1 ) Ap,T (2x1 − 1)p (1)
p=0

where YjE denotes V E and η, and (M + 1) is the optimal number of parameters. The
coefficients Ap,T in Eq. 1 are regressed by employing the least-squares fit method. The
corresponding standard deviation σ (Y ) was calculated by Eq. 2:

i=N
 Yi,exp − Yi,calc
σ (Y ) =  (2)
i=1
N −M

where N and M are the numbers of data points and adjustable parameters, respectively.
In our previous paper [1], we have already reported the variation of excess molar vol-
umes (V E ) and the viscosity deviations η with the mole fraction of the solvent (x1 ). In
the same context, Desnoyers and Perron [4] showed that an examination of the trends of
the dependence of V E and η on x1 suggests that many of these systems are similar, and
the differences in interactions are significant mainly in solutions very rich in component 2.
From the treatment of excess thermodynamic quantities for liquid mixtures proposed by
Desnoyers and Perron [4], we can conclude that the excess quantity (V E or η) gives an
overall view of the origin of nonideality in the mixtures but still can be quite misleading,
especially for systems that show strong interactions at high dilution. Desnoyers and Perron
[4] suggested that, in agreement with the original statements of R-K [2], it is better to use
the ratio V /x1 (1 − x1 ) or η/x1 (1 − x1 ) for this purpose.
This ratio is the so called the experimental reduced R-K excess property QY,exp,T (x1 ) and
is expressed by Eq. 3,
 
QY,exp,T (x1 ) = YjE / x1 (1 − x1 ) (3)
where YjE denotes V E or η. Also, excess thermodynamic quantities have the advantage
of illustrating the sign and magnitude of their nonideality, but the reduced R-K functions
V E /x1 (1 − x1 ) or η/x1 (1 − x1 ) give a much better handle on the origin of this nonide-
ality. We note that the reduced Redlich–Kister excess property is more sensitive than the
corresponding direct excess property V E or η to interactions that occur at low concen-
trations [4, 11]. We note that Redlich and Kister treated data with a power series, putting
greater weight on data near 0.5 mole fraction [4]. As Desnoyers et al. have shown, this is
not always the best approach for mixtures having specific interactions such as association at
J Solution Chem (2012) 41:1334–1351 1337

Table 1 Redlich–Kister expression given by Eq. 1 and Legendre polynomials given by Eq. 4 as a function
of polynomial order p

Polynomial Redlich–Kister, Legendre polynomial,


order, p (2x1 − 1)p , Eq. 1 Lp (2x1 − 1), Eq. 4

0 1 1
1 2x1 − 1 2x1 − 1
2 4(x12 − x1 + 1/4) 6(x12 − x1 + 1/6)
3 8(x13 − 3/2 · x12 + 3/4 · x1 − 1/8) 20(x13 − 3/2 · x12 + 3/5 · x1 − 1/20)
4 16(x14 − 2 · x13 + 9/6 · x12 − 1/2 · x1 + 1/16) 70(x14 − 2 · x13 + 9/7 · x12 − 2/7 · x1 + 1/70)

Table 2 Values of the Legendre polynomial constants ap for the reduced Redlich–Kister excess volume
and viscosity deviation (Eq. 4), QY,T (x1 ), at various temperatures and the corresponding standard deviations
σ (Y ) for DMA (1) + FA (2) mixtures

T /K a0 a1 a2 a3 a4 σ (Y )

V E / 10−6 m3 ·mol−1
298.15 −2.30041 −0.77665 0.05588 −0.00575 −0.07842 0.00743
308.15 −2.41894 −0.71521 0.08201 −0.03098 −0.04681 0.00509
318.15 −2.79808 −0.58259 0.09807 −0.02816 0.00108 0.00295

η / mPa·s
298.15 4.84854 −5.68009 0.35106 1.11908 −0.09509 0.00878
308.15 3.40019 −3.79587 0.03691 0.77994 −0.11323 0.01106
318.15 2.54241 −2.79721 0.13420 0.47198 −0.01768 0.00212

low concentration. Therefore, elimination of the factor [x1 (1 − x1 )] in the reduced R-K ex-
cess function QY,exp,T (x1 ) (Eq. 3) removes this effect and gives a specific reduced function
QY,exp,T (x1 ) characterizing the viscosity or other properties that also gives evidence for the
existence of important interactions.
Moreover, the reduced R-K polynomials QY,exp,T (x1 ) were fitted with a least-squares
optimization procedure to a series of Legendre polynomials Lp (2x1 − 1) using the Origin-
Pro 7.5 program:

p=M
QY,exp,T (x1 ) = ap,T Lp (2x1 − 1) (4)
p=0

The five first terms of this equation with p = 0, 1, . . . , 4 are presented in Table 1.
We note that Legendre polynomials can be generated using Gram–Schmidt orthonor-
malization in the interval [−1, 1] with the weighting function 1 (unit function) [12]. The
Legendre polynomials are the basis for the set of polynomials that are appropriate for the
interval [−1, 1].
As mentioned by Peralta et al. [13–16], Legendre polynomials belong to the category
of orthogonal functions such as Fourier, Bessel, Hermite and Chebyshev, which have the
valuable feature that, for a continuous series of observations, the values of the coefficients
do not change appreciably as the number of terms in the series is increased (Tables 1 and 2).
This is an important property because if a physical meaning can be assigned to one of its
1338 J Solution Chem (2012) 41:1334–1351

Fig. 1 Experimental reduced


Redlich–Kister excess properties
QV ,T (x1 ) = V E /x1 (1 − x1 ) of
the excess molar volume for
DMA (1) + FA (2) mixtures
against the mole fraction x1 in
DMA at the temperatures:
("), 298.15 K; (!), 308.15 K;
(Q), 318.15 K

Fig. 2 Experimental reduced


Redlich–Kister excess properties
QV ,T (x1 ) = η/x1 (1 − x1 )
(mPa·s) of the viscosity deviation
for DMA (1) + FA (2) mixtures
against the mole fraction x1 in
DMA at the temperatures:
("), 298.15 K; (!), 308.15 K;
(Q), 318.15 K

coefficients, then its value remains practically constant for the various orders of fits. For the
case of discrete measurements, such as the determination of volumes of mixing and viscosi-
ties of mixtures, the values of the coefficients vary, but only slightly [13–16]. In addition,
as shown in Table 1, the series of Legendre polynomials have the important characteristic
that the structure of its first five terms is practically the same as that of the first five terms
of the Redlich–Kister expression. The mathematical procedure to transform a power expan-
sion, such as that of Redlich–Kister, into an orthogonal series has been described in detail
by Tomiska [17–19]. Tomiska provides the iteration formula for Legendre or Chebyshev’s
series of any order as well as proof that the procedure is independent of the conversion
coefficients from the actual excess property [13–16].
For the system studied in this work, the experimental reduced excess molar volumes and
viscosity deviations QY,exp,T (x1 ), Eq. 3, are plotted in Figs. 1 and 2 versus the mole fraction
x1 of DMA. Changes in curvature found in both Figs. 1 and 2 are due to hydrogen bond-
ing between unlike molecules of the mixtures leading to strong intermolecular correlations.
J Solution Chem (2012) 41:1334–1351 1339

Fig. 3 Experimental relative


reduced Redlich–Kister excess
properties Qrel,V ,T (x1 ) for the
ratio = QV ,T (x1 )/V of the
excess molar volume for
DMA (1) + FA (2) mixtures
against the mole fraction x1 in
DMA at the temperatures:
("), 298.15 K; (!), 308.15 K;
(Q), 318.15 K

The decrease in QV ,exp,T (x1 ) with increasing temperature (Fig. 1) is explained by consid-
ering the differences in the molar volumes of the two liquids at different temperatures, and
this raises the possibility of the smaller sized molecules of FA fitting into the voids created
by the larger molecules of DMA.

3.1.2 The Reduced Redlich–Kister Viscosity Deviation

Although DMA + FA mixtures are nonelectrolyte solutions, we can adopt an expansion


equivalent to the Debye–Hückel and Jones–Dole expressions [20–24] for very dilute non-
electrolyte solutions where the first parameter a1 (of Eq. 5) is related to solute–solute inter-
actions while the second parameter a2 is related to solute–solvent interactions. Generally,
the experimental reduced R-K excess viscosity (Eq. 3) Qη,exp,T (x1 ) increases exponentially
and diverges at infinite dilution (as x1 → 0) in several binary mixtures [4–8, 25–28].
 1/2 
η = η2 1 + a1 x1 + a2 x1 + · · · (5)
We note that, at higher temperatures, there will be a competition between molecular in-
teractions and thermal agitation. On the other hand, the absence of a divergence of the
experimental reduced R-K viscosity deviation (Fig. 2) at infinite dilution of FA in DMA
(x1 → 1) gives evidence for a negligible a1 parameter. In fact, for nonelectrolyte solutions,
the absence of solute–solute interactions leads us to consider that the a1 value is always
zero. We can also state that the large separation of curves near x1 ≈ 0, and their closeness
near x1 ≈ 1, suggest that the solute–solvent parameter a2 (Eq. 5), which is equivalent to
the B parameter of the Jones–Dole equation [20–24], is much larger for DMA in FA than
for the reverse case. Like the relative deviation η/η, we consider the relative reduced
R-K function, QY,rel,T (x1 ) = QY,T (x1 )/Y . The plots of QY,rel,T (x1 ) versus mole fraction x1
(Figs. 3 and 4) show that the curves draw close together and QY,rel,T (x1 ) seems to depend
very slightly on temperature. We conclude that the nearness and then the larger separation
of curves in Fig. 2 at the two limits of infinite dilution do not depend only on the value of the
solute–solvent parameter a2 , but are also affected by temperature. We remark that the differ-
ences between their Arrhenius activation energies of viscosity Ea can play an important role
1340 J Solution Chem (2012) 41:1334–1351

Fig. 4 Experimental relative


reduced Redlich–Kister excess
properties Qrel,η,T (x1 ) for the
ratio = Qη,T (x1 )/η of the
viscosity deviation for DMA (1)
+ FA (2) mixtures against the
mole fraction x1 in DMA at the
temperatures: ("), 298.15 K;
(!), 308.15 K; (Q), 318.15 K

in this effect. Note that the relative reduced R-K function QY,rel,T (x1 ) is also a good tool,
like the reduced R-K function QY,T (x1 ), for interpreting different types of interactions. We
add that the minimization of the temperature effect shows a decrease of the relative reduced
R-K excess molar volume QV ,rel,T (x1 ) against mole fraction x1 in Fig. 3, while the reduced
R-K excess molar volume QV ,T (x1 ) shows an increase against mole fraction x1 in Fig. 1.
This effect is confirmed by Figs. 2 and 4, in which the temperature affects the FA-rich region
more than the DMA-rich one.
In addition, values of reduced R-K functions QY,T (xi ) at infinite dilution represent values
equivalent to the partial excess physical magnitudes at infinite dilution (x1 → 0+ ) [29],
which can be also calculated from the adjustable parameters using

Qη,T (x1 = 0) = A0,T − A1,T + · · · + (−1)p · Ap,T + · · · + (−1)n · An,T


= a0,T − a1,T + · · · + (−1)p · ap,T + · · · + (−1)n · an,T
   
= (η1 − η2 ) · 1 + ∂ ln(η − η2 )/∂x1 T ,x =0 (6)
1

and

Qη,T (x1 = 1) = A0,T + A1,T + · · · + Ap,T + · · · + An,T


= a0,T + a1,T + · · · + ap,T + · · · + an,T
   
= (η1 − η2 ) · 1 − ∂ ln(η − η2 )/∂x1 T ,x (7)
1 =1

3.1.3 The Limiting Excess Molar Volumes

As discussed in the literature [13–16], the variation of QV ,exp,T (x1 ) = V E /x1 (1 − x1 ) (Eq. 3)
with composition was used in each case to test the quality of the data; this function is ex-
tremely sensitive to experimental errors, particularly in the dilute ranges. In addition, its
values at infinite dilution represent the values of the partial excess molar volume at infi-
E,∞
nite dilution, V i [29], which can be also calculated from the adjustable parameters using
Eqs. 8 and 9:
J Solution Chem (2012) 41:1334–1351 1341

Table 3 Values of the limiting


/ 10−6 m3 ·mol−1 / 10−6 m3 ·mol−1
E,∞ E,∞
excess partial molar volume at T /K V1 V2
infinite dilution calculated from
the Redlich–Kister expression 298.15 −1.5405 −3.1054
(Eq. 1) or Legendre expression 308.15 −1.6375 −3.1299
(Eq. 4) for DMA (1) and FA (2)
at different temperatures 318.15 −2.0882 −3.3097

QV ,T (x1 = 0) = A0,T − A1,T + · · · + (−1)p · Ap,T + · · · + (−1)n · An,T


= a0,T − a1,T + · · · + (−1)p · ap,T + · · · + (−1)n · an,T
E,∞ ∞
=V1 = V 1 − V1 (8)
and

QV ,T (x1 = 1) = A0,T + A1,T + · · · + Ap,T + · · · + An,T


= a0,T + a1,T + · · · + ap,T + · · · + an,T
E,∞ ∞
=V2 = V 2 − V2 (9)
for the R-K expression and for the Legendre polynomial. In Eqs. 8 and 9, Vi is the molar
E,∞
volume of pure component ‘i’. The values of V i are listed in Table 3. We note that Eqs. 6
and 7 or 8 and 9 yield the same limiting values of QY,T .
As discussed in literature [4, 26–30], it can also readily be shown that, for many proper-
ties such as enthalpies, heat capacities, volumes, compressibilities and expansibilities, this
quantity (QY,T (x1 )) is directly related to the apparent molar quantities of both components:

(Y2,ϕ − Y2 ) Y1,ϕ − Y1
QY,T (x1 ) = Y E /x1 (1 − x1 ) = = (10)
x1 x2
This function is therefore equivalent to an apparent molar quantity over whole mole frac-
tion range, and its extrapolation to x1 = 0 and x1 = 1 {parameters sum QY,T (x1 = 0) and
E,∞
QY,T (x1 = 1) of Eqs. 8 and 9} will give the two excess partial molar quantities Y i (where
i = 1 or 2). These two limiting parameters are of fundamental importance since they are, by
definition, measures of the solute–solvent interactions of both components.
It is therefore important to determine precisely the R-K interaction parameters, and es-
pecially those corresponding to x1 = 0 and x1 = 1. Desnoyers and Perron [4] suggested that
one of the simplest ways of determining these parameters is from a plot of Y E /x1 (1 − x1 ) (or
dY E /dx1 for Gibbs energies). Note that the R-K excess quantity (Eq. 1) cannot appropriately
represent the composition dependence of the viscosity.
To this fact, we add that the partial molar volumes can be obtained in three other man-
ners:
(i) A local extrapolation by fitting the curves QV ,T (x1 ) separately in each rich region of
one component (1) or (2). The values obtained are more reliable than those given from
Eqs. 8 and 9.
(ii) The fit of the molar volume V (x1 ) against the mole fraction (x1 ) gives directly the
partial molar volume V i of the component ‘i’ by the following equation:

∂V (xj )
V i = V (xi ) − xj · (11)
∂xj T ,p
1342 J Solution Chem (2012) 41:1334–1351

When we proceed to operations at the limits of Eq. 11 at infinite dilution (x1 → 0+ or


x1 → 1− ), we can easily obtain the values of the excess partial molar volume at infinite
E,∞
dilution of component ‘i’ in the other one (V i ):

E,∞ ∞ ∂V (x1 )
V 1 = V 1 − V1 = −(V1 − V2 ) + (12)
∂x1 T ,p,x1 =0

and

E,∞ ∂V (x1 )
V2 = V2∞ − V2 = (V1 − V2 ) − (13)
∂x1 T ,p,x1 =1

(iii) We note that in recent work, Belda [31] proposed a new empirical correlation equation
for four properties (density, viscosity, surface tension, and refractive index), Eq. 14,
which introduces a correcting factor to linearity as a homographic function acting upon
the mole fraction of one component of the binary mixture (x1 ):

1 + m1 (1 − x1 )
YB (x1 ) = Y2 + (Y1 − Y2 ) · x1 · (14)
1 + m2 (1 − x1 )
where YB (x1 ) is the mixture property, and m1 and m2 are the two introduced adjustable
empirical parameters.
In previous papers [26, 28] we discussed the validity of this equation for some physico-
chemical properties of the 1,4-dioxane–water and isobutyric acid–water mixtures, and we
concluded that this equation can give better results than the Redlich–Kister equation with
the same number of free parameters. We have also given physical significance [27] to their
corresponding parameters (m1 and m2 ). In fact, when we proceed to some operations of
limits and derivations on the Belda equation, by inserting Eq. 14 into Eqs. 12 and 13, we
can easily obtain the values of partial excess property at infinite dilution of component ‘i’
E,∞
in the other component (Y i ) through the following equations:

E,∞ ∞ m1 − m2
V 1 = V 1 − V1 = · (V1 − V2 ) (15)
1 + m2
and
E,∞ ∞
V2 = V 2 − V2 = (m1 − m2 ) · (V1 − V2 ) (16)
We note that in the case of DMA + FA mixtures, all of these three techniques give practically
E,∞
the same values of the two limiting excess partial molar volumes (V i ) as those calculated
by the Redlich–Kister parameters (Eqs. 8 and 9, Table 3).

3.1.4 Viscosity Temperature Dependence

It is found that the temperature dependence of viscosity can be fitted with an Arrhenius
equation as:

η = AS exp(Ea /RT ) (17)


where R, Ea and AS are, respectively, the gas constant, the Arrhenius activation energy and
the pre-exponential (entropic) factor of the Arrhenius equation for the mixtures. Equation 17
can be rewritten in the logarithmic form:

ln η = ln AS + Ea /RT (18)
J Solution Chem (2012) 41:1334–1351 1343

Fig. 5 Logarithm of shear


viscosity ln(η) for the system of
DMA (1) + FA (2) mixtures
versus the reciprocal of the
absolute temperature at some
fixed mole fractions x1 in the
temperature range (298.15 to
318.15) K: ("), x1 = 0.0;
(!), x1 = 0.0700;
(Q), x1 = 0.2053;
(P), x1 = 0.3563;
(2), x1 = 0.5011;
(1), x1 = 0.6500;
(F), x1 = 0.7948; and
(E), x1 = 1.0

Fig. 6 Arrhenius activation


energy Ea and enthalpy of
activation of viscous flow H ∗
for DMA (1) and FA (2) mixtures
versus the mole fraction x1 of
N,N-dimethylacetamide in the
temperature range (298.15 to
318.15) K: ("), Ea (kJ·mol−1 );
(!), H ∗ (kJ·mol−1 )

However, the plot of the logarithm of shear viscosity ln(η) against the reciprocal of absolute
temperature (1/T ) is practically linear (Fig. 5) and the Arrhenius parameters Ea and AS are
thus independent of temperature over the studied temperature range (298.15 to 318.15) K.
Using both graphical and least-squares fitting methods, the slope of the straight line is equal
to Ea /R and the intercept is equal to ln(AS ). Calculated values of Ea are reported in Table 4
and are plotted against the mole fraction of DMA (1) in Fig. 6.
Nevertheless, values of ln AS (x1 ) permit us to calculate the pre-exponential factor (AS ) in
µPa·s, which are reported in Table 4. Since the AS values are closely related to the viscosity
of the system in the vapor phase [32], we observe high values of viscosity in the DMA-
rich region and we can assume that bonding of the DMA molecules is more correlated and
ordered in the vapor state. We can predict that in vapor–liquid equilibrium, the values of the
1344 J Solution Chem (2012) 41:1334–1351

Table 4 Arrhenius activation energy Ea (kJ·mol−1 ), entropic factor of Arrhenius AS (10−6 Pa·s), and the
enthalpy of activation of viscous flow H ∗ (kJ·mol−1 ) for {DMA (1) and FA (2)} mixtures as a function of
the mole fraction for the R-K equation (x1 ) over the temperature range (298.15 to 318.15) K

x1 Ea / kJ·mol−1 AS / 10−6 Pa·s H ∗ / kJ·mol−1

0.0000 16.41 3.95 15.769


0.0350 17.25 3.05 16.622
0.0700 17.97 2.44 17.349
0.1110 18.57 2.04 17.958
0.1500 19.07 1.74 18.470
0.2053 19.47 1.53 18.870
0.2414 19.62 1.45 19.027
0.3009 19.65 1.42 19.044
0.3563 19.52 1.44 18.906
0.4177 19.09 1.60 18.458
0.4580 18.73 1.74 18.095
0.5011 18.15 2.05 17.504
0.5474 17.57 2.35 16.921
0.5972 16.81 2.87 16.148
0.6500 15.85 3.73 15.182
0.7092 14.81 4.87 14.129
0.7500 14.03 6.00 13.334
0.7948 13.26 7.29 12.554
0.8910 11.70 10.82 10.974
1.0000 9.884 17.30 9.1447

mixing enthalpy must have high positive values in this region. In addition, the DMA + FA
interaction has an H-bound character [1] and the observed maximum of Ea (Fig. 6) reflects
the strong energy of hydrogen bonds between molecules [33].
We can add that, by inspection of data from Table 4, the entropic factor AS starts out with
elevated values in the very DMA-rich region (AS = 17.30 at x1 = 1.0), while in the FA-rich
region it takes on low values (AS = 3.95 at x1 = 0.0). This is due to the presence of an
FA–FA complex and to the modification of the geometry and stability of the system [1, 34].
On the other hand, in the case of the liquid phase we can add that the absolute reaction
rate theory of Eyring and John [35] and Ali et al. [36] relates the kinematic viscosity with
the Gibbs energy (G∗ ) of activation of viscous flow:

hNA ρ G∗
η= exp (19)
M RT
where η, h, R, NA , ρ and M are the dynamic viscosity of binary mixtures, Plank’s constant,
universal gas constant, Avogadro’s number, density and average molar mass of the mixture
(Eq. 21), respectively, and:

G∗ = H ∗ − T S ∗ (20)
M = x1 (M1 − M2 ) + M2 (21)
where M1 and M2 are the molar masses of the pure component, DMA and FA, respectively.
J Solution Chem (2012) 41:1334–1351 1345

Fig. 7 Correlation between the


Arrhenius activation energy
Ea (kJ·mol−1 ) of viscosity and
the logarithm of the Arrhenius
entropic factor ln AS for
DMA (1) and FA (2) mixtures in
the temperature range (298.15 to
318.15) K

By assuming that the activation parameters H ∗ and S ∗ [36, 37] are independent of
temperature, we obtained for each mixture composition (x1 , x2 ) these parameters when ln(η ·
M/(h · NA · ρ)) was plotted against 1/T . Using both graphical and least-squares fitting
methods, the slope that is equal to H ∗ /R and the intercept or ordinate that is equal to
−S ∗ /R. We can then use Eq. 20 to determine the activation parameters H ∗ and S ∗ .
Values of the enthalpy of activation of viscous flow, H ∗ , are presented in Table 4.
Inspection of the Ea values and those of the enthalpy (H ∗ ) of activation of viscous flow
(Table 4, Fig. 6) shows that the Ea and H ∗ values are very closely related, which is also
observed for 1,4-dioxane + water mixtures in previous works [27, 32]. In this context, the
variation of the Arrhenius activation energy Ea versus molar fraction (x1 ) gives evidence
for the eventual change in the complex structure or cluster formed in binary liquid mixtures
from the temperature effect and/or mixture composition [27, 32, 38, 39]. Thus, in the DMA-
rich region, the rapid decrease of Ea shows a changing of structure for FA.
In fact, it is shown in Fig. 6 that the Ea values increase with increasing mole fraction of
DMA in the formamide-rich region below the composition x1 = 0.3009, showing an effect
on Ea resulting from strong interactions between the FA and DMA molecules in the vicinity
of x1 = 0.3009. After that, Ea (x1 ) decreases with increases of the mole fraction of DMA (1).
This fact suggests that structural order is not much destroyed by the activation process;
consequently bonds among the associated molecules are not broken very much to form
smaller clusters (3DMA:7FA). We conclude that there are two association-type structures
limited by a composition around x1 ≈ 0.3 of DMA. These distinct behaviors are clearly
shown when the correlation between disorder and order is plotted in Fig. 7 (enthalpy factor
against the corresponding entropy factor). The reduced Redlich–Kister function QEa (x1 )
for Ea /x1 (1 − x1 ) is plotted in Fig. 8, which shows a continuous decrease with the mole
fraction of DMA (1).
In the same context as for the reduced Redlich–Kister function, we consider the Gibbs
energy of activation of viscous flow. In fact, on the basis of the theory of absolute reaction
rates, the excess Gibbs energies (G∗E ) of activation of viscous flow were calculated [33,
40, 41] as follows:

 
ηMρ2 η1 M1 ρ2
G∗E = RT · ln − x1 · ln (22)
η2 M2 ρ η2 M2 ρ1
1346 J Solution Chem (2012) 41:1334–1351

Fig. 8 Experimental reduced


Redlich–Kister excess properties
QRK,Ea (x1 ) (Eq. 3) for the ratio
Ea /x1 (1 − x1 ) of the deviation
energy for DMA (1) + FA (2)
mixtures against the mole
fraction x1 in DMA (1) and
FA (2) mixtures in the
temperature range (298.15 to
318.15) K

Fig. 9 Experimental relative


reduced Redlich–Kister excess
properties Qrel,exp,Y,T (x1 )
for the ratio
GE∗ /{x1 (1 − x1 )G∗ } of the
excess Gibbs of activation of
viscous flow (Eq. 9) for DMA (1)
+ FA (2) mixtures against the
mole fraction x1 in DMA at the
temperatures: ("), 298.15 K;
(!), 308.15 K; (Q), 318.15 K

where ηi and ρi are the dynamic viscosity and the density of the pure component ‘i’. By
considering Eq. 3, we can deduce the reduced Redlich–Kister excess Gibbs energy function
Qrel,G∗ ,T (x1 ) = QG∗ ,T (x1 )/G∗ and show the relative one Qrel,G∗ ,T (x1 ) in Fig. 9. How-
ever, the monotonous decrease of Qrel,G∗ ,T (x1 ) without any peculiar behavior shows that
changes of the solvent structure and changes of the different existing types of interactions
with composition are gradual.

3.2 Herráez Equation

In a recent paper [3], Herráez et al. proposed a new empirical correlation equation (Eqs. 23
and 26) that introduces a correcting polynomial (Eqs. 24 and 27) as an exponential acting on
the mole fraction of one of the mixture’s components. We note that the viscosity excesses
calculated with this model generally yield satisfactory results for many studied mixtures
showing monotonic variations in viscosity with mole fraction, but exhibits inferior perfor-
mance when the data exhibits a maximum or minimum [3]:
J Solution Chem (2012) 41:1334–1351 1347

Fig. 10 Variation of the


experimental Herráez exponent
P12,Herráez (x1 ) (Eq. 25) against
the mole fraction x1 in
DMA (1) + FA (2) mixtures at
the temperatures: ("), 298.15 K;
(!), 308.15 K; (Q), 318.15 K

P (x1 )
η(x1 ) = η2 + (η1 − η2 ) · x1 12,T (23)
where P12,T (x1 ) is a power polynomial with order (n) and (n + 1) adjustable parameters
Bp,T :

p=n
p
P12,T (x1 ) = Bp,T · x1 (24)
p=0

Hence, the Herráez Pn,T (x1 ) polynomials of Eq. 24 can be inspected experimentally and
graphically (Fig. 10) using Eq. 25:
 η (x1 )−η2 
ln exp,T
η1 −η2
P12,exp,T (x1 ) = (25)
ln x1
where η1 and η2 are the dynamic viscosity of pure components (1) and (2), respectively, and
ηexp,T (x1 ) is the dynamic viscosity of the mixture at mole fraction x1 and temperature T for
x1 over the interval [0, 1].
On the other side of the model and when it is mathematically possible, we can write:
P (x2 )
η(x2 ) = η1 + (η2 − η1 ) · x2 21,T (26)
where P21,T (x2 ) is a power series polynomial with order (n) and (n + 1) adjustable parame-

ters Bp,T :


p=n
 p
P21,T (x2 ) = Bp,T · x2 (27)
p=0

Hence, the Herráez Pn,T (x2 ) polynomials of Eq. 27 can be analyzed experimentally and
graphically (Fig. 11) using Eq. 28:
 η (x2 )−η1 
ln exp,T
η2 −η1
P21,exp,T (x2 ) = (28)
ln x2
where η1 and η2 are the dynamic viscosity of pure components (1) and (2), respectively, and
ηexp,T (x2 ) is the dynamic viscosity of the mixture at mole fraction x2 and temperature T for
x2 over the interval [0, 1].
1348 J Solution Chem (2012) 41:1334–1351

Fig. 11 Variation of the


experimental Herráez exponent
P21,Herráez (x2 ) (Eq. 28) for
DMA (1) + FA (2) mixtures
against the mole fraction x2 in
DMA (1) and FA (2) mixtures at
the temperatures: ("), 298.15 K;
(!), 308.15 K; (Q), 318.15 K

Table 5 Variation of the Bp,T constants for the Herráez exponent polynomial Pij,T (xi ) (Eqs. 25 and 28)
with temperature, and the corresponding standard deviation σ (η) for {DMA (1) and FA (2)} mixtures, at the
three studied temperatures

T /K B0 B1 B2 B3 B4 σ (η)

298.15 0.97149 −4.55685 −9.79381 89.0697 −160.538 0.00628


308.15 0.96353 −2.10267 −40.8738 214.598 −320.378 0.00829
318.15 0.89599 11.3394 −241.744 1039.97 −1327.44 0.01135

T /K B0 B1 B2 B3 B4 σ (η)

298.15 0.99581 −0.02673 0.01468 0.00838 −0.01351 0.00228


308.15 0.99241 −0.05045 0.09403 −0.12475 0.05952 0.00229
318.15 0.98871 −0.08103 0.17482 −0.23427 0.10824 0.00139

However, at infinite dilution (xi → 0+ ), the Pexp,T (xi ) values converge to a surprising
single point (Pexp,T (0) = 1.0) that is independent of temperature (Figs. 10 and 11), showing

a fixed value of B0,T or Bp,T corresponding to the first monomial of Pij,T (xi ). We can
ascertain this when we inspect the B0,T values in Table 5. Note that the same conclusion
was observed in our previous work [5–8, 27] investigating the viscosity in isobutyric acid +
water and 1,4-dioxane + water mixtures. We have concluded that B0 is a universal exponent
characterizing the preponderant type of interaction at infinite dilution.
In the case of dilute solutions of FA (2) in DMA (1), Eq. 5 can be rewritten as:
 1/2 
η = η1 1 + b1 x2 + b2 x2 + · · · (29)
Hence, in the extended conformal solution (ECS) theory [42, 43], Qexp,T (x1 = 1) is the
regular viscosity term and is denoted by η21 [22, 23, 42, 43], and we can deduce that the b2
coefficient for a nonelectrolyte binary solution as given by Nakagawa [22] is:

b2 = (η2 − η1 ) + η21 /η1 (30)
where η21 is the famous reduced Redlich–Kister function Qexp,η,T (x1 = 1), and the sub-
scripts 1 and 2 denote DMA and FA respectively. The b2 coefficient is divided in the two
J Solution Chem (2012) 41:1334–1351 1349

parts: the first is Bid which is based on the contribution of ideal mixture (term in parenthesis
in Eq. 30), and the second Bn coefficient is based on the net interaction between the solute
(FA) and solvent (DMA) [22, 23].
Nevertheless, we can add that in the case of nonelectrolyte solutions, when the Falken-
hagen parameter [24] (solute–solute interaction) is null or vanishingly small (a1 or b1 = 0),
and considering Eqs. 5 and 29, we can rewrite Eqs. 25 and 28 in new limiting asymptotic
expansions at high dilution solution of FA (2) in DMA (1) or the reverse, and we then obtain:
2
ln η1η−η · a2 + o(x1 )
P12,exp,T (x1 ) = 1 + 2
(31)
ln x1
and

ln η1
η2 −η1
· b2 + o(x2 )
P21,exp,T (x2 ) = 1 + (32)
ln x2
from which we can conclude that (limx1 →0+ Pexp,T (x1 ) = 1) and at very high dilution where
the solute–solvent interaction is dominant. Note that the universal exponent B0 of the Her-
ráez polynomial tends to the exponent of the term of Eq. 5.

4 Conclusion

Experimental densities and viscosities for DMA + FA binary mixtures at three different tem-
peratures (298.15, 308.15 and 318.15) K have been reported earlier [1]. The corresponding
reduced Redlich–Kister excess functions calculated from these two properties are presented
here along with a comparison of the direct excess properties of earlier investigation [1]. In
addition, the relative reduced Redlich–Kister equation has been introduced in order to re-
duce the temperature effect and can also be a good tool, like the reduced Redlich–Kister
function, for interpreting different types of interactions. Limiting excess partial molar vol-
umes at infinite dilution in the case of DMA + FA binary mixture at these three different
temperatures were deduced through different techniques and make a correlation with the val-
ues calculated by the Redlich–Kister parameters. In the present work, Arrhenius parameters
of viscosity for DMA + FA mixtures were investigated at (298.15, 308.15 and 318.15) K.
The pre-exponential entropic factor, equivalent to the viscosity at infinite temperature, is
closely related to that of the same system in the vapor phase. Also, the Arrhenius activation
energy is almost equal to the enthalpy of activation of viscous flow and it is correlated with
its enthalpy of vaporization. Correlation between the two Arrhenius parameters for DMA +
FA mixtures can give evidence for the existence of distinct composition regions with differ-
ent behaviors. Therefore, the results discussed so far reveal that in the DMA + FA system,
strong hetero-association is present through multiple hydrogen bonding among the polar
groups of the unlike molecules, along with a possible interstitial accommodation of FA in
the DMA structure.
The reduced R-K function relative to the Arrhenius activation energy, which is equivalent
to an apparent molar property, is more sensitive than the corresponding excess property to
interactions and gives more information. In the case of viscosity data having a maximum
as a function of mole fraction, the two investigated equations show deficient performance
for their corresponding second and third polynomial degrees. This divergence is more pro-
nounced in the Redlich–Kister model than in the Herráez model. Starting with fourth order,
the two models begin to better reproduce the experimental data points. The Redlich–Kister
polynomial is more sensitive to the number of experimental points whereas the Herráez
1350 J Solution Chem (2012) 41:1334–1351

model offers a good smoothed interpolation without any oscillation between data points,
even in the case where there are few experimental measurements. It is observed that the
viscosity deviations calculated with the Herráez model generally yield satisfactory results
for many studied mixtures showing monotonous variation in viscosity values with mole
fraction. In the case of some nonelectrolyte mixtures, the net absence of solute–solute inter-
actions leads to a fixed value of the Herráez constant of 1.0.

References

1. Das, D., Ray, S.K., Hazra, D.K.: Excess molar volumes and viscosity deviations in binary mixtures
of N,N-dimethylacetamide with formamide and N,N-dimethylformamide at 298.15 K, 308.15 K and
318.15 K. J. Indian Chem. Soc. 80(4), 385–390 (2003)
2. Redlich, O., Kister, A.T.: Algebraic representation of thermodynamic properties and the classification of
solutions. Ind. Eng. Chem. 40, 345–348 (1948)
3. Herráez, J.V., Belda, R., Diez, O., Herráez, M.: An equation for the correlation of viscosities of binary
mixtures. J. Solution Chem. 37, 233–248 (2008)
4. Desnoyers, J.E., Perron, G.: Treatment of excess thermodynamic quantities for liquid mixtures. J. Solu-
tion Chem. 26, 749–755 (1997)
5. Ouerfelli, N., Kouissi, T., Zrelli, N., Bouanz, M.: Competition of correlation viscosities equations in
isobutyric acid + water binary mixtures near and far away from the critical temperature. J. Solution
Chem. 38, 983–1004 (2009)
6. Cherif, E., Ouerfelli, N., Bouaziz, M.: Competition between Redlich–Kister and adapted Herráez equa-
tions of correlation conductivities in isobutyric acid + water binary mixtures near and far away from the
critical temperature. Phys. Chem. Liq. 49, 155–171 (2011)
7. Ouerfelli, N., Kouissi, T., Iulian, O.: The relative reduced Redlich–Kister and Herráez equations for
correlation viscosities of 1,4-dioxane + water mixtures at temperatures from 293.15 K to 323.15 K.
J. Solution Chem. 39, 57–75 (2010)
8. Ouerfelli, N., Iulian, O., Bouaziz, M.: Competition between Redlich–Kister and improved Herráez equa-
tions of correlation viscosities in 1,4-dioxane + water binary mixtures at different temperatures. Phys.
Chem. Liq. 48, 488–513 (2010)
9. Covington, A.K., Dickinson, T.: Physical Chemistry of Organic Solvent Systems, p. 5. Plenum,
New York (1973)
10. Das, D., Das, B., Hazra, D.K.: Viscosities of some tetraalkylammonium and alkali salts in N,N-dimethyl-
acetamide at 25 °C. J. Solution Chem. 32, 85–91 (2003)
11. Besbes, R., Ouerfelli, N., Latrous, H.: Density, dynamic viscosity, and derived properties of binary mix-
tures of 1,4-dioxane with water at T = 298.15 K. J. Mol. Liq. 145, 1–4 (2009)
12. Abramowitz, M., Stegun, I.A. (eds.): Handbook of Mathematical Functions with Formulas, Graphs, and
Mathematical Tables, 9th edn. Dover, New York (1972)
13. Peralta, R.D., Infante, R., Cortez, G., Ramírez, R.R., Wisniak, J.: Densities and excess volumes of binary
mixtures of 1,4-dioxane with either ethylacrylate, or butylacrylate, or methylacrylate, or styrene at T =
298.15 K. J. Chem. Thermodyn. 35, 239–250 (2003)
14. Peralta, R.D., Infante, R., Cortez, G., Elizalde, L.E., Wisniak, J.: Density, excess volumes and partial
volumes of the systems of p-xylene + ethyl acrylate, butyl acrylate, methyl methacrylate, and styrene at
298.15 K. Thermochim. Acta 421, 59–68 (2004)
15. Wisniak, J., Peralta, R.D., Infante, R., Cortez, G., Lopez, R.G.: Densities, isobaric thermal compressibili-
ties and derived thermodynamic properties of the binary systems of cyclohexane with allyl methacrylate,
butyl methacrylate, methacrylic acid, and vinyl acetate at T = (298.15 and 308.15) K. Thermochim.
Acta 437, 1–6 (2005)
16. Wisniak, J., Villarreal, I., Peralta, R.D., Infante, R., Cortez, G., Soto, H.: Densities and volumes of mixing
of the ternary system toluene + butyl acrylate + methyl methacrylate and its binaries at 298.15 K.
J. Chem. Thermodyn. 39, 88–95 (2007)
17. Tomiska, J.: Calculation of the thermodynamics of ternary systems based upon experimental data of
e.m.f. measurements. Calphad 5, 81–92 (1981)
18. Tomiska, J.: Zur konversion der Anpassungen Thermodynamischer Funktionen mittels einer Reihe Leg-
endre’scher Polynome und der Potenzreihe. Calphad 5, 93–102 (1981)
19. Tomiska, J.: Mathematical conversions of the thermodynamic excess functions represented by the
Redlich–Kister expansion and by the Chebyshev polynomial series to power series representations and
vice-versa. Calphad 8, 283–294 (1984)
J Solution Chem (2012) 41:1334–1351 1351

20. Ouerfelli, N., Bouanz, M.: A shear viscosity study of cerium(III) nitrate in concentrated aqueous solu-
tions at different temperatures. J. Phys., Condens. Matter 8, 2763–2774 (1996)
21. Jones, G., Dole, M.: The viscosity of aqueous solutions of strong electrolytes with special reference to
barium chloride. J. Am. Chem. Soc. 51, 2950–2964 (1929)
22. Nakagawa, T.: Is viscosity B coefficient characteristic for solute–solvent interaction? J. Mol. Liq. 63,
303–316 (1995)
23. de Ruiz Holgado, M.E., de Schaefer, C.R., Araneibia, E.L.: Viscosity study of 1-propanol with polyethy-
lene glycol 350 monomethyl ether systems at different temperatures. J. Mol. Liq. 79, 257–267 (1999)
24. Falkenhagen, H.: Theorie der Elektrolyte. Hirzel, Leipzig (1971)
25. Ouerfelli, N., Bouanz, M.: Excess molar volume and viscosity of isobutyric acid + water mixtures near
and far away from the critical temperature. J. Solution Chem. 35, 121–137 (2006)
26. Ouerfelli, N., Messaâdi, A., Bel Hadj H’mida, E., Cherif, E., Amdouni, N.: Validity of the correlation—
Belda equation for some physical and chemical properties in isobutyric acid + water mixtures near and
far away from critical temperature. Phys. Chem. Liq. 49(5), 655–672 (2011)
27. Ouerfelli, N., Barhoumi, Z., Besbes, R., Amdouni, N.: The reduced Redlich–Kister excess molar Gibbs
energy of activation of viscous flow and derived properties in 1,4-dioxane + water binary mixtures from
293.15 K to 309.15 K. Phys. Chem. Liq. 49, 777–800 (2011)
28. Ouerfelli, N., Iulian, O., Besbes, R., Barhoumi, Z., Amdouni, N.: On the validity of the correlation—
Belda equation for some physical and chemical properties in 1,4-dioxane + water mixtures. Phys. Chem.
Liq. 50, 54–68 (2012)
29. Van Ness, H.C., Abbott, M.M.: Classical Thermodynamics of Nonelectrolyte Solutions. McGraw-Hill,
New York (1982)
30. Perron, G., Couture, L., Desnoyers, J.E.: Correlation of the volumes and heat capacities of solutions with
their solid–liquid phase diagrams. J. Solution Chem. 21, 433–443 (1992)
31. Belda, R.: A proposed equation of correlation for the study of thermodynamic properties (density, vis-
cosity, surface tension and refractive index) of liquid binary mixtures. Fluid Phase Equilib. 282, 88–99
(2009)
32. Ouerfelli, N., Barhoumi, Z., Iulian, O.: Viscosity Arrhenius activation energy and derived partial molar
properties in 1,4-dioxane + water binary mixtures from 293.15 K to 323.15 K. J. Solution Chem. 41,
458–474 (2012)
33. Erdey Gruz, T.: Transport Phenomena in Aqueous Solutions. AH PB, London (1958)
34. Lu, J.f., Zhou, Z.y., Wu, Q.-y., Zhao, G.: Density functional theory study of the hydrogen bonding inter-
action in formamide dimer. J. Mol. Struct., Theochem 724, 107–114 (2005)
35. Eyring, H., John, M.S.: Significant Liquid Structures. Wiley, New York (1969)
36. Ali, A., Nain, A.K., Hyder, S.: Ion–solvent interaction of sodium iodide and lithium nitrate in N,N-di-
methylformamide + ethanol mixtures at various temperatures. J. Indian Chem. Soc. 75, 501–505 (1998)
37. Leaist, D.G., MacEwan, K., Stefan, A., Zamari, M.: Binary mutual diffusion coefficients of aqueous
cyclic ethers at 25 °C. Tetrahydrofuran, 1,3-dioxolane, 1,4-dioxane, 1,3-dioxane, tetrahydropyran, and
trioxane. J. Chem. Eng. Data 45(5), 815–818 (2000)
38. Iulian, O., Ciocîrlan, O.: Viscosity and density of systems with water, 1,4-dioxane and ethylene glycol
between (293.15 and 313.15) K. I. Binary systems. Rev. Roum. Chim. 55, 45–53 (2010)
39. Guettari, M., Gharbi, A.: A correspondence between Grunberg–Nissan constant d  and complex varieties
in water/methanol mixture. Phys. Chem. Liq. 49, 459–469 (2011)
40. Bearman, R.J., Jones, P.F.: Statistical mechanical theory of the viscosity coefficients of binary liquid
solutions. J. Chem. Phys. 33, 1432–1438 (1960)
41. Chevalier, J.L., Petrino, P., Gasto-Bonhomme, Y.: Estimation method for the kinematic viscosity of a
liquid-phase mixture. J. Chem. Eng. Sci. 43, 1303–1309 (1988)
42. Longuet-Higgins, H.C.: The statistical thermodynamics of multicomponent systems. Proc. R. Soc.
Lond., Ser. A, Math. Phys. Sci. 205, 247–269 (1951)
43. Matsubayashi, N., Nakahara, M.: Dynamics in regular solution: concentration and viscosity dependence
of orientational correlation of a benzene molecule. J. Chem. Phys. 94, 653–661 (1991)

View publication stats

You might also like