Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Molecular Liquids 342 (2021) 117504

Contents lists available at ScienceDirect

Journal of Molecular Liquids


journal homepage: www.elsevier.com/locate/molliq

Effect of phosphorus P-p bonding on the volumetric properties and


vapor-liquid equilibrium of phosphorus trichloride-benzene liquid
mixtures
Bingwen Long ⇑, Jiapo Song, Keke Yao, Yigang Ding
Engineering Research Center of Phosphorus Resources Development and Utilization of Ministry of Education, Key Laboratory for Green Chemical Process of Ministry of Education,
Hubei Key Laboratory for Novel Reactor and Green Chemical Technology, School of Chemical Engineering and Pharmacy, Wuhan Institute of Technology, Wuhan 430074, China

a r t i c l e i n f o a b s t r a c t

Article history: New density and vapor–liquid equilibrium data of PCl3-benzene binary system at atmospheric pressure
Received 21 July 2021 (99.7–101.3 kPa) were carefully measured to provide crucial information for their industrial reaction and
Revised 2 September 2021 separation. The experimental mixture densities show excellent linear dependence on both solution com-
Accepted 6 September 2021
position and temperature and a generalized empirical correlation is proposed with an average relative
Available online 09 September 2021
deviation of 0.13%. The specific attractive P-p interaction between the phosphorus atom of PCl3 and
the p cloud of benzene seems to be quite strong in solution, leading to large negative excess volumes.
Keywords:
But the p-p antibonding effect from benzene dimers are predominant in PCl3 highly dilute solution,
Phosphorus trichloride
Benzene
where positive excess volume and excess partial molar volumes of PCl3 were observed. Experimental
Density bubble points of the binary mixture were corrected to those under 101.325 kPa. Two distinctive methods
Vapor–liquid equilibrium of coexistence equation and activity coefficient models of Wilson and NRTL equations were applied to
P-p interaction model the VLE data, and they yielded very consistent results. Although no azeotrope were found, this
Modeling mixture exhibited quite low relative volatility.
Ó 2021 Elsevier B.V. All rights reserved.

1. Introduction points of PCl3 and benzene, the further separation of them is quite
challenging. However, the fundamental physicochemical and
Phosphorus trichloride (PCl3) and benzene are the most impor- phase equilibrium properties of this binary mixture and their the-
tant feedstocks to manufacture phosphorus flame-retardants and oretical modeling, which are highly demanded for the reliable
photo-initiators, such as dichlorophenyl phosphine (DCPP) and design and optimal operation of their separation, transportation
diphenyl phosphine chloride (CDPP). While CDPP is one of the and storage, are not available.
most important intermediates for phosphorus photo-initiators, From a theoretical perspective, any deviation from ideal solu-
ligands of catalyst and organophosphorous flame retardant [1], tion behavior is considered as the joint results of various intra/in-
the industrial production of CDPP uses the Friedel-Crafts reaction termolecular interactions between molecules of different size,
of PCl3 and benzene with anhydrous aluminum chloride [2] or shape and polarity [5–7]. It is well established that benzene molec-
Lewis acid ionic liquid such as [BuPy]Cl–xAlCl3 [3] as the catalyst. ular contains electron rich p cloud that could accept proton to form
A recent reaction modeling study [4] has shown that the vapor–liq- molecular complexes with the proton donor by hydrogen bonding.
uid equilibrium (VLE) has significant influence on the catalytic effi- For example, the formation of X-Hp bonding complexes have
ciency and product selectivity as well as the hydrodynamics inside been demonstrated in binary liquid mixtures of benzene with
the reactor and hence a full and accurate thermodynamic descrip- HCl and with chloroform through both quantum chemistry compu-
tion of the reactants and products is a must. After the reaction, tations [8] and nuclear magnetic resonance (NMR) [9] and Raman
most unreacted PCl3 and benzene must be recovered from the liq- spectroscopy measurements [10]. This X-Hp bonding has
uid system by atmospheric distillation as mixture and then vac- resulted in the formation of eutectic mixture of benzene and chlo-
uum distillation is applied to separate residual PCl3 and benzene roform at 85℃ and benzene mole fraction of 0.735. On the other
from raw CDPP [2]. Due to the high volatility and close boiling hand, as chlorine is quite electronegative, the three chloride atoms
on chloroform direct away from the benzene ring, causing an pos-
⇑ Corresponding author.
itive excess volume over the entire composition upon mixing them
E-mail address: Bingwen_long@163.com (B. Long).
[11,12]. Apart from the hydrogen atom, phosphorus atom can also

https://doi.org/10.1016/j.molliq.2021.117504
0167-7322/Ó 2021 Elsevier B.V. All rights reserved.
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

act as the p electron acceptors through the P-p bonding. The P-p
bonding is a new class of non-covalent intermolecular interaction,
which was firstly reported from an ab initio computation study by
Klinkhammer and Pyykko in 1995 [13]. The p-type bonding inter-
action of PCl3 with benzene was explored by Xu [14] using DFT cal-
culations, which showed that benzene could provide high p-
electron energy capacity and form stable complex with PCl3. More
recently, the PCl3-benzene dimer was experimentally detected by
Ramanathan using matrix isolation infrared spectroscopy and the
further ab initio computations confirmed that the experimentally
detected dimer is stabilized through the strong non-covalent phos-
phorus bonded P-p interactions [15]. Although the phosphorus
bonded P-p interaction has been recognized computationally,
experimental studies of its effect on the macroscopic thermophys-
ical properties and fluid phase equilibrium relationship are still
rare.
In summary, accurate determination of the thermodynamic
Fig. 1. Densities of benzene at different temperatures measured in this work
properties of binary mixtures of PCl3 and benzene would provide (points) and the recommended equation calculations (line) as the best fit to the
not only fundamental data for their industrial reaction and separa- critically evaluated experimental data from literature.
tion but also useful information to extend our understanding of the
phosphorus bonded P-p interaction formed in them. In this contri- 1.57464 g/cm3 and those at other temperatures are listed in
bution, the measurements of the densities and vapor–liquid equi- Table 1.
librium of PCl3-benzene binary mixtures of various compositions
and temperatures under atmospheric pressure were carefully per- 2.2. Apparatus and procedures
formed, as they are sensitive indicators of the behavior of mixed
liquid systems and good thermodynamic tools to explore the accu- 2.2.1. Density measurements
racy of liquid mixture theories. As expected, deviations to ideal The volumetric properties of the mixtures of phosphorus
solution behavior have been observed and hence the obtained data trichloride (PCl3) + benzene were derived from their experimental
were correlated with robust thermodynamic models and further densities. So, a serials of binary liquid mixtures were prepared
discussions were undertaken to probe the nature of the observed gravimetrically based on the mole fraction of PCl3 in a mixing
nonideality. chamber under N2 atmosphere to avoid their contact with air
and atmospheric moisture. The two components of desired
amounts were weighted individually on an electronic analytical
2. Experimental balance (AUW220D, accuracy ±0.01 mg, Shimadzu, Japan) and then
were quickly subject to adequate mixing in sealed vials. The exact
2.1. Materials compositions of the solutions were calculated from the added mass
of each component and the uncertainty of the mole factions of the
Phosphorus trichloride (PCl3) and anhydrous benzene were pur- mixtures was estimated to be better than ±0.0002. All the vials
chased from Zhengzhou Paini Chemicals Co. and Aladdin Co., were immediately tightly sealed after the solution sample were
respectively. The obtained anhydrous benzene was of HPLC grade syringed in so as to minimize the sample evaporation.
with a stated purity greater than 99.9% and was used without fur- Density measurements were carried out right after the prepara-
ther purification. The stated purity of PCl3 was greater than 98% tion of solutions by using an automatic vibrating U-tube densito-
and it was purified by fractional distillation, in which only the frac- meter (DMA 4500 M, Anton Paar, Austria). The detailed
tions with boiling points in between 348 and 350 K were collected experimental procedure can be referred to our previous work
for subsequent mixture preparation. All the liquids were tightly [19,20]. Special care was taken to ensure that all the samples did
sealed in dark bottles and stored in dry box to prevent moisture not contact air during the injection step and then they were kept
and air contamination. in a completely sealed U tube all through the subsequent density
The purity of the chemicals used in this work was further measurement. For each measurement, the apparatus automatically
checked by measuring their densities at different temperatures equilibrated the injected liquid sample at the pre-setting tempera-
and compared with those found in literature. The density of pure ture with an accuracy and stability of ±0.01 K through a built-in
liquid at different temperatures can serve as a good indicator of Peltier thermostat and then gave the density value after viscosity
its purity [16], because liquid density of pure component is very correction. Each sample was measured at least three times and
sensitive to temperature and the presence of impurities. More the uncertainty in the density measurement was better than ±0.0
important, liquid density can now be measured very fast and pre- 0005 g/cm3. The apparatus was periodically calibrated by ambient
cisely with commercial vibrating U-tube densitometer. The densi- dry air and ultra-pure water under atmospheric pressure according
ties of benzene at virous temperatures have been measured to the protocol recommended by the producer for acquiring the
extensively and Wilhoit et al. [17] have collected and critically best accuracy.
evaluated the experimental density data of benzene reported in lit-
erature since 1883. The most reliable density value at 293.15 K rec- 2.2.2. Vapor-liquid equilibrium
ommended by them is 0.87899 g/cm3 while our measured one is The vapor–liquid equilibrium (VLE) relations of the binary mix-
0.87896 g/cm3. Fig. 1 compares our measured densities for ben- ture of PCl3 and benzene were determined by measuring their
zene at various temperatures with those recommended by Wilhoit pressure, temperature and liquid composition (PTx) at equilibrium
et al. [17] and reasonably excellent agreement can be observed. For in a flow-type ebulliometer, as PTx measurement is quicker, easier,
PCl3, however, there is no such data source to refer. The density and more precise compared to measuring PTxy data for highly vola-
data of PCl3 documented in CRC’ Handbook is 1.574 g/cm3 at tile mixtures at low pressure [21]. The ebulliometer used in this
294.15 K [18], while our experimental value at 293.15 K is work is structurally the same as that used by Kato and Tanaka,
2
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

Table 1
Experimental densities (q) of (PCl3 (1) + benzene(2)) binary mixtures over the entire composition and temperatures (T) from 283.15 to 313.15 K at atmospheric pressure.

x1 q/(gcm3)
283.15 K 288.15 K 293.15 K 298.15 K 303.15 K 308.15 K 313.15 K
0.0000 0.88965 0.88432 0.87896 0.87361 0.86823 0.86284 0.85746
0.0499 0.92326 0.91775 0.91222 0.90667 0.90112 0.89554 0.88999
0.0999 0.95849 0.95279 0.94705 0.94131 0.93554 0.92976 0.92398
0.1499 0.99403 0.98813 0.98220 0.97625 0.97028 0.96429 0.95828
0.1999 1.03003 1.02393 1.01763 1.01164 1.00546 0.99927 0.99305
0.2498 1.06567 1.05937 1.05302 1.04665 1.04043 1.03385 1.02742
0.2997 1.10155 1.09504 1.08835 1.08191 1.07531 1.06869 1.06204
0.4000 1.17310 1.16619 1.15920 1.15220 1.14516 1.13808 1.13097
0.4994 1.24419 1.23688 1.22949 1.22204 1.21456 1.20707 1.19951
0.5999 1.31702 1.30930 1.30149 1.29364 1.28575 1.27783 1.26988
0.6997 1.38972 1.38191 1.37371 1.36548 1.35722 1.34891 1.34058
0.8004 1.46256 1.45435 1.44576 1.43718 1.42853 1.41984 1.41109
0.9000 1.53124 1.52196 1.51345 1.50449 1.49547 1.48639 1.47724
0.9494 1.56316 1.55355 1.54494 1.53577 1.52658 1.51736 1.50802
0.9696 1.57533 1.56593 1.55693 1.54766 1.53838 1.52906 1.51964
1.0000 1.59324 1.58397 1.57464 1.56520 1.55583 1.54640 1.53692

 
who have used such ebulliometer to successfully determine the q= g=cm3 ¼ ð0:000805x1 þ 0:001075ÞðT=K Þ þ 0:939018x1
vapor–liquid equilibrium of a number of polar systems such as
þ 1:193001
alcohols with tetrahydrofuran [22], amines [23] and alkoxides
[24] with very high accuracy. At the start of each VLE measure-
ment, around 40 cm3 liquid mixture of desired composition was 0  x1  1; 283:15  ðT=K Þ  313:15 ð1Þ
charged into the boiling flask of the ebulliometer and heated with
an electric cartridge heater. The glass heat-transfer surface was where x1 is the mole fraction of PCl3 in solution. Eq. (1) gives an
frosted to ensure a stable boiling and a drying tube containing cal- average relative deviation of only 0.13% to the 112 experimental
cium chloride was placed on the top of the condenser to prevent density data over the entire composition range and temperatures
moisture contamination. After the mixture was boiled, the current from 283.15 to 313.15 K. The deviation was found to be indepen-
of the heater was carefully adjusted to allow good and stable flash dent on temperature but randomly distributed with x1, indicating
of the vapor–liquid mixture. The boiling point was measured by a the density is more sensitive to solution composition. Fig. 2 shows
Pt-100 Type thermometer (CENTER-376, Qunte, China) with an the density data of the mixtures determined at 283.15 K
uncertainty of ±0.01 K. The exact experimental pressure were mea- covering the entire composition. The excellent q  x relationship
sured by an electronic barometer (DYM3-01, Yongzhi Instrument (linear correlation coefficient R2 = 0.9999) allows it to be an ideal
Co., China) with an uncertainty of ±0.01 kPa. Equilibrium between tool to correct the liquid compositions for the subsequent vapor–
the vapor and liquid phases was assumed to be achieved when the liquid equilibrium measurements.
flow of the condensed liquid was stabilized and the boiling point
was constant with a variation less than 0.02 K for at least
30 min. Then the composition of the liquid phase at equilibrium 3.2. Derived volumetric properties
were corrected by measuring their density at 283.15 K and the liq-
uid composition uncertainty was better than 0.001 in mole frac- The nonideality of a liquid mixture is usually characterized by
tion. As a test of the accuracy of the above apparatus and its excess thermodynamic properties [5]. Hence, the excess molar
procedure for vapor–liquid equilibrium measurement, the vapor volumes, V EM of the binary mixtures with different compositions
pressures of pure PCl3 and benzene measured by us were
100.07 kPa at 348.09 K and 99.69 kPa at 352.66 K, respectively,
while the NIST database gives very close values of 104.27 kPa at
348.5 K and 99.62 kPa at 352.32 K. Then the measurements were
performed for PCl3-benzene mixtures over the entire composition
range.

3. Results and discussion

3.1. Density of the binary mixture

The experimental densities (q) of the binary liquid mixtures of


phosphorus trichloride (PCl3) and benzene at different composi-
tions and at temperatures from 283.15 to 313.15 K under atmo-
spheric pressure are summarized in Table 1. In general, the
densities for the mixtures of certain composition always decrease
linearly with an increase of temperature, while those for the mix-
tures at certain temperature show very good linear relationship
with the mole fraction of PCl3. Hence, an attempt to generalize
the experimental density data was made by the following Fig. 2. Densities of binary mixtures of PCl3 (1) + benzene (2) at 283.15 K and the
equation, straight line is the linear regression of the experimental density data.

3
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

at certain temperature are calculated from their densities by the


following equation:

1 1 1 1
V EM ¼ x1 M 1 ð  Þ þ x2 M 2 ð  Þ ð2Þ
qM q1 qM q2
where x and M are the mole fraction and molar mass of each
component, respectively. The subscript 1, 2 and M denote PCl3,
benzene and their mixture, respectively. The calculated V EM are
summarized in Table 2 and those at 288.15 K are visually shown
in Fig. 3 against PCl3 mole fraction. Further, the dependence of VE
on solution composition is quantitatively represented by the
Redlich-Kister polynomial equation:

X
n X
n
V E ¼ x1 x 2 Ai ðx1  x2 Þi ¼ x1 ð1  x1 Þ Ai ð2x1  1Þi ð3Þ
i¼0 i¼0

where n is the degree of polynomial and Ai are the coefficients,


which were determined simultaneously by using multiple linear
regression method. Analysis of variance (ANOVA) results showed Fig. 3. Excess molar volumes (VE) for binary mixture of PCl3 (1) + benzene (2) at
that Eq. (3) with n = 4 could give significant enough correlations 288.15 K and the curve is calculated by the Redlich-Kister polynomial equation.
to all the systems according to the F-test at 95% confidence level
and hence the corresponding best Ai values are summarized in
more negative until PCl3 mole fraction reaches 0.8, afterwards it
Table 3 together with the standard deviation d to characterize
increases sharply with PCl3 mole fraction though still being nega-
the goodness of the correlations. d is defined as:
tive. The large and negative value of V EM in general is most likely
2  2 312 due to molecular packing effect, strong intermolecular attractive
E E
6Xm V exp  V calc 7 force or both. The free volume and interstitial accommodation con-
d¼4 5 ð4Þ
mn tribution that tend to decrease V EM is considered to be minor as the
1
molar volumes of the two components are very close (87.21 and
where m is the number of experimental points at each working 88.87 cm3mol1 for PCl3 and benzene respectively at 293.15 K).
On the other hand, the strong attractive P-p interaction between
temperature, while V Eexp and V Ecalc are the experimental excess
the phosphorus atom of PCl3 and the electron rich p cloud of ben-
molar volumes in Table 2 and those calculated with Eq. (3), respec-
zene tends to form PCl3-benzene complex with the most stable
tively. The obtained d in Table 3 are all less than 0.007 cm3mol1
structure of the phosphorus atom facing towards the plane of the
and the |d/V Emin | for all the mixtures are less than 0.01, which is con- benzene ring with a distance of 3.26 Å, shorter than the distance
sistent with the precision attainable with the instrument used. The of between the preferred T-shaped dimer of two benzene mole-
VE curve calculated by Eq. (3) at 288.15 K is also graphically cules (4.96 Å) [25]. This P-p interaction was found to be even
depicted in Fig. 3, in which fairly excellent agreement has been stronger than the hydrogen bonding one under the same condi-
observed, suggesting an adequate and continuous representation tions [15]. Besides, the three electronegative chloride atoms on
of the experimental VE data that allows the derivation of other vol- PCl3 could have Cl-p interaction with benzene ring by accepting
umetric properties. its p electrons. Imai [26] has shown that Cl-p interaction is attrac-
It can be seen from Fig. 3 that the V EM , with the exception at PCl3 tive in nature and mainly from the dispersion force which is even
very dilute region, is generally negative, as an indication of volume stronger than the CH-p interaction. Such Cl-p interaction would
contraction upon mixing the two components. The V EM curve is increase the molecular deformation and tends to decrease the sys-
visually asymmetric and skewed towards higher mole fraction of tem volume so that the V EM for binary mixtures of chloroform-
PCl3. With the increase of PCl3 content in the solution, V EM becomes benzene [11] is much less positive than that of dichloromethane-

Table 2
Excess molar volumes (VE) of (PCl3 (1) + benzene(2)) binary mixtures over temperatures (T) from 283.15 to 313.15 K at atmospheric pressure.

x1 VE/(cm3mol1)
283.15 K 288.15 K 293.15 K 298.15 K 303.15 K 308.15 K 313.15 K
0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
0.0499 0.0828 0.0820 0.0805 0.0808 0.0797 0.0791 0.0759
0.0999 0.0297 0.0277 0.0256 0.0247 0.0238 0.0227 0.0217
0.1499 0.0478 0.0507 0.0538 0.0555 0.0569 0.0584 0.0579
0.1999 0.1541 0.1576 0.1471 0.1644 0.1666 0.1694 0.1704
0.2498 0.2222 0.2263 0.2297 0.2327 0.2491 0.2366 0.2377
0.2997 0.2983 0.3029 0.2952 0.3103 0.3122 0.3142 0.3152
0.4000 0.3812 0.3862 0.3880 0.3913 0.3913 0.3913 0.3893
0.4994 0.4505 0.4557 0.4574 0.4594 0.4574 0.4573 0.4533
0.5999 0.5535 0.5586 0.5615 0.5644 0.5644 0.5644 0.5634
0.6997 0.6532 0.6789 0.6834 0.6904 0.6934 0.6964 0.6994
0.8004 0.7012 0.7258 0.7317 0.7427 0.7477 0.7527 0.7567
0.9000 0.5392 0.5227 0.5533 0.5643 0.5683 0.5723 0.5743
0.9494 0.3406 0.3143 0.3475 0.3553 0.3584 0.3633 0.3639
0.9696 0.2114 0.2003 0.2144 0.2194 0.2204 0.2224 0.2214
1.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

4
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

Table 3
Redlich-Kister coefficients and standard deviations for representation of the excess molar volumes of (PCl3 + benzene) binary mixtures at temperatures from 283.15 to 313.15 K.

T/K A0 (cm3mol1) A1 (cm3mol1) A2 (cm3mol1) A3 (cm3mol1) A4 (cm3mol1) 103d (cm3mol1)


283.15 1.8098 1.5108 3.3236 3.9031 2.7411 6.60
288.15 1.8186 1.7168 3.8793 3.4392 3.8098 5.56
293.15 1.8336 1.6832 3.6226 3.7824 3.0158 4.90
298.15 1.8405 1.6336 3.8667 3.9641 3.2397 5.92
303.15 1.8333 1.6026 4.0737 4.0477 3.4712 4.99
308.15 1.8340 1.6340 4.0514 4.0404 3.3527 6.66
313.15 1.8225 1.6553 4.1731 4.0141 3.4725 6.59

benzene [27]. Therefore, the positively cooperation of the P-p and


Cl-p interactions that are mainly originated from the electrostatic
interactions and dispersion interactions respectively, should be
responsible for the observed negative V EM . Nevertheless, we think
that the P-p interaction should be the dominant and decisive
one, as the V EM of benzene with most halogen derivatives such as
chloroform, carbon tetrachloride and chlorobenzene are all posi-
tive [28].
The positive V EM at PCl3 very dilute region should be attributed
to the break of the self-aggregation of benzene molecules that tend
to form dimers in the pure state due to the dispersion and electro-
static quadrupole–quadrupole interactions (aromatic p-p interac-
tions) [29] but the benzene dimer should bind less strongly than
the PCl3-benzene complex which is increasingly formed and lead-
ing the contraction of solution volume as the gradual addition of
PCl3. The experimental data in Table 2 also show that V EM overall
becomes more negative as the increase of temperature, however,
the impact of temperature on V EM is much less notable and the min-
Fig. 4. Excess partial molar volumes of binary mixture of PCl3 (1) + benzene (2) at
imum V EM decreases only by 7.9% from 283.15 to 313.15 K. 283.15 K.
Further, the partial molar volumes (V  i ) and the excess partial
molar volumes (V  i  V 0 ) of PCl3 and benzene in solution can
E ¼ V
i i
be estimated by applying the obtained Redlich-Kister correlations, abrupt changes of V  E at both ends suggest the occurrence of more
i
which could provide a more direct indication of composition remarkable interactions between the unlike molecules in dilute
impact on the solution volume. solutions, in which solute molecules are isolated from each other
and only surrounded by vast majority of solvent molecules and thus
E solute–solvent interactions play a central role. Therefore, by letting
V1 E ¼ V 1  V 0 ¼ V E þ x2 dV  V 0 ¼ ð1  x1 Þ2
1
dx1 1 x1 = 0 in Eq. (5) and x1 = 1 in Eq. (6), the excess partial molar volumes
X4 X 4 at infinite dilution (V E;1 ) of PCl3 (V1 E;1 ¼ P4 Ai ð1Þi ) and benzene
A ð2x1  1Þi þ 2x1 ð1  x1 Þ2 i¼0 i Ai ð2x1  1Þi1
i¼0 i
ð5Þ P
i i¼0

(V2 E;1 ¼ 4i¼0 Ai ) were estimated. The calculated V  E;1 at different


i

temperatures are plotted in Fig. 5, in which V  E;1


of PCl3 are all posi-
E i
 2  V 0 ¼ V E þ x1 dV  V 0
V2 E ¼ V tive whereas those of benzene are negative and both of them are
2 2
dx2
X4
¼ x1 2 i¼0 Ai ð2x1  1Þi  2x1 2 ð1
X4
 x1 Þ i¼0 i Ai ð2x1  1Þi1 ð6Þ

 E of PCl3 and benzene in solutions of


Fig. 4 shows the calculated V i

different compositions at 283.15 K. It has been observed that the V E


1
of PCl3 decreases very sharply from positive to negative upon its ini-
tial addition into benzene till its mole fraction reached 0.2, after-
wards it increases slowly back to zero till the end, while V  E of
2
benzene keeps almost unchanged at around zero with the increase
of PCl3 until the mole fraction of PCl3 exceeded 0.7, after which the
 E of benzene decreases much more rapidly to a large negative value
V 2

around 8 cm3mol1. The negative of V of PCl3 at intermediate to


E
1
PCl3-rich composition region means when PCl3 was added to the
mixture, the solution’s volume increase is actually smaller than
the adding PCl3 volume, confirming the attractive P-p interaction
 E , adding benzene into
with solution volume contraction. While for V 2
benzene-rich solution, this effect is much insignificant due to the Fig. 5. Excess partial molar volumes at infinite dilution of PCl3 and benzene in
antibonding p-p interactions of benzene dimers. Nevertheless, the respective binary mixtures of PCl3 + benzene at different temperatures.

5
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

independent of temperature with mean values of 3.14 and equation and excess Gibbs energy based activity coefficient equa-
7.90 cm3mol1, respectively. The positive V  E;1 of PCl3 is likely tions are used and compared. The coexistence equation relates
1
due to the strong p-p interactions between majority benzene mole- pressure, temperature, heats of vaporization and fugacities of the
cules that tends to keep the T-shape benzene dimer structure and equilibrated two phases to satisfy the Gibbs-Duhem equation
thus counteracts the phosphorus P-p bonding interactions, which [31]. At constant pressure, the equation is given by [31]:
however is the principal interactions in solution of benzene at infi- " #
dy1 dlnðcV1 =cV2 Þ DHVap
 E;1 of benzene.
nite dilution in PCl3, resulting in large negative V ¼ y1 ðy1  1Þ þ m
ð10Þ
2 dT dT ðy1  x1 ÞRT 2

3.3. Vapor-liquid equilibrium Vap


DHVap
m ¼ x1 D H 1 þ x2 DHVap
2 ð11Þ
The experimental P-T-x data of binary mixtures of PCl3 + benzene where xi and yi are the mole fractions of component i in liquid
of different compositions are summarized in Table 4. To obtain iso- and vapor phase respectively. DHVap and cVi are the molar heat of
i
baric vapor–liquid equilibrium data, the raw boiling points were
vaporization and vapor activity coefficient of component i. DHVap
corrected to those under 101.325 kPa using the following equa- i

tions [22]: at different temperatures are calculated from the empirical corre-
  lations given by Yaws [32]
dT
T cor ¼ T exp þ ð101:325  Pexp Þ ð7Þ DHVap ¼ Að1  T=T Ci Þn ð12Þ
dP P¼101:325 i

      The component-dependent constants A and n of PCl3 and ben-


dT dT dT zene are listed in Table 6 together with their critical temperatures
¼ x1 þ x2 ð8Þ
dP P¼101:325 dP 1;P¼101:325 dP 2;P¼101:325 (TCi) . Therefore, the dew point curve of y1  T can be obtained by
numerical integration of Eq. (10), provided the ratio of cVi can be
The dependence of pure components’ bubble point on pressure
  properly estimated. Further, the truncated virial equation of state
at 101.325 kPa, dT
dP i;P¼101:325
, were estimated from the well estab-
was used to estimate the vapor activity coefficient term in Eq. (10):
lished Antoine equation:
cV1 d12 Pð1  2y1 Þ
B ln ¼ ð13Þ
log10 ðP=kPaÞ ¼ A  ð9Þ cV2 RT
T=K þ C
in which A, B and C are the Antoine constants that should be fit- d12 ¼ 2B12  B11  B22 ð14Þ
ted to the experimental vapor pressure data for specific compo-
The second virial coefficients of Bij are only functions of temper-
nent. The Antoine constants for PCl3 were fit to the experimental
ature, and that of pure benzene, B22 were correlated from the
vapor pressure data stored in NIST database while those for ben-
experimental data given by Bottom and Spurli [33], while that of
zene were directly taken from the handbook edited by Dykyj
PCl3 B11 and the cross virial coefficients of B12 were calculated from
et al. [30]. These parameters are summarized in Table 5. Fig. 6
the correlations developed by Tarakad and Danner [34] who pro-
shows the experimental and calculated vapor pressures of PCl3
vided size-shape and polar corrections to the Pitzer-Tsonopoulos’
and benzene, and it suggests that relative volatility between them
general second virial coefficient correlation for non-polar com-
are rather small at around atmospheric pressure. Then, the cor-
pounds [35]. To simplify the subsequent differential calculation,
rected boiling points under 101.325 kPa can be calculated by Eqs.
the calculated d12 (cm3/mol) by Eq. (14) were correlated as a
(7) and (8) and they are listed in Table 4.
dlnðcV1 =cV2 Þ
To determine the composition of the equilibrated vapor phase three-order polynomial function of temperature so that dT
that is difficult to measure accurately, two methods of coexistence in Eq. (10) can be readily estimated.

Table 4
Experimental boiling-point data of (PCl3(1) + benzene (2)) a binary mixtures and the estimated vapor composition by the coexistence equation (Eq. (10)), Wilson and NRTL model.

x1 Texp/(K) Pexp/(kPa) Tcor/(K) ycal


1

Eq. (10) Wilson NRTL


0.000 352.66 99.69 353.18 0.000 0.000 0.000
0.085 352.13 99.70 352.65 0.094 0.102 0.102
0.129 351.82 99.71 352.34 0.149 0.151 0.152
0.153 351.76 100.02 352.18 0.176 0.177 0.178
0.160 351.66 100.01 352.08 0.194 0.184 0.185
0.192 351.56 100.51 351.82 0.239 0.218 0.219
0.233 351.38 100.51 351.64 0.270 0.261 0.263
0.277 351.16 100.52 351.42 0.308 0.308 0.309
0.325 350.80 100.51 351.06 0.372 0.356 0.356
0.397 350.60 100.74 350.79 0.422 0.431 0.430
0.475 350.27 100.85 350.42 0.493 0.511 0.510
0.551 349.93 100.91 350.06 0.568 0.588 0.586
0.591 349.77 101.31 349.77 0.633 0.625 0.623
0.647 349.49 101.05 349.58 0.677 0.683 0.681
0.673 349.45 101.31 349.45 0.709 0.708 0.706
0.706 348.97 100.12 349.35 0.735 0.741 0.739
0.790 348.71 100.17 349.08 0.807 0.824 0.823
0.889 348.36 100.19 348.72 0.911 0.919 0.919
1.000 348.09 100.07 348.49 1.000 1.000 1.000

a: x1 is the liquid mole fraction of PCl3; Texp is the experimental boiling point and Pexp is the experimental atmospheric pressure; Tcor is the boiling points corrected to
101.325 kPa. ycal
1 is the mole fraction of PCl3 in the equilibrated vapor phase as evaluated from the three models.

6
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

Table 5
The Antoine constants of PCl3 and benzene used in Eq. (9).

Chemical A B C T range (dT/dP)|p=101.325


a
PCl3 6.70537 1632.930 0.119 340–440 0.31689
b
Benzene 5.98523 1184.236 55.623 278–376 0.32051

a: The constants were obtained by fitting Eq. (9) to the experimental vapor pressure data stored in NIST database;
b: The constants were directly taken from ref. [30]

Fig.6. Vapor pressure of PCl3 and benzene at different temperatures. The points are
literature data stored in NIST database while the curves are calculated by the Fig. 7. Experimental boiling point curve of PCl3-benzene at 101.325 kPa and the
Antoine equation (Eq. (7)). calculated dew point curves with the coexistence equation (solid), Wilson model
(dashed) and NRTL model (dotted).

d12 ¼ 1:3309  104 T 3  0:1482T 2 þ 54:86T  6775:93 ð15Þ By doing this, the vapor composition yi can be determined
simultaneously because of the composition summation constraint
Hence, the vapor composition y1 at given temperature T can be P
of yi ¼ 1. This method finds wide and very successful applica-
obtained straightforward by numerical integration of Eq. (10) using
tions in correlation of VLE data but it should be undertaken in a
the fourth-order Runge-Kutta method in descending temperature
regression way because most activity coefficient models contain
sequence. As shown by Barrier and Adler [31], the accuracy of y1
system-dependent unknown model parameters [36]. The Wilson
estimated by the coexistence equation method is generally better
and NRTL(Non-Random Two Liquid) models are among the most
than 0.01 with maximum deviation of 0.02 for five binary systems
commonly used activity coefficient models due to their high accu-
of polar compounds including benzene, ammonia and acetone. The
racy, fair simplicity, and good extensibility. The Wilson Equation is
calculated vapor compositions of PCl3, y1 at each experimental
given as [36]:
point are listed in Table 4 and the entire y1  T curve is plotted !
in Fig. 7. It can be seen that, although no azeotrope has been found X X x K
lnci ¼ ln xj Kij þ1 P m mi ð17Þ
j¼1 xj Kmj
for this binary mixture computationally, the saturated vapor and
j¼1 m¼1
liquid boundaries are quite close especially at the two ends, sug-
gesting an infeasible complete separation of them by conventional 
Vj Dkij
distillation. Kij ¼ exp  ð18Þ
Vi RT
The activity coefficient equation method, on the other hand,
employs a composition dependent model to describe the liquid- in which Vi is the molar volume of pure component i and Dkij are
phase activity coefficients cLi and regulates the model parameters the two unknown binary energy parameters to be determined. The
to best fit the experimental equilibrium properties such as pres- expression of the NRTL model is [36]:
sure or temperature under the phase-equilibrium criteria of Eq. P  P 
s
j¼1 ji xj Gji
X xj Gij m¼1 xm smj Gmj
(16): lnci ¼ P þ P sij  P ð19Þ
   m¼1 xm Gmi j¼1 m¼1 xm Gmj m¼1 xm Gmj
V i P  Psi
xi cLi Psi /si exp ¼ P/Vi yi ð16Þ
RT Dg ji
sji ¼ ð20Þ
RT

Table 6
Critical constants and acentric factor of PCl3 and benzene and the constants used in Eq. (12)

Chemical TC/K PC/Mpa VC/(cm3/mol) x A n


PCl3 563.15 5.670 260 0.2344 45.9178 0.4229
Benzene 562.05 4.895 256 0.210 43.7047 0.3747

7
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504



Gji ¼ exp asji ð21Þ sion, the y1 given by the two activity coefficient models are a little
bit lower than that given by the coexistence equation at benzene-
where a is an empirical parameter relevant to the non-
rich region, while slightly higher at PCl3-rich region.
randomness of system and was set as a constant of 0.3 in this work,
and Dgij are the unknown binary energy parameters. In this work,
/si , the fugacity coefficient of pure component i under its saturated 4. Conclusions
pressure Psi
and that in mixed gas phase /Vi
in Eq. (16) were also
Experimental determinations and theoretical modeling of volu-
estimated by the truncated virial equation of state
metric properties and vapor–liquid equilibrium at atmospheric
pressure were performed for the binary mixture of PCl3 and ben-
Bii zene over the entire composition range. The obtained density data
ln/si ¼ ð22Þ
RT of the binary mixture at constant temperature showed excellent
linear relationship with the PCl3 mole fraction which allows the
mixture density as a good tool to predict solution composition.
PðBii þ y2j d12 Þ The densities of the binary mixture at certain composition also
ln/Vi ¼ ð23Þ decrease with the increase of temperature in a perfect linear
RT
way. So, a generalized empirical correlation for the mixture density
and Psi and Vi at each experimental temperature T was calcu- from 283.15 to 313.15 K and over the entire composition is pro-
lated by Eq. (9) and the modified Rackett equation [32], respec- posed which is simple and accurate enough for quick and reliable
tively. Therefore, the calculated equilibrium pressure and vapor engineering calculations with an average relative deviation of only
composition should satisfy Eq. (16) and the constrain of 0.13%. Excess volumes of the binary mixture calculated from the
y1 + y2 = 1 at the same time. Further, the energy parameters of measured density data were negative for solutions of PCl3 mole
the Wilson and NRTL models were optimized to best fit the calcu- fraction from around 0.2 to 1 as a result of dominant strong attrac-
lated pressure to the experimental one at each experimental tem- tive P-p interaction between the phosphorus atom of PCl3 and the
perature using the Levenberg–Marquardt method [37]. The electron rich p cloud of benzene, whereas the excess volumes are
obtained Dk12 and Dk21 of the Wilson model are 3210.89 and small positives for solutions of PCl3 mole fraction from 0 to around
2066.71 J/mol respectively with an average relative deviation 0.2, which is considered as the strong p-p interactions of benzene
(ARD) of the calculated equilibrium pressure from the experimen- dimer to counteract the phosphorus P-p bonding interactions. The
tal ones of 0.31%, while the best Dg12 and Dg21 of the NRTL model P-p bonding and p-p antibonding effects were further confirmed
are 3565.36 and 4753.75 J/mol respectively with a slightly lower by the estimation of excess partial molar volumes of PCl3 and ben-
ARD of 0.27%. The relative residual distributions of the calculated zene at infinite dilution, which are found to be positive and nega-
equilibrium pressure by the two models are shown in Fig. 8, which tive respectively and both are independent of temperature.
can be a good alternative reliability test for VLE data without mea- Ebulliometric measurements of PTx data for PCl3-benzene sys-
sured vapor composition [21]. It can be seen that most relative tem were carefully performed over the entire composition range.
residual values of the two models are less than ±0.5% without sys- Two methods of coexistence equation and activity coefficient are
tematic errors except for the data for solutions of very high PCl3 applied to estimate the vapor composition at equilibrium. Two
content, where the experimental uncertainties in T and x were activity coefficient models of Wilson and NRTL equations were
higher due to the high volatility of PCl3. But they still do not exceed used and the respective model parameters were optimized to best
1.5% even when the mole fraction of PCl3 was above 0.9. Moreover, fit the experimental vapor pressure data. Pressure residual distri-
the calculated vapor composition y1 at each experimental temper- bution diagrams showed that the measurements have no signifi-
ature by the Wilson and NRTL models are listed in Table 4 and the cant systematic errors. Albeit the two methods are
entire y1  T curves are also plotted in Fig. 7 for comparison. computationally in distinctive integration and regression modes,
Though different procedures were taken to derive y1, the disparity they gave very close results. Although no azeotrope has been found
between the three models is actually quite minor, confirming their for this binary mixture, the two-phase area in the vapor–liquid
quite good consistency. Nonetheless, due to the nature of regres- phase diagram is very narrow due to the strong attractive P-p
interaction, suggesting their full separation by conventional distil-
lation is practically infeasible.

CRediT authorship contribution statement

Bingwen Long: Conceptualization, Formal analysis, Writing –


original draft, Writing – review & editing, Supervision. : . Jiapo
Song: Investigation, Formal analysis, Resources, Visualization.
Keke Yao: Investigation, Formal analysis, Visualization. Yigang
Ding: Supervision, Project administration.

Declaration of Competing Interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared
to influence the work reported in this paper.

Acknowledgments
Fig. 8. Calculated equilibrium pressure residual distribution over the entire
composition of PCl3-benzene mixture by the Wilson and NRTL activity coefficient Financial supports from National Natural Science Foundation of
models. China (No. 22078250) are gratefully acknowledged.
8
B. Long, J. Song, K. Yao et al. Journal of Molecular Liquids 342 (2021) 117504

References [19] Z. Deng, Y. Xia, B. Long, Y. Ding, Volumetric properties of diisopropyl ether
with acetone at temperatures from 283.15 K to 323.15 K: An experimental and
theoretical study, J. Mol. Liq. 243 (2017) 257–264.
[1] Y.J. Li, X.Y. Gu, J. Zhao, P. Jiang, J. Sun, T. Wang, Flame retardancy effects of
[20] B. Long, Y. Ding, Probing the intermolecular attractive interactions of binary
phosphorus-containing compounds and cationic photoinitiators on
mixtures of formic acid + methanol or water via volumetric studies, J. Mol. Liq.
photopolymerized cycloaliphatic epoxy resins, J. Appl. Polym. Sci. 131 (2014)
206 (2015) 137–144.
40011.
[21] R.P. Vasanta, V.R. Kamireddi, Experimental binary vapor–liquid equilibrium
[2] H. Duan, X. Xu, K. Leng, G. Guo, Q. Yu, X. Li, Y. Han, J. Gao, Z. Wang, AlCl3-
data of morpholine with methanol, 1-propanol, and 2-ethoxyethanol, J. Chem.
catalyzed C-H phosphination of benzene: A mechanistic study, Appl. Catal., A
Eng. Data 66 (2021) 178–188.
611 (2021) 117943.
[22] Y. Yoshikawa, A. Takagi, M. Kato, Indirect determination of vapor-liquid
[3] Z. Wang, L. Wang, Preparation of dichlorophenylphosphine via friedel–crafts
equilibriums by a small ebulliometer. Tetrahydrofuran-alcohol binary
reaction in ionic liquids, Green Chem. 5 (2003) 737–739.
systems, J. Chem. Eng. Data 25 (1980) 344–346.
[4] A. Aguirre, M.F. Neria d’Angelo, The role of vapor-liquid equilibria during the
[23] M. Kato, H. Tanaka, Vapor-liquid equilibrium determination with a flow-type
fischer-tropsch synthesis: A modeling study, Chem. Eng. Sci. 233 (2021)
ebulliometer for six binary systems made of alcohol and amine, J. Chem. Eng.
116394.
Data 34 (1989) 203–206.
[5] B. Long, S. Zhu, Y. Ding, Antisolvent effect on the crystallization of fosfomycin
[24] M. Kato, H. Tanaka, Ebulliometric measurement of vapor-liquid equilibria for
phenylethylamine by acetone: Solubility measurement and thermodynamic
four binary systems: Methanol + silicon tetramethoxide, methanol + silicon
mechanism analysis, AICHE J. (2021) e17140.
tetraethoxide, ethanol + silicon tetramethoxide, and ethanol + silicon
[6] B. Long, S. Chen, Y. Xia, Z. Wang, Y. Ding, Insight into the solubility and solution
tetraethoxide, J. Chem. Eng. Data 34 (1989) 206–209.
thermodynamics of fosfomycin phenylethylamine in water and ethanol for its
[25] M.O. Sinnokrot, E.F. Valeev, C.D. Sherrill, Estimates of the ab initio limit for pi-
cooling crystallization, J. Mol. Liq. 317 (2020) 113967.
pi interactions: The benzene dimer, J. Am. Chem. Soc. 124 (2002) 10887–
[7] B. Long, Z. Li, T. Li, Y. Ding, Characterization of the solubility of m-phthalic acid
10893.
in co-solvent mixture of acetone and ethanol at different temperatures, J. Mol.
[26] Y.N. Imai, Y. Inoue, I. Nakanishi, K. Kitaura, Cl–p interactions in protein–ligand
Liq. 296 (2019) 111759.
complexes, Protein Sci. 17 (2008) 1129–1137.
[8] I. Rozas, I. Alkorta, J. Elguero, Unusual hydrogen bonds: Hp interactions, J.
[27] Y. Sun, L. Su, H. Wang, Volumetric properties of binary liquid mixtures:
Phys. Chem. A 101 (1997) 9457–9463.
Application of the prigogine–flory–patterson theory to excess molar volumes
[9] L.W. Reeves, W.G. Schneider, Nuclear magnetic resonance measurements of
of dichloromethane with benzene or toluene, J. Chem. Thermodyn. 41 (2009)
complexes of chloroform with aromatic molecules and olefins, Can. J. Chem. 35
1154–1161.
(1957) 251–261.
[28] R. Battino, Volume changes on mixing for binary mixtures of liquids, Chem.
[10] H. Nomura, S. Koda, Y. Miyahara, A study on molecular reorientational motion
Rev. 71 (1969) 5–45.
of benzene molecule and intermolecular interaction in the mixture of
[29] M. Pitoňák, P. Neogrády, J. R̆ezáč, P. Jurečka, M. Urban, P. Hobza, Benzene
chloroform and benzene by raman spectroscopy, J. Chem. Phys. 65 (1976)
dimer: High-level wave function and density functional theory calculations, J.
4339–4342.
Chem. Theory Comput. 4 (2008) 1829–1834.
[11] M.L. Kijevčanin, S.P. Šerbanović, I.R. Radović, B.D. Djordjević, A.Ž. Tasić,
[30] J. Dykyj, J. Svoboda, R.C. Wilhoit, M. Frenkel, K.R. Hall, Vapor pressure and
Volumetric properties of the ternary system ethanol + chloroform + benzene
antoine constants for hydroncarbons, and sulfur, selenium, tellurium, and
at temperature range (288.15–313.15) K: Experimental data, correlation and
halogen containing organic compounds, Springer-Verlag, Berlin Heidelberg,
prediction by cubic EOS, Fluid Phase Equilib. 251 (2007) 78–92.
Berlin, 1999.
[12] R.P. Rastogi, J. Nath, R.R. Misra, Thermodynamics of weak interactions in liquid
[31] H.E. Barner, S.B. Adler, Calculation of solution nonideality from binary t-x data,
mixtures of chloroform and aromatic hydrocarbons, J. Chem. Thermodyn. 3
Ind. Eng. Chem. Res. 12 (1973) 71–75.
(1971) 307–317.
[32] C.L. Yaws, Thermophysical properties of chemicals and hydrocarbons, 2nd,
[13] K.W. Klinkhammer, P. Pyykko, Ab initio interpretation of the closed-shell
Elsevier Inc., Waltham, MA, 2014.
intermolecular e.Cntdot. Cntdot. Cntdot. E attraction in dipnicogen (H2E-EH2)
[33] G.A. Bottomley, T.H. Spurling, Measurement of the temperature variation of
2 and dichalcogen (HE-EH)2 hydride model dimers, Inorg. Chem. 34 (1995)
virial coefficients. II. The second virial coefficients of benzene, Aust. J. Chem. 19
4134 4134.
(1966) 1331–1341.
[14] H.Y. Xu, W. Wang, J.W. Zou, X.L. Xu, Theoretical calculations of p-type
[34] R.R. Tarakad, R.P. Danner, An improved corresponding states method for polar
pnicogen bonds in the triad intermolecular complexes, J. Theor. Comput.
fluids: Correlation of second virial coefficients, AICHE J. 23 (1977) 685–695.
Chem. 13 (2015) 1450068.
[35] C. Tsonopoulos, An empirical correlation of second virial coefficients, AICHE J.
[15] N. Ramanathan, K. Sankaran, S. Kalyanasundaram, PCl3-C6H6 heterodimer:
20 (1974) 263–272.
Evidence for p. . .p phosphorus bonding at low temperatures, Phys. Chem.
[36] B. Long, J. Li, R. Zhang, L. Wan, Solubility of benzoic acid in acetone, 2-propanol,
Chem. Phys. 18 (2016) 19350–19358.
acetic acid and cyclohexane: Experimental measurement and thermodynamic
[16] B. Long, Z. Wang, H. Yang, Y. Ding, Intermolecular interactions of 1,2-
modeling, Fluid Phase Equilib. 297 (2010) 113–120.
diethoxyethane with toluene: An insight from surface and volumetric
[37] B. Long, Y. Xia, Z. Deng, Y. Ding, Understanding the enhanced solubility of 1,3-
properties at different temperatures, J. Mol. Liq. 249 (2018) 1–8.
benzenedicarboxylic acid in polar binary solvents of (acetone+water) at
[17] R.C. Wilhoit, X. Hong, M. Frenkel, K.R. Hall, Densities of aromatic hydrocarbons,
various temperatures, J. Chem. Thermodyn. 105 (2017) 105–111.
Springer, New York, 1998.
[18] W.M. Haynes, CRC handbook of chemistry and physics, 94th, CRC Press, Boca
Raton, FL, 2014.

You might also like