Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Review

pubs.acs.org/IECR

Chemical and Technical Aspects of the Synthesis of Chlorohydrins


from Glycerol
E. Santacesaria,* R. Vitiello, R. Tesser, V. Russo, R. Turco, and M. Di Serio
Dipartimento di Scienze ChimicheUniversity of Naples FEDERICO II − NICL − Naples Industrial Chemistry Laboratory,
Complesso di M.te S. Angelo, Via Cintia, 80126 Napoli, Italy

ABSTRACT: In the synthesis of biodiesel via the transesterification of vegetable oils, 10 wt % of glycerol is obtained as
byproduct. This means that, by increasing the biodiesel production, the glycerol availability also increases and its cost goes down
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

more and more. In order to consume the large amount of glycerol derived from biodiesel production in a profitable way, only two
Downloaded via UNIV OF NAPLES FEDERICO II on December 12, 2019 at 10:06:34 (UTC).

strategies are possible: (i) use glycerol as raw material to produce fuel additives and (ii) use glycerol as raw material to produce
commodities. In the present work, we have briefly considered the first aspect while focusing, in particular, on the second
opportunity by reviewing the production of chlorohydrins by glycerol hydrochlorination with HCl. Chlorohydrins are important
intermediates in the production of epichlorohydrin used to produce epoxy−resins. The advantages of producing chlorohydrins
by starting from glycerol instead of propenethat is, the classical routewill be discussed. The glycerol hydrochlorination
reaction is catalyzed by carboxylic acids, and in this work, we describe (i) the reaction conditions normally adopted; (ii) the
behavior of different catalysts proposed in the literature (concerning activity and selectivity); (iii) the reaction mechanism; (iv)
the kinetic laws, reported by different authors, along with the related parameters; and (v) the role of mass transfer. A brief
discussion on the best reactors for performing the reaction and some information about the different processes used to produce
epichlorohydrin starting from glycerol will also be reported. Some catalysts, other than carboxylic acids, have also been briefly
reviewed, although they have not been used in industrial plants until recently.

1. INTRODUCTION that could potentially be obtained as byproduct from biodiesel


Biodiesel production has strongly increased recently as one of plants, only two acceptable strategies can be followed:1 (i) the
the possible substitutes of diesel from petroleum. In perspec- production in large amount of oxygenated additives for biofuels
tive, biodiesel could become convenient if the price of petro- and (ii) the use of glycerol as raw material for obtaining com-
leum goes up, as a consequence of both an increase in the con- modities. However, any economical forecast about the con-
sumption and a decrease in the availability. This energy source, venience of using glycerol as raw material in both of the
covering a relatively small segment of the human energetic mentioned fields is a very complicated matter, because, while,
needs, will be in concurrence also with diesel that can be biodiesel production is often sustained by government subsidies
obtained with more- or less-complicated processes also from for strategic reasons, glycerol is rigorously subjected to the
coal, natural gas, or biomasses. As known, crude glycerol is an market laws. As it will be seen, glycerol can be used in some
important byproduct of biodiesel synthesis, corresponding to important processes as a raw material alternative to propene;
∼10 wt % of the produced biodiesel.1 Clearly, by increasing the therefore, the cost of propene can be considered as a reference
biodiesel production, the availability of glycerol also increases point for establishing the economic convenience of using
and, consequently, its price goes down more and more. In the glycerol as raw material for producing fuel additives or chem-
past, glycerol has found applications in several fields, such as icals. Ultimately, glycerol obtained as a byproduct of biodiesel
foods, cosmetics and personal care, pharmaceuticals and drugs, by following the conventional route, based on the use of a
polyethers/polyols, explosives, alkyd resins, triacetin, deter- homogeneous alkaline catalyst, is impure and requires an
gents, cellophane, and tobacco industries. However, the expensive purification procedure before any use as raw material.
mentioned markets are almost saturated and cannot absorb For this reason, many studies have been devoted worldwide to
the great amounts of glycerol coming from the increasing the development of new heterogeneous catalysts, promoting
production of biodiesel. Glycerol often is burned by the the transesterification reaction with the attainment of pure
biodiesel producers to obtain the energy to be used inside the glycerol.1
biodiesel production process. Therefore, in the past decade, Many different substances, synthesized from glycerol, can be
intense research activity has been developed worldwide to find used as blending components for fuels, such as ethers (glycerol
new profitable uses for glycerol, and many different proposals isobutylethers),4−6 esters (triacetin),7 acetals,8,9 and ketals.9
can be found in some reviews on the topic that recently
appeared in the literature.2,3 Clearly, finding new profitable uses Special Issue: Massimo Morbidelli Festschrift
for glycerol is also useful for decreasing the biodiesel prod-
uction costs. Therefore, several possible new uses of glycerol as Received: October 2, 2013
feedstock are briefly summarized in the next section. Revised: December 2, 2013
1.1. New Potential Routes of Glycerol Chemical Accepted: December 4, 2013
Transformation. Considering the large amounts of glycerol Published: December 4, 2013

© 2013 American Chemical Society 8939 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

In particular, etherification to obtain a mixture of di- and tri- an intermediate in epichlorohydrin synthesis. The production
isobutyl ethers (GTBE) is the most promising reaction. GTBE of epichlorohydrin occurs through the following reac-
is a good additive for diesel (both fossil and biodiesel) and tion steps:
also for gasoline as an octane booster. In diesel and biodiesel
(7.5 wt %), it will lead to a reduction in the emissions of
particulates, NOx, and unburned hydrocarbons. Moreover,
blending diesel or biodiesel with GTBE also reduces the vis-
cosity, cloud point, and pour point, but it reduces the calorific
power somewhat.
As previously mentioned, another possible use of the big
amounts of glycerol potentially available in the world is the
production of commodities. Examples of this type are (i) the
hydrochlorination of glycerol to obtain chlorohydrins,10−12
which are useful intermediates for producing epychlorodrin
to be used in the production of epoxy resins; and (ii) the
dehydration to acrolein, followed by the oxidation to acrylic
acid,13,14 via the following two-step reaction:

Instead of NaOH, Ca(OH)2 also can be used. Epichlorohy-


drin is an important raw material for the production of some
polymers such as epoxide resins, synthetic elastomers, and
sizing agents for the papermaking industry. Starting from
glycerol, the reaction proceeds through a first hydrochlorination
reaction, primarily forming 1-chloro-2,3-propanediol (1-MCH)
and water, along with small amounts of the isomer 2-chloro-
1,3-propanediol (2-MCH) (monochlorohydrins). This is
The second reaction step of reaction scheme described followed by a second hydrochlorination step from which 1,3-
by reaction 1 is a well-known technology. In fact, acrylic acid DCH is obtained as the main product, together with modest
is commonly produced by propene in two oxidation steps: amounts of 1,2-chloro-3-propanol (1,2-DCH). Another inter-
the first giving acrolein and the second one giving acrylic esting aspect is that, in some cases, crude glycerol could
acid using two different catalysts. Therefore, starting from probably be used in the reaction, with evident economic ad-
glycerol, the second step is exactly the same as that observed vantage, although this possibility has not been studied enough.
for propene.
Only one paper has been published on the subject by Kruper
In both of the described examples, glycerol substitutes for
propene as raw material and clearly the convenience of the et al.39 However, it can be foreseen that some problems will
new processes is related to the costs of, respectively, propene arise in continuous plants for the necessity of a purge during
and glycerol. Other reactions are obviously possible and it the recycle of the catalyst for eliminating the impurities
has also been reported that glycerol can be thermochemically contained in crude glycerol coming from biodiesel production
converted to propylene glycol,15−17 acetol, or a variety of other (normally potassium or sodium salts).
products. 18 An aqueous phase reforming process that Different carboxylic acids can be used as catalysts. The
transforms glycerol to hydrogen also has been developed.19 hydrochlorination of glycerol is very selective in giving 1,3-
Another possibility is the biological conversions of crude DCH. In contrast, starting from propene and following the
glycerol, because glycerol represents a good feedstock in traditional technology, a mixture of 1,3-DCH and 1,2- DCH
various fermentation processes. For example, glycerol has (30:70%) is obtained. This is an important advantage of the
been used in the fermentation of Anaerobiospirillum succinici- process via glycerol, because 1,3-DCH is much more reactive
producens for the production of succinic acid,20 or it can be than 1,2-DCH and, consequently, the plants for obtaining
converted to citric acid by using the yeast Yarrowia lipolytica. epychlorohydrin, in this case, are much smaller in size and,
In particular, it has been reported that this organism produces hence, less expensive.
the same amount of citric acid when grown on glucose or on In this review, the activity and selectivity shown by different
raw glycerol.21 The scheme described by Figure 1 shows the carboxylic acids used as catalysts in promoting the reaction and
most relevant reactions in which glycerol can be involved as a the effect of the catalyst concentration have been examined and
new building block.1
compared; then, the reaction mechanisms suggested in the
1.2. Production of Chlorohydrins from Glycerol
literature by different authors have been compared and dis-
Hydrochlorination. In this work, we focused our attention
to the glycerol hydrochlorination process for producing chloro- cussed; at last, the importance of using gaseous hydrochloric
hydrins with the aim to review the “state of the art” of this acid and the role of both the HCl pressure and the temperature
technology. The strong industrial interest for this production on the reaction rate have been examined and discussed. In
route is confirmed by the fact that recently some big companies particular, all the kinetic aspects of the reaction, in relation
have announced plans to commercialize their technology to with the suggested reaction mechanisms, will be discussed in
manufacture epichlorohydrin starting from glycerol.22 Studies detail. Finally, some suggestions will be given about the process
of glycerol hydrochlorination have focused mainly on the scaleup and the possible structure of the hydrochlorination
production of 1,3-dichloro-2-propanol (1,3-DCH), which is reactors.
8940 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Figure 1. Glycerol as a new building block; examples of the reactions involving glycerol as feedstock are shown.

2. COMPARISON OF THE NEW PROCESS VIA


GLYCEROL FOR PRODUCING EPICHLOROHYDRIN
WITH THE TRADITIONAL PROCESS VIA PROPENE
The process for the synthesis of epichlorohydrin is a rather old
process. The classical process starts from propene and, through
a first step, consisting of a high-temperature chlorination, allyl
chloride is obtained;23 subsequently, combining allyl chloride
with hypochlorous acid gives glycerol dichlorohydrins isomers,
and, finally, the reaction of the glycerol dichlorohydrins with
sodium hydroxide or calcium hydroxide leads to epichlorohy-
drin. The main reactions involved in the epichlorohydrin
synthesis23 are shown in Figure 2. As it can be seen, from the
second step of this process, a mixture of dichlohydrins is
obtainedmore precisely, the reaction product is a mixture
composed of 1,2-DCH (70%) and 1,3-DCH (30%). As already
mentioned, the reactivity of 1,2-DCH, which is the most
abundant component, is much lower and the difference in the
reactivity has already been explained, based on the reaction
mechanism.24 This gives the following disadvantages: (i) an
increase in reactor size (reactive distillation column) is nec-
essary to obtain satisfactory conversion of the less-reactive Figure 2. Reaction scheme for the traditional process via propene.
reactant and (ii) the formation of undesired byproduct occurs,
as a consequence of the long residence time.24,25 This process (i) The reaction is slow because of the abundant amount of
also has the disadvantage of using an abundant amount of water present in the system, coming from both the
chlorine. The availability of great amounts of glycerol, at low aqueous HCl and the water produced by the reaction;
cost, stimulated the development of alternative processes, based (ii) Separation of the dichlorohydrins from the reaction
on the use of this substance as raw material instead of propene.
mixture is difficult; and
It must be recognized that the process based on glycerol
hydrochlorination, to obtain chlorohydrins, has been known for (iii) Acetic acid, used as a catalyst, can be lost from the
a long time.26−33 The old process involved the reaction of reaction mixture by evaporation, since its boiling point is
glycerol with aqueous hydrochloric acid at high concentration, relatively low (117 °C).
using acetic acid as a catalyst. The reaction was performed at The disadvantages listed above discouraged the possibility of
80−100 °C. The process, which is based on the use of aqueous further studies in this direction. In contrast, new studies have
HCl, is affected by many drawbacks, such as the following: been directed toward processes using gaseous HCl instead of
8941 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Figure 3. Overall reaction scheme from glycerol to epichlorohydrin, proposed by Santacesaria et al.10−12,24,25

aqueous solutions and new catalysts (less volatile than acetic 3. CATALYSTS PROPOSED FOR PROMOTING THE
acid, such as carboxylic acids of higher molecular weight). In GLYCEROL HYDROCHLORINATION REACTION
this way, it is possible to keep the concentration of the catalyst
inside the reactor constant during the reaction.10−12 As already seen in the scheme reported in the previous section
The use of organic acids less volatile than acetic acid allows (Figure 3), the reaction between glycerol and gaseous HCl
operation at higher temperatures, thus increasing the reaction gives monochlorohydrins (mainly 1-MCH and small amounts
rate significantly. The greater cost of the less-volatile organic of 2-MCH) in a first step. Subsequently, hydrochlorination
acids, with respect to acetic acid, is probably compensated proceeds and the dichlorohydrins (mainly 1,3-DCH and small
by the minor loss of catalyst for volatilization during the amounts of 1,2-DCH) are produced. In Table 1, some physico-
process. The use of gaseous HCl also allows easier recovery of
the dichlorohydrins produced from the reaction mixture, Table 1. Properties of the Components Involved in the
because 1,3-DCH and 1,2-DCH are both more volatile than Glycerol Hydrochlorination Reaction
both the unreacted glycerol and the intermediate monochloro-
hydrins 1-MCH and 2-MCH. Gaseous HCl can be fed to the molecular solubility boiling
glycerol mass density in water point
reaction system either pure or diluted with an inert gas.10−12 derivate [g/mol] [g/cm3] refractive index at 25 °C (°C)
The process for producing dichlorohydrins, according to this glycerol 92.09 1.261 1.475 soluble 290
method, is performed starting from pure or crude glycerol, as a α-MCH 110.54 1.322 1.480 soluble 213
byproduct of biodiesel production, treated with gaseous β-MCH 110.54 1.303 1.473 high 248.5
hydrochloric acid, in the presence of a carboxylic acid. α,γ-DCH 128.99 1.364 1.483 15.6 174
The pressure of HCl has a positive effect on both the α,β-DCH 128.99 1.360 1.4835−1.4855 12.7 182
reaction rate and selectivity toward 1,3-DCH, and it is usually acetic acid 60.05 1.050 1.3716 soluble 117
maintained in the range of 1−10 bar. The complete conversion
of glycerol, however, involves a reaction time that is dependent
on the pressure of HCl, the adopted temperature, and the chemical properties of the chlorinated derivatives of glycerol
catalyst concentration. This time varies from 2 h to 24 h, de- and of glycerol itself are reported. Some properties of acetic
pending on the adopted operative conditions. However, dif- acid, which is the most commonly employed catalyst, also are
ferent patents34−41 and publications10−12,42−49 have appeared reported in the same table.
in the literature on this subject in recent years, starting from As it can be seen, acetic acid as a catalyst has a relatively low
2005. In particular, Santacesaria et al.,10−12 based on experi- boiling point (117 °C), which is near the temperature normally
mental observations, suggested the reaction scheme reported in adopted for this reaction, and a loss of catalyst by evaporation
Figure 3, in which all of the occurring successive-parallel can be predicted.
reactions are numbered. As it can be seen, glycerol reacts with The catalytic performances of acetic acid sometimes are
HCl in two steps, giving monochlorohydrins in the first step also reported by different authors, in comparison with other
and then dichlorohydrins in the second one. It is interesting to carboxylic acids. Tesser et al.,12 for example, have used acetic
point out that very small amounts of 2-DCH and 1,2-DCH are acid as a catalyst with a concentration of 8 mol %, by operating
obtained and, if the reaction time is long enough, 1,3-DCH is in a hastelloy semibatch stirred reactor at 5.5 bar and 100 °C.
obtained with very high yields (85%−95%). Ultimately, both Under these conditions, acetic acid gives place to a good
1,2- and 1,3-dichlorohydrins gives epichlorohydrins. This last performance, more precisely, after 4 h of reaction, glycerol was
reaction is promoted by a basic environment, that is, by completely converted, mainly to 1,3-DCH (89.37 mol %). Bell
contacting dichlorohydrins with an aqueous solution of et al.42 also used acetic acid as a catalyst. They have studied the
Ca(OH)2 or NaOH. reaction at a pressure of 7.6 bar and temperature of 110 °C with
Some patents have also been published in the literature 5 mol % acetic acid, obtaining a total conversion of glycerol
in which an inert organic solvent is used; this solvent is after less than 4 h of reaction. The mixture of the reaction
nonmiscible with water and is souble in dichlorohydrins. In this products contained ∼93 mol % dichlorohydrins and 6 mol %
type of process, the reaction is performed at temperatures monochlorohydrins. In this case, at the end of the reaction, the
below the boiling point of the mixture (usually less than concentration of 2-MCH was higher than that of 1-MCH,
110 °C).28 because 2-MCH is much less reactive to further chlorination
8942 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Table 2. Summary Table Reporting the Performances of Some Catalysts Used in the Literaturea
catalyst reaction 1-MCH 2-MCH 1,3-DCH 1,2-DCH glycerol
[organic acids] ref pKa reaction condition time [h] [mol %] [mol %] [mol %] [mol %] conversion [%]
acetic acid 10 4.79 reflux 3 63.01 5.98 29.82 0.33 99.1
malonic acid 10 2.92 reflux 3 55.85 7.24 35.87 0.46 99.4
levulinic acid 36 4.78 reflux 3 60.04 6.98 32.20 0.37 99.6
citric acid 36 3.15, 4.77 reflux 3 70.45 6.49 17.61 0.26 94.8
succinic acid 36 4.24, 5.64 reflux 2.5 60.88 6.84 30.05 0.37 98.1
propionic acid 36 4.79 reflux 3 49.88 8.77 41.00 0.35 100
caprilic acid 34 4.89 reflux 3 57.2 39.7 96.9
adipic acid 34 4.43, 5.41 reflux 3 7.4 82.3 89.7
pivalic acid 10 4.94 reflux 3 9.19 1.65 1.11 0.00 11.9
benzoic acid 10 4.2 reflux 3 12.38 1.27 0.00 0.00 13.7
oxalic acid 41 1.38 reflux 3 69.95 0.01 70.0
fumaric acid 41 3.03, 4.44 reflux 3 93.07 1.02 94.9
tartaric acid 41 3.07 reflux 3 85.86 0.00 85.9
maleic acid 41 3.15, 5.13 reflux 3 92.06 1.74 93.8
acetic acid 12 4.79 semibatch 4.5 bar 4 0.70 7.66 89.37 2.85 100
monochloroacetic acid 12 2.85 semibatch 4.5 bar 4 69.98 6.37 10.09 0.33 86.5
dichloroacetic acid 12 1.48 semibatch 4.5 bar 4 41.94 4.71 1.99 0.00 48.0
trichloroacetic acid 10 0.7 semibatch 4.5 bar 4 26.64 3.08 1.47 0.00 30.6
trichloroacetic acid 12 0.7 reflux 3 15.08 2.01 0.00 0.00 17.1
a
Some runs have been made under a flowing stream of HCl (24 g min−1) by refluxing the other components, some others in semibatch conditions
feeding gaseous HCl at constant pressure. The reaction temperature was always 100 °C.

than 1-MCH; therefore, the amount of 2-MCH formed in the A recent patent37 proposed hydrochlorination via the use of a
initial phase of the reaction (5%−6%) does not change any reactor column consisting of a vertical cylinder with an external
further with time. Kruper et al.39 also studied the hydro- recirculation of the liquid reaction mixture. The column was
chlorination reaction of glycerol with gaseous hydrochloric acid filled with glycerol (97.5%), water (0.5%), and acetic acid as a
in the presence of acetic acid as a catalyst. The test was catalyst (2%); such a mixture was recirculated with a flow rate
conducted at a pressure of 5.6 bar and temperature of 93 °C for of 5.0 kg/h. Gaseous HCl was fed from the bottom of the
90 min. After this time, the reaction products obtained were reactor column with a flow rate of 4.6 kg/h. In the external
1,3-DCH (92.6 mol %) and 1,2-DCH (1.7 mol %); both mono- recirculation line, a vacuum rectification column had been
chlorohydrins (4.4 mol %) and unreacted glycerol (1.0 mol %) inserted, downstream from the reactor, from which a stream
also were observed. These results were obtained by using a composed of a mixture of dichlorohydrins, water formed during
purified commercial glycerol. Kruper et al.39 also studied the the reaction, and unreacted HCl were separated, using a flow
possibility to conduct the chlorination reaction with gaseous rate of 9.3 kg/h. The distillation residue is then recycled to
HCl, in the presence of acetic acid as a catalyst, but using crude the reactor. Finally, a purge of the residue of the distillation
glycerol coming directly from biodiesel plants. The test was products, containing unwanted byproducts, was collected in a
conducted at a pressure of 8.3 bar and temperature of 120 °C tank for waste disposal. Krafft et al. reported, in another recent
for 90 min. The reaction products, discharged from the reactor, patent,34 the use of glacial acetic acid as a catalyst. The authors
was a liquid with a suspension of a white solid. After filtration, used glycerol and acetic acid (10 mol %) with gaseous hydro-
the filtrate was analyzed by gas chromatography (GC), and the chloric acid at 110 °C. They used, for 2 h, a HCl flow rate of
products were 1,3-DCH (95.3 wt %), 1,2-DCH (2.6 wt %), 5.2 mol/h, then a flow rate of 3.8 mol/h for an additional 100 min,
2-acetoxy-1,3-dichloropropane (0.7 wt %), 1-acetoxy-2,3-di- and finally a flow rate of 1.3 mol/h for 317 min. They obtained,
chloropropane (0.1 wt %), and acetoxychloropropanols in this way, a glycerol conversion of 99.1% and 1,3-DCH
(0.87 wt %). Luo et al.57 studied the formation of dichloro- resulted, also in this case, as the main product (>75 mol %).
hydrins from glycerol and HCl, in the presence of acetic acid Also, Siano et al.36 reported the use of acetic acid as a catalyst in
as a catalyst, but using an aqueous HCl solution. They their patent. The test was carried out at 100 °C by flowing, at
performed the reaction in a well-stirred batch reactor with a atmospheric pressure, gaseous HCl in the reaction environ-
capacity of 500 cm3. The authors evaluated the trend, relative ment. A second addition of catalyst during the reaction was
to the time of reagent and product concentrations, and made to compensate for the loss of catalyst caused by the
observed that the amount of glycerol decreased, reaching a stripping effect of HCl during the reaction. The duration of the
minimum after ∼100 min, while 1-MCH increased, reaching a reaction was 5 h, but already in the first hour, a high rate of 1,3-
maximum at the same time and then was reacted by forming DCH formation was observed. After 2 h of reaction, the
mainly 1,3-DCH and therefore decreased in concentration. conversion of glycerol was almost complete, but the amount of
After 100 min, the concentration of glycerol remained low but 1,3-DCH formed was 20 mol %; meanwhile, after 5 h of
constant. In all the described cases, the observed behavior, reaction, the amount of 1,3-DCH formed was ∼70 mol %. The
characterized by a maximum monochlorohydrin concentration, advantage of operating in a flowing stream of gaseous HCl is
clearly means that monochlorohydrins are intermediates of the related to the high volatility of 1,3-DCH under conditions that
formation of dichlorohydrins, confirming the reaction scheme allow one to recover this product by condensation from the
presented in Figure 3. flowing stream. However, as mentioned previously, acetic acid
8943 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Table 3. Results Reported by Schreck et al.,35 Using Different Catalystsa

8944 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

Table 3. continued

8945 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

Table 3. continued

8946 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

Table 3. continued

a
The results are related to 4 h of reaction for runs performed at a temperature of 110 °C and HCl pressure of 7.48 atm, catalyst concentration =
3 mol %, with respect to loaded glycerol (330 mmol) containing 167 mmol water.

is subjected to some drawbacks, such as the catalyst loss by Schreck et al.,35 in their patent work, tested many catalysts.
evaporation during the reaction, limiting the reaction temper- Among these, the most active, showing a glycerol conversion
ature to less than 100−110 °C. In order to overcome these of >90 mol % in 4 h, were hexanoic acid, 4-trimethylammo-
problems, many researchers have investigated the possibility of niumbutyric acid, 4-dimethylbutyric acid, 4-aminobutyric acid,
finding new catalysts to substitute in place of acetic acid. Siano glycine, glycolic acid, lactic acid, 4-aminophenylacetic acid,
et al.,36 for example, studied the use of malonic acid as a 4-hydroxyphenylacetic acid, 4-methylvaleric acid, heptanoic
possible catalyst and compared its performance with that of acid, ε-caprolactone, and γ-butyrolactone. Therefore, all these
acetic acid. This catalyst was studied under different operating catalysts have activities comparable to that of acetic acid, but
conditions (flowing stream of HCl and total reflux) and at not all are as selective to 1,3-DCH. All the results obtained by
different temperatures. More precisely, the tests were con- these authors,35 with the different tested catalysts, are reported
ducted at temperatures of 80−110 °C, and from the results, it in Table 3.
was observed that the yield in 1,3-DCH, after 3 h of reaction, It is interesting to observe from the data reported in Table 3
increased by increasing the temperature, reaching a maximum that some catalysts are not carboxylic acids and some others are
of ∼57 wt % at 110 °C. The same authors36 also tested levulinic carboxylic acids but also contain other more- or less-vicinal
acid, operating at 100 °C under total reflux, with 8 mol % functional groups. Those functional groups sometime have a
catalyst and a flowing stream of HCl (50 NL/h). Under these dramatic effect on the activity and selectivity. For example,
conditions, after 3 h of reaction, the conversion of glycerol was benzoic acid (see Table 2) is not a good catalyst, despite a pKa
almost complete, but the yield in 1,3-DCH was slightly above value very near to that of acetic acid; a low activity is also
30 mol %, with 1-MCH being the main product (60 mol %). shown by 2-aminobenzoic acid (see Table 3) and no activity is
Citric acid was also found to be a good catalyst of glycerol shown by 2-methylaminobenzoic acid. In contrast, phenylacetic
chlorination.36 It was used again at a concentration of 8 mol %, acid has shown a high activity. Another interesting observation
temperature of 100 °C, and a flowing stream of HCl of 50 NL/h, is that some catalysts that exhibit good activity could be derived
for 3 h. After 3 h of reaction, the residual glycerol was slightly from renewable sources, such as levulinic acid, for example.
above 5 mol %, and the yield in 1-MCH was >70 mol %. This Bell et al.42 have also studied the performances of less-
shows that this catalyst is more selective toward 1-MCH, with the common carboxylic acids, such as 3-methylvaleric and 3,3-
second step of chlorination being very slow. The same authors also dimethylbutanoic acid, giving a high HCl consumption rate,
used succinic acid and propionic acid as possible catalysts. The while 2-trimethylammoniumacetic acid chloride has showed a
operative conditions were the same as those of the previous tests. reaction rate slightly above that of blank run (without catalyst).
Using succinic acid, it has been observed almost complete con- Bell et al.42 compared these carboxylic acids with acetic acid,
version of glycerol in 150 min of reaction and the collected and a reaction rate higher than that of acetic acid was never
1,3-DCH was more than 30 mol %. Propionic acid has shown an observed.
activity comparable to that of acetic acid and after 3 h of reaction, a Some carboxylic acids that can be characterized by a high
complete conversion of glycerol and a 1,3-DCH yield of 41 mol % boiling point, such as adipic acid and caprylic acid, have been
is observed. The same results have been reported in more detail by proposed as catalysts by Krafft et al.34 In this case, working at
Tesser et al.10−12 All of the aforementioned results are summarized 120 °C, with HCl in solution, a conversion of glycerol above
in Table 2, together with the performances reported in the litera- 95% was observed, using caprylic acid as a catalyst with a
ture, also using other catalysts for a useful comparison. 1-MCH selectivity of 57.2% and a 1,3-DCH selectivity of 39.7%.
8947 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Working at 130 °C, with azeotropic HCl in solution, a glycerol (see Table 3). However, the compounds used are probably
conversion of >95% was obtained, using caprylic acid (96.7%) reactive substances that give the true catalyst in situ and
and adipic acid (99.4%), but a very high selectivity to 1,3-DCH normally are all organic substances. In contrast, Lee et al.44
was obtained in the presence of adipic acid (82.3%), greater investigated, for the same reaction, some inorganic compounds
than that in the presence of caprylic acid (60.3%). as catalysts. They studied, first of all, the behavior of a com-
Tesser et al.12 also tested different carboxylic acids. All cat- mercial polyoxometallate catalyst H3PW12O40. This catalyst was
alysts were tested at 100 °C, under conditions of total reflux used after a thermal pretreatment at 300 °C for 2 h. The
except for HCl, with a HCl stream of 24 g/min and with a reaction between glycerol and HCl solution was carried out, at
catalyst concentration of 8 mol %, as reported in Table 2. 100 °C, for times between 5 h and 30 h. H3PW12O40 showed a
Among the catalysts tested, the most active were acetic acid, total conversion of glycerol in all the experimental runs, and the
malonic acid, propionic acid, succinic acid, citric acid, and main products formed were dichlorohydrins. However, small
levulinic acid. In conclusion, it has been shown by several amounts of acrolein, dichloropropane, propanediol, and
authors that there are many different catalysts that have high dichloroethane were obtained as byproducts. The authors44
boiling points that can advantageously replace acetic acid. It is attributed the total conversion of glycerol and the high
interesting to point out that some carboxylic acids are very selectivity to DCH to the fact that the Brönsted acid sites of
selective toward monochlorohydrins, in particular, 1-MCH, as the catalyst are particularly active in promoting the glycerol
explained in a recent patent by Di Serio et al.41 These authors chlorination reaction. The same authors45 have studied the
reported the use of dicarboxylic or hydroxycarboxylic acids as reaction using many other commercial heteropolyacids cata-
catalysts. All tests were carried out also in this case under reflux lysts, such as H3PMo12−XWXO40 (X = 0−12), H4SiMo12−XWXO40
using a vertical condenser. The reaction was carried out for 3 h, (X = 0−12), H3+XPW12−XVXO40 (X = 0−3), and H3+XPMo12−XVXO40
at 100 °C, using a catalyst concentration of 8 mol %. From the (X = 0−3). All catalysts were tested after a thermal treat-
results obtained, it can be seen that maleic and fumaric acids ment at 300 °C for 2 h. H3PMo12−XWXO40 (X = 0−12) and
show similar catalytic behavior, with a conversion of glycerol H4SiMo12−XWXO40 (X = 0−12) showed 100% glycerol conver-
slightly above 90 mol %, with an almost total selectivity to 1-MCH. sion. Even in these cases, monochlorohydrins are intermediate
Also tartaric and oxalic acids, after 3 h at 100 °C, using a products formed in great amounts but, again, small amounts of
catalyst concentration of 8 mol %, have shown a selectivity near acrolein, dichloropropane, propanediol, and dichloroethane
to 100% toward 1-MCH, although the conversion of glycerol were obtained as byproducts.
was not total (∼70 mol %). Considering that the tungsten-containing HPA are more
As previously mentioned, some catalysts selectively promote acidic than the molybdenum-containing HPA catalysts,46−50 it
the first hydrochlorination step, mainly toward the formation of can be inferred that the acid property of HPA catalysts plays a
1-MCH, while, dichlorohydrins start to form only after the very important role in determining the selectivity to DCH. For
complete conversion of glycerol. Clearly, the hydrochlorination example, H3PMo12−XWXO40 (X = 0−12) showed a higher
mechanism for these type of catalysts would be different, as will selectivity to DCH than H4SiMo12−XWXO40 (X = 0−12); this
be discussed later. In conclusion, an attempt could be made to can be explained by the fact that the catalyst containing phos-
classify the efficiency of the different catalysts tested by con- phorus as a heteroatom is even more acidic than the cor-
sidering the influence of the following parameters: responding HPA with silicon. H3+XPW12−XVXO40 (X = 0−3),
(i) the acidity strength, as in the case of the series, “acetic and H3+XPMo12−XVXO40 (X = 0−3) both showed 100%
acid, mono-chloroacetic, dichloroacetic, and trichloro glycerol conversion in the reaction. The reactions was carried
acetic acids”, tested by Tesser et al.12 by considering their out in a liquid-phase batch reactor (200 mL); 12.6 g of glycerol
pKa value; (reactant), 78.9 g of aqueous HCl solution (37 wt %, chlo-
(ii) the length and branching of the alkyl chain of the mono- rination agent), 20 g of H2O (reaction medium), and 15 g of
carboxylic acids; HPA catalyst were charged into a batch reactor. After the
(iii) the presence of different functional groups near the homogeneous solution was heated to 110 °C with vigorous
carboxylic group; and stirring (450 rpm), nitrogen was fed into the reactor to keep
(iv) the presence in the molecules of more than one the reaction pressure at 10 bar. The catalytic reaction was
carboxylic group (dicarboxylic, tricarboxylic). carried out at 110 °C for 20 h with vigorous stirring.45 In both
HPA catalysts containing vanadium, the selectivity to DCH
These aspects will be considered and discussed in more detail decreases as the amount of vanadium is increased. However,
in another section that will be devoted to the reaction mech- H3+xPW12−xVxO40 (x = 0−3) catalysts showed a selectivity for
anism. However, we can summarize here the most important DCH higher than H3+xPMo12−xVxO40 (x = 0−3) at the same
experimental observations: level of vanadium substitution. The above results can be
(i) Some catalysts are active in glycerol hydrochlorination attributed to the acidic property of HPA-substituted catalysts
but promote only the first step to monochlorohydrins, decreasing in the following order:
while others promote both successive hydrochlorination
steps to, respectively, monochlorohydrins and dichloro- tungsten‐containing HPA > molybdenum‐containing HPA
hydrins; > vanadium‐containing HPA
(ii) Some dicarboxylic acids, such as adipic acid, are more
active than monocarboxylic acids, provided that the alkyl The above results strongly support the observation that the
chain between the two carboxylic acids is not too short to acidic properties of HPA catalysts play a key role in deter-
give a steric hindrance effect or too long to decrease its mining the selectivity toward DCH. The more acidic the HPA
solubility in glycerol. catalyst, the more active the catalyst toward the formation of
Lastly, we have seen that, in some cases, catalysts different DCH from glycerol. Song et al.51 studied the effect of the
from carboxylic acids have been used with satisfactory results operating conditions on the production of DCH. They studied,
8948 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Figure 4. Glycerol hydrochlorination reaction mechanism suggested by Tesser et al.10−12

for example, the influence of the stirring rate on the formation that the selectivity to DCH was continuously increased by
of DCH in the presence of H3PW12O40 as a catalyst. The increasing the amount of catalyst. In conclusion, the strong
obtained results showed that the stirring rate is very important Brönsted acid sites of H3PW12O40 catalyst favorably contribute
for an efficient formation of DCH in this solvent-free gas to glycerol chlorination to obtain 1,3-DCH.
(hydrochloric acid)−liquid (glycerol/chlorohydrins) system.
However, the selectivity to DCH was almost constant at the 4. REACTION SCHEME, REACTION MECHANISM, AND
highest stirring rates (≥600 rpm), suggesting that mass transfer KINETICS
between gaseous HCl and glycerol could become a key factor in
4.1. Hydrochlorination Reaction Mechanism. Glycerol
this reaction system. For this reason, the authors have stirred all
hydrochlorination is a reaction that has been investigated many
of the reaction experiments at 900 rpm, in order to avoid any
times in the past century, but only recently have some different
mass-transfer limitations. Another important aspect is the
reaction mechanisms and reaction conditions been proposed.
reaction temperature. The authors51 studied the effect of the Since the reaction takes place in the presence of a carboxylic
reaction temperature on the formation of DCH from glycerol acid as a catalyst, Santacesaria et al.10−12 hypothesized, in agree-
and gaseous HCl. In all of the experimental runs, which were ment with the previous literature,53,54 that, in the presence of a
performed at different temperatures, glycerol conversion was strong acid environment, because of HCl, the reaction occurs
100% but the selectivity to DCH gradually increased by through an initial esterification. The reaction mechanism proposed
increasing the reaction temperature until reaching a constant by Tesser et al.10 is represented schematically in Figure 4.
value of ∼98% at temperatures of >150 °C, while the selectivity The first step of the scheme depicted in Figure 4 is a nucleo-
to MCH gradually decreased by increasing the reaction philic addition, in which one glycerol hydroxyl attacks the
temperature. It is interesting to note that trichlorohydrins protonated carbonyl group. The first step is then followed by
(TCHs), which are a chlorination product of DCH, was formed the formation of an oxonium group and, subsequently, by the
in a very small amount only at very high reaction temperature. addition of a chloride ion that leads to the formation of a
The dependence of product selectivity on the reaction pressure monochlorohydrin. The last step is a nucleophilic substitution
showed the same trend as those of the reaction temperature. In SN2 and occurs mainly in the α-position that is much more
other words, selectivity to DCH increased by increasing both favored, with respect to the β-position. According to the
pressure and temperature. The same authors51 also reported proposed mechanism, the product obtained in a larger amount
that, through the addition of a substance that absorbs water, is always 1-MCH. In conclusion, according to the authors, the
such as silica gel, the reaction favorably moves toward the overall reaction, in practice, is a nucleophilic substitution of a
formation of DCH.51,52 This means that water has a negative carboxylic group with a chloride ion. It is well-known that
effect on glycerol chlorination. Finally, the same authors,51, carboxylic groups are better leaving groups than the hydroxyl
using H3PW12O40 as a catalyst model, controlled the effect of groups. For this reason, carboxylic acids are good catalysts for
the amount of catalyst on the formation of DCH and noted this reaction, because, in an acidic environment, they easily
8949 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

form esters and can then be substituted by chlorine more easily


than the hydroxyl groups. But the acidity of the carboxylic acids
cannot be too strong to avoid the fact that the corresponding
esters are too stable to complete the reaction. As a matter of
fact, by comparing the results reported in the last four examples
reported at the bottom of Table 2, it is possible to observe that
acetic acid is much more active and selective than monochloro-
acetic acid and the activities and selectivities decrease more and
more for respectively dichloro and trichloro acetic acids. A
useful parameter to be considered in the mentioned cases is the
pKa. As it can be seen, pKa gradually decreases by introducing
more chlorine in the acetic acid molecule, that is, monochloro-
acetic acid, dichloroacetic acid, and trichloroacetic acid have
gradually increasing acidity, with respect to acetic acid, and this
is detrimental for the hydrochlorination reaction. This is probably
due to the increased stability of the corresponding ester. Figure 6. Evolution with time of the experimental products
The reaction mechanism described above is in agreement distribution for the run with tartaric acid at T = 100 °C. Open box
with the experimental evidence that the amount of 1-MCH that (□) corresponds to 1,3-DCH. Points are experimental data, while the
is formed is always greater than the amount of 2-MCH. More- solid lines are simulations except for 1,3-DCH.
over, 2-MCH does not react further to give 1,2-DCH. The
absence of two vicinal OH groups, in the case of 2-MCH, keq1

probably prevents the formation of the intermediate oxonium, glycerol + catalyst HoooI ester1 + water (4)
thus hindering the second chlorination reaction. In contrast, the
keq2
1-MCH can undergo a further chlorination with a mechanism 1‐MCH + catalyst HoooI ester2 + water (5)
similar to that shown above, forming mainly 1,3-DCH, accompanied
by small quantities of 1,2-DCH. The reaction scheme suggested If Keq1 ≈ Keq2, the formation of monochlorohydrins and
by Tesser et al.,10 based on the experimental observation is dichlorohydrins follows a reaction-in-series mechanism, accord-
given by Figure 5. ing to that proposed. In contrast, if Keq1 ≫ Keq2, the catalyst
gives mainly monochlorohydrin, because the catalyst gives
mainly the ester1 as an intermediate. However, also a strong
difference in the two direct kinetic constants or a significant
influence of the HCl mass-transfer limitation, in the initial
period of the reaction, can contribute to determining the selec-
tivity to MCH shown by some catalysts. However, it is then
interesting to observe that catalysts having pKa ≥4 are normally
Figure 5. Scheme of the hydrochlorination reactions (see Santacesaria selective to dichlorohydrin, while the catalysts with pKa values
et al.10−12). in the range of 1.2−3 are more selective to monochlorohydrins.
Finally, the selectivity to monochlorohydrins shown by
different catalysts41 opens a perspective to the industrial prod-
From the data reported in the previous section, it is possible uction of 1-MCH and, consequently, also to the production of
to observe that some catalysts are highly selective in promoting glycidol that both could become building blocks for other
the formation of monochlorohydrins (in particular, 1-MCH). interesting syntheses.
It seems that, in those cases, the second hydrochlorination step As seen previously, the more-acidic carboxylic acids, such as
is prevented until almost all of the glycerol is converted to trichloroacetic acid, are not active in the reaction. Clearly, to
monochlorohydrins, as can be seen in Figure 6, relative to the promote the reaction, it is necessary to form esters that are not
behavior of tartaric acid used as a catalyst.11 too stable to allow the subsequent reaction steps. However, as
This particular behavior requires an explanation based on recently shown by Tesser et al.,12 the pKa value is not the only
the reaction mechanism. The first step of the reaction, that is, factor influencing the selectivity, because the presence of more
the formation of an ester between glycerol and the carboxylic than one carboxylic group in the same molecule and/or the
acid used as catalyst is surely an equilibrium reaction more or presence of other functional groups could be important. More
less shifted to the right. The successive hydrochlorination deepened mechanistic studies are clearly necessary for a fine-
again requires the formation of an ester between a mono- tuning of the interpretation of the selectivity−structure
chlorohydrin and the carboxylic acid, and obviously this reac- relationship. An improvement is represented by the application
tion also is an equilibrium reaction. Tesser et al.10−12 suggested of the Taft equation.12 It has been shown, by applying the Taft
that when the equilibrium constant of the first equilibrium equation to the homologous seriesacetic acid, monochloro-
step of esterification is much greater than the second one, acetic acid, dichloroacetic acid, and trichloroacetic acidthat
only the first reaction can occur, while the second one is both the polar and steric effect are important in determining
prevented, because all of the loaded catalyst is involved in the catalyst behavior, in terms of activity and selectivity.
the most favorable equilibrium until the concentration of Other two different alternative mechanisms have been
the main reactant (glycerol) becomes very low. In con- proposed in the literature.42,55,57−59 Bell et al.42 suggested a
clusion, the reaction mechanism remains the same, as shown mechanism, derived from an old paper,55 in which esterification
in Figure 4 or 5 but the following esterification reactions can is again the first step but, after the ester formation, they suggest
have different equilibrium constants: the formation of a tautomeric cyclic molecule that, in the acidic
8950 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Figure 7. Mechanism of glycerol hydrochlorination catalyzed by carboxylic acids, according to Bell et al.42,55

Figure 8. A bimolecular SN2 mechanism.

environment, is converted to an acetoxonium cationic ring, Moreover, the chlorination reactions of Figure 5 (reactions 1
which is destabilized by a chloride ion and gives monochlor- and 3) were considered irreversible. From their kinetic runs, it
ohydrins after the hydrolysis of the corresponding ester, as it can be observed that the presence of water has a detrimental
can be seen in the scheme depicted in Figure 7. effect on the reaction rate as recently confirmed by Dmitriev
The further chlorination step would occur with the same et al.56
mechanism and mainly gives 1,3-DCH. This mechanism is Ling et al.60 proposed a reaction scheme slightly different
intriguing but, as it has been seen, requires a final hydrolysis to from that proposed by Tesser et al.,10 reported in Figure 5,
close the catalytic cycle; that is, according to this mechanism, assuming that the reaction is reversible only as reaction 3. They
the presence of water would be important for favoring the last studied the kinetics of glycerol chlorination in the presence of
indispensable reaction step. In contrast, it has been demon- different catalysts, such as acetic acid, propionic acid, malonic
strated that aqueous HCl is less active than gaseous anhydrous acid, succinic acid, and adipic acid, and they studied the
HCl and that the presence of water negatively affects both the behavior of adipic acid in particular. The authors collected the
reaction rate and the yields.56 experimental data using the gas chromatography−mass spec-
Luo et al.57,58 and Lim et al.59 both studied the kinetics of troscopy (GC-MS) analysis method. Although these authors
glycerol hydrochlorination using aqueous HCl (37 wt %) and announced the proposal of a new mechanism in their work,
assumed a mechanism similar to that proposed by Tesser they interpreted all their experimental data with pseudo-first-
et al.10 but not considering any possible intermediate species, order kinetics laws and the reaction scheme has been presented
with the exclusion of the esters; that is, they considered only as a reaction mechanism.
the double-sequence esterification−chlorination, applying More recently, Salmi and co-workers61 reinterpreted the data
a second-order kinetics law to all of the reactions that are from Tesser et al.10 and their own experimental results with a
occurring and a pseudo-steady-state condition to the ester more general and detailed mechanism in which all the possible
compounds. intermediate species are considered. Intermediates species con-
This mechanism can be considered a direct bimolecular sidered are as follows: E1 (the first ester between glycerol and
nucleophilic substitution SN2, in which Cl− anion substitute in the catalyst), E2 (the second ester between MCH and the
one step RCOO− in the ester molecule (Figure 8): catalyst), I1+ (the first ionic intermediate (protonated E1)), and I2+
8951 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

(the second ionic intermediate (protonated E2)). The proposed where catalyst+ is the protonated catalyst, I1+ the first reaction
mechanism is the same as that depicted in Figure 4, but the kinetic intermediate, ester+ the ester in protonated form, and I2+ the
approach is based on the quasi-steady-bstate hypothesis applied to second reaction intermediate in protonated form. At this point,
the four mentioned intermediate species. The authors applied their it is easy to write the kinetic laws for each elementary step
model to the runs performed by Tesser et al.10 in the presence of reaction as
malonic acid and to their run made in the presence of acetic acid.
4.2. Reaction Kinetics. 4.2.1. Kinetics of Glycerol r1* = k1[Cat][H+] r4* = k4[ester][H+]
Hydrochlorination. As seen in the scheme depicted in Figure 3,
r2* = k 2[Gly][Cat+] r5* = k5[ester +]
the synthesis of epichlorohydrin from glycerol involves several
reaction steps, and for a correct plant design, all of the kinetic r3* = k 3[I1*] r6* = k6[I+][Cl−]
laws and related parameters of the eight reactions appearing in
Figure 3 must be known. For this purpose, a detailed kinetic Then, following the rate-determining step (RDS) approach,
investigation performed in the kinetic regime is required. Until two different kinetic models were applied. In the first model
now, different papers have been published regarding the (Model A), the oxonium ion formation (a three-membered
kinetics of (i) the formation of chlorohydrins, starting from ring, see the mechanism described in Figure 4) was considered
glycerol and gaseous HCl, in the presence of a carboxylic acid as the RDS; in the second one (Model B), the ester formation
as catalysts10−12,60,61 or using a concentrated aqueous HCl was assumed to be the RDS. By mathematical manipulation, it
solution;57−59 (ii) the formation of epichlorohydrin from is possible to define, for each hydrochlorination reaction, in
dichlorohydrins and degradation of this product, as a reactions (6−9), a kinetic law expression for Model A was as
consequence of the ring-opening reaction or other undesired follows:
side reactions.24,25 However, it is opportune to focus, first of all,
only on the reactions involved in the hydrochlorination that, ⎡ ⎛ C HCGly ⎞ K ⎤
according to Santacesaria et al.,10−12 are those reported in the r1 = CC⎢K1⎜⎜ ⎟⎟ − 1 (C1‐MCH)⎥
⎢⎣ ⎝ C W ⎠ K E1 ⎥⎦ (16)
scheme described by Figure 5. The experimental data collected
on these reactions by different authors have been interpreted
with kinetic models that differ for the adopted reaction mecha- ⎛ C HCGly ⎞
r2 = K 2CC⎜ ⎟
nism and/or for the number of physical phases considered. ⎝ CW ⎠ (17)
Tesser et al.,10 for example, have interpreted all their experi-
mental data with a pseudohomogeneous kinetic model. The ⎡ ⎛C C ⎞ K ⎤
authors have studied in detail the kinetics of the hydro- r3 = CC⎢K3⎜⎜ H 1‐MCH ⎟⎟ − 3 (C1,3‐DCH)⎥
chlorination reaction in the presence of malonic acid as a ⎢⎣ ⎝ C W ⎠ K E3 ⎥⎦ (18)
catalyst, determining for this catalyst the best performing
kinetic laws and related parameters. Then, they applied the ⎛C C ⎞
same kinetic laws to many other catalytic systems and r4 = K4CC⎜ H 1‐MCH ⎟
determining the related kinetic parameters for a fixed temperature ⎝ CW ⎠ (19)
of 100 °C. It is possible to recognize, in the scheme described For Model B, assuming that the ester formation is the RDS, the
by Figure 5, a sequence of four reactions: following kinetic expressions resulted:
k1
glycerol + HCl HooI 1‐MCH + H 2O ⎡ K ⎤
k −1 (6) r1 = CC⎢K1C HCGly − 1 (C1‐MCHC W )⎥
⎢⎣ K E1 ⎥⎦ (20)
k2
glycerol + HCl → 2‐MCH + H 2O (7)
r2 = K 2CCC HCGly (21)
k3
1‐MCH + HCl ⇌ 1,3‐DCH + H 2O
k ‐3 (8) ⎡ K ⎤
r3 = CC⎢K3C HC1‐MCH − 3 (C1,3‐DCHC W )⎥
k4 ⎢⎣ K E3 ⎥⎦ (22)
1‐MCH + HCl → 1,2‐DCH + H 2O (9)
Each reaction reported here (reactions 6−9) corresponds to r4 = K4CCC HC1‐MCH (23)
a sequence of elementary reaction steps and we can write, in a
simplified way, for each hydrochlorination step, the following The equations representing Model A (eqs 16−19) or Model
sequence of elementary reactions: B (eqs 20−23) have been used by the authors10 to solve the
material balance equations that, for the liquid phase, in a batch
step 1) catalyst + H+ → catalyst+ (10) reactor, by assuming a chemical kinetic regime, can be written
as follows:
step 2) glycerol + catalyst+ → I1+ (11)
dnG dn1‐MCH
= (− r1 − r2)VR = (r1 − r3 − r4)VR
step 3) I1+ → ester + H 2O + H+ (12) dt dt
dn2‐MCH dn1,3‐DCH
step 4) ester + H+ → ester + (13) = (r2)VR = (r3)VR
dt dt
step 5) ester + → I 2+ + catalyst (14) dn W dn1,2‐DCH
= (r1 + r2 + r3 + r4)VR = (r4)VR
dt dt
step 6) I 2+ + Cl− → 1‐MCH (15) (24)

8952 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

Figure 9. Parity plot experimental versus calculated data for a kinetic run performed with malonic acid at T = 100 °C. Comparison between Model A
and Model B (from Tesser et al.10).

Here, ni is the moles of component i in the liquid phase and Equation 27 has been applied to the runs performed with
VR is the volume of the reacting liquid mixture. The details malonic acid as a catalyst, and a ratio of Kchlor/Kest ≫ 1 resulted.
for solving the ordinary differential equation (ODE) system This suggests that the esterification step is much slower than
described by eqs 24 are reported in ref 10. All the experimental the chlorination one, confirming the assumption that ester for-
data were submitted to mathematical regression analysis,62 and mation is the rate-determining step (RDS) and Model B would
Model B exhibited the best results, with regard to fitting the be the most reliable. In conclusion, the esterification of glycerol
available data, as shown in Figure 9. with a carboxylic acid used as catalyst seems to be the RDS of
The two models have also been submitted to a different approach the overall hydrochlorination reaction.
based on the “steady-state approximation”. At this purpose, the Tesser et al.12 very recently have implemented their Model B
catalytic cycle for the chlorination of glycerol can be expressed also considering the role of the gas−liquid mass transfer of
according to the following simplified scheme (Figure 10): HCl. At this purpose, they adopted the Whitman double-film
theory assuming, as a first approximation, that no resistance to
mass transfer is given by the gas-side film. In this way, it is
possible to write a mass-transfer rate expression for HCl in the
liquid film as follows:
JHCl (mol/cm 3 min) = k1a([HCl]* − [HCl])

= β([HCl]* − [HCl]) (28)


where [HCl]* is the HCl equilibrium concentration at the gas/
liquid interphase, evaluated as
Figure 10. Catalytic cycle according to Tesser et al.10
[HCl]* = HHClP (29)
The two following reaction rate expressions can be written as HHCl is the Henry constant of the HCl solubility, correspond-
overall esterification reaction: ing to

rest = KestCC*CGly solubility


(25) HHCl =
P (30)
overall chlorination reaction:
with P being the pressure. The solubility in a multicomponent
rchlor = KchlorC E*C H (26) reaction mixture such as that considered here can be calcu-
lated as
By applying the steady-state approximation to the ester
i
concentration, and taking into account for the catalyst the solubility = ∑ xiKHCl (31)
material balance CC = CC* + CE* considering that the catalyst is
partitioned between the original form of carboxylic acid and the A complete mass balance can be written as follows.
ester. An overall reaction rate equation for the product forma- Liquid-Phase Mass Balance.
tion can be derived:
dnHCl
⎡ ⎤ = −r1VR − FGly
dt (32)
⎢ 1 ⎥
rchlor = KchlorCCCGC H⎢ ⎥ dn1‐MCH
⎣ ( )
⎢ CG + Kchlor C H ⎥
K est ⎦ (27) dt
= ( +r1 + r3 + r4)VR − F1‐MCH
(33)

8953 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

Table 4. Kinetic Constants for Reactions 6−9, and Corresponding Arrhenius Parametersa
temp, T [°C] k1 k2 k3 k4 KE1b KE2c
80 7667 ± 940 450 ± 41 714 ± 227 8±3 3846 194
90 11704 ± 1272 764 ± 60 1109 ± 307 13 ± 5 3064 167
100 13274 ± 1692 1089 ± 87 1784 ± 407 26 ± 7 2470 146
110 19433 ± 2216 1465 ± 123 2383 ± 532 32 ± 9 2015 128
120 27411 ± 2861 2215 ± 170 2179 ± 685 31 ± 13 1660 113
reaction 1 reaction 2 reaction 3 reaction 4
Ea [kJ mol−1] 35.2 ± 0.3 44.3 ± 0.2 34.9 ± 0.8 42.1 ± 1.0
ln A 20.9 ± 9 21.3 ± 0.7 18.6 ± 2.2 16.5 ± 2.8
a
Kinetic constants are expressed in units of cm6/(mol2 min). Data have been collected by using malonic acid as catalyst. Data taken from Tesser
et al.10 bEquilibrium constants for reaction 6. cEquilibrium constants for reaction 8.

Table 5. Kinetic Constants, at T = 100 °C, for Various Catalystsa


Kinetic Constants [cm6/(mol2 min)]
catalyst k1 k2 k3 k4
acetic acid 34619 ± 4012 2342 ± 198 1576 ± 302 17 ± 4
malonic acid 13274 ± 1692 1089 ± 87 1784 ± 407 26 ± 7
citric acid 3307 ± 387 247 ± 21 269 ± 41 0.5 ± 0.2
levulinic acid 16905 ± 1801 1315 ± 119 1421 ± 85 17 ± 6
succinic acid 13549 ± 1290 1028 ± 95 1354 ± 78 14 ± 5
propionic acid 25545 ± 2871 2265 ± 201 1197 ± 91 12 ± 4
tartaric acidb 21712 ± 2561 2096 ± 201
a
Data taken from Tesser et al.10 bTartaric acid is selective to 1-MCH.

dn1,3‐DCH kinetic constants obtained for other catalysts are reported in


= +r3VR − F1,3‐DCH Table 5.
dt (34)
In the previous section, we have seen that some catalysts are
dn2‐MCH selective in producing monochlorohydrins, because the second
= +r2VR − F2‐MCH
dt (35) hydrochlorination step seems to be prevented. As mentioned
previously, Santacesaria et al.11 have suggested that this is
dn1,2‐DCH mainly due to a difference in the two equilibrium constants of
= +r4VR − F1,2‐DCH
dt (36) esterification of glycerol and monochlorohydrins (reactions 4
and 5, respectively). By introducing this change, in the pre-
dn H 2 O viously described model, it is possible on a theoretical basis to
= ( +r1 + r2 + r3 + r4)VR − FH2O
dt (37) also correctly simulate the kinetic runs performed in the
presence of catalysts that are selective in producing mono-
dnHCl chlorohydrins instead of dichlorohydrins. However, it is difficult
= ( −r1 − r2 − r3 − r4)VR + JHCl VR − FHCl
dt (38) to discriminate if this is the only reason for the observed
with selectivity or if other factors can contribute toward such as a
strong difference in the kinetic constant between first and
r1 = k1[Cat][Gly][HCl] − k −1[Cat][H 2O][1‐MCH] second hydrochlorination or a mass-transfer limitation for HCl
migration from the gas phase to the liquid phase. This difficulty
r2 = k 2[Cat][Gly][HCl]
is related to the long time necessary, in those cases, to collect
r3 = k 3[Cat][1‐MCH][HCl] − k −3[Cat][H 2O][1,3‐DCH] reliable kinetic data for the second hydrochlorination step for
the low conversion level toward 1,3-DCH. Therefore, the
r4 = k4[Cat][1‐MCH][HCl] kinetic analysis has been restricted only to the first hydro-
(39) chlorination step. Figure 6 shows an example simulation of a
Gas-Phase Mass Balance. kinetic run performed in the presence of tartaric acid, while the
corresponding kinetic parameters are reported in Table 5.
dn HClgas
= FHCl IN − JHCl VR Another kinetic approach has recently been proposed by
dt (40) Ling et al.,60 based on a reaction scheme, such as that reported
IN in Figure 5, but assuming only reaction 3 to be reversible. They
FHCl = KP(PSET − PTOT) (41) particularly studied the kinetics of glycerol chlorination in the
Clearly, the mass-transfer parameters kLaL depends on both presence of adipic acid, proposing an unreliable model in which
the device used to perform the reaction and the adopted all the reactions are considered to be pseudo-first order.
operative conditions (pressure, temperature, type of catalyst, Consequently, the obtained fittings are not satisfactory, as can
and concentration). All the kinetic parameters, collected by be seen in Figure 11. On the other hand, the work is interesting
Santacesaria et al.10−12 by applying the described Model B to for the experimental results reported on both the kinetic behavior
the catalyst malonic acid, are reported in Table 4, while the of adipic acid and the comparison of glycerol conversion in the
8954 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

presence of different catalysts, such as acetic acid, propionic acid,


malonic acid, succinic acid, and adipic acid. The kinetic parameters
found by Ling et al.,60 along with their pseudo-first-order model,
are collected in Table 6.
We attempted to simulate a run reported in the work of Ling
et al.60 Unfortunately, in that work, the initial amount of
glycerol and the amount of catalyst used (adipic acid) were
not reported. To make the simulation, we assumed 100 g of
glycerol, 5 mol % catalyst, a temperature of 110 °C, and a
pressure of 1 bar. The simulation has been performed with the
most complete kinetic model proposed by Tesser et al.,12 and
the results obtained can be appreciated in Figure 11, where the
results are compared with the results obtained with the pseudo-
first-order model suggested by Ling et al.60 The best-fitting
parameters of the Tesser et al.12 model used for the simulation
of the data reported in Figure 11 are as follows: k1 = 5.30 × 105, Figure 11. Comparison of the performance of the kinetic models
k2 = 2.62 × 104, k3 = 2.54 × 104, k4 = 32.1, k−1 = 3.03 × 10−11, respectively proposed by Tesser et al.12 and Ling et al.60 The kinetic
k−3 is negligible, and β = KLa = 0.12 (see eqs 28−41). run has been performed by Ling et al.60
Dmitriev et al.63 studied the kinetics of both esterification
and hydrochlorination of glycerol separately. Acetic acid has
been used in the esterification of glycerol as a reagent, while, as
Table 6. Pseudo-First-Order Kinetic Constants for the
always, acetic acid is used in hydrochlorination as a catalyst.
Hydrochlorination of Glycerol in the Presence of Adipic
The reactions considered for esterification are
Acida
Pseudo-First-Order Kinetic Constants [min−1]
temp, T [°C] k1 × 102 k2 × 104 k3 × 103 k4 × 105
90 1.23 3.16 2.42 3.12
100 1.35 4.59 4.18 8.70
110 2.01 6.59 5.39 11.10
120 2.56 9.07 5.03 11.37
a
Runs performed in flowing HCl at atmospheric pressure.60

Keq1 = 1.89 (47)

⎡ −(53000 ± 3000) ⎤
k 2 (L2 s2/mol2) = 1.60 × 104 exp⎢ ⎥
⎣ RT ⎦
(48)
Keq2 = 1.00 (49)
From these parameters, it is possible to observe that
reaction 42 is much faster than reaction 43. However, these
data are not reliable, because the presence of HCl as a
catalyst also induces the hydrochlorination reaction. Another
observation is related to the negative role of water in both
reactions. In the presence of water, the esterification reaction
In both cases, HCl has been introduced in a relatively small rate is lower and the concentration of esters at equilibrium
amount to act as a catalyst. Kinetic data of esterification have also is lower.
been interpreted with the kinetic laws: The same authors also studied the hydrochlorination reac-
tions, assuming a bimolecular nucleophilic substitution mech-
⎛ CGly AcetateC H2O ⎞ anism SN2 to interpret the results. They considered all of the
r1 = k1C HCl ⎜⎜CGlyCAcetic Acid − ⎟⎟ reactions of the scheme reported in Figure 5 irreversible and,
⎝ Keq1 ⎠ (44) therefore, used pseudo-second-order kinetic laws equations.
The reactions were studied in the absence and the presence
⎛ C1‐MCHC H2O ⎞ of acetic acid as a catalyst, observing that the reaction rate in
r2 = k 2C HCl ⎜⎜C1‐MCHCAcetic Acid − ⎟⎟
the absence of acetic acid is only 3% of that in the presence
⎝ Keq2 ⎠ (45) of this catalyst. By applying their bimolecular model, they
The kinetic parameters determined by the authors are found that kinetic constants were strongly affected by the
ratio CHCl/CH2O and introduced this term, in an empirical
⎡ −(58500 ± 3000) ⎤
k1 (L2 s2/mol2) = 6.29 × 105 exp⎢ ⎥ way, inside the kinetic law to interpret the kinetic data. For
⎣ RT ⎦ the formation of 1-MCH and 2-MCH, the following
(46) relations have been proposed:
8955 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Table 7. Reinterpretationa of the Kinetic Data from Tesser et al.10 with the Kinetic Model Proposed by Salmi and Co-workers61
T = 80 °C T = 90 °C T = 100 °C T = 110 °C T = 120 °C
k3′ [L mol−1min−1] 9.36 × 10−4 3.98 × 10−3 1.26 × 10−2 4.30 × 10−2 1.70 × 10−1
k4′ [L mol−1min−1] 5.45 × 10−5 2.64 × 10−4 1.01 × 10−3 3.30 × 10−3 1.54 × 10−2
k′7 [L mol−1min−1] 6.55 × 10−7 5.03 × 10−6 5.75 × 10−6 3.29 × 10−5 7.77 × 10−5
k′8 [L mol−1min−1] 1.81 × 10−4 5,48 × 10−4 1.25 × 10−3 2.23 × 10−3 2.89 × 10−3
β′ [mol L−1] 0.57 2.05 6.33 17.50 50.70
δ′ [mol L−1] 1.43 2.62 4.06 6.91 10.30
a
The best-fitting parameters are reported in the table.

⎡ −(102000 ± 3000) ⎤ ⎛ C HCl ⎞


(10.5 − 0.023T )
k 8′cCatc1‐MCHc HCl 2
r3 = 2.1 × 109 exp⎢ ⎥ × ⎜⎜ ⎟⎟ CGlyC HCl r8 =
⎣ RT ⎦ ⎝ CH O ⎠ D78 (57)
2

(50) where r3 represents the rate of 1-MCH formation, r4 the rate


of 2-MCH formation, r7 the rate of 1,2-DCH formation, and r8
⎡ − (132000 ± 4000) ⎤ ⎛ C HCl ⎞
(14.7 − 0.034T )

r4 = 2.2 × 1012 exp⎢ ⎥ × ⎜⎜ ⎟ CGlyC HCl the rate of 1,3-DCH formation. D34 = CCatCW + (α′CW + β′)
⎣ RT ⎦ ⎝ C H O ⎟⎠ CHCl and D78 = CCatCW + (γ′CW + δ′)CHCl. By assuming
2

(51) CW ≈ 0, because water is continuously removed, the four


expressions become
while, for the formation of 1,3-DCH and 1,2-DCH,
r3 = k 3*CCatCGlyC HCl r4 = k4*CCatCGlyC HCl
⎡ −(86000 ± 1000) ⎤
11
r5 = 1.24 × 10 exp⎢ ⎥
⎣ RT ⎦ r7 = k 7*CCatCMCHC HCl r8 = k 8*CCatCGlyC HCl (58)
⎛C ⎞ (25.63 − 0.056T )
where k*i corresponds to k3′/β′, k′4/β′, k′7/δ′, and k′8/δ′.
× ⎜⎜ HCl ⎟⎟ CGlyC HClCAcetic Acid They successfully applied their model to the kinetic data
⎝ C H 2O ⎠ (52) collected by Tesser et al.10 for the runs performed at different
temperatures in the presence of malonic acid, and the cor-
⎡ −(91500 ± 1000) ⎤ responding kinetic parameters are reported in Table 7. Table 8
r6 = 1.38 × 1010 exp⎢ ⎥
⎣ RT ⎦
Table 8. Activation Energies and Pre-exponential Factors for
⎛ C ⎞(13.36 − 0.026T ) the Hydrochlorination Reactions of Glycerol Catalyzed by
× ⎜⎜ HCl ⎟⎟ CGlyC HClCAcetic Acid Malonic Acida
⎝ C H 2O ⎠ (53)
r3 r4 r7 r8
In contrast with Tesser et al.10 and others, these authors Ea [kJ/mol] 147.5 159.4 132.1 80.7
concluded that the esterification reaction is fast and must be ln A 32.3 33.5 20.0 8.1
always considered at equilibrium and hydrochlorination reac- a
A reinterpretation by Salmi and co-workers61 of kinetic runs
tions are much slower. However, these results are not supported performed by Tesser et al.10
by experimental data, because the kinetic data of hydrochlorination
are not reported. Moreover, we have seen that the esterification shows the activation energies and pre-exponential factors
data are not reliable for the presence of HCl as a catalyst, which determined from the parameters reported in Table 7. The
can become a reagent of the hydrochlorination in this reaction same model has also been used by the authors to interpret
system. their own data of a run performed in the presence of acetic acid
Salmi and co-workers61 improved the kinetic approach of at 105 °C, and the corresponding parameters are reported in
Tesser et al.10 to describe the glycerol hydrochlorination reac- Table 9.
tions. They assumed a similar mechanism but applied the principle
of quasi-steady state approximation to all of the intermediate
Table 9. Kinetic Parameters Obtained by Salmi and
species appearing in the mechanism, that is, the esters E1 and E2
Co-workers61 for the Hydrochlorination Reactions of
and the charged species I1+ and I2+. After opportune mathe-
Glycerol, Catalyzed by Acetic Acid (11 mol %), for Runs
matical manipulations, the authors obtained the following set of
Performed at 105 °C
kinetic law expressions:
parameter value
k 3′cCatcGlyc HCl 2 k3′ 2.16 × 10−2 [L mol−1min−1]
r3 =
D34 (54) k4′ 1.39 × 10−3 [L mol−1min−1]
k′7 1.29 × 10−9 [L mol−1min−1]
k′8 1.01 × 10−3 [L mol−1min−1]
k4′cCatcGlyc HCl 2
r4 = β′ 17.4 mol L−1
D34 (55) δ′ 1.91 mol L−1

k ′c c c 2 The mechanism assumed by Salmi and co-workers61 is quite


r7 = 7 Cat 1‐MCH HCl similar to that proposed by Tesser et al.,10 being different in
D78 (56) only the kinetic model approach. The kinetic model of Salmi
8956 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

and co-workers61 is more complex and rigorous, but many


simplifications have been introduced by the authors probably
for limiting, at an acceptable level, the number of kinetic
parameters to be determined from the experimental data.
Therefore, to use the most rigorous model, a large number of
kinetic runs is necessary. It is not possible to make a direct
comparison between the kinetic constants reported in Table 7
and those reported in Table 4, because the units are different;
however, it is possible to evaluate a ratio between the
corresponding kinetic constants. It is interesting to note that
these ratios are quite similar. The conclusion is that the two
models are practically equivalent in simulating the experimental
data collected at different temperatures, relative to malonic acid.
The models proposed by other authors are less reliable.
Lastly, it is worth remembering that Luo et al. and Lim
et al.58−60 studied the kinetics of glycerol hydrochlorination in Figure 12. Plots of HCl absorbed by a mixture of 85 wt % glycerin,
9 wt % water, and 7 wt % acetic acid at 90 °C at different pressures
the presence of acetic acid but using concentrated aqueous
(from ref 42).
HCl. They adopted the bimolecular substitution S N 2
mechanism of Figure 8 and considered only esterification as
equilibrium reactions. On the basis of this mechanism, all the As previously mentioned, the pressure of HCl affects both
reactions are considered to be pseudo-second-order. The the hydrochlorination reaction rate and the yield toward the
kinetic parameters determined by Luo et al.59 are summarized formation of DCHs. This behavior can be simulated by using
in Table 10. Clearly, the kinetic behavior in the presence of the more complete model suggested by Tesser et al.,12 also
considering HCl solubility and mass-transfer limitations.
Table 10. Kinetic Constants Obtained by Luo et al.59 for For this purpose, Figure 13 shows a simulation of two runs
the Reactions Occurring in Glycerol Hydrochlorination performed by Bell et al., at 90 °C, in the presence of 7 wt % of
Performed with HCl in Aqueous Phase, Using Acetic Acid
as a Catalyst
kinetic constant activation energy
[mol L−1 min−1] [kJ/mol] reaction
k1 = 1542.3 exp(−5702.7/T) 47.4 glycerol consumption
k2 = 175.6 exp(−4918.8/T) 40.9 1-MCH formation
k3 = 5.8 exp(−4381.2/T) 36.4 1,3-DCH formation
k4 = 40.34 exp(−7171.32/T) 59.6 2-MCH formation
k5 = 27.83 exp(−4950.5/T) 41.2 1,2-DCH formation

aqueous HCl cannot be directly compared with that in the


presence of pure gaseous HCl; however, it is possible to
conclude qualitatively that the presence of water decreases the
hydrochlorination reaction rate.
As has been seen, in the kinetic model proposed by different
authors, glycerol and monochlorohydrin hydrochlorination
reactions are considered to be first order, with respect to Figure 13. Simulation of the runs performed by Bell et al.42 at different
HCl concentration, but using gaseous HCl; this concentration pressures. The other experimental conditions are summarized in the
depends on the solubility of HCl in the reaction environment. legend of Figure 12.
This solubility obviously changes with pressure, temperature,
and composition of the liquid mixture. Unfortunately, there is acetic acid, at different pressures of 2.76 and 5.52 bar, respe-
not a work in the literature in which this aspect has been ctively. The plot reports the evolution of the HCl consumption
examined deeply enough. The experimental observation is that as a function of time in the used semibatch reactor. The kinetic
pressure has a dramatic effect on both reaction rate and final parameters giving the best fit in these simulations are k1 =
yields of dichlorohydrins. Bell et al.42 reported the evolution 31234, k2 = 3040, k3 = 2578, k4 = 9.9 × 10−6, k−1 = 1.18 ×
with time of the number of moles of HCl adsorbed per mole of 10−12, k−3 is negligible, and β = KLa = 0.99 (see eqs 28−41).
loaded glycerol for different pressures (from 1 bar to 5.5 bar). Tesser et al.12 also made and simulated some kinetic runs at
The obtained results are reported in Figure 12. As can be seen different HCl pressures, but using monochloroacetic acid as a
by increasing the pressure, the reaction rate increases very catalyst; the results are reported in Figure 14. The trends
much but the HCl uptake also increases; this last observation observed by Bell et al.42 for the effect of pressure on activity and
means that hydrochlorination reactions are limited by the yield are also confirmed in this case.
equilibrium. As a matter of fact, by operating at 120 °C with 4.2.2. Kinetics of Epichlorohydrin Synthesis from Dichloro-
2 wt % of acetic acid, at 1.36 atm of HCl pressure, a glycerol hydrins. The kinetics of epichlorohydrin synthesis with
conversion of ∼50% is observed; at 2.04 atm, the glycerol related side reactions (reactions 5−8 of Figure 3) has
conversion is >85%; and at 3.40 atm, the glycerol is completely been studied in detail many years ago by Carrà et al.24 The
converted.42 reaction occurs by dehydrochlorination starting from both
8957 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

the kinetic data related to the synthesis of epichlorohydrin with


a pseudo-first-order kinetic law satisfactorily.
Epoxides are reactive substances and, in a basic environ-
ment, epichlorohydrin is subjected to an undesired ring-
opening reaction, decreasing the yield, and occurring with
the following mechanism, which has been suggested by
Patai:64

Figure 14. Glycerol conversion as a function of time for runs 2, 5, and


6 of Table 2 for the monochloroacetic acid catalyst.

1,3- and 1,2-MCH in an aqueous basic environment created by Therefore, the kinetic law for the ring opening (rro) will be
dissolving in water Ca(OH)2 or NaOH. The main occurring rro = k roK ii[EPOX][OH−] ≈ keff [EPOX] (65)
reactions are:
1,2‐DCH → EPY + HCl (59) Again, it is possible to assume a pseudo-first-order kinetic law.
On the basis of all the experimental data collected, Carrà et al.24
1,3‐DCH → EPY + HCl (60) isolated the following reaction scheme:
Clearly, the developed HCl neutralize the basic catalyst used.
Therefore, if NaOH is used, the pH changes rapidly and the
reaction rate slows down. By using Ca(OH)2, which is poorly
soluble in water, a buffer effect is operative, because of the
presence of both Ca(OH)2 and CaCl2. As a consequence, the
OH − concentration changes smoothly with the DCH
conversion, according to the relationship
[OH−]3 + XC0[OH−]2 − 2KS = 0 (61)
0
where X is the conversion, C the initial DCH concentration,
and KS the solubility product of Ca(OH)2 in water. Reactions
59 and 60 can be classified as ring closure reactions, and both
occur with the following mechanism (as reported in ref 24 and
references therein):
The kinetic parameters obtained for the pseudo-first-order
reactions are reported in Table 11. The same table also gives

Table 11. Kinetic Constants for Dehydrochlorination


Reactions and Epichlorohydrin Degradation in an Alkaline
Environment Created by Ca(OH)2a
reagent pre-exponential factor [s−1] activation energy [cal/mol]
This mechanism can be regarded as an internal nucleo-
1,3-DCH 1.00 × 107 11718
philic substitution (SN2), preceded by a base-catalyzed
1,2-DCH 6.40 × 108 16984
dissociation equilibrium. By assuming the ring closure
1-MCH 3.74 × 107 13200
(rrc) to be the RDS, it is possible to write the following
EPY 4.92 × 108 18852
rate law: a
− −
Data taken from Carrà et al.24
rrc = k rcK i[MCH][OH ] = k′[MCH][OH ] ≈ kapp[MCH]
(63) the kinetic parameters for glycidol formation, by assuming a
where [MCH] is the reagent concentration, Ki is the equi- direct conversion from epichlorohydrin to glycidol, given that
librium constant for the formation of the ionic intermediate ion, the dehydrochlorination of 1-MCH is very fast. It is very important
k is the true reaction rate constant, while kapp is the apparent to point out that 1,3-DCH is much more reactive than 1,2-DCH;
kinetic constant. By assuming a constant [OH−] value, cor- therefore, the residence times for these two reactants are very
responding to the Ca(OH)2 solubility, it is possible to interpret different.
8958 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research Review

Figure 15. A simplified scheme of glycerol hydrochlorination.

5. GLYCEROL HYDROCHLORINATION REACTORS The simplest and most economical solution would be a bubble
AND PROCESSES column. Clearly, the optimal choice is related to the reaction
Until now, three companies (Dow Chemicals, Solvay EPICEROL conditions, that is, temperature, pressure, and catalyst
Technology, and CONSER SpA ECH-EF = Eco Friendly)) have concentration determine the reaction rate and, therefore, the
developed their own process for producing dichlorohydrins from HCl consumption. However, the choice is complicated by the
glycerol and HCl, using different catalysts (different carboxylic corrosiveness of HCl in the presence of water that is formed
acids or mixture of them),22 and some plants have recently been as a byproduct of the reaction. It is well-known that few
constructed or are still under construction in China (Solvay, materials are compatible with HCl in the presence of moisture
DOW Chemicals, CONSER SpA), Thailand (Solvay), and (for example, glass, Hastelloy, Teflon, enameled steel);
Korea (Dow Chemicals). All of these plants need the following, therefore, some of the aforementioned possibilities could be
as unit operations: (i) a well-mixed gas−liquid reactor, in which realized with difficulty. The problem of the compatibility of
glycerol, containing the catalyst in an opportune concentration, the materials with moisturized HCl is important also for the
is contacted with gaseous HCl (under moderate pressure, 1−10 stripping unit, and very probably, a neutralization is necessary
bar) at the reaction temperature (100−120 °C); (ii) a heat after this unit, to eliminate any trace of this reagent before the
exchanger to remove the reaction heat, given that the reaction separation by distillation of the dichlorohydrins.
is moderately exothermic (the heat exchanger can also be Finally, it is important to point out that the process via
located inside the reactor); (iii) a stripping unit, to separate the glycerol has an important advantage, with respect to the process
unreacted HCl and, eventually, also the water produced as a
via propene, because the hydrochlorination reaction is highly
consequence of the occurred hydrochlorination reactions;
selective toward the formation of 1,3-DCH. To better empha-
(iv) the separation units having the scope of recovering pure
dichlorohydrins as feedstock for the successive step of size the aforementioned advantage, a simulation has been made
epichlorohydrin synthesis. Dichlorohydrins are much more by Santacesaria et al.11 for comparing the yields of the two
volatile than monochlorohydrins and glycerol and a processes. A reactive distillation column has been simulated
distillation column is probably sufficient for this separation. with a commercial process simulation package (Chemcad 5.2)
The residue of the column will contain monochlorohydrins, for two different feed streams: the first stream was a mixture
small amounts of glycerol, and the catalyst (a carboxylic acid) with the traditional composition (30% 1,3-DCH and 70% 1,2-
if its boiling point is high; otherwise, the catalyst must be DCH), coming from a via propene process, while the second
separated beforehand (for example, in the stripping unit). The one was constituted by pure 1,3-DCH. The simulation of the
residue can be recycled in the gas−liquid reactor. Some epichlorohydrin reactive column has been performed by
undesired byproducts can be formed, and a purge is probably introducing the kinetic expression and parameters taken from
necessary. Based on the list of the previously described unit Carrà et al.24 and summarized in Table 11, using the model
operations, a simplified scheme of general validity can be sketched published by Carrà et al.25 In both cases, the reactive column
(Figure 15). has been simulated by assuming the following: total pressure,
The operation units of this scheme are all conventional sys- 1 bar; reboiler heat duty, 17 000 kcal/h; 15 theoretical
tems that do not require particular mention, with the exclusion plates; liquid holdup, 0.02 L/stage. The comparison between
of the reactor unit requiring (i) a system providing a large gas/
the two different feeds to the column is reported in Figure
liquid interface area, to avoid HCl mass-transfer limitation;
16, where epichlorohydrin yields are plotted as a function of
(ii) a good heat exchange, to eliminate the reaction heat
maintaining the desired temperature; and (iii) a good mixing the reflux ratio adopted in the reactive column. The yield is
system, to warrant a good dispersion of the gas in the liquid the ratio between the distilled epichlorohydrin and the
phase and a uniform composition. There are some different dichlorohydrins used in the feed. As it can be seen, a higher
solutions that can be adopted to satisfy the aforementioned epichlorohydrin yield is always obtained at low reflux ratios
requisites. A large gas/liquid interface area can be realized by with a feed that is constituted by pure 1,3-DCH. Moreover,
using, for example, a falling film reactor with gaseous HCl in order to obtain a satisfactory yield in epichlorohydrin by
flowing in countercurrent, a spray tower loop reactor, a Venturi feeding a 30%/70% mixture, a very high reflux ratio must be
loop reactor, a bubble tray column, or a perforated plate column. used, with, obviously, a greater consumption of energy.
8959 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962
Industrial & Engineering Chemistry Research


Review

REFERENCES
(1) Santacesaria, E.; Martinez Vicente, G.; Di Serio, M.; Tesser, R.
Main technologies in biodiesel production: State of the art and future
challenges. Catal. Today 2012, 195, 2−13.
(2) Pagliaro, M.; Rossi, M. The Future of Glycerol: New Usages for a
Versatile Raw Material; RSC Green Chemistry Book Series; Royal
Society of Chemistry (RSC): Cambridge, U.K., 2008.
(3) Zhou, C. H.; Beltramini, J. N.; Fan, Y. X.; Lu, G. Q.
Chemoselective catalytic conversion of glycerol as a biorenewable
source to valuable commodity chemicals. Chem. Soc. Rev. 2008, 35,
527−549.
(4) Behr, A.; Obendorf, L. Development of a Process for the Acid-
Catalyzed Etherification of Glycerine and Isobutene Forming
Glycerine Tertiary Butyl Ethers. Eng. Life Sci. 2003, 2, 185−189.
(5) Jaecker-Voirol, A.; Durand, I.; Hillion, G.; Delfort, B.; Montagne,
Figure 16. Comparison between the epichlorohydrin yields of two
different feeds plotted as a function of the reflux ratio adopted in the X. Glycerin for New Biodiesel Formulation. Oil Gas Sci. Technol. 2008,
reactive column.25 63, 395−404.
(6) Di Serio, M.; Casale, L.; Tesser, R.; Santacesaria, E. Development
of a Process for the Acid-Catalyzed Etherification of Glycerine and
6. CONCLUSIONS Isobutene Forming Glycerine Tertiary Butyl Ethers. Energy Fuels 2010,
24, 4668−4672.
The synthesis of epichlorohydrin starting from glycerol instead of (7) Melero, J. A.; Van Grieken, R.; Morales, G.; Paniagua, M. Acidic
propene occurs through two steps: the preparation of Mesoporous Silica for the Acetylation of Glycerol: Synthesis of
dichlorohydrins by glycerol hydrochlorination and the successive Bioadditives to Petrol Fuel. Energy Fuels 2007, 21, 1782−1791.
step of dichlorohydrins dehydroclorination to epichlorohydrin. (8) Silva, P. H. R.; Gonçalves, V. L. C.; Mota, C. J. A. Glycerol acetals
The second step is well-known, because it is the same for both of as anti-freezing additives for biodiesel. Bioresour. Technol. 2010, 101,
the aforementioned processes. The only difference is that, starting 6225−6229.
from propene, a mixture of the two dichlorohydrins (1,2-DCH (9) Crotti, C.; Farnetti, E.; Guidolin, N. Alternative intermediates for
(70%) and 1,3-DCH (30%)) is obtained and used as feedstock for glycerol valorization: iridium-catalyzed formation of acetals and ketals.
the successive dehydrochlorination step, while, starting from Green Chem. 2010, 12, 2225−2231.
glycerol, 1,3-DCH is the main component obtained (more than (10) Tesser, R.; Santacesaria, E.; Di Serio, M.; Di Nuzzi, G.; Fiandra,
95%). This is an advantage of the glycerol route: the residence V. Kinetics of Glycerol Chlorination with Hydrochloric Acid: A New
Route to αγ-Dichlorohydrin. Ind. Eng. Chem. Res. 2007, 46, 6456−
time can be reduced and, hence, the reactor volume for the
6465.
dehydrochlorination step also is reduced. (11) Santacesaria, E.; Tesser, R.; Di Serio, M.; Casale, L.; Verde, D.
In this review, we have collected all data available from the New Process for Producing Epichlorohydrin via Glycerol Chlorination.
literature, for what concerns the glycerol hydrochlorination Ind. Eng. Chem. Res. 2010, 49, 964−970.
reaction, intensively studied in the last years. These data are related (12) Tesser, R.; Di Serio, M.; Vitiello, R.; Russo, V.; Ranieri, E.;
to the catalysts that can be used to promote the reaction, their Speranza, E.; Santacesaria, E. Glycerol Chlorination in Gas-Liquid
performances, the kinetics of the reactions involved in the Semibatch Reactor: An Alternative Route for Chlorohydrins
hydrochlorination with the related mechanisms, and some aspects Production. Ind. Eng. Chem. Res. 2012, 51, 8768−8776.
that are particular to this gas−liquid system. All this information is (13) Atia, H.; Armbruster, U.; Martin, A. Dehydration of glycerol in
fundamental for the choice of the most suitable reactor. The gas phase using heteropolyacid catalysts as active compounds. J. Catal.
catalysts used until now are carboxylic acids with the preference for 2008, 258, 71−82.
compounds of high boiling point, which allows one to recycle the (14) Katryniok, B.; Paul, S.; Bellière-Baca, V.; Rey, P.; Dumeignil, F.
catalyst without problems. Many kinetic data are available but are Glycerol dehydration to acrolein in the context of new uses of glycerol.
rather scattered, because they have been collected on different Green Chem. 2010, 12, 2079−2098.
(15) Dasari, M. A.; Kiatsimkul, P. P.; Sutterlin, W. R.; Suppes, G. J.
catalysts and interpreted with different kinetic models, based on
Low-pressure hydrogenolysis of glycerol to propylene glycol. Appl.
different reaction mechanisms. Further work is necessary to
Catal., A 2005, 281, 225−231.
identify the most-reliable catalytic mechanism and the best catalyst. (16) Alhanash, A.; Kozhevnikova, E. F.; Kozhevnikov, I. V.
For example, the use of polyoxometallates as a catalyst is a new Hydrogenolysis of glycerol to propanediol over Ru: Polyoxometalate
promising route that is worth of deeper examination. Lastly, more bifunctional catalyst. Catal. Lett. 2008, 12, 307−311.
investigation about the role of mass transfer of HCl and of HCl (17) Chiu, C. W.; Dasari, M. A.; Sutterlin, W. R.; Suppes, G. J.
solubility in the reaction environment would be necessary for Removal of residual catalyst from simulated biodiesel’s crude glycerol
correct reactor modeling.


for glycerol hydrogenolysis to propylene glycol. Ind. Eng. Chem. Res.
2006, 45, 791−795.
AUTHOR INFORMATION (18) Johnson, D. T.; Taconi, K. A. The glycerin glut: Options for the
Corresponding Author value-added conversion of crude glycerol resulting from biodiesel
*E-mail: elio.santacesaria@unina.it. production. Environ. Progress 2007, 26, 338−348.
(19) Cortright, R. D.; Davda, R. R.; Dumesic, J. A. Hydrogen from
Notes
catalytic reforming of biomass-derived hydrocarbons in liquid water.
The authors declare no competing financial interest.


Nature 2002, 418, 964−967.
(20) Lee, P. C.; Lee, W. G.; Lee, S. Y.; Chang, H. N. Succinic acid
ACKNOWLEDGMENTS production with reduced by-product formation in the fermentation of
Thanks are due to CONSER SpA and Eurochem Engineering anaerobiospirillum succiniciproducens using glycerol as a carbon
srl for funding the research. source. Biotechnol. Bioeng. 2001, 72, 41−48.

8960 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

(21) Papanikolaou, S.; Aggelis, G. Lipid production by Yarrowia for recovering dichlorohydrins. U.S. Patent US 2010/0137652A1,
lipolytica growing on industrial glycerol in a single-stage continuous 2010.
culture. Bioresour. Technol. 2002, 82, 43−49. (41) Di Serio, M.; Santacesaria, E.; Tesser, R. (Eurochem
(22) (a) Ullmann’s Encylopedia of Industrial Chemistry, 7th ed.; Engineering srl). Process for monochlorohydrins production from
Wiley−VCH: Weinheim, Germany, 2006. (b) Dow Chemical glycerol and hydrochloric acid. World Patent WO 2008/132770 A1.
Company. C&E News 2006 (Aug. 14), 3. (c) Solvay. EPICEROL (42) Bell, B. M.; Briggs, J. R.; Campbell, R. M.; Chambers, S. M.;
Process; http://www.solvaychemicals.com/EN/Sustainability/Issues_ Gaarenstroom, P. D.; Hippler, J. G.; Hook, B. D.; Kearns, K.; Kenney,
Challenges/EPICEROL.aspx. (d) CONSER SpA. ECH-EF Process J. M.; Kruper, W. J.; Schreck, D. J.; Theriault, C. N.; Wolfe, C. P.
(Eco Friendly); http://www.conserspa.com/frame_pr-de.htm. (e) de Glycerin as a Renewable Feedstock for Epichlorohydrin Production.
Guzman, D. Growing glycerine-to-ECH plants. ICIS Green Chemistry The GTE Process. Clean 2008, 36, 657−661.
Jan. 20, 2011; Available via the Internet at: http://www.icis.com/ (43) Lee, S. H.; Park, D. R.; Kim, H.; Lee, J.; Jung, J. C.; Woo, S. Y.;
blogs/green-chemicals/2011/01/growing-glycerine-to-ech-plant/. Song, W. S.; Kwon, M. S.; Song, I. K. Korean J. Chem. Eng. 2008, 25
(23) Sienel, G.; Rieth, R.; Rowbottom, K. T. Epoxides. In Ullmann’s (5), 1018−1021.
Encyclopedia of Industrial Chemistry; Wiley−VCH: Weinheim, (44) Lee, S. H.; Park, D. R.; Kim, H.; Lee, J.; Jung, J. C.; Woo, S. Y.;
Germany, 2005 (DOI: 10.1002/14356007.a09_531). Song, W. S.; Kwon, M. S.; Song, I. Catal. Commun. 2008, 9, 1920−
(24) Carrà, S.; Santacesaria, E.; Morbidelli, M.; Schwarz, P.; Divo, C. 1923.
Synthesis of Epichlorohydrin by Elimination of Hydrogen Chloride (45) Lee, S. H.; Park, D. R.; Kim, H.; Lee, J.; Jung, J. C.; Park, K. M.;
from Chlorohydrins. 1. Kinetic Aspects of the Process. Ind. Eng. Chem. Cho; Song, I. K. React. Kinet. Catal. Lett. 2008, 94, 71.
Process. Des. Dev. 1979, 18 (3), 424−427. (46) Misono, M. Heterogeneous Catalysis by Heteropoly Com-
(25) Carrà, S.; Santacesaria, E.; Morbidelli, M.; Schwarz, P.; Divo, C. pounds of Molybdenum and Tungsten. Catal. Rev. Sci. Eng. 1987, 29,
Synthesis of Epichlorohydrin by Elimination of Hydrogen Chloride 269.
from Chlorohydrins. 2. Simulation of the Reaction Unit. Ind. Eng. (47) Hill, C. L.; Prosser-McCartha, C. M. Homogeneous catalysis by
Chem. Process. Des. Dev. 1979, 18 (3), 428−433. transition metal oxygen anion clusters. Coord. Chem. Rev. 1995, 143,
(26) (a) Ger. Patent 197308, 1906. (b) Ger. Patent 197309, 1906. 407−455.
(c) Ger. Patent 238341, 1908. (48) Mizuno, N.; Misono, M. Heterogeneous Catalysis. Chem. Rev.
(27) Hill, A. J.; Fischer, E. J. A; Synthesis of beta-chloro-allyl chloride. 1998, 98 (1), 199−218.
J. Am. Chem. Soc. 1922, 44, 2582−2595. (49) Okuhara, T.; Mizuno, N.; Misono, M. Catalytic Chemistry of
(28) Britton, E. G.; Heindel, R. X. Preparation of glycerol Heteropoly Compounds. Adv. Catal. 1996, 41, 113−252.
dichlorohydrin. U.S. Patent 2,144,612, 1939. (50) Kozhevnikov, I. V. Catalysis by Heteropoly Acids and
(29) Britton, E. C.; Slagh, H. R. Glycerol dichlorohydrin. U.S. Patent Multicomponent Polyoxometalates in Liquid-Phase Reactions. Chem.
2,198,600, 1940. Rev. 1998, 98 (1), 171−198.
(30) Clarke, H. T.; Hartman, W. W. Epichlorohydrin. Org. Synth., (51) Song, S. H; Lee, S. H.; Park, D. R.; Kim, H.; Woo, S. Y; Song,
Coll. 1923, 3, 47. W. S.; Kwon, M. S.; Song, I. K. Direct preparation of dichloropropanol
(31) Conant, J. B.; Quayle, O. R. Glycerol α-monochlorohydrin. Org. from glycerol and hydrochloric acid gas in a solvent-free batch reactor:
Synth., Coll. 1922, 2, 33. Effect of experimental conditions. Korean J. Chem. Eng. 2009, 26 (2),
(32) Conant, J. B.; Quayle, O. R. Glycerol α,γ-dichlorohydrin. Org. 382−386.
Synth., Coll. 1922, 2, 29. (52) Lee, S. H.; Song, S. H.; Park, D. R.; Jung, J. C.; Song, J. H.; Woo,
(33) Thompson, W. P. Recovery of aliphatic chlorohydrins having S. Y.; Song, W. S.; Kwon, M. S.; Song, I. K. Solvent-free direct
from 3 to 5 carbon atoms from aqueous solution. Br. Patent GB 931,211, preparation of dichloropropanol from glycerol and hydrochloric acid
1963. gas in the presence of H3PMo12−XWXO40 catalyst and/or water
(34) Krafft, P.; Gilbeau, P.; Gosselin, B.; Claessens, S. (Solvay). absorbent. Catal. Commum. 2008, 10, 160−164.
Process for producing dichloropropanol from glycerol, the glycerol coming (53) Graham Solomons, T. W. Organic Chemistry, 7th ed.; John
eventually from the conversion of animal fats in the manufacture of Wiley & Sons: New York, 1999; pp 741−744.
biodiesel. PCT Patent WO 054167 A1, 2005. (54) March, J. Advanced Organic Chemistry, 5th ed.; Wiley−-
(35) Schreck, D.; Kruper, W.; Varjian, R. D.; Jones, M. E.; Campbell, Interscience: New York, 2001; pp 308−312.
R. M.; Kearns, K.; Hook, B. D.; Briggs, J. R.; Hippler, J. G. (Dow (55) Boschan, R.; Winstein, S. The Role of Neighboring Groups in
Global Tech., Inc.). Conversion of multihydroxylated-aliphatic hydro- Replacement Reactions. XXI. Front-Side Participation of the Acetoxy
carbon or ester thereof to a chlorohydrin. PCT Patent WO 2006/020234 Group. Catalytic Effect of Acetic Acid on the Reaction of Glycols with
and Eur. Patent EP 1 771 403 B1, 2007. Hydrogen Chloride. J. Am. Chem. Soc. 1956, 78, 4921−4925.
(36) Siano, D.; Santacesaria, E.; Fiandra, V.; Tesser, R.; Di Nuzzi, G.; (56) Dmitriev, S. G.; Zanaveskin, N. L. Hydrochlorination of
Di Serio, M.; Nastasi, M. (Eurochem Engineering srl). Process for the GlycerolThe Role of the Water on the Process. J. Chem. Chem. Eng.
production of α,γ-dichlorohydrin from glycerin and hydrochloric acid. 2011, 5, 1179−1182.
WO 111810 A2, 2006; U.S. Patent US 2009/0062574, 2009; Eur. (57) Luo, Z. H.; You, X. Z.; Li, H. R. Direct Preparation Kinetics of
Patent EP 1879842 B1, 2012; Chin. Patent ZL 2006, 80012864.9, 1,3-Dichloro-2-propanol from Glycerol Using Acetic Acid Catalyst.
2006. Ind. Eng. Chem. Res. 2009, 48, 446−452.
(37) Kubicek, P.; Sladek, P.; Buricova, I. (Spolek Pro Chemickou). (58) Luo, Z. H.; You, X. Z.; Li, H. R. A kinetic model for glycerol
Method of preparing dichloropropanols from glycerine. World Patent chlorination in the presence of acetic acid catalyst. Korean J. Chem.
WO2005/021476 and U.S. Patent US 2007/0167659, 2007. Eng. 2010, 27 (1), 66−72.
(38) Santacesaria, E.; Di Serio, M.; Tesser, R. (Eurochem (59) Lim, J. H.; Song, W. S.; Woo, S. Y.; Lee, D. H. Kinetic model of
Engineering srl). Process for monochlorohydrins production from glycerol glycerol chlorination with hydrochloric acid. Korean J. Chem. Eng.
and hydrochloric acid. PCT Patent WO 132770 A1, 2008. 2010, 27 (3), 785−790.
(39) Kruper, W. J., Jr.; Arrowood, T.; Bell, B. M.; Briggs, J. ; (60) Ling, D. Q.; Lu, J.; Wang, M. X.; Liang, S.; Zhang, W.; Ren, J.;
Campbell, R. M.; Hook, B.; Nguyen, A.; Theriault, C.; Fitschen, R. Ouyang, C. P. Investigation of the kinetics and mechanism of the
Batch, semi-continuous or continuous hydrochlorination of glycerol glycerol chlorination reaction using gas chromatography−mass
with reduced volatile chlorinated hydrocarbon by-products and spectrometry. J. Serb. Chem. Soc. 2010, 75 (1), 101−112.
chloroacetone level. U.S. Patent US 7,906,690 B2, 2011. (61) De Araujo Filho, C. A.; Salmi, T.; Bernas, A.; Mikkola, J. P.
(40) Tirtowidjojo, D.; Merenov, A. S.; Kneupper, C. D.; Hook, B. D.; Kinetic Model for Homogeneously Catalyzed Halogenation of
Metha, A. L. (Dow Chemical Co.). Multi-stage process and apparatus Glycerol. Ind. Eng. Chem. Res. 2013, 52, 1523−1530.

8961 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962


Industrial & Engineering Chemistry Research Review

(62) Macey, R.; Oster, G. Berkeley Madonna’s Modeling and Analysis


of Dynamic Systems, release 8.3; University of CaliforniaBerkeley:
Berkeley, CA, 2001.
(63) Dmitriev, S. G.; Zanaveskin, N. L. Synthesis of Epichlorohydrin
from Glycerol. Hydrochlorination of Glycerol. Chem. Eng. Trans. 2011,
24, 43−48.
(64) Patai, S. The Chemistry of Functional Groups; Interscience: New
York, 1967.

8962 dx.doi.org/10.1021/ie403268b | Ind. Eng. Chem. Res. 2014, 53, 8939−8962

You might also like