You are on page 1of 27

119

6 Solventless Chlorination of Glycerol to


Dichloropropanols


120

6.1 INTRODUCTION
Biodiesel production and availability of glycerol as the co-product in ~10% by weight
have been much talked about subjects. Huge quantities of bioglycerol will be
available in future if the correct projections materialize and thus bioglycerol needs to
be valorized to arrive at economically favorable cost of biodiesel (Stelmachowski,
2011; Fan et al., 2010; Yadav et al., 2012
a,b,c,d
). Although glycerol itself has wide
applications in foods, paint, pharmaceutical, cosmetics, soap, toothpaste industries,
the amount of glycerol produced will be massive and hence its new application per se
or value added derivatives must be produced. Some of them include conversion of
glycerol to acrolein, citric acid, lactic acid, dihydroxyacetone, 1,3-propanediol, 1,3-
and 1,2-dichloropropanols, epichlorohydrin, and a host of ethers.

Dichloropropanols are valuable chemicals which can be produced from glycerol.
They are used as intermediates in epichlorohydrin production. Epichlorohydrin is
enormously versatile intermediate used in wide variety of applications.
Approximately 76% of the worlds consumption of epichlorohydrin is used to make
epoxy resins (Bruin, 1959; Brojer et al., 1965). It has also been used to cure
propylene-base rubbers, as a solvent for cellulose esters and ethers, surface active
agent used in cosmetics and shampoos, as a stabilizer in rubber exhibiting resistance
to extreme temperatures applications in automotive and aircraft parts, seals gaskets,
and agricultural products such as insecticides, bactericides and fungicides and in
resins with high wet-strength for the paper industry (Weissermel and Arpe, 1997).
Epichlorohydrin is prepared industrially by various processes which include high
temperature chlorination of propylene to allyl chloride, followed by hypo-chlorination
of allyl chloride to give dichloropropanols and then by dehydrochlorination with
caustic (Scheme 6.1, Route 1). However, this process has considerable drawbacks
which include sacrificial use of chlorine and complications due to the industrial use
and generation of hypochlorous acid and formation of unwanted chlorinated
hydrocarbons including trichloropropane, chlorinated ethers and oligomers (Scheme
6.2) (Weissermel and Arpe, 1997; Green et al., 2000; Aoki et al., 2011; Nagato et al.,
1987; Kasai et l., 1992; Stephen and Maria, 1991; Kaisha, 1969; Kawabe et al., 1978;
Grimsby, 1986; Grimsby, 1987; Mass and Petrus, 1994; Dettloff and Null, 2005).
Another known process involves insertion of oxygen in allylic position of allyl
chloride using hydrogen peroxide over titano-silicate catalyst (Scheme 6.1, Route 2)
121

(Clerici and Ingallina; 1993; Bhaumik et al., 1994; Gao et al., 1996; Thomas et al.,
1992; Dartt and Davis, 1994). Propylene is converted to allyl alcohol by oxidative
acetoxylation to allyl acetate followed by hydrolysis in the Showa Denko process
commercialized in 1980s. Allyl alcohol is then chlorinated in aqueous HCl to give
glycerol dichlorohydrin, followed by dehydrochlorination using a base to give
epichlorohydrin (Scheme 6.1, Route 3) (Aoki, 1999). Dows patented process starts
with propylene oxidation to acrolein in first step, followed by chlorination to 2, 3-
dichloropropanal in second step, which is further hydrogenated to give 2, 3-
dichloropropanol (Scheme 6.1 Route 4). However, in this case 2,3-dichloropropanol
(2,3-DCP) is predominant which is 10 fold less reactive than 1,3- dichloropropanol
(1,3-DCP) in the process of preparing epichlorohydrin (Kenneth et al., 1958). Both
1,3- and 1,2- DCPs are the starting materials for epichlorohydrin manufacture among
which 1,3-dichloropropanol is almost one order of magnitude more reactive to form
epichlrohydrin. Thus, selectivity to 1,3-DCP is of paramount importance for
commercial operation to reduce reactor size and time. According to an Asahi patent,
acetone produced in the Hock process for phenol can be chlorinated to give
dichloroacetone. Upon hydrogenation of dichloroacetone, as per a related Mitsubishi
patent, dichloropropanol is formed (Scheme 6.1, Route 5). Dehydrochlorination with
a base affords epichlorohydrin (Denttloff and Null, 2005). All the processes for
epichlorohydrin production go through dichloropropanols as intermediates. Glycerol
can be converted to dichloropropanol by reacting it with hydrochloric acid in presence
of organic carboxylic acids as the catalyst followed by dehydrochlorination with a
base to give epichlorohydrin; it offers many economical and environmental
advantages over conventional method (Gosselin et al., 2005; Schreck et al., 2006;
Gilbeau, 2006; Siano et al., 2006; Krafft et al., 2007; Krafft and Gilbeau, 2008; Mehta
et al., 2010).

Substantial work has been done for the conversion of glycerol to dichloropropanols.
All patents describe use of homogeneous organic acids such as carboxylic acid and its
derivatives, dicarboxylic or polycaboxylic acid with 2-10 carbon atoms, an anhydride,
an acid chloride, an ester, a lactone, a lactam, an amide, a metal organic compound
and a metal salt (Gosselin et al., 2005; Schreck et al., 2006; Gilbeau, 2006; Siano et
al., 2006; Krafft et al., 2007; Krafft and Gilbeau, 2008; Mehta et al., 2010). Processes
using homogeneous organic acids as catalysts suffer from considerable drawbacks
122

such as, difficulties in separation of product 1, 3 -dichlorohydrin from corrosive
reaction mixture containing water, organic acid, dissolved HCl; loss of the catalyst
during the reaction due to the low boiling organic acid catalyst such as acetic acid;
and decrease in the rate of reaction due to introduction of water in the reaction
mixture by means of aqueous hydrochloric acid use, which prolongs reaction time.

Scheme 6.1: Various industrial routs to synthesize epichlorohydrin

Scheme 6.2: Side reaction of scheme 4.1 route 1

Lee et al. solved this problem by using heteropolyacids s catalyst (Lee et al., 2008)
but their yields are very low and solubility in reaction mixture leads to separation
problems. Hence it is needed to develop an efficient and reusable heterogeneous
catalytic system.

123

In present study, chlorination of glycerol was studied with hydrogen chloride gas by
using MUICaT-5, which is modified version of zirconia catalyst. The catalyst showed
excellent activity, selectivity and stability for dehydration of glycerol to acrolein as
well as in chlorination of glycerol to dichloropropanol (Yadav et al., 2010
a,b
; Yadav
and Surve, 2011; Tesser et al., 2007). The reaction parameters were optimized for
chlorination reaction in semi-batch process at atmospheric pressure with continuous
flow of HCl gas. Reaction was also studied in semi-batch process at super-
atmospheric pressure in autoclave with HCl gas pressure and kinetic model has been
developed for the same. These results are novel.

6.2 EXPERIMENTAL
6.2.1 Chemicals and catalysis
All chemical were procured from reputed firm and used without further purification:
Glycerol (LR), (S.D. Fine Chem., Mumbai, India), 3-chloro-1,2-propanediol (Sigma-
Aldrich,USA) hexadecyl amine (M/s. Spectrochem Ltd., Mumbai, India). Hydrogen
chloride gas (M/s. A.A. Traders, Mumbai)

6.2.2 Catalyst Synthesis
MUICaT-5 prepared by method reported by Yadav et al. (Yadav et al., 2010
a,b
).

6.2.2.1 Synthesis of Hexagonal mesoporous silica (HMS)
5 g Dodecyl amine was dissolved in 41.8 g of ethanol and 29.6 g of distilled water.
20.8 g of tetraethyl orthosilicate was added under vigorous stirring. The addition of
ethanol improved the solubility of the template. The reaction mixture was aged for 18
h at 30 C. The clear liquid above the white colored precipitate was decanted and the
precipitate HMS was dried on a glass plate. The template was removed by calcining
the resulting material at 650 C in air for 3 h.

6.2.2.2 Synthesis of MUICaT-5 Catalyst
Zirconium oxychloride (2.39 g) and aluminum nitrate (0.11 g) were dissolved in
distilled water. This solution was then added by incipient wetness technique to 5.0 g
of precalcined HMS. The resulting solid material is then dried in an oven at 110 C
for 3 hours. The dried material was hydrolysed by ammonia gas and washed with
deionised water until a neutral filtrate is obtained, then it is filtered, dried in an oven
124

for 24 hours at 110 C. The generation of super acidic centers in to this material is
made by grinding tungustic acid (0.1078 g) with 0.9 g of zirconium hydroxide and
aluminum hydroxide on HMS, followed by hydrothermal treatment and then it is
calcined at 750 C for 3 hours.

6.2.3 Experimental Setup
6.2.3.1 Semi-batch process at atmospheric pressure
Semi-batch experiments at atmospheric pressure were studied in a 5 cm internal
diameter, fully baffled mechanically agitated glass reactor 100 cm
3
total capacity with
reactive distillation assembly. The reactor was kept in an isothermal oil bath whose
temperature could be maintained at a desired value. It was equipped with a 6 bladed-
turbine impeller and equi-spaced baffles. Gaseous hydrogen chloride was fed directly
to the bottom of the reactor through classical dispersing device at a constant flow rate.

6.2.3.2 Semi-batch process at super atmospheric pressure
Semi-batch experiments were also carried out at super atmospheric pressure in a Parr
Autoclave made of non corrosive Hastalloy-C with 100 cm
3
capacity which was with
magnetically driven. It consisted of a four-pitched blade turbine impeller, temperature
indicator controller, speed controller, pressure controller and thermocouple. Gaseous
hydrogen chloride from a cylinder was fed directly to the bottom of the reactor and its
flow maintained with a flow controller at a desired pressure.

6.2.4 Reaction procedure
6.2.4.1 Semi-batch process at atmospheric pressure
In a typical experiment, 40 g (0.43 mol) of glycerol was fed to the reactor. The reactor
was kept in an isothermal oil bath adjusted to the desired temperature and 2 g of
catalyst (5% w/w) was then added. Initial sample was withdrawn and anhydrous
hydrogen chloride gas was passed through the bottom of the reactor at a desired flow
rate. Speed of agitation was kept at 1100 rpm. Further samples were drawn at periodic
intervals up to 10 h.

6.2.4.2 Semi-batch process at super atmospheric pressure
Glycerol 40 g (0.43 mol) was charged along with the catalyst 2 g (5% w/w) in the
reactor. The reactor was heated up to the desired temperature and anhydrous hydrogen
125

chloride gas fed to reactor at the desired pressure. Initial sample was withdrawn and
speed of agitation was kept at 1100 rpm. Further samples were drawn at periodic
intervals up to 4 h. In this process, chlorination proceeds through formation of -
monohloropropanediol (-MCPD) and subsequently 1,3- and 1,2-dichloropropanols
are formed. Therefore, chlorination of -monohloropropanediol (MCPD) was also
studied independently to establish the kinetics of both mono and dichlorination of
glycerol. In a typical experiment -monochloropropanediol 40 g (0.36 mol) charged
along with the catalyst 1.2 g (3% w/w) in the reactor. The reactor was heated up to the
desired temperature and anhydrous hydrogen chloride gas was fed to the reactor at a
desired pressure. Initial sample was withdrawn and speed of agitation was kept at
1100 rpm. Further samples were drawn at periodic intervals up to 4 h.

6.2.5 Method of analysis
Samples were centrifuged to remove traces the catalyst and weighed quantity reaction
mixture (containing dissolved HCl gas) was dissolved in methanol and neutralized by
passing through sodium carbonate bed. External standard method was used to
calculate conversion and yield; n-hexanol was used as the external standard.
Standardized samples were analyzed by using gas chromatography (Chemito, GC
1000 using BP-20 capillary column (0.25 i.d., 30 m length) and FID. Reaction
products were confirmed by GC-MS (Perkin-Elmer Clarus 500 Model,) with BP-20
capillary column (0.25 i.d., 30 m length).

6.3 RESULT AND DISCUSSION
Scheme 4.3 depicts the reaction pathways in chlorination of glycerol. Both - and -
monochloropropanediols (MCPD) are formed.
6.3.1 Semi-batch process at atmospheric pressure
6.3.1.1 Proof of absence of external mass transfer resistance
The effect of speed of agitation was studied in the range of 8001200 rpm at a catalyst
loading 5 wt% of glycerol under otherwise similar conditions. The glycerol
conversion and dichlorohydrin selectivity were practically the same at these speeds
(Figure 6.1), which indicated the elimination of mass transfer resistance in studied
ranges of speed.
126


Figure 6.1: Effect of speed of agitation
Glycerol, 40 g (0.43 mol), HCl gas flow rate, 40 ml/min, reaction temperature, 160
C, catalyst loading, 0.06 g/cc, - Glycerol conc. at 900 rpm, 1,3-DCP conc. at
900 rpm, -Glycerol conc. at 1000 rpm, 1,3-DCP conc. at 1000 rpm, -
Glycerol conc. at 1200 rpm, 1,3-DCP conc. at 1200 rpm

6.3.1.2 Effect of HCl gas flow
Effect of anhydrous HCl gas flow was studied in the range of 20-50 cm
3
/min on yield
of dichloropropanols under otherwise similar conditions (Figure 6.2). As HCl gas
flow was increased, the selectivity towards dichloropropanol increased but this effect
was observed up to 40 cm
3
/min suggesting mass transfer limitation or deficiency of
HCl in the reaction mass. Since mass transfer resistance was absent, it was the
deficiency of HCl. Further increase in flow rate did not affect selectivity of
dichloropropanols and led to formation of other dehydration products (Figure 6.3). As
the reaction is reversible in nature, presence of water in reaction mixture decreases
rate of reaction. Increase in HCl gas flow rate strips out water formed in the reaction
0
10
20
30
40
50
60
70
80
90
100
0 1 2 3 4 5 6 7 8 9 10 11 12 13
C
o
n
c
e
n
t
r
a
t
i
o
n

(
%
)

Time (h)
127

which further increases reaction rate and subsequently selectivity of
dichloropropanols.

Figure 6.2: Effect of HCl gas flow
Glycerol, 40 g (0.43 mol), reaction temperature, 160 C, catalyst loading, 0.06 g/cc,
speed of agitation, 1000 rpm, Glycerol conc. at 30ml/min, 1,3-DCP conc. at
30 ml/min, Glycerol conc. at 40 ml/min, 1,3-DCP conc. at 40 ml/min,
Glycerol conc. at 50 ml/min, 1,3-DCP conc. at 50 ml/min
0
10
20
30
40
50
60
70
80
90
100
0 1 2 3 4 5 6 7 8 9 10 11 12
C
o
n
c
e
n
t
r
a
t
i
o
n
(

%
)

Time (h)
128


Figure 6.3: Effect of HCl gas flow on selectivity of DCP; Reaction time- 10 h
Glycerol, 1,3-DCP, MCPD, other

6.3.1.3 Effect of catalyst loading
Catalyst loading at different catalyst concentration varied over a range of 0.03 to 0.1
g/cc with respect to the total reaction volume at 160 C. In the absence of external
mass transfer resistance, the rate of reaction is directly proportional to the catalyst
loading based on the entire liquid-phase volume. Percentage Concentrations of
glycerol and dichloropropanol versus time are plotted, which show effect of catalyst
loading on concentration of glycerol and dichloropropanol (Figure 6.4). The rate of
formation of monochloropropanediol from glycerol is very fast and hence there is no
significant effect of catalyst loading on this step-1 (Scheme 6.4). However, the rate of
formation of dichloropropanol from -monochloropropanediol (Scheme 6.5) is very
slow hence increasing catalyst loading, increases concentration of dichloropropanol.
This effect is consistent up to 0.06 g/cm
3
of catalyst loading; the loading above this
shows decrease in selectivity due to formation of other byproducts.
0
10
20
30
40
50
60
70
80
90
100
20 30 40 50
C
o
n
v
e
r
s
i
o
n
/
s
e
l
e
c
t
i
v
i
t
y

(
%
)

HCl gas flow(ml/min)
129


Figure 6.4: Effect of catalyst loading
Glycerol, 40 g (0.43 mol), reaction temperature, 160 C, HCl gas flow rate, 40
ml/min, speed of agitation, 1000 rpm, Glycerol conc. at 0.02 g/cc, 1,3-DCP
conc. at 0.02 g/cc, Glycerol conc. at 0.06 g/cc, 1,3-DCP conc. at 0.06 g/cc,
Glycerol conc. at 0.1 g/cc, 1,3-DCP conc. at 0.1 g/cc

6.3.1.4 Effect of temperature
Effect of temperature was carried over MUICaT-5, with 2 wt% of catalyst
concentration, over a range of 140-160 C, under otherwise similar conditions, in the
absence of external mass transfer resistance. The initial rate of reaction is directly
proportional to the temperature. Percentage concentration of glycerol and
dichloropropanol versus time are plotted (Figure 6.5). The rate of formation of
monochloropropanediol from glycerol is very fast, hence there is no significant effect
of temperature on this step-1. However, the rate of formation of dichloropropanol
from monochloropropanediol (Scheme 6.5) is very slow hence increasing temperature
leads to an increase conversion to dichloropropanol. This effect is consistent up to 160
C; the temperature above this shows decrease in selectivity due to formation of other
byproducts such as trichloropropanol.
0
10
20
30
40
50
60
70
80
90
100
0 1 2 3 4 5 6 7 8 9 10 11 12
C
o
n
c
e
n
t
r
a
t
i
o
n
(
%
)

Time (h)
130


Figure 6.5: Effect of temperature
Glycerol, 40 g (0.43 mol), HCl gas flow rate, 40 ml/min, speed of agitation, 1000
rpm, catalyst loading, 0.06 g/cc, Glycerol conc. at 140 C, 1,3-DCP conc. at
140 C, Glycerol conc. at 150C, 1,3-DCP conc. at 150 C, Glycerol
conc. at 160 C, 1,3-DCP conc. at 160 C

6.3.1.5 Concentration profile
The concentrations profile at optimized reaction conditions are shown in Figure 6.6. It
shows complete conversion of glycerol to monochloropropanediol in 6 h. Rate of
formation of dichloropropanol from monochlorpropanediol is very slow and, hence it
is the rate determining step.
0
10
20
30
40
50
60
70
80
90
100
0 1 2 3 4 5 6 7 8 9 10 11 12
C
o
n
c
e
n
t
r
a
t
i
o
n

(
%
)

Time (h)
131


Figure 6.6: Concentration profile for batch process at atmospheric pressure
Glycerol, -MCPD, -MCPD, 1,3-DCP, 1,2-DCP,
Others

6.3.2 Semi-batch process at super atmospheric pressure
As described above rate of formation of monochloropropanediol from glycerol
(Scheme 6.4) is very fast, hence there is no significant effect of any parameter.
However, the rate of formation of dichloropropanol from monochloropropanediol
(Scheme 6.5) is very slow and hence it is the rate determining step. The semi-batch
process at super atmospheric pressure was studied for establishing the kinetics of
chlorination of monochloropropanediol to dichloropropanol.

6.3.2.1 Proof of absence of external mass transfer resistance
The effect of speed of agitation was studied in the range of 8001200 rpm at a catalyst
loading 3 wt % of monochloropropanediol under otherwise similar conditions, which
corresponded to 0.039 g/cm
3
of the reaction volume. The monochloropropanediol
conversion and dichlorohydrin selectivity were practically the same at these speeds
0
10
20
30
40
50
60
70
80
90
100
0 1 2 3 4 5 6 7 8 9 10 11 12
C
o
n
c
e
n
t
r
a
t
i
o
n

(
%
)

Time (h)
132

(Figure 6.7), which suggests the elimination of mass transfer resistance in the range of
speeds employed.

Figure 6.7: Effect of speed of agitation
-MCPD, 40 g (0.36 mol), reaction pressure, 7 bar, reaction temperature, 150 C,
catalyst loading, 0.039 g/cc, 800 rpm, 1000 rpm, 1200 rpm

6.3.2.2 Effect of HCl gas pressure
Effect of anhydrous HCl gas pressure was studied in the range of 5-atm on yield of
dichloropropanols under otherwise similar conditions (Figure 6.8). As HCl gas
pressure increases, the selectivity towards dichloropropanol increases. However, this
effect was consistent up to 7 atm HCl pressure; further increase in pressure did not
affect selectivity of dichloropropanols and led to the formation of other dehydration
products. Plot of initial rate vs HCl pressure shows linear increase in initial rate of
reaction with increase in HCl pressure. (Figure 6.9) This is due to higher
concentration of HCl dissolved in the reaction medium.




0
10
20
30
40
50
60
70
80
90
0 20 40 60 80 100 120 140 160 180 200 220 240
C
o
n
v
e
r
s
i
o
n
(
%
)

Time (min)
133



Figure 6.8: Effect of HCl gas pressure
-MCPD, 40 g (0.36 mol), reaction temperature, 150 C, catalyst loading, 0.039 g/cc,
speed of agitation, 1000 rpm, 5 bar, 6 bar, 7 bar, 9 bar

Figure 6.9: Plot of initial rate vs HCl gas pressure
0
10
20
30
40
50
60
70
80
90
0 20 40 60 80 100 120 140 160 180 200 220 240
C
o
n
v
e
r
s
i
o
n

(
%
)

Time (min)
y = 1.2353x
R = 0.9912
0
1
2
3
4
5
6
7
8
9
10
0 1 2 3 4 5 6 7
-
r
A

E
-
0
6

Pressure (Bar)
134

6.3.2.3 Effect of catalyst loading
The effect of catalyst loading on -monochloropropanediol conversion was studied
from 0.01 to 0.09 g/cm
3
at 160 C (Figure 6.10). In the absence of external mass
transfer resistance, the rate of reaction is directly proportional to the catalyst loading
based on the entire liquid-phase volume. It was observed that as the concentration of
catalyst was increased up to 0.039g/cm
3
, the conversion increased, beyond which there
is no substantial increase. This would mean that the number of sites available is more
than that required. The reaction shows 52% conversion in 4 h time at when no catalyst
was used. This would indicate that acidity provided by dissolved HCl also acted as
catalyst. Thus, the reaction consists of both homogeneous and heterogeneous centers
and this is discussed separately.

Figure 6.10: Effect of catalyst loading
-MCPD; 40 g (0.36 mol), reaction pressure 7 bar, reaction temperature, 150 C,
speed of agitation, 1000 rpm, 0 g/cc, 0.013 g/cc, 0.026 g/cc, 0.039 g/cc,
0.09 g/cc.

0
10
20
30
40
50
60
70
80
90
100
0 20 40 60 80 100 120 140 160 180 200 220 240
C
o
n
v
e
r
s
i
o
n

(
%
)

Time (min)
135

6.3.2.4 Effect of temperature
Effect of temperature on -monochloropropanediol conversion was carried over
MUICaT-5, with 0.039 g/cm
3
catalyst concentration at 7 atm HCl pressure, over a
range of 140-160 C, under otherwise similar conditions (Figure 6.11). This effect is
consistent up to 150 C; the temperature above this shows decrease in selectivity due
to formation of other byproducts.

Figure 6.11: Effect of temperature on monochloropropanediol conversion
-MCPD, 40 g (0.36 moles), reaction pressure, 7 bar, speed of agitation, 1000 rpm,
catalyst loading, 0.039 g/cc, 130 C, 140 C, 150 C, 160 C

6.3.3 Reaction mechanism and concentration profile
Figure 6.12 shows complete concentration profile of all reaction components at super
atmospheric pressure. It shows complete conversion of glycerol to
monochloropropanediol in 3 h. Rate of formation of dichloropropanol from
monochlorpropanediol is very slow and is the, rate determining step.
0
10
20
30
40
50
60
70
80
90
100
0 20 40 60 80 100 120 140 160 180 200 220 240
C
o
n
v
e
r
s
i
o
n

(
%
)

Time(min)
136


Figure 6.12: Concentration profile for batch process at super atmospheric pressure
Glycerol, -MCPD, -MCPD, 1,3-DCP, 1,2-DCP

6.3.4 Catalyst reusability
The catalyst was filtered from the reaction mass and refluxed with methanol in order
to remove any adsorbed material. It was dried at 120 C before every use. There was
loss of catalyst due to attribution during filtration. The loss in the catalyst quantity in
every cycle, was made up by adding required amount of fresh catalyst. Catalyst
activity remains constant after 3
rd
reuse.

6.3.5 Reaction mechanism and kinetics
Chlorination of glycerol to dichloropropanol consists of two step reactions (Scheme
6.3). According to observed concentration profile in Figure 6.12, the 1
st
step i.e.
glycerol to -MCPD is fast and 2
nd
step i.e. -MCPD to 1,3-DCH is slow. It is rate
determining step. In agreement with experimental observations in the 1
st
step there is
a marginal difference in catalyzed and un-catalyzed reaction rates, because HCl itself
acts as the catalyst to give -MCPD. Mechanism of un-catalyzed reaction of glycerol
0
10
20
30
40
50
60
70
80
90
100
0 50 100 150 200 250
C
o
n
v
e
r
s
i
o
n

(
%
)

Time (min)
137

and hydrochloric acid to give -MCPD and -MCPD can be demonstrated in (Scheme
6.4 and 6.5). Nucleophillic substitution occurs at less substituted carbon atom of
oxonium intermediate to give chloride ion substitution at position to give -MCPD
predominantly. Although - substitution can occur it is less favorable. According to
previous report, the observed concentrations are very low (Tesser et al., 2007). Whil e
2
nd
step is also un-catalyzed reaction, there is a significant difference in catalyzed and
un-catalyzed reaction rates as per experimental observation shown in (Figure 6.10).
Since the 2
nd
step is rate determining step the mechanism of catalyzed and
uncatalyzed reaction can be explained as follows.

Uncatalyzed reaction mechanism can be same as that of 1
st
step reaction mechanism
of uncatalyzed reaction of glycerol to give -MCPD (Scheme 6.5). In the mechanism
for catalyzed reaction, -MCPD and HCl are adsorbed on adjacent catalyst sites.
Proton on catalyst site attacks terminal OH group followed by dehydration and
formation of oxonium intermediate. Oxonium intermediate is then attacked by Cl
-
ion
on terminal carbon atom (Scheme 6.6).

In agreement with experimental observation, the amount of -MCPD formed in 1
st

step is very low and remains constant throughout the reaction, which shows that -
MCPD does not react further to give 1,2-DCH. This can be explained on the basis of
proposed reaction mechanism of 2
nd
step. According to it, the formation of oxonium
ion is not possible in the absence of vicinal hydroxyl group since the -position is
already substituted by Cl
-
ion in - MCPD. This explanation is valid for catalyzed
reaction mechanism in agreement with the observed experimental data.
138


Scheme 6.3: Reaction mechanism chlorination of glycerol to dichloropropanol



Scheme 6.4: Mechanism of un-catalyzed reaction of glycerol and hydrochloric acid to
give -MCPD and -MCPD

139


Scheme 6.5: Mechanism of un-catalyzed reaction of -MCPD and hydrochloric acid
to give 1,3-DCP and 1,2-DCP



Scheme 6.6: Mechanism of catalyzed reaction of -MCPD and hydrochloric acid to
give 1,3-DCP and 1,2-DCP

140

6.3.5.1 Development of kinetic model
Glycerol (A) and hydrochloric acid (B) undergo uncatalyzed reaction to give -
MCPD(C) and -MCPD (D). -MCPD (C), -MCPD(D) and HCl (B) adsorb on the
adjacent catalyst sites(S) and only -MCPD (C) undergo catalyzed reaction to form
1,3-DCP(E) and 1,2-DCP(F), which then desorb from catalyst sites. The reaction
mechanism is given by Scheme 6.6.

The studies of the effects of the foregoing parameters on conversion and rates of
reactions suggested that the reaction was free from internal diffusion and external mass
transfer resistances and hence intrinsic kinetic equations could be written as follows.
1
K
A B C D + +
1
K
1
C D
1
K
1
D C D
1
K
1
C
1
(1)
.
Kc
C S C S +
Kc
C S
Kc
C S .
Kc
(2)
.
B
K
B S BS +
B
K
B
BS
B
K
B
BS .
B
K
B
(3)
2
. . .
K
C S BS E S S + +
2
K
2
E S
2
K
2
E S E S E S E S . E S
2
K
2
E S
2
. (4)
3
. . .
K
C S BS F S S + +
3
K
3
F S
3
K
3
F S F S F S F S . F S
3
K
3
F S
3
. (5)
1/
.
E
K
E S E S +
1/
E
K
E
E S
1/
E
K
E
S E S
1/
E
K
E
(6)
1/
.
F
K
F S F S +
1/
F
K
F
F S
1/
F
K
F
F S F S
1/
F
K
F
(7)
Step 4 is rate determining steps
1 C D A B
C C K C C = (8)
.
2 2 2 '
.
C S C S C
C C S C S
C K C C C K C C = = (9)
.
2 2 2 '
.
B S B
B B S B S B S
C K C C C K C C = = (10)
. 3 . . F S S C S B S
C C K C C = (11)
. E S E E S
C K C C = (12)
. F S F F S
C K C C = (13)
The overall rate of chlorination of -MCPD (C) is
'
2 . . 2 . c C S B S E S S
r k C C k C C = (14)
The first term on RHS represents rate of adsorption of -MCPD and HCl and second
term on RHS represents rate of formation of 1,3-DCP. Eq. 14 can be simplified as.
' ' 2 ' 2
2 2
C S S
c B C B E E
r k K K C C C k K C C =
2 ' ' '
2 2
C
c S B C B E E
r C k K K C C k K C ( =

(15)
141

The site balance can be represented as,
. . . . . t C S D S B S E S F S S
C C C C C C C = + + + + + (16)
' ' '
C D
t C S D S B B S E E S F F S S
C K C C K C C K C C K C C K C C C = + + + + +
' ' '
1
C D
t S C D B B E E F F
C C K C K C K C K C K C ( = + + + + +

(17)
From eq. 17 rate of reaction can be given as,
' ' ' 2
2 2
' ' '
1
C
C D
B C B E E t
c
C D B B E E F F
k K K C C k K C C
r
K C K C K C K C K C
(

=
( + + + + +

(18)
For initial rate of reaction
( )
' ' 2
2
' ' '
1
C
initial
C
B C B t
c
C D D B B
k K K C C C
r
K C K C K C
=
+ + +
(19)
As adsorption constants are very low,
' ' 2
2
initial C
c B t C B
r k K K C C C = (20)
0 0
B C
C C >>
' ' 2
2
initial C
c B t C
r k K K C C = (21)
As - MCPD is produced in the reactor by a uncatalyzed reaction, the overall reaction
will have a rate constant which is summation of uncatalyzed reaction rate constant
and catalyzed reaction rate constant, hence the equation will be.
initial
c P C
r k C = (22)
Where
P
k (Pseudo reaction rate constant)
. . P uncat cat
k k k W = +
W : catalyst loading (g/cc)
k
uncat
: Un-catalyzed reaction rate constant
k
cat
: Catalyzed reaction rate constant
C
P C
initial
dC
k C
dt
| |
=
|
\ .
(23)
(1 )
c
P C
dX
k X
dt
= (24)
| | ln 1
C P
X k t = (25)
Thus plots were made in consonance with the equation (25) at different temperature
considering only the initial rate of reaction. The plot shows the straight lines passing
142

through origin (Figure 6.13). The slope of these lines gives the value of k
P
. A plot of
ln (k
P
) vs 1/T was made at different temperatures (Figure 6.14). The slope of the plot
gives value of activation energy to be 10.39 kcal/mol.

The plots were also made in consonance with equation (25) at different catalyst
loading considering only initial rate of reaction. The plot shows straight line passing
through origin (Figure 6.15) .The slope of these lines give the value of k
P
. The plot of
k
P
vs w (catalyst loading) was obtained to get a straight line which doesnt pass
through origin (Figure 6.16). It is so because the un-catalyzed reaction is also possible
which evident from plot of (dX/dt) vs w (Figure 6.17). The slope of the line in the plot
of k
P
vs w gives intrinsic kinetic constant for catalyzed to be 0.5 cm
3
/gcat

s

and
intercept on Y axis gives the uncatalyzed reaction rate constant to be 0.004 s
-1
.

Figure 6.13: Plot of -ln (1-X
C
) vs time at different temperature
130 C, 140 C, 150 C, 160 C



y = 0.0125x
R = 0.9948
y = 0.0175x
R = 0.995
y = 0.0246x
R = 0.9978
y = 0.0293x
R = 0.9913
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 5 10 15 20 25
-
l
n
(
1
-
X
C
)

Time (min)
143


Figure 6.14: Plot of ln (k
P
) vs 1/T


Figure 6.15: Plot of -ln (1-X
C
) vs time at different catalyst loadings
0.013 g/cc, 0.026 g/cc, 0.039 g/cc
y = -5232.4x + 8.5848
R = 0.9881
-4.5
-4.3
-4.1
-3.9
-3.7
-3.5
0.0023 0.00235 0.0024 0.00245 0.0025
l
n

k

1/T
y = 0.0111x
R = 0.9878
y = 0.0176x
R = 0.9983
y = 0.0246x
R = 0.9978
0
0.1
0.2
0.3
0.4
0.5
0 5 10 15 20
-
l
n
(
1
-
X
C
)

Time (min)
144


Figure 6.16: Plot of ln (k
P
) vs catalyst loading (w)


Figure 6.17: Plot of (dX/dt) vs catalyst loading (w)

y = 0.5x + 0.0043
R = 0.998
0
0.005
0.01
0.015
0.02
0.025
0.03
0 0.01 0.02 0.03 0.04
l
n

K
P

Catalyst Loading (w) (g/cc)
y = 90.769x + 5.58
R = 0.9847
5
6
7
8
9
10
0 0.01 0.02 0.03 0.04 0.05
I
n
i
t
i
a
l

r
a
t
e

(
-
d
X
/
d
t
)

E
-
0
5

Catalyst loading (w) (g/cc)
145

6.4 CONCLUSION
A green process has been developed for the chlorination of glycerol with hydrogen
chloride gas by using MUICaT-5 as the solid acid catalyst. The reaction parameters
such as catalyst loading, speed of agitation, HCl gas flow rate, HCl pressure and
temperature were optimized for the glycerol chlorination reaction in a semi-batch
process at atmospheric pressure with continuous flow of HCl gas and also in semi -
batch process at super-atmospheric pressure in autoclave with HCl gas pressure. A
kinetic model has been developed for semi-batch process at super-atmospheric
pressure. The reaction is intrinsic kinetically controlled and follows pseudo-first order
kinetics. The activation energy was found out to be 10.39 kcal/mol. The glycerol
conversion and 1,3-DCP selectivity were found out to be 100% and 70%, respectively
in semi-batch process at atmospheric pressure in 10 h. In semi-batch process at super-
atmospheric pressure glycerol conversion and 1,3-DCP selectivity found out to be
100% and 72%, respectively in 4 h. The catalyst is heterogeneous and can be easily
separated from reaction mixture by simple filtration and can be reused several times.
Unreacted monochloropropanediols can be recycled back for chlorination to give 1,3-
DCP.

You might also like