You are on page 1of 8

MOLLIQ-04384; No of Pages 8

Journal of Molecular Liquids xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Journal of Molecular Liquids


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

New tetrapropylammonium bromide-based deep eutectic solvents:


Synthesis and characterizations
Baba Jibril ⁎, Farouq Mjalli, Jamil Naser, Zaharaddeen Gano
Petroleum and Chemical Engineering Department, College of Engineering, Sultan Qaboos University, P.O. Box 33, Al Khoud 123, Muscat, Oman

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

Article history: Deep eutectic solvents are analogues of ionic liquids that are environmentally benign, have a range of desirable
Received 2 March 2014 liquid properties and are easy to prepare. In this study, new tetrapropylammonium bromide (TPAB)-based
Received in revised form 26 July 2014 deep eutectic solvents (DESs) are introduced. Different hydrogen bond donors (HBDs) — ethylene glycol,
Accepted 4 August 2014
triethylene glycol and glycerol were used at different molar ratios of TPAB:HBD (1:2.5–1:5). Effects of tempera-
Available online xxxx
ture and the TPAB:HBD ratios on the DESs were investigated. Solvents of different physicochemical properties
Keywords:
and wide liquid ranges with minimum melting point temperatures of 249.75, 253.95 and 257.05 K were obtained
Deep eutectic solvent for TPAB:ethylene glycol (1:4), TPAB:triethylene glycol (1:3) and TPAB/glycerol (1:3) respectively. The samples
Tetrapropylammonium bromide were further characterized in the temperature range of 293.15 to 353.15 K and the property ranges measured
Ethylene glycol were density (1.096–1.218 g/cm3), viscosity (6.02–1510.0 cP), surface tension (41.9–53.2 mN m−1), conductiv-
Triethylene glycol ity (167.2–11,500.0 μS cm−1), refractive index (1.4466–1.4908) and pH (4.837–7.290). These demonstrate that
Glycerol the samples have excellent potentials for different applications. Relationships between these properties and tem-
Properties perature were explored. Depending on the property, the effect of temperature was modeled using either simple
linear or Arrhenius-based models. The properties of these new DESs could be predicted at different temperatures
for potential applications. The potential of tunable properties makes it possible to employ DESs as a reaction me-
dium, electrochemical process medium, solvent and absorbent. Therefore, these properties must be measured
and modeled for possible use in process simulation package databases.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction considered as advanced ionic liquids [8]. They have been tested in
different fields, especially in electrochemical applications such as
There has been an increase in research interest in ionic liquids due to metal deposition and as electrolytes [9,10]. It was demonstrated
the wide range of possible constituents in their synthesis, their tunable that DESs can be effective media for chemical reactions due to their
properties and immense potential applications [1]. However, in many dissolution or catalyst effects [11,12]. Recently, a DES prepared
cases, the synthesis of ionic liquids (IL) requires the use of organic sol- using choline chloride and urea was used as a medium in catalytic
vents and supply of heat [2,3]. These and the associated waste disposal reactions and turned out to be more effective than the traditional
raise the cost of IL among other challenges [4]. This led to the recent solvents [13]. Another catalytic effect using choline chloride-based
focus on ionic liquid analogues called deep eutectic solvents (DESs), DES was reported in the depolymerization/dehydration of a biopoly-
which were initially demonstrated by Abbott and coworkers in 2003 mer with product yields above 50% [14]. Due to the high viscosity of
[5]. DESs consist of a salt and a hydrogen bond donor (HBD). A large some reported DESs, there are important questions on understand-
number of the possible salts – organic and inorganic (hydrated or anhy- ing the system's hydrodynamics that need to be addressed [4].
drous) – and HBDs make it possible to report many DESs with physico- There are issues related to possible contamination of reaction prod-
chemical properties similar to those of ILs [4]. In addition, DESs exhibit ucts and corrosion by halogen-based DES. This is important due to
high thermal and chemical stabilities and are biodegradable. Another a recent report of adverse effect of chlorine to enzymes in catalytic
advantage is that DESs are generally considered non-toxic, although biochemical reactions [15] and DES possible degradation at the reac-
recent reports have shown that more research is needed to have more tion conditions [16,17]. Such challenges may limit DES potential in
categorical conclusions on the toxicity [6,7]. catalysis applications [18].
These properties of DESs and their advantage over other ionic Therefore, there have been attempts to modify reported DESs or
liquids indicate their great potential in applications and are therefore introduce new ones. Choline chloride/urea DES showed an eutectic
freezing point of 285.15 K. A change from urea to thiourea or chlorine
⁎ Corresponding author. Tel.: +968 24142582. to fluorine led to drastic changes of the freezing point to 342.15
E-mail addresses: baba@squ.edu.om, byjibril@gmail.com (B. Jibril). or 274.15 K respectively [5]. In a first pulsed field gradient nuclear

http://dx.doi.org/10.1016/j.molliq.2014.08.004
0167-7322/© 2014 Elsevier B.V. All rights reserved.

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
2 B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx

Triethylene glycol

Tetrapropylammonium
Bromide
a) Organic salt b) Hydrogen bond donors
Fig. 1. Chemical structures of the DES constituents used in this work.

magnetic resonance study of choline chloride-based DESs, it has been different TPAB/HBD molar ratios are shown in Table 1. An incubator
shown that the molecular structure of the hydrogen bond donor has shaker (Brunswick Scientific Model INNOVA 40R) was used to mix the
strong effects on the mobility of the whole DES samples [19]. In a recent salt and the HBD. Each DES mixture was mixed at 400 rpm and
study, choline chloride-based DESs were formed with renewable HBDs 353.15 K for 2 h until a homogeneous transparent colorless liquid was
such as itaconic and tartaric acid with freezing points of 330.15 and formed. DES samples were synthesized at atmospheric pressure and
320.15 K respectively. Changing this system to a ternary by addition of under tight control of moisture content. Samples were kept in well-
glycerol led to a drastic decrease in the freezing point and other proper- sealed glass vials after preparation and fresh samples were used for
ties [20]. Therefore, DES properties such as diffusivity, ionic conductivity analysis to avoid structural change or environmental effects on their
and viscosity could be tuned based on the selection of the HBD or co- physical properties. The properties measured were density, viscosity,
HBD. In addition, the selection of the salt, HBD and salt/HBD ratios, the surface tension, conductivity, reflective index and pH. The temperature
salt symmetry and the nature of HBD are important parameters for range considered for all measured physical properties was 293.15–
determining the final properties of the DES. 353.15 K. A further detail on the preparations and characterizations is
Based on the foregoing, we found it of interest to study new deep given earlier [21].
eutectic solvents based on tetrapropylammonium bromide (TPAB) as
quaternary ammonium salt and different HBDs. The HBDs considered 3. Results and discussion
are ethylene glycol, triethylene glycol and glycerol. These may lead to
DESs of different physicochemical properties with potentials for differ- The deep eutectic solvents were prepared using TPAB as the organic
ent applications. The study focuses on the synthesis and characteriza- salt. Effects of variations of the HBD, TPAB/HBD ratio and temperature
tion of the new DESs. The characterization was done by measuring on the physicochemical properties of the samples were studied.
density, viscosity, surface tension, conductivity, reflective index and Table 1 displays the TPAB/HBD ratios, their abbreviations and freezing
pH of the synthesized DESs at different TPAB:HBD molar ratios and tem- points for different ratios. The individual constituents of the DES (and
peratures. In order to explore possible range of applications, regressions their freezing points) are tetrapropylammonium bromide (543.15 K),
of these properties with temperature were performed using simple ethylene glycol (260.3 K), triethylene glycol (266.2 K) and glycerol
mathematical models. This systematic study of the physical properties (290.5 K). As shown in the table, the DES samples exhibited eutectic
of ionic liquid-analogue called deep eutectic solvent is important due points at TPAB/HBD ratio of about 1:4. By adding 4 mol of ethylene gly-
to potentials of using them in process simulation package databases. col to 1 mol of TPAB a freezing point of as low as 249.75 K was observed.
Currently, ionic liquid and deep eutectic solvent property data are not Such decrease in freezing; lower than either of the constituents, is asso-
available in the database of the commercial process simulators. The ciated with sequestration effect of HBD around the bromide ions. Due to
new DES characterized could be added to such simulators. The mathe- hydrogen bonding, the anion is shielded from the quaternary ammoni-
matical model may be used to extrapolate the properties. For separation um cation, thereby weakening the electrostatic attractions between the
and heat transfer process equipment design; properties such as viscos- cation and anion. With such weak interactions, the lattice energy
ity, surface tension, density and pH are important. The results reported decreased, thus leading to lower freezing point. As expected, a HBD
in this work could be used to estimate the properties for design and/or with more linear chain shows less decrease in the freezing point [4].
operation of the process equipment. For TPAB–GLY, the lowest freezing point of 257.05 K was obtained.
Both results suggest easier packing arrangements of the TPAB–TEG
2. Experimental
Table 1
2.1. Chemicals used Compositions and abbreviations of the studied DESs.

Molar ratio Abbreviation Freezing point, K


For the preparations of the DESs, the chemicals used are tetra-
TPAB:ethylene glycol
propylammonium bromide (TPAB), ethylene glycol, triethylene glycol
1:3 DES–EG3 259.85
and glycerol (with purity N 98%) all supplied by Merck Chemicals 1:4 DES–EG4 249.75
(Darmstadt, Germany). The chemical structures of these chemicals are 1:5 DES–EG5 255.65
shown in Fig. 1.
TPAB:triethylene glycol
1:2.5 DES–TEG2.5 261.15
2.2. Preparation and characterization of TPAB-based deep eutectic solvents 1:3 DES–TEG3 253.95
1:4 DES–TEG4 263.55

The DESs were prepared with TPAB as the organic salt. The HBD used TPAB:glycerol
were ethylene glycol, triethylene glycol and glycerol. The TPAB:HBD of 1:2 DES–GLY2 258.15
molar ratios 1:3, 1:4, and 1:5 were used. In the case of TPAB:triethylene 1:3 DES–GLY3 257.05
1:4 DES–GLY4 268.45
glycol the ratios were 1:2.5, 1:3, and 1:4. DES samples synthesized in

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx 3

1.15 48
TPAB : EG (1 : 3)
TPAB : EG (1 : 3)
1.14 TPAB : EG (1 : 4)
47 TPAB : EG (1 : 4)
TPAB : EG (1 : 5) TPAB : EG (1 : 5)

Surface Tension, mN/m


46
1.13
Density, g/cm3

45
1.12

44
1.11
43
1.10
42
a
1.09 a
290 300 310 320 330 340 350 360 41
300 310 320 330 340 350 360
Temperature, K
Temperature, K
1.15 47
TPAB : TEG (1 : 2.5) TPAB : TEG (1 : 2.5)
TPAB : TEG (1 : 3) TPAB : TEG (1 : 3)
1.14 TPAB : TEG (1 : 4) 46 TPAB : TEG (1 : 4)

Surface Tension, mN/m


Density, g/cm3

1.13 45

44
1.12

43
1.11
42
b
1.10 b
290 300 310 320 330 340 350 360 41
Temperature, K 300 310 320 330 340 350 360
Temperature, K
1.23 54
TPAB : GLY (1 : 2)
TPAB : GLY (1 : 3) TPAB : GLY (1 : 2)
1.22
TPAB : GLY (1 : 4) TPAB : GLY (1 : 3)
52 TPAB : GLY (1 : 4)
1.21
Surface Tension, mN/m
Density, g/cm3

1.20
50
1.19

1.18 48
1.17

1.16 46
c
1.15 c
290 300 310 320 330 340 350 360 44
Temperature, K 300 310 320 330 340 350 360
Temperature, K
Fig. 2. Variations of densities with temperature for DESs of different hydrogen bond
donors; (a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol. Fig. 3. Variations of surface tension with temperature for different hydrogen bond donors;
(a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol.

Table 2 Table 3
Density–temperature model parameters and regression coefficients. Surface tension–temperature model parameters and regression coefficients.

DESs a b R2 DESs a (mN m−1) b R2

DES–EG3 1.3379 −6.7857E−04 0.9997 DES–EG3 71.90 −0.0849 0.9958


DES–EG4 1.3371 −6.8214E–04 0.9996 DES–EG4 73.20 −0.0877 0.9923
DES–EG5 1.3312 −6.6429E−04 0.9967 DES–EG5 73.28 −0.0871 0.9956
DES–TEG2.5 1.3522 −7.0000E−04 1.0000 DES–TEG2.5 73.56 −0.0894 0.9864
DES–TEG3 1.3512 −7.0000E−04 1.0000 DES–TEG3 70.84 −0.0814 0.9911
DES–TEG4 1.3567 −7.1790E−04 1.0000 DES–TEG4 65.91 −0.0669 0.9570
DES–GLY2 1.3887 −6.6430E−04 0.9996 DES–GLY2 83.17 −0.1060 0.9907
DES–GLY3 1.4027 −6.6430E−04 0.9997 DES–GLY3 79.45 −0.0883 0.9893
DES–GLY4 1.4127 −6.6430E−04 0.9999 DES–GLY4 80.72 −0.0900 0.9907

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
4 B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx

100 systems at all salt/HBD ratios were measured at different temperatures


TPAB : EG (1 : 3) that ranges from 293.15 to 353.15 K. The results of density variations
TPAB : EG (1 : 4)
TPAB : EG (1 : 5)
with temperature for the three DES systems at all ratios are shown in
80
Fig. 2. The densities of the individual constituents of the DESs at
293.15 K are; TPAB (1.15 g/cm3), ethylene glycol (1.11 g/cm3),
Viscosity, cP

60 triethylene glycol (1.10 g/cm3) and glycerol (1.26 g/cm3). No signifi-


cant change in density was observed for DES–EG and DES–TEG as a
result of changing the salt:HBD ratio. However, the change in density
40 for DES–GLY was more pronounced when the salt:HBD was changed
from 1:3 to 1:5. The effect of temperature on DES density was explored.
The density decreased linearly with increasing temperature. This behav-
20
ior of density with temperature was reported for other DES systems
a [21]. With increase in temperature (293.15–353.15 K), the values de-
0 creased linearly (from 1.139 to 1.098 g/cm3). In the same temperature
290 300 310 320 330 340 350 360 range, for DES–TEG the density decreases from 1.147 to 1.105 g/cm3.
Temperature, K For both DES–EG and DES–TEG no significant differences in values of
densities were observed with change in the ratios. More significant
variation in density with DES–GLY ratio was observed. The values
140
change from 1.194 to 1.218 g/cm3 for the ratios 1/2.5 to 1/4. For DES–
TPAB : TEG (1 : 2.5)
TPAB : TEG (1 : 3) GLY ratio of 1/4, with increase in temperature (293.15–353.15 K), the
120
TPAB : TEG (1 : 4) density decreases linearly from 1.218 to 1.178 g/cm3. The decreases in
100 densities with respect to temperature are the results of volume increase
due to thermal expansion and gains in both vibrational and translational
Viscosity, cP

80 energy of the DES ensembles and free ions. At 293.15 K (ratio = 1:3),
the values of density obtained are higher than typical values of choline
60 chloride:GLY [24]. This agrees with the suggestion above that the pack-
ing is easier in the case of TPAB–GLY than choline chloride. The densities
40 were fitted using a simple linear model as follows:

20 ρ ¼ a þ b ðTÞ ð1Þ
b
0 where ρ is the density in g/cm3, T is the temperature in Kelvin, and a and
290 300 310 320 330 340 350 360
b are constant parameters that depend on the DES molar ratios and
Temperature, K nature of the HBD. The values of a and b are shown in Table 2. The
results show strong linear relations between the density and tempera-
1600
ture for all samples. Again, as indicated above; the results of DES–EG
TPAB : GLY (1 : 2)
1400 TPAB : GLY (1 : 3) and DES–TEG are similar, though the rate of change in density with tem-
TPAB : GLY (1 : 4) perature is higher in the case of DES–TEG. This may be associated with
1200
the longer chain and higher electron density of the TEG than EG. Thus,
1000 the change in temperature has more significant effect on the DES–
Viscosity, cP

TEG. DES–GLY exhibited lower rate of change of density with tempera-


800
ture. This is due to relatively higher number of hydrogen bonding sites
600 which restricts its translational and vibrational movements or expan-
sion with increase in temperatures.
400

200 5. Surface tension

0
c Surface tension is very useful in characterizing solvents from differ-
ent constituents; because in the body of a given sample, molecular
290 300 310 320 330 340 350 360 ensembles or free ions move about responding to both long-range and
Temperature, K short-range forces from each other. A balance among the forces and
the existence of ions is determined by factors such as salt/HBD ratio,
Fig. 4. Variations of viscosity with temperature for different hydrogen bond donors; the nature of HBD and temperature. As shown in Fig. 3, for both EG
(a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol.

Table 4
Viscosity–temperature model parameters and regression coefficients.
and TPAB–GLY than TPAB–EG. The relative lower value for TPAB–EG
than TPAB–GLY is similar to earlier observations [22,23]. However, the DESs μo (cP) E/R R2
freezing points reported here are much higher than typical values of DES–EG3 7.3944E −05 4104 0.998
207.15 and 233.15 K obtained for choline chloride–EG [22] and choline DES–EG4 1.3557E −04 3806 0.998
DES–EG5 2.3765E −04 3577 0.999
chloride–GLY [23] DESs respectively.
DES–TEG2.5 9.8545E −05 4112 0.998
DES–TEG3 1.1727E −04 4040 0.998
4. Density DES–TEG4 1.2704E −04 3957 0.998
DES–GLY2 1.4164E −07 6772 0.998
Density is an important property for determining solvent diffusion DES–GLY3 3.4147E −07 6407 0.998
DES–GLY4 2.1557E −07 6536 0.998
and miscibility with other liquids. The densities of all synthesized DES

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx 5

and TEG, at 303.15 K, the surface tension of about 46 mN m− 1 was increase in kinetic energy. The surface tensions were fitted using a
obtained. The values did not exhibit significant change by changing simple linear model as follows:
the TPAB/HBD ratio [3–5]. For each ratio, it shows linear decrease (46–
42 mN m−1) with increase in temperature (303.15–353.15 K). When γ ¼ a þ b ðTÞ ð2Þ
the HBD was changed to GLY, the surface tension shows more signifi-
cant change (51–53 mN m−1) for DES–GLY ratios of 1/2 to 1/4. At the where γ is the surface tension in mN m−1, T is the temperature in
ratio of 1/2, the value varies (51 to 46 mN m− 1) in the temperature Kelvin, and a and b are constant parameters. The values of a and b are
range. These are similar to values obtained for choline chloride–glycerol shown in Table 3. The results show strong linear relation between the
system [24,25]. The decreases in surface tension with increases in tem- surface tension and temperature for all samples. Although for DES–EG
perature are due to weaker interactions among the constituents due to and DES–TEG the results are similar, the rate of decrease with tempera-
ture is higher in the former. DES–GLY exhibited higher values of surface
tension and relatively higher rate of decrease with temperature than the
other two samples.

6. Viscosity

Similar to the observation in the case of surface tension, the viscosity


is strongly affected by both changes in the HBD, TPAB/HBD ratio and
temperature. As shown in Fig. 4, for DES–EG of 1/3, at 303.15 K the
value is 58.2 cP. It decreases to 32.8 cP for the ratio of 1/5. It also
decreases drastically from 86.1 to 8.6 cP with temperature (293.15–
353.15 K). When EG was replaced with TEG, higher values of viscosities
were observed. For DES–TEG of 1/2.5, at 303.15 K the value is 78.2 cP. It
decreases to 61.8 cP for the ratio of 1/4. It also decreases drastically from
117.0 to 11.8 cP with temperature (293.15–353.15 K). For TPAB/GLY
a much higher viscosity was observed. As shown in the figure, at room
temperature the viscosity is 1510.0 cP at DES–GLY = 1/2, the value de-
creases to 980.0 at the ratio of 1:4. The sample exhibited more drastic
decrease in viscosity (1510.0 to 33.0 cP) with temperature (293.75 to
352.75 K). The low viscosity values obtained here are very useful for po-
tential applications. Many DESs have viscosities higher than 100 cP at
room temperature [4]. This is due to strong interactions such as hydro-
gen bonding and electrostatic attractions among the DES components.
As shown here, the selection of HBD and its ratio with the correspond-
ing salt is very important in designing low viscosity DES samples.
In order to quantify the effects of the HBD, the variation of viscosity
with temperature was modeled using an Arrhenius equation:

μ ¼ μ o exp½Em =RT ð3Þ

where μ is the viscosity, μo is a pre-exponential constant, Eμ is the activa-


tion energy, R is the gas constant, and T is the temperature in Kelvin. μo
b and Eμ are model parameters that depend on the nature of the HBDs.
The parameters are shown in Table 4. As shown in the table, there is a
strong relation (R2 = 0.998) between the viscosity and temperature
for all samples. The value of Eμ/R exhibit general decrease with increase
of TPAB/HBD ratio. Longer chain length (EG vs TEG) and higher number
of hydrogen bonding sites (EG vs GLY) lead to higher activation energy.

7. Conductivity

Fig. 5 shows the conductivities of the DESs for different TPAB/HBD


ratios and temperatures (293.15–353.15 K). The conductivity exhibits

Table 5
Conductivity–temperature model parameters and regression coefficients.

DESs ko (μS cm−1) E/R R2

DES–EG3 1.493E + 07 −2545 0.965


DES–EG4 1.553E + 07 −2518 0.990
c DES–EG5 1.677E + 07 −2541 0.958
DES–TEG2.5 2.509E + 07 −3027 0.976
DES–TEG3 2.444E + 07 −3048 0.990
DES–TEG4 1.383E + 07 −2866 0.986
DES–GLY2 1.373E + 09 −4585 0.979
DES–GLY3 1.890E + 09 −4724 0.997
Fig. 5. Variations of conductivities with temperature for DESs of different hydrogen bond
DES–GLY4 1.967E + 09 −4741 0.994
donors; (a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol.

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
6 B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx

general increase with either increase in the ratios or temperatures. For 25]. The effect of temperature on the conductivity was modeled using
DES–EG at 293.15 K, the value increases from 2180 to 2950 μS cm−1, an Arrhenius equation:
for ratios of 1/3 to 1/5. For DES–TEG, lower value with less level of
increase was observed at similar conditions. This is associated with κ ¼ κo exp½−Eκ =RT ð4Þ
relatively less effect of temperature on the long chain of TEG compared
with EG. This decreases the effective movement of conducting species in where κ is the viscosity, κo is a pre-exponential constant, Eκ is the acti-
the sample. The conductivities exhibited typical corresponding trend vation energy of conductivity, R is the gas constant, and T is the temper-
with that of viscosity. The lower is the viscosity the higher is the conduc- ature in Kelvin. The model parameters are shown in Table 5. The effect
tivity due to higher mobility of charge carrying species in the DESs [24, of temperature on the conductivity is different between short and
long chain HBDs. The rate of increase of conductivity with temperature
is about 15% higher in the case of DES–TEG than DES–EG. This suggests
that as the temperature increases, species gain energy, thereby weaken-
ing the hydrogen bonding between triethylene glycol and the ions.
1.480
Similarly, for DES–GLY the rate of increase in conductivity is about
TPAB : EG (1 : 3)
TPAB : EG (1 : 4)
37% higher than that in DES–TEG. Again, as indicated above, the lower
1.475
TPAB : EG (1 : 5) conductivity of DES–GLY than the other two arises from relatively
higher number of hydrogen bonding sites. With increase in tempera-
1.470
ture, such bonds become weaker thereby showing higher rates of
Refractive Index

increase.
1.465

1.460
8. Refractive index and pH

1.455 Refractive index (RI) is a useful physical property that can be utilized
to characterize new solvents due to its strong association with the
1.450 materials' electrical permittivity and magnetic permeability. There are
limited reports of the refractive index of ionic liquids in general and
a DES in particular. Fig. 6 shows the variations of the RI with temperature
1.445
290 300 310 320 330 340 350 360 for all DES systems. All samples exhibit strong linear decrease in RI with
Temperature, K increase in either the HBD composition or temperature. For DES–EG
(ratio = 1/3), the value decreases by about 3% in the temperature of
1.480 293.15–353.15 K. For the same conditions, the values for DES–TEG
TPAB : TEG (1 : 2.5) change by about 1%. Refractive index-temperature exhibited a strong
1.475
TPAB : TEG (1 : 3) linear relationship and fitted using the following simple linear model.
TPAB : TEG (1 : 4)
Refractive Index

1.470 RI ¼ a þ b ðTÞ ð5Þ

1.465 where T is temperature in K, and a and b are constants that vary


according to the type of DES. The values of the constants are shown
1.460 in Table 6. All the DES samples showed similar rate of change in RI
with temperature.
The pH values of the samples are shown in Fig. 7. It is important to
1.455
estimate the pH of new solvents to understand their corrosion, dissolu-
b tion, and catalytic and other properties useful for applications. General-
1.450 ly, for a non-aqueous solvent, pH depends on the chemical potential of
290 300 310 320 330 340 350 360
hydrogen and temperature. The chemical potential is affected by the
Temperature, K presence of cations and anions. The degrees of association such as
1.495 hydrogen bonding of these ions with other species in the solvent deter-
TPAB : GLY (1 : 2) mine the value of the pH. At room temperature, pH of 6.3 was observed
1.490 TPAB : GLY (1 : 3) for DES–EG (at the ratio of 1/3). No significant change was observed
TPAB : GLY (1 : 4) when the ratio was changed to 1/4. At the ratio of 1/5, the pH value
was 7.3. The result suggests non-linear relation between the hydrogen
Refractive Index

1.485
chemical potential and DES–EG ratio. It implies that below the eutectic

1.480
Table 6
1.475 Refractive index–temperature model parameters and regression coefficients.

DESs a b R2
1.470 DES–EG3 1.5657 −3.10E−04 0.9999
DES–EG4 1.5571 −3.00E−04 0.9996
c DES–EG5 1.5557 −3.10E−04 0.9999
1.465 DES–TEG2.5 1.5623 −3.00E−04 0.9898
290 300 310 320 330 340 350 360
DES–TEG3 1.5641 −3.00E−04 0.9969
Temperature, K DES–TEG4 1.5656 −3.20E−04 0.9997
DES–GLY2 1.5692 −2.65E−04 0.9963
DES–GLY3 1.5722 −2.85E−04 1.0000
Fig. 6. Variations of refractive indices with temperature for DESs of different hydrogen
DES–GLY4 1.5616 −2.61E−04 0.9908
bond donors; (a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol.

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx 7

point there is a higher degree of association of the ions with HBD. This is Table 7
limited when higher amount of HBD was added. At a constant ratio, as pH–temperature model parameters and regression coefficients.

temperature increases the association increases thereby decreasing DESs a b R2


the pH values. When the HBD was changed from EG to TEG, generally DES–EG3 8.7910 −8.000E−03 0.989
lower pH values were observed. For DES–GLY similar pH values were DES–EG4 8.6140 −7.000E−03 0.973
obtained. Contrary to the observation in the case of EG and TEG; glycerol DES–EG5 10.8100 −1.200E−02 0.991
exhibited a decrease in pH with the increase in DES–GLY ratio. This DES–TEG2.5 6.6463 −5.220E−03 0.954
DES–TEG3 6.7218 −5.050E−03 0.993
suggests that as the amount of glycerol increases, more hydrogen bond-
DES–TEG4 6.6549 −5.060E−03 0.989
ing is available. Thus, the chemical potential of hydrogen decreases. The DES–GLY2 8.3678 −6.610E−03 0.977
pH–temperature exhibited a strong linear relationship and fitted to the DES–GLY3 6.5821 −2.085E−03 0.965
following simple linear model. DES–GLY4 7.0024 −3.857E−03 0.968

7.4

7.2 TPAB : EG (1 : 3)
TPAB : EG (1 : 4)
TPAB : EG (1 : 5) pH ¼ a þ b ðTÞ ð6Þ
7.0

6.8 where T is temperature in K, and a and b are constants that depend on


the type of DES. The values of the constant are shown in Table 7. DES–
6.6
EG samples show a rate of decrease of pH with temperature of about
pH

6.4 0.008. The value decreases to 0.005 for DES–TEG. This indicates the
relatively stronger interaction of long chain TEG than EG. Similarly,
6.2
when the HBD was replaced with glycerol, the rate shows further
6.0 decrease. This is associated with stronger interaction of the HBD with
ions due to higher number of bonding sites.
5.8
a
5.6 9. Conclusion
290 300 310 320 330 340 350 360
Temperature, K In this study a new set of tetrapropylammonium bromide
(TPAB)-based deep eutectic solvents (DESs) was prepared using
5.3 three hydrogen bond donors at different salt:HBD molar ratios. The
TPAB : TEG (1 : 2.5) chain length and the number of hydrogen bonding functional groups
TPAB : TEG (1 : 3)
5.2 TPAB : TEG (1 : 4)
of the hydrogen bond donors (HBD) showed strong effects on the
DES physicochemical properties. DES–GLY at a ratio of 1/3, exhibited
5.1 a wide liquid range, with a low eutectic point of about 306.65 K lower
than the HBD freezing point. The samples were further characterized
in the temperature range of 293.15 to 353.15 K. Important solvent
pH

5.0
physicochemical properties including density, surface tension, vis-
cosity, conductivity, refractive index and pH were measured. The
4.9 effects of temperature as well as the salt:HBD of the DESs were dem-
onstrated. Wide property ranges were observed — density (1.096–
4.8 1.218 g/cm3), viscosity (6.02–1510.0 cP), surface tension (41.9–
53.2 mN m− 1), conductivity (167.2–11,500.0 μS cm− 1), refractive
4.7
b index (1.4466–1.4908) and pH (4.837–7.290). Property–tempera-
290 300 310 320 330 340 350 360 ture relationships were fitted to either simple linear models or
Temperature, K Arrhenius' law. The ranges of the properties indicate that these
new deep eutectic solvents have a strong potential in applications
6.6 such as metal dissolutions and reaction media. Currently, ionic liquid
TPAB : GLY (1 : 2) and deep eutectic solvent physical properties are not available in the
6.4 TPAB : GLY (1 : 3) database of the commercial process simulators. The new DES proper-
TPAB : GLY (1 : 4) ties could be added to such simulators.

6.2
Acknowledgments
pH

6.0 We appreciate the financial support of the College of Engineering,


Sultan Qaboos University, Muscat, Oman.
5.8
References
5.6 [1] Q. Zhang, S. Zhang, Y. Deng, Green Chem. 13 (2011) 2619–2637.
[2] R. Rogers, K. Seddon, S. Volkov (Eds.), Springer, 2003.

5.4
c [3] S.B. Phadtare, G.S. Shankarling, Green Chem. 12 (2010) 458–462.
[4] Q. Zhang, K. De, O. Vigier, S. Royer, F. Jerome, Chem. Soc. Rev. 41 (2012) 7108–7146.
290 300 310 320 330 340 350 360 [5] A.P. Abbott, G. Capper, D.L. Davies, R.K. Rasheed, V. Tambyrajah, Chem. Commun.
Temperature, K (2003) 70–71.
[6] M. Deetlefs, K.R. Seddon, Green Chem. 12 (2010) 17–30.
[7] A. Romero, A. Santos, J. Tojo, A. Rodrıguez, J. Hazard. Mater. 151 (2008) 268–273.
Fig. 7. Variations of pH with temperature for DESs of different hydrogen bond donors; [8] V. De Santi, F. Cardellini, L. Brinchi, R. German, Tetrahedron Lett. 53 (2012)
(a) ethylene glycol, (b) tri-ethylene glycol and (c) glycerol. 5151–5155.

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004
8 B. Jibril et al. / Journal of Molecular Liquids xxx (2014) xxx–xxx

[9] E. Gomez, P. Cojocaru, L. Magagnin, E. Valles, J. Electroanal. Chem. 658 (2011) 18–24. [19] C. D'Agostino, R.C. Harris, A.P. Abbott, L.F. Gladden, M.D. Mantle, Phys. Chem. Chem.
[10] P. Cojocaru, L. Magagnin, E. Gomez, E. Valles, Mater. Lett. 65 (2011) 3597–3600. Phys. 13 (48) (2011) 21383–21391.
[11] Y. Jung, R.A. Marcus, J. Am. Chem. Soc. 129 (2007) 5492–5502. [20] Z. Maugeri, P. Domínguez de Marıa, RSC Adv. 2 (2012) 421–425.
[12] F. Jutz, J. Andanson, A. Baiker, Chem. Rev. 111 (2011) 322–353. [21] F.S. Mjalli, J. Naser, B.Y. Jibril, S.S. Al-Hatmi, Z.S. Gano, Thermochim. Acta 575 (2014)
[13] Y.A. Sonawane, S.B. Phadtare, B.N. Borse, A.R. Jagtap, G.S. Shankarling, Org. Lett. 12 135–143.
(2010) 1456–1459. [22] M. Hayyan, F.S. Mjalli, M.A. Hashim, I.M. AlNashef, Fuel Process. Technol. 91 (2010)
[14] H. Zhao, J.E. Holladay, H. Brown, Z.C. Zhang, Science 316 (2007) 1597–1600. 116–120.
[15] H. Zhao, G.A. Baker, S. Holmes, Org. Biomol. Chem. 9 (2011) 1908–1916. [23] K.A. Shahbaz, F.S. Mjalli, M.A. Hashim, I.M. AlNashef, J. Appl. Sci. 10 (2010)
[16] D. Linberg, M. de la Fuente Revenga, M. Widersten, J. Biotechnol. 147 (2010) 3349–3354.
169–171. [24] A.P. Abbott, R.C. Harris, K.S. Ryder, J. Phys. Chem. B 111 (2007) 4910–4913.
[17] D. Yue, Y. Jia, Y. Yao, J. Sun, Y. Jing, Electrochim. Acta 65 (2012) 30–36. [25] A.P. Abbott, D. Boothby, G. Capper, D.L. Davies, R.K. Rasheed, J. Am. Chem. Soc. 126
[18] F. Rantwijk, R.A. Sheldon, Chem. Rev. 107 (2007) 2757–2785. (2004) 9142–9147.

Please cite this article as: B. Jibril, et al., Journal of Molecular Liquids (2014), http://dx.doi.org/10.1016/j.molliq.2014.08.004

You might also like