You are on page 1of 8

Fuel Processing Technology 120 (2014) 40–47

Contents lists available at ScienceDirect

Fuel Processing Technology


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

Lactic acid production from glycerol using CaO as solid base catalyst
Lu Chen, Shoujie Ren, X. Philip Ye ⁎
Department of Biosystems Engineering and Soil Science, The University of Tennessee, 2506 E.J. Chapman Drive, Knoxville, TN 37996, USA

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

Article history: In the valorization of glycerol as byproduct of biodiesel production, although recent progress in glycerol conver-
Received 2 September 2013 sion to lactic acid using homogeneous chemocatalysis showed promising high yield, the used high alkalinity
Received in revised form 18 November 2013 entails high corrosiveness to reactors and problematic downstream separations. In this study, five solid base
Accepted 19 November 2013
catalysts were screened for converting glycerol to lactic acid with the aim to ease corrosiveness, focusing on
Available online xxxx
inexpensive CaO as a promising solid base. Process conditions were systematically investigated for optimization.
Keywords:
The highest yield of lactic acid achieved was 40.8 mol% with a glycerol conversion of 97.8 mol% at the optimum
Biodiesel conditions using refined glycerol. Similar conversion rate and lactic acid yield were also obtained in the conver-
Glycerol sion of crude glycerol using CaO if the water content in crude glycerol is lower than 10%. CaO exhibited lower
Calcium oxide activation energy in converting glycerol to lactic acid compared to homogeneous NaOH catalyst. Corrosiveness
Lactic acid to reactor using CaO was proven much lower than that using homogeneous NaOH catalyst. CaO as catalyst for
Corrosiveness both biodiesel production and subsequent crude glycerol conversion to lactic acid was investigated, revealing
Solid base catalysts its potential industrial applications for the production of both biodiesel and lactic acid.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction unfriendly purification of the products resulting in waste gypsum [15].


In chemical conversion route, Kishida et al. [20] first reported a high
Crude glycerol is a main byproduct in biodiesel production formed in LA yield of 90 mol% from aqueous glycerol (0.33 M) at hydrothermal
the transesterification of triglycerides, with an output of approximately conditions of 300 °C and 90 min with a 1.25 M NaOH as homogeneous
10% by weight based on vegetable oil used [1]. The U.S. production of catalyst. Shen et al. [21] further investigated the effects of eight alkali
biodiesel was 967 million gallons in 2011 [2]. With the growth of bio- metals and alkaline-earth metals on hydrothermal conversion of glycer-
diesel industry, large amount of crude glycerol output has created a ol to lactic acid, concluding that alkali-metal hydroxides were more
glut in the glycerol market and consequently significant price decrease effective than alkaline-earth-metal hydroxides for the catalysis; the
for glycerol [3]. It is noteworthy to point out that glycerol has been best LA yield of 90 mol% was obtained at 300 °C with NaOH or KOH as
selected as one of the top 12 platform chemicals from biomass [4] and a homogeneous catalyst. To increase productivity by using a higher
the conversion of glycerol to value-added chemicals have been inten- glycerol concentration of 2.5 M, Ramirez-Lopez et al. [6] reported an
sively investigated in recent years [5–14]. 84.5 mol% LA yield achieved at 280 °C and 90 min with a 1.1 NaOH to
Lactic acid (LA) is a widely used chemical in food and food related glycerol molar ratio. However in these studies, the low LA productivity
industries [15], and chemical industries to form other useful chemicals due to low glycerol concentration and high corrosiveness due to high
and materials such as pyruvic acid, acrylic acid, and polylactic acid alkali concentration render the processes industrially unattractive, and
polymer (PLA) which is known to be one of the best materials for biode- it was also noted that the product existed as lactate via the instant
gradable plastics [16]. Previous reports show that lactic acid can be reaction of lactic acid with the base catalysts. This consumption of
produced through both fermentation of carbohydrates and chemical base requires a high concentration of OH − in order to catalyze the
conversion starting from glycerol. In fermentation processes which conversion of a high concentration of glycerol, which increases the
currently dominate the industrial production of lactic acid, although corrosiveness. Previous study [6] and our preliminary trials showed
high yield of lactic acid (~ 90 wt.% based on dextrose equivalent of that high concentration of OH − (N1 M) caused severe corrosion to
carbohydrates) can be achieved, very low LA productivity in the range stainless steel reactors, and high reaction temperature as well as high
of 0.005–0.08 g · min−1 L−1 is also evident [17]. Several problems are pressure further aggravate the corrosion, making the processes difficult
further associated with fermentation, such as high cost of culture to be used in industrial production. Therefore, to make the process
media due to the specific requirement of lactic acid producing bacteria converting glycerol to lactic acid more practical for industrial applica-
[18], product inhibition [19], and the uneconomical and environmentally tion, it is key to develop suitable catalysts with less corrosiveness.
A solid base is a substance that has the tendency to accept protons
⁎ Corresponding author. Tel.: +1 865 974 7129; fax: +1 865 974 4514. (Brønsted base) or donate electron pairs (Lewis base). Potential advan-
E-mail address: xye2@utk.edu (X.P. Ye). tages of using solid base catalysts include low corrosiveness, catalyst

0378-3820/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.fuproc.2013.11.019
L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47 41

reusability, and environmental friendliness by lowering reaction (Ea) using CaO as solid base catalyst was also estimated based on the
temperature, simplifying catalyst/product separation, and reducing results. In order to find out better reaction conditions, higher molar
chemical waste [22,23]. Solid bases are expected to have increasing ratios of catalyst to glycerol at 0.3, 0.4 and 0.5 were then investigated
importance in the fine chemical industries. Among solid base catalysts, at fixed reaction temperature and time of 290 °C and 150 min,
calcium oxide (CaO) has a strong basicity and economic advantages respectively.
such as the low cost and easy preparation. CaO as solid base catalyst Because water is the major impurity in crude glycerol, the effect of
has been studied in many processes especially in biodiesel production water content in feedstock was also investigated by adding water to
[24–28]. However to the best of our knowledge, CaO as solid base refined glycerol at reaction temperature of 290 °C and reaction time of
catalyst in glycerol conversion to lactic acid has not been reported. 150 min with a fixed molar ratio of CaO to glycerol at 0.3. The water
Objective of this study is to investigate the conversion of glycerol to contents tested were 2, 5, 10, 20 and 30 wt.% in comparison with 0%
lactic acid using solid bases, focusing on the inexpensive calcium oxide (pure refined glycerol).
as a promising solid base. Process conditions such as reaction tempera- In order to determine the corrosiveness of solid base catalyst of CaO
ture and time, catalyst structure and loading, and water content in in comparison with that of homogeneous NaOH, we replicated the
refined and crude glycerol were systematically studied. Direct use of experiment with the highest LA yield described in Kishida et al. [20],
industrial crude glycerol was also investigated in order to shed light which was conducted at 300 ºC for 90 min with 1.25 M NaOH as cata-
on the effects of impurities contained in crude glycerol. CaO as the lyst, and compared it with CaO catalyzed glycerol conversion with the
catalyst for both transesterification of soybean oil to biodiesel and the best LA yield conducted at 290 ºC for 150 min with a CaO to glycerol
subsequent crude glycerol to lactic acid were investigated. The corro- molar ratio of 0.3. The concentration of Fe3+ ions in the mixture after
siveness of CaO to reactor in the glycerol conversion was also evaluated reaction was selected as an indicator of corrosiveness to the stainless
in comparison with that of NaOH. steel reactor, which was measured using atomic absorption spectrosco-
py (AAnalyst™ 700, PerkinElmer, Inc.).
2. Materials and methods
2.4. Structure–function of CaO on lactic acid production
2.1. Materials and catalyst preparation
Four types of CaO catalysts including CaO powder (Thermo Fisher
Base catalysts or the precursors and support used in this study were Scientific), CaO calcinated from CaCO3 (Thermo Fisher Scientific) at
powders of CaO, Ca(OH)2, hydrotalcite (Mg6Al2(CO3)(OH)16·4(H2O)), 910 °C in a muffle furnace for 3 h, CaO nanopowder (Strem Chemicals),
CaCO3, CaO nanopowder, and alumina. All the above and standard and regenerated CaO used in this study were investigated to determine
chemicals of refined glycerol (99.5% purity), lactic acid (90% solution if their structures affect the glycerol conversion and lactic acid produc-
in water), and sulfuric acid (98%) were purchased from Thermo Fisher tion. Regenerated CaO was obtained by the following procedure. This
Scientific (Waltham, MA) except for hydrotalcite (Sigma-Aldrich, was just meant to see the effect of regeneration, although better regen-
St. Louis, MO) and CaO nanopowder (Strem Chemicals, Newburyport, eration methods need to be developed. Stoichiometric amount of
MA). Deionized water (DI water) was used throughout all reactions. sodium hydroxide solution was added into the product mixture to
Five solid base catalysts were first screened, including CaO, Ca(OH)2, convert the calcium lactate to calcium hydroxide which precipitated
hydrotalcite, hydrotalcite dehydrated at 450 ºC with N2 flow for 4 h, out from the solution. Then the mixture went through filtration to
and alumina loaded with 35 wt.% KNO3. CaO and CaO nanopowder collect the precipitation which mainly consisted of Ca(OH)2. The precip-
were calcined at 910 °C for 3 h before being used in the reaction. Alumi- itation slurry was dried in a convective oven at 110 °C overnight. At the
na loaded with KNO3 was fabricated following the method reported by end, the collected solids were calcinated at 910 °C in a muffle furnace
Xie et al. [29]. For catalyst screening, all the reactions were conducted at for 3 h, in order to convert the Ca(OH)2 back to CaO and burn off any
300 °C for 90 min with 1.25 M catalyst and 0.33 M aqueous solution of deposited organic molecules. Glycerol conversion using these catalysts
glycerol. was performed at the reaction temperature of 290 °C for 150 min
with the catalyst to glycerol molar ratio of 0.3.
2.2. Reaction procedure The structure and surface of these catalysts were studied using X-ray
diffraction (XRD) and scanning electron microscopy (SEM) (Zeiss 1525,
Conversion of glycerol to LA with solid base catalysts were carried Carl Zeiss SMT Inc.). XRD diffractograms were acquired on a Bruker AXS
out in a salt bath heater, as previously described [30]. A stainless steel D8 Advance X-ray diffractometer using Cu-Ka radiation (0.15418 nm).
(SS316) batch reactor with OD of 15.9 mm (5/8 in.), ID of 11.05 mm The tube voltage and current was 40 kV and 40 mA, respectively.
(wall thickness of 0.095 in.), and length of 105 mm was used. Total The scanning step size was 0.04° with a rate of 1 s per step. The
volume of the reactor is 10 mL. The reactants can be heated to desired diffractogram was collected for the 2θ range of 2°–80°.
temperatures in about 90 s from room temperature.
Glycerol and catalyst with a pre-calculated molar ratio based on 2.5. Conversion of crude glycerol to lactic acid using CaO
experimental design were well mixed and placed into the reactor.
Before sealing the Swagelok® caps, the mixture was degassed by purg- To investigate whether the impurities in crude glycerol might affect
ing with nitrogen. Then the loaded reactor was immersed into the the conversion of glycerol to lactic acid using CaO, three types of crude
preheated salt bath for designated time lengths. After the reaction, the glycerol from biodiesel industry with different compositions were test-
reactor was quenched into an ice-water bath immediately, and the ed. Approximate analysis of these crude glycerol samples are shown in
product sample was collected for analyses. Table 1. The reactions were carried out at 290 °C for 150 min with a
0.3 molar ratio of CaO to glycerol.
2.3. Kinetics investigation using CaO
2.6. Production of biodiesel and lactic acid from soybean oil using CaO
Refined glycerol without water addition was first tested. According
to our preliminary experiments, the molar ratio of CaO powder to glyc- Biodiesel production from soybean oil and subsequent glycerol
erol at 0.2 was first selected to investigate the effects of reaction temper- conversion to lactic acid using CaO as catalyst were investigated. Two
ature and time on the glycerol conversion and lactic acid yield. The methods of biodiesel production were utilized and the reactions were
reaction temperatures was chosen at 280 °C, 290 °C and 300 °C and carried out in a Parr® reactor (model 4560, Parr Instrument Company,
the reaction time ranged from 45 min to 240 min. Activation energy IL). In the first method, the reaction was carried out at 65 °C for 3 h,
42 L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47

Table 1 where Ybiodiesel is the yield of biodiesel (wt.%), mproduced is the mass of
Properties of crude glycerol. biodiesel produced, and mTheoreticalMax is the mass of theoretically maxi-
ID# Glycerol Water Soap content Methanol pH mum biodiesel yield calculated based on the stoichiometric equation of
(wt.%) (wt.%) (wt.%) (wt.%) biodiesel production from the vegetable oil.
CG1 88.5 ± 1.18a 7.4 ± 1.54 2.4 ± 0.6 1.3 ± 0.4 7.0
CG2 32.1 ± 2.00 4.4 ± 2.71 60.8 ± 1.4 0.0 ± 0.0 12.0
3. Results and discussion
CG3 20.4 ± 0.93 12.0 ± 2.26 53.1 ± 2.5 10.3 ± 3.8 12.2
a
Standard error of three samples. 3.1. Screening of solid base catalysts

Results of glycerol conversion and lactic acid yield using five differ-
ent solid base catalysts are shown in Table 2. CaO powder and alumina
with a molar ratio of methanol to oil at 12:1, 8 wt.% CaO powder as
loaded with KNO3 showed glycerol conversion greater than 78%, but
catalyst based on oil weight, and 2.03 wt.% water (based on methanol)
Ca(OH)2, hydrotalcite and dehydrated hydrotalcite exhibited much
which was same as the method reported by Liu et al. [27]. In the other
less activity in the dehydrogenation of glycerol to LA. The low observed
method, CaO powder was first modified with trimethylchlorosilane
yields of lactic acid for all reactions might be due to the mild basicity of
(TMCS) according to a method reported by Tang et al. [26]; the mixture
the catalysts, except for CaO which obviously reacted with water to
of methanol to oil (with 15:1 molar ratio) and 5 wt.% modified CaO was
form Ca(OH)2, decreasing its basicity. Although alumina loaded with
heated to 65 °C for 3 h.
35 wt.% KNO3 resulted in the highest glycerol conversion, the low selec-
After biodiesel production, the excess methanol was distilled off
tivity of this catalyst rendered only 3.8 mol% yield of LA. Furthermore,
under vacuum, and the remaining products formed three layers after
the fabrication and recycle of a supported catalyst are more costly
centrifugation; from top to bottom layers were biodiesel, glycerol, and
than those of straightforward use of CaO. Relatively, CaO powder
mixture of solid CaO with a small amount of glycerol, in agreement
showed the best performance. Therefore, the performance and reaction
with that described by Liu et al. [27]. The byproduct of crude glycerol
kinetics of using CaO were systematically examined in the following
together with CaO was separated by a separatory funnel, and the
sections of 3.2 to 3.5.
mixture was used for the subsequent lactic acid production by heating
It is noteworthy to point out that at the end of the reactions, lactic
to 290 °C for 150 min.
acid existed in the form of lactate which was converted to lactic acid
by adding H2SO4 before HPLC analysis. So, in this study the lactic acid
2.7. Sample analysis
is used to express the results.

The product mixture was washed out with DI water and the pH
was adjusted to 3–4 using diluted sulfuric acid. Then the sample 3.2. Kinetics results using refined glycerol and CaO
was centrifuged at 4700 ×g for 15 min, and the clear liquid collected
from upper layer after centrifugation was passed through an ion- A typical HPLC chromatogram of the product analysis from glycerol
exchange column packed with DOWEX 50WX8-400 resin (Sigma Al- conversion using CaO catalyst is depicted in Fig. 1, showing major
drich) to remove dissolved metal ions. The pretreated samples were chemicals detected.
then analyzed on a HPLC (Waters® 410, Waters Corporation) In the conversion of refined glycerol without water addition, both
equipped with a Shodex SH1011 (300 × 8 mm) column and a refrac- reaction temperature and time had significant effects on the glycerol
tive index detector. The column temperature was set at 60.0 °C, and conversion and lactic acid yield (Fig. 2). Obviously, glycerol conversion
the mobile phase used was 0.5 mM H 2SO4 in DI water with a flow increased with the increase of reaction temperature and time. The
rate of 0.6 mL/min. highest lactic acid yield at ~ 28 mol% with CaO to glycerol ratio of 0.2
Glycerol conversion, lactic acid yield, selectivity to lactic acid, and were observed at the reaction temperatures of 290 °C and 300 °C, and
yield of biodiesel were calculated according to the following equations: reaction time around 90 min; similar lactic yield was also observed at
280 °C but with much longer reaction time of 205 min. The maximum
nfeed −nremained yield was achieved at a shorter reaction time when the reaction temper-
X GL ðmol% Þ ¼  100% ð1Þ
nfeed ature increased, indicating that the reaction rate was accelerated at
higher temperatures.
where XGL is glycerol conversion (mol%), nfeed is the moles of starting At temperatures of 290 °C and above, lactic acid yield greatly
glycerol, and nremained is the moles of remaining glycerol in the sample decreased with the increase of reaction time after reaching the maxi-
quantified by HPLC. mum yield. Similar changes of lactic acid yield were also reported by
Ramirez-Lopez et al. [6] in the conversion of glycerol at alkaline

nLA
Y LA ðmol% Þ ¼  100% ð2Þ
nfeed

where YLA is the yield of lactic acid (mol%), nLA is the moles of lactic acid
produced.
Table 2
Results of solid base screening.a

nLA Description of solid base LA yield Glycerol conversion


SLA ðmol% Þ ¼  100% ð3Þ (mol%) (mol%)
nfeed −nremained
CaO 7.5 78.5
Ca(OH)2 4.6 27.6
where SLA is the selectivity to lactic acid (mol%). Hydrotalcite (Mg6Al2(CO3)(OH)16·4(H2O) 0.0 19.3
Hydrotalcite (dehydrated at 450 °C with N2 flow) 0.0 25.1
Alumina loaded with 35 wt.% KNO3 3.8 83.1
mproduced
Y biodiesel ðwt:% Þ ¼  100% ð4Þ
a
All reactions were conducted at 300 °C for 90 min with 1.25 M catalyst and 0.33 M
mTheoreticalMax aqueous glycerol solution.
L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47 43

Fig. 1. Typical HPLC chromatogram of product analysis.

hydrothermal conditions. This decrease of lactic acid yield at reaction


temperatures ≥ 290 °C is due to the degradation of calcium lactate,
Fig. 3. Effect of CaO loading on glycerol conversion and lactic acid yield; all reactions were
which may have higher activation energy than that for the generation conducted at 290 °C for 150 min; error bar indicates the standard error of four samples.
of lactic acid from glycerol and is accelerated at higher temperatures.

3.2.1. Effect of CaO loading sufficient for this process. It is important to point out that the formation
To examine the effect of catalyst loading, we fixed reaction temper- of calcium lactate is nontrivial because calcium lactate is much more
ature at 290 °C and reaction time at 150 min to achieve high yield of heat-resistant than lactic acid to withstand the reaction conditions
LA and limit the degradation, and the results are shown in Fig. 3. without being quickly decomposed.
With the increase of CaO to glycerol molar ratio from 0.2 to 0.3, glyc-
erol conversion greatly increased from 81% to 97.8%, but further increas-
ing the CaO loading did not result in significant changes. Similar trend 3.2.2. Effect of water content in feed
was also observed for the LA yield, which increased from 29.4% to The effect of water addition to glycerol feed is shown in Fig. 4. Glyc-
40.8% with the increasing molar ratio from 0.2 to 0.3, but there were erol conversion showed decreasing trend with the increase of water
no obvious changes when the catalyst loading was further increased. content from 0 to 10%, but this decreasing trend was not observed
In this reaction system, the produced lactic acid would instantly with further increasing water content. Overall, the glycerol conversion
react with CaO due to acid–base neutralization, forming calcium lactate rates were higher than 85% at all the levels of water content tested.
and consuming the catalyst. With the reaction going on, less CaO will be However, the water content showed significant effect on LA yield
available to catalyze the desired dehydrogenation of glycerol to lactic which was revealed by using Tukey's test at significance level of 0.05.
acid. This explains why a lower molar ratio of 0.2 resulted in lower LA When the water content was lower than or equal to 5%, LA yield was
yield. Our results suggest that a CaO to glycerol molar ratio at 0.3 is not significantly affected by water addition (p N 0.05 for 0%, 2%, and
5% water contents, Tukey's test). But when the water content was 10%
100 or higher, LA yield significantly decreased compared with that when
the water content was lower than 5% (p b 0.05, Tukey's test). When
Glycerol conversion (mol%)

80 water exists in the system, CaO would react with water forming
Ca(OH)2. Although Ca(OH)2 is slightly soluble in water acting as a
60 homogeneous base catalyst in the glycerol conversion, its basicity is
much weaker than CaO. Therefore, the catalytic activity decreased,
40 resulting in less conversion of glycerol and lower LA yield. Based on
the CaO loading in this study, CaO would be totally converted into
20 Ca(OH)2 when the water content is more than 5%.
280°C 290°C 300°C
0
0 50 100 150 200 250
Time (min)
35
Lactic acid yield (mol%)

30

25

20

15

10

5
280°C 290°C 300°C
0
0 50 100 150 200 250
Time (min)

Fig. 2. Glycerol conversion and lactic acid yield as function of reaction temperature and Fig. 4. Effect of water content in feed on glycerol conversion and lactic acid yield; all reac-
time with 0.2 molar ratio of CaO to glycerol; error bar indicates the standard error of tions were conducted at 290 °C for 150 min with CaO to glycerol molar ratio of 0.3; error
four samples. bar indicates the standard error of four samples.
44 L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47

Previous report pointed out that lactic acid can be degraded to carbon Table 4
dioxide and other chemicals such as acetic acid, especially at tempera- Performance of four types of CaO in glycerol conversion to lactic acid.

tures higher than 280 °C [6]. In this study, gas bubbles were observed Catalysts Glycerol Lactic acid Lactic acid
when sulfuric acid was added to the product mixture. This might indicate conversion yield selectivity
that lactic acid decomposed to carbon dioxide and then reacted with (mol%) (mol%)a (mol%)

Ca(OH)2 to form CaCO3, further decreasing the catalyst activity. CaO powder 97.8 ± 2.22b 40.8 ± 2.36A 41.7 ± 3.26
Analysis of our experimental data showed that the conversion of CaO calcinated from CaCO3 96.7 ± 2.83 37.2 ± 0.28AB 38.5 ± 1.40
Regenerated CaO 95.2 ± 2.85 38.7 ± 1.95A 40.6 ± 2.01
glycerol catalyzed by CaO follows a first order kinetics with a activation
CaO nanopowder 94.1 ± 6.92 35.8 ± 1.56B 38.0 ± 1.17
energy (Ea) of 103 kJ/mol, lower than that of 174 kJ/mol reported by
a
Kishida et al. [31] in a kinetic study on the conversion of glycerol to lactic Means with different letters are significantly different from each other (p ≤ 0.05,
Tukey's test).
acid at hydrothermal conditions catalyzed by homogeneous NaOH b
Standard error of four samples.
catalyst.
In the evaluation of corrosiveness to reactor, the measured Fe3 +
concentration in the mixture after CaO catalyzed conversion was for CaO powder, but the intensity of these characteristic peaks for CaO
3.06 ppm, much lower than that of 9.55 ppm when NaOH was used as nanopowder was much lower, indicating that the crystalline portion
catalyst. in these CaO catalysts might play an important role in promoting lactic
acid production.
3.3. Crude glycerol conversion using CaO Fig. 5B shows SEM micrographs of the four types of CaO. The CaO
powder has spherical shape with particle size around 1 μm, while the
Results for the conversion of crude glycerol at 290 °C for 150 min regenerated CaO exhibits similar shape and particle size but the parti-
with a 0.3 molar ratio of CaO to glycerol are shown in Table 3. cles were obviously aggregated during the regeneration process. The
Glycerol conversion and lactic acid yield from CG1 are very close to CaO calcinated from CaCO3 shows a porous structure with larger aggre-
the case of converting refined glycerol with small amount of water ad- gate particle size. Although the CaO nanopowder has a high specific
dition (2 wt.% and 5 wt.%) as shown in Fig. 4. The reaction using CG2 surface area (≥90 m2/g) ideal for catalytic reaction, it also has a much
showed similar glycerol conversion to the case of converting refined larger aggregate size (mean aggregate size = 4 μm according to manu-
glycerol with small amount of water addition (2 wt.% and 5 wt.%) but facturer specifications) that might hinder the dispersion of the catalyst
higher lactic acid yield. From Table 1, it can be seen that the pH of CG2 into glycerol. This possible poor dispersion would result in insufficient
is 12, indicating that the alkali residue from biodiesel production actual- mass transfer of glycerol to the catalyst, failing to fully utilize the high
ly helped the production of lactic acid by increasing OH\ strength in the surface area.
reaction mixture. This observation may help in the handling of crude
glycerol in biodiesel production if the crude glycerol is to be used for 3.5. Production of biodiesel and lactic acid from soybean oil using CaO
the production of lactic acid. Although the pH of CG3 is close to that of
CG2, the high water content (12 wt.%) might have caused deactivation Results for biodiesel production from soybean oil and subsequent
of CaO, as discussed in Section 3.2.2, resulting lower glycerol conversion glycerol conversion to lactic acid using CaO as catalyst are shown in
and lactic acid yield. Table 5. The yields of biodiesel and lactic acid using unmodified CaO
were 78.3 wt.% and 49.7 mol% respectively, while the TMCS modified
3.4. Structure–function of CaO on lactic acid production CaO rendered higher biodiesel yield (86.5 wt.%) but significantly
lower lactic acid yield (15 mol%), indicating that the modification favors
Results for evaluating the four types of CaO catalysts are shown in the biodiesel production by facilitating mass transfer between oil and
Table 4. All the reactions were conducted at 290 °C for 150 min with catalyst through altering CaO surface to be more hydrophobic. However,
the catalyst to glycerol molar ratio of 0.3, using refined glycerol as the subsequent production of LA requires a strong base with hydrophilic
feed without water addition. Similar glycerol conversion and lactic surface to facilitate the mass transfer between glycerol and the catalyst.
acid yield were observed for CaO powder, CaO calcinated from CaCO3, The TMCS modification might have also weakened the basicity of CaO.
and regenerated CaO. However, the reaction using CaO nanopowder re- Future research should focus on the development of amphiphilic base
sulted in the lowest glycerol conversion and lactic acid yield, which is catalysts that would be favorable to both biodiesel production and sub-
surprising because we have chosen a CaO nanopowder with ideal spec- sequent glycerol conversion to lactic acid. Furthermore, it is noteworthy
ifications for glycerol conversion: specific surface area (BET) ≥ 90 m2/g, to point out that the high glycerol conversion using both unmodified
average pore diameter = 110 Å, total pore volume ≥ 0.2 cm3/g, CaO (100 mol%) and TMCS modified CaO (90.5 mol%) was due to the
and crystallite size ≤ 20 nm (according to technical data from the high molar ratio of CaO to glycerol (about 1.6 for unmodified CaO and
manufacturer). 0.9 for TMCS modified CaO). Our results suggest the potential of using
The crystalline phases and structure differences among the four a single base catalyst for integrated production of biodiesel and lactic
types of CaO were studied using X-ray diffraction (XRD) and scanning acid, increasing values of products from renewable resources.
electron microscopy (SEM). In the XRD diffractogram (Fig. 5A), peaks
at the angles of 33°, 37°, 54°, 65°, and 67° were used to characterize 3.6. Reaction pathway of lactic acid production from glycerol using CaO
CaO. At these positions, intensity of the characteristic peaks for CaO cal-
cinated from CaCO3 and regenerated CaO was slightly lower than that It may be controversial to claim CaO as a heterogeneous catalyst
because CaO is considered soluble in glycerol according to technical
data from Thermo Fisher Scientific (Waltham, MA), and it has also
Table 3 been reported that CaO reacts with glycerol forming calcium glycerides
Crude glycerol conversion and lactic acid yield using CaO.
in the temperature range from 10 °C to 150 °C [25,28,32]. However, in a
Crude glycerol ID# Glycerol conversion Lactic acid yield Lactic acid selectivity study of glycerol etherification to di- and triglycerol using CaO as a
(mol%) (mol%) (mol%) heterogeneous catalyst, Ruppert et al. [33] found that colloidal CaO
CG1 93.7 ± 2.21a 40.3 ± 2.36 43.0 ± 2.7 particles of about 50–100 nm could be spontaneously generated during
CG2 87.9 ± 1.21 52.3 ± 3.17 59.5 ± 1.4 reaction, revealing both hetero- and homogeneous nature of CaO
CG3 79.5 ± 1.08 18.4 ± 1.15 23.1 ± 1.3 catalyst in glycerol. We observed that the above possible dissolution
a
Standard error of four samples. or reactions were slow and occurred at lower temperatures. Subjecting
L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47 45

Fig. 5. X-ray diffraction patterns (A) and scanning electron micrographs (B) of four types of CaO: (a) CaO powder; (b) CaO calcinated from CaCO3; (c) regenerated CaO; (d) CaO
nanopowder.

the reaction mixture to a higher temperature would change reaction catalyst, the reaction pathways are proposed as depicted in Fig. 6. The
kinetics favorable to LA production. entire process initiates with the conversion of glycerol to glyceralde-
Reaction pathways for the alkali-catalyzed conversion of glycerol in hyde via glyceroxide ion; CaO catalyst plays an important role in
aqueous solution were previously proposed [6,20]. Using CaO as solid advancing the formation of glyceroxide ion and promoting hydrogen

Table 5
Biodiesel and lactic acid production using CaO as catalyst.

Biodiesel production Lactic acid production

MeOH: Catalyst Temperature Time Biodiesel yield Temperature Time Lactic acid yield Glycerol conversion
oil (°C) (min) (wt.%) (°C) (min) (mol%) (mol%)

12:1 CaO 65 180 78.3 290 150 49.7 100.0


(unmodified)
15:1 TMCS 65 180 86.5 290 150 15.0 90.5
modified CaO
46 L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47

(A)

(B)

Fig. 6. Reaction pathways for glycerol conversion to lactic acid catalyzed by CaO.

abstraction. At a basic condition, glyceraldehyde is dehydrated to 2- CaO performed the best. Kinetic study focusing on CaO demonstrated
hydroxypropenal, which readily converts to pyruvaldehyde via keto- that glycerol conversion and lactic acid yield were significantly af-
enol tautomerization (no catalyst is necessary in this step). Hydrogen fected by reaction temperature and time, the CaO catalyst loading,
addition may occur to both 2-hydroxypropenal and pyruvaldehyde crystalline structure of CaO, and water content in the reactants. At
using in situ generated H2, and propylene glycol is formed as the result; the optimum condition (reaction temperature of 290 °C, reaction
the small amount of propylene glycol formed, as shown in Fig. 1, indicat- time of 150 min, molar ratio of CaO to glycerol at 0.3), lactic acid
ing that the CaO catalyst is not favorable to the hydrogenolysis. yield reached 40.8 mol% with a 97.8 mol% glycerol conversion
Therefore, pyruvaldehyde may preferably undergo benzilic acid rear- using refined glycerol without water addition. Water content higher
rangement with the assistance of CaO, and calcium lactate is formed than 10% in reactants had negative effects on the LA yield, due to its
down this path (Fig. 6A). reaction with CaO forming Ca(OH)2 that has weaker basicity. Appli-
Unfortunately, base-assisted formation of glyceroxide ion may also cation of the optimum condition to the conversion of crude glycerol
branch into oligomerization (Fig. 6B), although the oligomerization is rendered similar glycerol conversion and LA yield if the water con-
favored at lower temperatures in the absence of water, as reported in tent of crude glycerol was lower than 10%, and we also found that
previous studies [33,34]. Furthermore, instead of benzilic acid rear- the alkali residue in crude glycerol as a byproduct of biodiesel pro-
rangement (intramolecular reaction), pyruvaldehyde may also undergo duction has a positive effect on the LA yield. The solid base of CaO
intermolecular reaction, leading to higher molecular weight chemicals. exhibited lower activation energy in converting glycerol to lactic
These oligomers might be less soluble or tightly adsorbed on the catalyst acid compared to homogeneous alkali of sodium hydroxide. Analy-
surface, contributing to the loss of carbon balance. sis of Fe3 + concentration in reaction mixtures showed that the cor-
Based on the results of this study, the LA productivity can reach to rosiveness to stainless steel reactor in glycerol conversion using CaO
~10 g min−1 L−1, compared to that of 0.005–0.08 g min−1 L−1 achieved solid base is much lower compared to that using homogeneous
using fermentation technology [17] and ~0.3 g min−1 L−1 obtained by NaOH as catalyst.
Kishida et al. [20] using NaOH as homogeneous catalyst at hydrothermal Combination of biodiesel production from soybean oil and the
conditions. However, we have to admit that the process developed in this subsequent crude glycerol conversion to LA, both catalyzed by a CaO
study also produces calcium lactate, similar to the fermentation route. catalyst, suggests that CaO can serve as a single catalyst in the produc-
Further development in downstream processing is needed to avoid the tion of both biodiesel and lactic acid with potential to increase benefits
costly multiple purification steps such as esterification–distillation– for a biodiesel plant, although further development is needed to make
hydrolysis encountered in the fermentation route. CaO catalysts with amphiphilic surfaces.

4. Conclusions Acknowledgments

Different solid bases were tested as potential catalysts for the This work was supported by the United Soybean Board and the U.S.
conversion of glycerol to lactic acid, and we found the inexpensive Department of Agriculture HATCH project No. TEN00428.
L. Chen et al. / Fuel Processing Technology 120 (2014) 40–47 47

References [17] M. Dusselier, et al., Lactic acid as a platform chemical in the biobased economy: the
role of chemocatalysis, Energy Environ. Sci. 6 (5) (2013) 1415–1442.
[1] M.A. Dasari, et al., Low-pressure hydrogenolysis of glycerol to propylene glycol, [18] H. Oh, et al., Lactic acid production through cell-recycle repeated-batch bioreactor,
Appl. Catal. A Gen. 281 (1–2) (2005) 225–231. Appl. Biochem. Biotechnol. 105 (2003) 603–613.
[2] EIA, Biofuels Issues and Trends, U.S.D.o.E. Independent Statistics and Analysis by U.S. [19] S.K. Singh, S.U. Ahmed, A. Pandey, Metabolic engineering approaches for lactic acid
Energy Information Administration (EIA), Oct. 2012, Editor. 2012, Independent Sta- production, Process Biochem. 41 (5) (2006) 991–1000.
tistics and Analysis by U.S. Energy Information Administration (EIA), U.S. Depart- [20] H. Kishida, et al., Conversion of glycerin into lactic acid by alkaline hydrothermal
ment of Energy, Oct. 2012. http://www.eia.gov/biofuels/issuestrends/pdf/bit.pdf. reaction, Chem. Lett. 34 (11) (2005) 1560–1561.
[3] D.T. Johnson, K.A. Taconi, The glycerin glut: options for the value-added conversion [21] Z. Shen, et al., Effect of alkaline catalysts on hydrothermal conversion of glycerin into
of crude glycerol resulting from biodiesel production, Environ. Prog. 26 (4) (2007) lactic acid, Ind. Eng. Chem. Res. 48 (19) (2009) 8920–8925.
338–348. [22] M. Hara, Environmentally benign production of biodiesel using heterogeneous
[4] T. Werpy, G. Petersen, Top value added chemicals from biomass, in: N.R.E.L.N. Pacific catalysts, Chemsuschem 2 (2) (2009) 129–135.
Northwest National Laboratory (PNNL), Office of Biomass Program (EERE) (Eds.), [23] T. Mallat, A. Baiker, Selectivity enhancement in heterogeneous catalysis induced by
Results of Screening for Potential Candidates From Sugars and Synthesis Gas, reaction modifiers, Appl. Catal. A Gen. 200 (1–2) (2000) 3–22.
vol. 1, Pacific Northwest National Laboratory (PNNL), National Renewable Energy [24] M. Kouzu, et al., Active phase of calcium oxide used as solid base catalyst for
Laboratory (NREL), Office of Biomass Program (EERE), 2004. transesterification of soybean oil with refluxing methanol, Appl. Catal. A Gen. 334
[5] M. Balaraju, et al., Catalytic hydrogenolysis of biodiesel derived glycerol to (1–2) (2008) 357–365.
1,2-propanediol over Cu–MgO catalysts, Catal. Sci. Technol. 2 (9) (2012) 1967–1976. [25] M. Kouzu, et al., Heterogeneous catalysis of calcium oxide used for
[6] C.A. Ramirez-Lopez, et al., Synthesis of lactic acid by alkaline hydrothermal conversion transesterification of soybean oil with refluxing methanol, Appl. Catal. A Gen. 355
of glycerol at high glycerol concentration, Ind. Eng. Chem. Res. 49 (2010) 6270–6278. (1–2) (2009) 94–99.
[7] A. Dibenedetto, et al., Converting wastes into added value products: from glycerol to [26] Y. Tang, et al., Highly active CaO for the transesterification to biodiesel, Bull. Chem.
glycerol carbonate, glycidol and epichlorohydrin using environmentally friendly Soc. Ethiop. 25 (1) (2011) 37–42.
synthetic routes, Tetrahedron 67 (6) (2011) 1308–1313. [27] X. Liu, et al., Transesterification of soybean oil to biodiesel using CaO as a solid base
[8] Y. Liu, et al., From glycerol to allyl alcohol: iron oxide catalyzed dehydration and catalyst, Fuel 87 (2) (2008) 216–221.
consecutive hydrogen transfer, Chem. Commun. (Camb.) 46 (8) (2010) 1238–1240. [28] M. Kouzu, Eco-friendly production of biodiesel by utilizing solid base catalysis of
[9] J.R. Ochoa-Gómez, et al., Synthesis of glycerol carbonate from glycerol and dimethyl calcium oxide for reaction to convert vegetable oil into its methyl esters, in: T. Yao
carbonate by transesterification: catalyst screening and reaction optimization, Appl. (Ed.), Zero-Carbon Energy Kyoto 2009, Springer, Japan, 2010, pp. 20–28.
Catal. A Gen. 366 (2) (2009) 315–324. [29] W.L. Xie, H. Peng, L.G. Chen, Transesterification of soybean oil catalyzed by potassi-
[10] Z. Yuan, et al., Biodiesel derived glycerol hydrogenolysis to 1,2-propanediol on um loaded on alumina as a solid-base catalyst, Appl. Catal. A Gen. 300 (1) (2006)
Cu/MgO catalysts, Bioresour. Technol. 101 (18) (2010) 7099–7103. 67–74.
[11] L. Cheng, L. Liu, X.P. Ye, Acrolein Production from Crude Glycerol in Sub- and [30] L. Cheng, et al., Investigation of rapid conversion of switchgrass in subcritical water,
Super-Critical Water, J. Am. Oil Chem. Soc. 90 (4) (2013) 601–610. Fuel Process. Technol. 90 (2) (2009) 301–311.
[12] L. Liu, X.P. Ye, J.J. Bozell, A comparative review of petroleum-based and bio-based [31] H. Kishida, et al., Kinetic study on conversion of glycerin to lactic acid by alkaline
acrolein production, Chemsuschem 5 (7) (2012) 1162–1180. hydrothermal reaction, Kagaku Kogaku Ronbun. 32 (6) (2006) 535–541.
[13] L. Cheng, X.P. Ye, Recent progress in converting biomass to biofuels and renewable [32] K. Fujii, W. Kondo, Calcium glyceroxides formed in the System of calcium
chemicals in sub- or supercritical water, Biofuels 1 (1) (2009) 109–128. oxide-glycerol. With 4 figures, Z. Anorg. Allg. Chem. 359 (5–6) (1968)
[14] L. Cheng, X.P. Ye, A DRIFTS study of catalyzed dehydration of alcohols by 296–304.
alumina-supported heteropoly acid, Catal. Lett. 130 (1–2) (2009) 100–107. [33] A.M. Ruppert, et al., Glycerol etherification over highly active CaO-based materials:
[15] R. Datta, S.P. Tsai, Lactic acid production and potential uses: a technology and new mechanistic aspects and related colloidal particle formation, Chem. Eur. J. 14
economics assessment, Fuels Chem. Biomass 666 (1997) 224–236. (7) (2008) 2016–2024.
[16] R.E. Drumright, P.R. Gruber, D.E. Henton, Polylactic acid technology, Adv. Mater. 12 [34] A. Martin, M. Richter, Oligomerization of glycerol — a critical review, Eur. J. Lipid Sci.
(23) (2000) 1841–1846. Technol. 113 (1) (2011) 100–117.

You might also like