Sunteți pe pagina 1din 13

Subscriber access provided by University of Massachusetts Amherst Libraries

Article
Production of methyl ethyl ketone from biomass
using a hybrid biochemical/catalytic approach
Alisha Multer, Nathan McGraw, Keith Hohn, and Praveen Vadlani
Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/ie3007598 • Publication Date (Web): 31 Aug 2012
Downloaded from http://pubs.acs.org on September 2, 2012

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 12 Industrial & Engineering Chemistry Research

1
2
3
Production of methyl ethyl ketone from biomass using a hybrid
4
5 biochemical/catalytic approach
6
7 Alisha Multer1, Nathan McGraw2, Keith Hohn2, and Praveen Vadlani2
8 (1) St. Edward’s University, Austin, TX 78704 (2) Kansas State University,
9 Manhattan, KS 66506
10
11
12 Abstract
13 The recent demand for sustainable routes to fuels and chemicals has led to an increased
14 amount of research in conversion of natural resources. A potential approach for
15 conversion of biomass to fuels and chemicals is to combine biochemical and chemical
16
processes. This research used microbial fermentation to produce 2,3-butanediol, which
17
18 was then converted to methyl ethyl ketone by dehydration over a solid acid catalyst. The
19 fermentation process was performed using the bacteria Klebsiella oxytoca (K.O). 2,3-
20 butanediol then dehydrated to form methyl ethyl ketone on a solid acid catalyst, the
21 proton form of ZSM-5, and heat. The goal was to determine the reaction kinetics of 2,3-
22 butanediol dehydration over ZSM-5, and to demonstrate the hybrid
23
24
biochemical/thermochemical approach for synthesizing chemicals from biomass. It was
25 found that ZSM-5 produced methyl ethyl ketone with high selectivity (greater than 90%),
26 and could convert fermentative 2,3-butanediol to methyl ethyl ketone. The reaction order
27 of 2,3-butanediol dehydration was found to be slightly large than one, and an activation
28 energy of 32.3 kJ/mol was measured.
29
30
31 Introduction
32 The recent demand for sustainable routes to fuels and chemicals has led to an increased
33 amount of research in conversion of natural resources. Fermentation can convert biomass
34 to a wide variety of products, frequently with high selectivity1. However, the reaction
35 rates are generally slow, and products are general highly dilute. Inorganic catalytic
36
37 routes offer higher reaction rates because higher temperatures can be used; however,
38 product selectivity can be an issue.
39
40 One potential approach for conversion of biomass to fuels and chemicals is to combine
41 biochemical and thermochemical/catalytic processes. With this approach, a process can
42
43
combine a fermentation route that gives high selectivity and productivity to a specific
44 product with an inorganic catalytic process to further upgrade this product. In this
45 contribution, the hybrid approach is applied to the synthesis of methyl ethyl ketone
46 (MEK) from biomass-derived carbohydrates. This process was chosen because methyl
47 ethyl ketone is a valuable chemical, with extensive use as a solvent and in painting and
48 coatings, because fermentation can produce high concentrations of 2,3-butanediol (2,3-
49
50 BD), and because dehydration of 2,3-BD to MEK can be achieved with high selectivity
51 over acid catalysts.
52
53 2,3-BD production is an attractive microbial fermentation because it can produce 2,3-BD
54 at a high concentration2,3. Several bacterial species are known to produce 2,3-BD from
55
56
pyruvate. Jansen et al.4 obtained a butanediol concentration of 12.63 g/l by fermentation
57 of media containing 50 g/l of xylose in a 7-L batch fermentor using K. oxytoca ATCC
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 12

1
2
3
8724. A maximum butanediol productivity of 1.35 g/l/h was obtained with an initial
4
5 xylose concentration of 100 g/l, when different xylose concentrations varying from 5-150
6 g/l were used. Qureshi and Cheryan5 obtained a butanediol yield of 0.5 g/g of glucose
7 with a maximum butanediol concentration of 84.2 g/l using K. oxytoca ATCC 8724. In
8 two- step fed batch fermentation the concentration of butanediol obtained was 85.5 g/l.6
9 obtained a final butanediol concentration of 95.5 g/l with a yield and productivity of 0.48
10
11
g/g of glucose and 1.71 g/l/h using K. oxytoca ME-UD-3 using a two stage agitation
12 speed control strategy under batch fermentation conditions. 2,3-BD synthesis occurs in
13 the late log phase/stationary phase of fermentation3.
14
15 2,3-BD produced via microbial fermentation can be further be converted to MEK by
16
dehydration over an acid catalyst. Such a hybrid approach has been suggested earlier, for
17
18 example by Tran and Chambers7. The chemistry of diol dehydration to a variety of
19 products has been extensively studied8-11, and has been reviewed12. The number of
20 studies looking specifically at 2,3-BD is fewer. Tran and Chambers studied the
21 dehydration of fermentative 2,3-BD over sulfonic acid-containing alumina and silica-
22 alumina catalysts. These researchers reported high conversion and selectivity, but noted
23
24
catalyst deactivation that they attributed to loss of sulfonic acid groups. Emerson and
25 coworkers studied dehydration 2,3-BD to MEK using sulfuric acid and used fermentation
26 broth in a limited number of experiments13. They reported that MEK was produced from
27 fermentation broth, but measured a slower reaction rate. This paper further describes the
28 reaction kinetics of 2,3-BD dehydration, measuring first-order kinetics in 2,3-BD. Bucsi
29
et al. studied the dehydration of diols, including 2,3-BD, over Nafion and NaHX zeolites
30
31 with the primary goal of understanding reaction pathways for producing different
32 products14. Torok and coworkers studied the catalytic dehydration of 2,3-BD and other
33 diols over heteropolyacids15. Manitto and coworkers studied dehydration of 2,3-BD
34 using an enzyme16.
35
36
37
38 While these previous studies provide some information on 2,3-BD dehydration, they are
39 generally lacking in detailed reaction kinetics, particularly for solid acid catalysts. In
40 addition, only a limited number of solid acid catalysts have been studied, some of which
41 were not found to be stable. This paper seeks to extend the body of knowledge on a
42
43
hybrid fermentation/catalytic approach for producing MEK from biomass. Towards this
44 goal, fermentation using Klebsiella oxytoca (K.O) is used to convert sugars to 2,3-
45 butanediol(2,3-BD). 2,3-BD is then catalytically dehydrated over a solid acid catalyst
46 (the proton form of ZSM-5, HZSM-5). The reaction kinetics of 2,3-BD on HZSM-5 are
47 measured, and the activity of HZSM-5 for dehydration of 2,3-BD found in fermentation
48 broth is measured.
49
50
51 Experimental
52
53 Preparation of 2,3-BD by Fermentation
54 Experiments were performed using synthetic glucose for the production of 2,3 BD in a
55
56
7L fermentor. 3500 ml fermentation medium was charged into the vessel and the whole
57 vessel (including pH and DO probes) was sterilized at 1210oC for 30 minutes. A 1000 ml
58
59
60
ACS Paragon Plus Environment
Page 3 of 12 Industrial & Engineering Chemistry Research

1
2
3
glucose solution was sterilized separately and the concentration was adjusted to to give
4
5 5%(w/v) in the fermentor. Klebsiella oxytoca (ATCC# 8724) was inoculated in 500 ml
6 growth media consisting of 10 g L-1 peptone, 5 g L-1 sodium chloride, and 3 g L-1 beef
7 extract. Inoculum was incubated at 37 °C and agitated at 130 rpm for 24 hours. The 3.5
8 liters fermentation media consisted of 2 g L-1 KH2PO4, 13.7 g L-1 K2HPO4, 3.3 g L-1
9 (NH4)2HPO4, 6.6 g L-1 (NH4)2SO4, 0.25 g L-1 MgSO4 · 7H2O, 0.05 g L-1 FeSO4 ·
10
11
7H2O, 0.001 g L-1 ZnSO4 · 7H2O, 0.001 g L-1 MnSO4 · H2O, 0.01 CaCl2 · 2H2O, 0.05
12 g L-1 EDTA. After sterilization, the vessel was cooled, the glucose solution and 500 ml
13 inoculum were added, and the fermentation was initiated. Concentrated sodium
14 hydroxide (3 molar) was used to maintain pH around 5.2-5.3. Samples were collected at
15 scheduled intervals and batch was run for 37 hours. Dry weight data were collected by
16
centrifuging 10 ml samples and collecting pellets. Pellets were then dried at 80 °C and
17
18 weighed.
19
20 Concentrations of sugars and butanediol in the fermentation broth were estimated using
21 HPLC with RID detector and phenomenex monosaccharide column. Degassed MilliQ
22 water was used as the mobile phase with a flow rate at 0.6 ml/min. The column and RID
23
24
temperatures were set at 83oC and 60oC, respectively. Samples were diluted 1:1 with
25 deionized water and passed through a filter before analysis. Samples were measured
26 against standards for glucose and 2,3-butanediol.
27
28 Some fermentation samples were used as the feed for 2,3-BD dehydration. The broth
29
was centrifuged to separate the solid and liquid phases prior to feeding it over the solid
30
31 acid catalyst.
32
33 2,3-BD conversion to MEK
34 Reaction kinetics of the dehydration of 2,3-butanediol to methyl ethyl ketone of HSZM-5
35 were measured in a packed bed reactor. A schematic of the reactor is shown in Figure 1.
36
37 An aqueous solution of 2,3-BD (Fisher Scientific, mixture of racemic and meso, 98%)
38 were pumped to the reactor using a Valco Series M liquid-handling pump. The reactant
39 was fed to a steel reactor where the catalysts (proton form of ZSM-5, HZSM-5, Zeolyst
40 International, Si/Al=50/50) was held in place by quartz wool on top of a wire mesh. The
41 reactor was heated with a furnace that surrounds the steel reactor. Products from the
42
43
reactor exited the reactor and flowed to a three-way valve. This valve directed products
44 either to a sample vial or to a collection flask for waste.
45
46 Prior to reaction, the catalyst was heated in a stream of nitrogen (introduced by a three-
47 way valve) to the reaction temperature. Samples were drawn twenty minutes after the
48 2,3-BD solution was introduced, which was found to be sufficient time to achieve steady
49
50 state. Three samples were collected for each condition studied.
51
52 Samples were diluted with ethanol(one milliliter of ethanol to three milliliters of sample),
53 which serves as initial standard, prior to injection into an SRI 8610 gc. 2,3-BD, MEK,
54 and the only side product (2-methylpropanal) were separated on a Chromosorb 101
55
56
column using a following temperature program was used: the staring temperature
57 was 100oC, and temperature was ramped by 40oC per minute to 190oC, and then
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 12

1
2
3
4
was ramped by 10oC per minute to 230oC where temperature was held for five
5 minutes. With this temperature program, the approximate peak retention times
6 were: ethanol – 5 minutes, 2-methyl propanal – 8.4 minutes, methyl ethyl ketone –
7 9.2 minutes, and 2,3-BD 16 minutes.
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 Figure 1. Schematic of packed bed reactor used to measure reaction kinetics of 2,3-
29
30
butanediol dehydration to methyl ethyl ketone on HZSM-5.
31
32 To measure reaction kinetics, the concentration of 2,3-BD in the feed, the reaction
33 temperature, and the total molar flow rate were varied. Conversion of 2,3-BD was
34 measured for three different concentrations of 2,3-BD (30 g/L, 50 g/L and 80 g/L) and
35
three different flow rates(90 ml/hr, 110 ml/hr, and 130 ml/hr) at 200oC to determine the
36
37 reaction rate and estimate reaction order In addition, the reactor was run with an inlet
38 concentration of 50 g/L at three different temperatures (175°C, 200°C and 225°C). For
39 each condition, three trials were conducted. These trials generally gave 2,3-BD
40 conversions within 10% of each other. The mass balance generally closed within 10%.
41
42
43
Prior to running kinetic trials, the temperature was held constant at 200oC while the total
44 flow rate was increased to determine whether external mass transfer limitations were
45 important. It was found that the conversion increased as flow rate was increased up to a
46 volumetric flowrate of 70 ml/hr. Because the catalyst mass was being held constant, this
47 indicated that external mass transfer limitations were important at flow rates below 70
48
ml/hr. Beyond 70 ml/hr, conversion decreased with increasing flow rate, the expected
49
50 trend. For this reason, all experiments were run at flow rates greater than 70 ml/hr.
51
52 HZSM-5 was pelletized prior to loading into the reactor. Catalyst powder was placed
53 into a Carver press and pressed into a pellet. The pellet was then broken and sieved to
54 give ~0.7mm particles. Approximately 0.25 grams was placed in the reactor for each
55
56
trial.
57
58
59
60
ACS Paragon Plus Environment
Page 5 of 12 Industrial & Engineering Chemistry Research

1
2
3
4
5 Calculation of Conversion and Selectivity
6
7 Conversion of 2,3-BD was calculated from the difference in the 2,3-BD molar flow rate
8 entering the reactor and in the products:
9
10
11 FAo − FA
X = (1)
12 FAo
13
14
15 Two products were detected following 2,3-BD dehydration: methyl ethyl ketone and 2-
16 methyl propanal. The selectivity of MEK was calculated as the ratio of the molar flow
17 rate of MEK divided by the total molar flow rate of products(methyl ethyl ketone and 2-
18 methylpropanal in our studies):
19
20
21 FMEK
S MEK = (2)
22 FMEK + Fprop
23
24
25
26 Data Analysis for Determining Reaction Kinetics
27
28 The mole balance for a packed bed reactor is17:
29
30
dX
FAo = − rA ' FA0 (3)
31 dW
32
33 where: FA0 is the inlet molar flow of the reaction (species A), X is the fractional
34
conversion of A, W is the catalyst weight, and rA’ is the reaction rate of species A per
35
36 gram catalyst.
37
38 This can be rewritten as:
39
40
dX
41 = −rA ' (4)
42 d (W / FAo )
43
44
45
46 Equation 4 shows that the reaction rate in a packed be reactor can be estimated from the
47 slope of X plotted as a function of W/FAo. In our experiments, the inlet molar flow rate
48 was varied for a fixed catalysts weight to change W/FAo. The measured conversion was
49 then plotted as a function of W/FAo so that reaction rate could be estimated.
50
51
52
Reaction rate, -rA, is a function of the concentration of A according to:
53
54 -rA=kCA α (5)
55
56 where: k is the reaction rate constant, CA is the concentration of species A, and alpha is
57
the order of the reaction.
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 12

1
2
3
4
5 The order of reaction was found using the differential method, where reaction rate was
6 plotted versus CA on a log-log plot. Data at three temperatures were fit with the
7 Arrhenius equation to give the activation energy and pre-exponential for the reaction.
8
9
10
11
12 Results and Discussions
13
14
15 Fermentation to 2,3-Butanediol
16
Figure 2 shows the profiles of glucose utilization, biomass (cell dry weight) and
17
18 butanediol production with time. Based on the experiments performed at the 7 L
19 fermentor level, we successfully repeated the results obtained at shake flask level (results
20 not shown). 2,3-butanediol yield from total glucose consumed is about 0.63 g/g. Further,
21 the initial batch fermentation was completed by 37 hours resulting in 1.1 g/l/hr
22 productivity (whereas it took 48 hours to attain the same value in shake flask studies).
23
24
25
26
27 70 3.5
28
Solulution Concentration

60 3

KO Concentration (g/L)
29
30 50 KO 2.5
31
32 40 2
(g/L)

33
34 30 1.5
2,3-Butanediol
35 20 1
36
37 10 Glucose 0.5
38
39 0 0
40 0 10 20 30 40
41
Time (hours)
42
43
44 Figure 2. Fermentation profile of 2,3 butanediol synthesis using Klebsiella oxytoca
45
46
47
48
2,3-Butanediol Dehydration Kinetics
49 Figure 3 shows the measured conversion of 2,3-BD as a function of W/FAo for three
50 different inlet concentrations. As expected, all three sets of data show a linear trend,
51 from which we can estimate reaction rate for the slope of the line.
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 7 of 12 Industrial & Engineering Chemistry Research

1
2
3
4 1
5 y = 1.89E-04x - 1.32E+00
0.9 2 y = 1.18E-04x - 1.30E+00
6 R = 9.97E-01 2
0.8 R = 9.95E-01
7
8 0.7
50 g/L
Conversion
Fractional

9 0.6
10 70 g/L 30 g/L
0.5
11
12 0.4
13 y = 3.96E-05x - 4.51E-01
0.3 2
14 R = 9.69E-01
15 0.2
16 0.1
17 0
18
0 5000 10000 15000 20000 25000 30000 35000
19
20 W/Fao (g/mol/sec)
21
22 Figure 3. Reaction kinetics for 2,3-butanediol dehydration to methyl ethyl ketone.
23 2,3-butanediol is plotted as a function of the ratio of catalyst weight to inlet molar
24 flow rate. The slope of each line gives the reaction rate.
25
26
27
28 The reaction order can be estimated by plotting reaction rate as a function of
29 concentration on a log-log plot. Because the data were obtained under conditions that
30 conversion was large, the concentration could not be assumed to be that of the inlet
31
concentration. Instead, the concentration in the packed bed was changing. We estimated
32
33 the concentration for each reaction rate from the conversion (i.e. CA=CAo(1-X)) measured
34 for the middle set of data points.
35
36 Figure 4 shows the log-log plot of reaction rate for different 2,3-BD concentrations. As
37 seen in this plot, a linear trend is noted with a slope of 1.226. This indicates a reaction
38
39 order of 1.226.
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 12

1
2
3
4 0
5 -2.5 -2 -1.5 -1 -0.5 0
6
-1
7
8
9 -2
10
ln (-rA)

11 -3
12 y = 1.226x - 3.0183
13 R2 = 0.9787
-4
14
15
-5
16
17
18 -6
19 ln (CA)
20
21
Figure 4. Log-log plot of reaction rate versus concentration of 2,3-butanediol. The
22
23 slope of the line gives the reaction order.
24
25 To determine the Arrhenius constants, the reaction rate constant was calculated at three
26 temperatures from the measure reaction rates by dividing reaction rate by 2,3-BD
27 concentration. In doing these, we assumed a first-order reaction, even though Figure 4
28
29
suggested a slightly higher reaction order. Figure 5 is an Arrhenius plot of the reaction
30 rate constant for 2,3-BD dehydration. This plot shows a slope of -3889.4, which
31 corresponds to an activation energy of 32.34 kJ/mol. The pre-exponential was found to
32 be 8.95 sec-1.
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 Figure 5. Arrhenius plot for 2,3-butanediol dehydration
59
60
ACS Paragon Plus Environment
Page 9 of 12 Industrial & Engineering Chemistry Research

1
2
3
4
5
6 Product Selectivity
7 Methyl ethyl ketone was always the major product from 2,3-BD dehydration, with
8 selectivity typically greater than 90%. Selectivity was nearly constant for the conditions
9 studied. Figure 6 shows the selectivity to MEK and 2-methylpropanal as a function of
10
11
temperature. As seen in this figure, the selectivities are essentially unchanged as
12 temperature is increased.
13
14
15 1
16 0.9
17 0.8 MEK
18
0.7
Selecitivity (%)

19
20 0.6
21 0.5
22 0.4
23
0.3
24
25 0.2 2-methylpropanal
26 0.1
27 0
28 440 450 460 470 480 490 500 510
29
30 Temperature (K)
31
32 Figure 6. Selectivity in 2,3-butanediol dehydration as a function of temperature.
33
34
35
36 Dehydration of 2,3-Butanediol Produced from Fermentation
37 The activity of ZSM-5 for dehydration of fermentative 2,3-BD was tested by pumping
38 fermentation broth through the catalyst bed. The concentration of the broth was
39 estimated using the gas chromatograph to be consistent with the data analysis used to
40 analyze the product. From this measure, the 2,3-BD concentration was only 2.63 g/L
41
42
(0.03 mol/L). This was much lower than can generally be formed via fermentation (see
43 Figure 2). This fermentation broth was sent through the reactor at 200°C at 130 mL/hr.
44 A conversion of 13% was noted after operating the reactor for one hour.
45
46 The long-term stability of the catalytic process operating with fermentation-derived 2,3-
47 BD was not assessed. Catalyst deactivation has been described by others when
48
49 fermentation broth is used7, so this deserves further study. We have found no evidence of
50 catalyst deactivation using pure 2,3-BD streams over 40 hours.
51
52 Comparison of Results to Literature
53 Emerson and coworkers measured reaction kinetics for 2,3-BD dehydration with sulfuric
54
55
acid13. They found that the reaction was first order in 2,3-BD. Neish and coworkers, on
56 the other hand, reported that 2,3-BD dehydration was 2nd order when run in the absence
57 of water18. Our results found a reaction order slightly above 1, suggesting agreement
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 12

1
2
3
with Emerson and coworkers. The activation energy measured in this paper (32.3
4
5 kJ/mol) was smaller than that reported by Emerson and coworkers (150.7 kJ/mol), most
6 likely due to difference between the heterogeneous catalyst used in this work and the
7 homogeneous acid used in their study.
8
9 Selectivity to MEK was over 90% in this work, in agreement with that reported by
10
11
Bucsi14 on supported heteropoly acids. Torok also reported high selectivity to MEK on
12 Nafion and NaHX zeolite15, though they reported dioxolanes as important side products
13 and reported MEK selectivities lower than we found(less than 80%). Using sulfuric acid,
14 Emerson reported MEK yields of over 90%, where their yield is equivalent to selectivity
15 in this paper. They suggested that aldol condensation limited MEK selectivity. These
16
results suggest that HZSM-5 is as selective or more selective than other reported
17
18 catalysts.
19
20
21 Conclusions
22 HZSM-5 is an active and selective catalyst for dehydration of 2,3-butanediol, and can be
23
24
used to dehydrate 2,3-butanediol produced through fermentation. Over 90% selectivity
25 to methyl ethyl ketone was produced on HZSM-5, with 2-methyl propanal as the only
26 other product detected. Dehydration kinetics were found to depend on 2,3-butanediol
27 concentration with a slightly higher reaction order than one (1.22). The activation energy
28 was estimated to be 32.3 kJ/mol.
29
30
31
32 Acknowledgements
33 Support for Alisha Multer from an NSF REU site grant (NSF EEC#0851799) is
34 gratefully acknowledged.
35
36
37 Dedication
38 This submission is dedicated to L.T. Fan, Distinguished Professor of Chemical
39 Engineering at Kansas State University. L.T.’s contributions in the field of biochemical
40 and reaction engineering are an inspiration, and the authors feel fortunate to have had the
41 opportunity to work with and learn from L.T.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
Page 11 of 12 Industrial & Engineering Chemistry Research

1
2
3
References
4
5
6 1. Lee, J.W.; Kim, H.U.; Choi, S.; Yi, J.; Lee, S.Y.; Microbial production of
7 building block chemicals and polymers. Curr. Opinion in Biotech. 2011, 22, 758.
8 2. Syu, M.J. Biological production of 2,3-butanediol. Appl. Microbiol. Biotechnol.
9 2001, 55, 10.
10
11
3. Celinska, E.; Grajek, W. Biotechnological production of 2,3-butanediol-Current
12 state and prospects. Biotechnol. Adv. 2009, 27, 715.
13 4. Jansen, N.B.; Flickinger, M.C.; Tsao, G.T. Production of 2,3-Butanediol from D-
14 Xylose by Klebsiella-Oxytoca Atcc 8724. Biotechnol. Bioeng. 1984, 26, 362.
15 5. Qureshi, N.; Cheryan, M. Effects of Aeration on 2,3-Butanediol Production from
16
Glucose by Klebsiella-Oxytoca J. Ferment. Bioeng. 1989, 67, 415.
17
18 6. Xiao, Z.; Ma, C.; Xu, P.; Lu, J.R. Acetoin Catabolism and Acetylbutanediol
19 Formation by Bacillus pumilus in a Chemically Defined Medium. Plos One 2009
20 4, e5627.
21 7. Tran, A. V.; Chambers, R. P., The dehydration of fermentative 2,3-butanediol into
22 methyl ethyl ketone. Biotechnol. Bioeng. 1987, 29, 343.
23
24
8. Molnar, Arpad; Felfoldi, K.; Bartok, M. Studies on the conversion of diols and
25 cyclic ethers -49: Stereochemistry of cyclodehydration of 1,4-diols on the action
26 of bronsted and lewis acids: a comprehensive study. Tetrahedron 1981, 37, 2149.
27 9. Molnar, A.; Bartok, M. Studies on the conversion of diols and cyclic ethers: I.
28 Investigation of part-steps in the dehydration and fragmentation of 1,3-diols on a
29
copper catalyst. J. Catal. 1981, 72, 322.
30
31 10. Sato, S.; Takahashi, R.; Sodesawa, T.; Honda, N. Dehydration of diols catalyzed
32 by CeO2. J. Mol. Catal. A: Chem. 2004, 221, 177.
33 11. Gotoh, H.; Yamado, Y.; Sato, S. Dehydration of 1,3 butanediol over rare earth
34 oxides. Appl. Catal. A.: Gen. 2010, 377, 92.
35 12. Barok, M.; Molnar, A. Dehydration of diols. The chemistry of functional groups.
36
37 In Ethers, Crown Ethers, Hydroxyl Groups and their Sulphur Analogues: Volume
38 2; S. Patai, Ed.; J. Wiley and Sons: Chichester, 1981; p. 721.
39 13. Emerson, R. R.; Flickinger, M. C.; Tsao, G. T., Kinetics of dehydration of
40 aqueous 2,3-butanediol to methyl ethyl ketone. Ind. Eng. Chem. Prod. Res. Dev.
41 1982, 21, 473.
42
43
14. Bucsi, I.; Molnár, Á.; Bartók, M.; Olah, G., Transformation of 1,2-diols over
44 perfluorinated resinsulfonic acids (nafion-H). Tetrahedron 1994, 50, 8195.
45 15. Torok, B.; Bucsi, I.; Beregszaszi, T.; Kaposci, I.; Molnar, A., Transformation of
46 diols in the presence of heteropoly acids under homogeneous and heterogeneous
47 conditions, J. Mol. Catal. A: Chem. 1996, 107, 305.
48
16. Manitto, P.; Speranza, G.; Fontana, G.; Galli, A., Stereochemistry and Fate of
49
50 Hydrogen Atoms in the Diol-Dehydratase-Catalyzed Dehydration of meso-
51 Butane-2,3-diol, Helv. Chim. Acta 2005, 81, 2005.
52 17. Fogler, H.S. Elements of Chemical Reaction Engineering, 4th Edition; Prentice
53 Hall: 2005.
54 18. Neish, A.C.; Haskell, V.C.; MacDonald, F.J. Production and Properties of 2,3-
55
56
Butanediol. 6. Dehydartion by Sulphuric Acid. Can. J. Res. 1945, 23B, 281.
57
58
59
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 12

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment

S-ar putea să vă placă și