Biobased Methacrylic Acid via Selective Catalytic ... - ACS Publications

Feb 23, 2017 - ABSTRACT: We report a biobased route to methacrylic acid via selective decarboxylation of itaconic acid utilizing catalytic ruthenium c...
0 downloads 0 Views 699KB Size
Subscriber access provided by University of Newcastle, Australia

Article

Bio-Based Methacrylic Acid via Selective Catalytic Decarboxylation of Itaconic Acid James C Lansing, Rex E. Murray, and Bryan R. Moser ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.6b02926 • Publication Date (Web): 23 Feb 2017 Downloaded from http://pubs.acs.org on March 1, 2017

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.

ACS Sustainable Chemistry & Engineering 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 33

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 Sustainable Chemistry & Engineering

Bio-Based Methacrylic Acid via Selective Catalytic Decarboxylation of Itaconic Acid James C. Lansing,1,2 Rex E. Murray1 and Bryan R. Moser1,*

1

United States Department of Agriculture Agricultural Research Service

National Center for Agricultural Utilization Research Bio-Oils Research Unit 1815 N. University St., Peoria, Illinois 61604, USA

2

United States Department of Energy

Oak Ridge Institute for Science and Education 1299 Bethel Valley Rd., Oak Ridge, Tennessee 37830, USA

[email protected]

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

1

ABSTRACT

2

We report a bio-based route to methacrylic acid via selective decarboxylation of itaconic acid

3

utilizing catalytic ruthenium carbonyl propionate in an aqueous solvent system. High selectivity

4

(> 90%) was achieved at low catalyst loading (0.1 mol %) with high substrate concentration (5.5

5

M) at low temperature (200 – 225 oC) and pressure (≤ 425 psig) relative to previous

6

contributions in this area. Direct decarboxylation of itaconic acid was achieved as opposed to

7

the conjugate base reported previously, thereby avoiding basification and acidification steps.

8

Also investigated was catalytic manganese (II) oxalate (5 mol %), but low yield (4.8%) and

9

evolution of carbon monoxide via oxalate decomposition was problematic. Attempts at

10

stabilization of the catalyst with triphenylphosphine were unsuccessful, but it exhibited greater

11

catalytic efficacy (14.0% yield) than the manganese catalyst (4.8% yield) at 5 mol %. Neither

12

carbon monoxide nor propylene (excessive decarboxylation) were detected during ruthenium-

13

catalyzed decarboxylation. In addition, co-solvents such as tetraglyme lowered vapor pressures

14

within the reaction vessel by > 100 psig while minimizing decomposition of starting acids. In

15

combination, these findings represent improvements over existing methodologies that may

16

facilitate sustainable production of methacrylic acid, an important petrochemically-based

17

monomer for the plastics industry.

18 19

KEYWORDS: Decarboxylation, Itaconic acid, Methacrylic acid, Ruthenium, Tetraglyme

20 21 22 23

2 ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

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 Sustainable Chemistry & Engineering

24

INTRODUCTION

25

Renewable materials are increasingly preferred over crude petroleum oil and its refinery products

26

as feedstocks for industrial chemicals.1 Various environmental, social, economic, legislative,

27

and geopolitical incentives motivate the commercial transition to bio-based chemicals.

28

Renewable alternatives are either direct drop-in replacements or structurally different but with

29

similar properties and performance. Direct drop-in replacements are preferred, as industry is

30

already familiar with and has accepted their structure, properties and performance. Examples

31

include bio-based ethylene and butadiene from ethanol, ethanol and succinic acid from glucose,

32

propylene, propylene glycol, syngas and epichlorohydrin from glycerol, diesel and jet fuel-range

33

hydrocarbons from plant oils, and aromatics from depolymerization of lignin.2-7

34

Methacrylic acid (2-methylpropenoic acid; MAA) and methyl methacrylate (MMA) are

35

important commodity monomers for numerous industrially significant plastics.8 The principal

36

application of MMA is homopolymerization to provide the lightweight, shatter-resistant,

37

thermoplastic poly(methyl methacrylate) (Plexiglas®) (PMMA) as a versatile alternative to glass.

38

MMA is also an essential component of copolymers found in surface coatings, paints, adhesives,

39

and emulsion polymers.8-10

40

The most significant petrochemical route to MAA/MMA is the acetone cyanohydrin

41

process (ACH) depicted in Figure 1. Acetone is initially reacted with hydrogen cyanide to form

42

an acetone cyanohydrin intermediate, which is then converted to methacrylamide sulfate upon

43

treatment with stoichiometric sulfuric acid at 140 oC. Production of MAA or MMA is achieved

44

by reaction of the sulfate with water or anhydrous methanol, respectively.11 Along with

45

production of MAA/MMA, an excess of ammonium bisulfate is obtained in a molar ratio of

46

1.5:1.8,9 Disadvantages of the ACH process include utilization of nonrenewable materials and

3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

47

stoichiometric amounts of harmful, toxic and corrosive reagents, generation of toxic

48

intermediates, and production of low-value ammonium bisulfate, which must be further reacted

49

with ammonia to yield ammonium sulfate for use as a fertilizer. Additional petrochemical routes

50

to MAA/MMA are reviewed elsewhere.8,9

51

Relatively few bio-based routes to MAA are reported. One such process is dehydration

52

and decarboxylation of citramalic (2-hydroxy-2-methylbutanedioc) acid at elevated temperatures

53

(250 – 400 oC) and pressures (450 – 4,000 psi; 31 – 276 bar) in the presence of catalytic sodium

54

hydroxide (NaOH).12 Application of this methodology to maleic (2Z-butenedioc) acid yields

55

acrylic (2-propenoic) acid (AA).12 The Alpha Process relies on carbon monoxide, methanol and

56

ethylene to yield MMA following a two stage sequence.13,14 Another methodology entails

57

dehydration and decarboxylation of citric acid to itaconic (2-methylidenebutanedioc) acid (IA),

58

followed by a second decarboxylation to afford MAA in the presence of stoichiometric bases

59

under near-critical and supercritical water conditions.15 However, accumulation of byproducts

60

such as 2-hydroxybutyric and crotonic [(E)-2-butenoic] acids (CA) at the expense of MAA was

61

problematic. In addition to low selectivity, high temperatures (> 350 oC) were required. In

62

another approach, selectivity was improved to over 90%, but high temperatures (245 – 270 oC)

63

and pressures (450 – 3000 psi; 31 – 207 bar) were needed, stoichiometric bases were used, and

64

byproducts such as CA, 2-hydroxybutyric acid and propylene were observed.16 Propylene arises

65

via decarboxylation of MAA and/or CA, thereby reducing the yield of the intended product.

66

More recently, either IA or citric acid was selectively decarboxylated using catalytic

67

palladium, platinum and ruthenium.17 Higher selectivity was achieved (up to 84%) and lower

68

reaction temperatures (200 – 250 oC) and pressures (550 psi; 38 bar) were utilized relative to

69

previous contributions (Figure 2). However, similar to previous reports, stoichiometric bases

4 ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

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 Sustainable Chemistry & Engineering

70

such as NaOH were utilized during the reaction. In such embodiments, NaOH reacts with the

71

starting acid to form a conjugate base, which is then decarboxylated to sodium methacrylate.

72

The rationale behind formation of monosodium salts is that the rate of decarboxylation at 280 –

73

330 oC and 4,000 psi (276 bar) is enhanced relative to the free acid and the disodium salt.18

74

However, subsequent acidification of the sodium salt is required to generate MAA, thereby

75

adding basification and acidification steps to the process (Figure 3). In essence, these

76

approaches convert the conjugate base of IA to sodium methacrylate, rather than IA directly to

77

MAA. Consequently, a sustainable technology is needed that avoids basification and

78

acidification steps through direct decarboxylation of the starting acid to MAA (Figure 3).

79

The objective of our study was to develop a sustainable route to MAA that is free from

80

the technical difficulties discussed above. Presented in Table 1 is a comparison of existing bio-

81

based routes to the present work. IA was chosen as the starting material since it is derived from

82

dehydration and decarboxylation of citric acid or from fermentation of simple sugars such as

83

glucose (Figure 4).19-21 Specific objectives included identifying catalysts efficient at direct

84

decarboxylation of IA to yield MAA, increasing throughput by increasing the concentration of

85

IA, and lowering reaction parameters such as temperature and pressure. In combination, these

86

changes would represent techno-economic improvements over existing methodologies, thus

87

yielding a sustainable route to MAA that is more amenable to commercialization.

88 89 90

EXPERIMENTAL SECTION Materials. Water for decarboxylation reactions and as a component of the mobile phase

91

for HPLC was ultra-pure (18 MΩ) obtained from a Barnstead (Lake Balboa, CA) Easy Pure II

92

RF/UV ultrapure water system. Prior to decarboxylation, ultra-pure water was degassed by a

5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

93

nitrogen sparge. CA (98%), IA (99%) and manganese (II) oxalate dihydrate (MnC2O4⋅2H2O;

94

99%) were purchased from Alfa Aesar (Ward Hill, MA). CCA (98%), MAA (99%), MSA

95

(98%), triphenylphosphine (PPh3; 99%), 4-Methoxyphenol (MeHQ; 99%), and trifluoroacetic

96

acid (CF3CO2H, TFA, 99%) were obtained from Sigma-Aldrich Corp (St. Louis, MO).

97

Ruthenium (I) dicarbonyl propionate was prepared according to literature precedent.22 All other

98

materials were obtained from Sigma-Aldrich Corp and used as received. Catalysts and IA were

99

stored prior to use in a nitrogen filled Innovative Technologies (Amesbury, MA) model IL-2GB

100 101

inert atmosphere glove-box with oxygen and water concentrations kept below 1 ppm. Catalytic Decarboxylation. Decarboxylations were performed using a Parr Instrument

102

Company (Moline, IL) 50 mL stainless steel (T316) reactor with a maximum allowable working

103

pressure of 3000 psi (207 bar) at 350 oC. The reactor was lined with a glass liner and equipped

104

with an internal thermocouple, cooling loop, variable speed overhead magnetic stirrer with a

105

standard shaft and impeller, heater assembly, and a Span (Waukesha, WI) overhead pressure

106

gauge rated up to 3,000 psi (207 bar). Internal reaction temperature and stirring rate (rpm) were

107

controlled by a Parr model 4848 reactor controller unit, which also provided internal vessel

108

pressure information.

109

In a typical experiment, IA (10.0 g; 76.9 mmol; 5.5 M in water), manganese (II) oxalate

110

(5 mol %) or ruthenium dicarbonyl propionate [(Ru(CO)2(CH3CH2COO)]n (0.1 mol %) and

111

MEHQ (5000 ppm) were weighed in the inert atmosphere glove box and placed in a glass liner.

112

A rubber septum was affixed to the glass liner, which was then removed from the glove box.

113

Ultra-pure and degassed water (14.0 mL) was added to the closed system using a syringe, and the

114

septum was removed before the liner was introduced into the reaction vessel. After the reactor

115

was closed, a headspace pressure of 400 ± 5 psig N2 was applied and the mixture stirred (500

6 ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

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 Sustainable Chemistry & Engineering

116

rpm) and heated to the desired reaction temperature. Typically, heating to the reaction

117

temperature was achieved in 20 min. The reaction time started when the desired temperature

118

was reached. Parameters such as time and temperature were varied to determine the optimum

119

conditions. In some instances, triphenylphosphine (PPh3) (0.5 – 5 mol %) was added to the

120

reaction mixture. After the allocated reaction time, the reactor was equilibrated to room

121

temperature (1 – 1.5 h) and the gas pressure was released. MAA was isolated via azeotropic

122

distillation and yield was reported as the percentage recovered of theoretical (6.63 g).

123

High Performance Liquid Chromatography. Samples (50 µL) collected in pairs were

124

diluted to a volume of 1.5 mL of 1:1 v/v methanol/water and analyzed in duplicate by HPLC.

125

Levulinic acid (1.0 M) was added to all samples as an internal standard. Analyses were

126

performed using an Agilent (Santa Clara, CA) 1100 series degasser, pump and auto-sampler

127

connected to a spectrophotometric ultraviolet diode array detector and a Sedere (Alfortville,

128

France) Sedex 85 LT-ELSD in series. Samples (5.0 µL) were injected onto a GL Sciences

129

(Tokyo, Japan) Inertsil ODS-3 column (5.0 µm; 4.6 mm × 250 mm). The mobile phase

130

consisted of a mixture of methanol/water with a flow rate of 1.0 mL/min. A gradient method

131

was employed with the following conditions: 85% (v/v) water initially, changing to 82% by 10

132

min, 78% at 11 min, 75% at 12 min, 80% at 49 min, 83% at 51 min, 85% at 53 min, hold at 85%

133

until 55 min. The water phase was spiked with 0.25% v/v TFA. Compounds were identified by

134

retention time comparison to reference standards using both ultraviolet detection at a wavelength

135

of 254 nm and the signal from the ELSD. Individual compounds were quantified by external

136

calibration with pure standards. Approximate retention times were: levulinic acid (6.6 min),

137

citraconic acid (9.4 min), itaconic acid (10.0 min), mesaconic acid (12.0 min), methacrylic acid

138

(19.6 min), and trans-crotonic acid (16.0 min).

7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

139

Micro Gas Chromatography. Gas analysis was performed in triplicate using a model

140

490 Micro GC by Agilent operating at a sample line temperature of 30 oC with helium as the

141

carrier gas. Gas samples were collected from the Parr reactor using a 3 oz. pressure reaction

142

bottle from Lab Crest-Andrews Glass (Vineland, NJ) and infused at 10 – 20 psi (0.7 – 1.4 bar)

143

into the instrument through a model 170 Lab Series membrane separator by Genie Filters (A+

144

Corporation, Gonzalez, LA). Analyses were isothermal with a duration of 3 min. Channel 1

145

(Cox 1 m) was used for characterization and its temperature and operating pressure was 80 o C

146

and 29 psi (2 bar). Compounds were identified by retention time comparison to reference

147

standards and approximate retention times were: nitrogen (0.37 min), carbon monoxide (0.47

148

min), carbon dioxide (2.2 min), and 1.0 min for hydrocarbons such as methane, ethane, propane,

149

and butane.

150

Spectroscopy. 1H-nuclear magnetic resonance (NMR) spectra were collected at 26.9 °C

151

using a Bruker Avance-500 spectrometer (Billerica, MA) operating at 500 MHz using a 5-mm

152

BBO probe. Chemical shifts (δ) are reported as parts per million (ppm) from tetramethylsilane

153

in CDCl3 (Cambridge Isotope Laboratories, Andover, MA).

154

Vapor pressure of solvent systems. In a typical experiment, the solvent (14 mL) of

155

interest (water; 1:1 mole ratio of tetraglyme to water, and tetraglyme) was introduced into the

156

reactor with a syringe. No other reactants or catalysts were added. The Parr reactor was

157

pressurized with nitrogen to ensure no leaks were present. The extra pressure was then relieved

158

such that the initial pressure of the reactor was 0 psig. The mixture was then stirred (500 rpm)

159

and heated (20 min) to the desired temperature. The recorded pressure (psig) was measured after

160

equilibration (40 min) at the desired temperature. Pressure was measured at 100, 150, 175, 200,

161

212, 225, and 250 oC (Figure 5).

8 ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33

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

162

ACS Sustainable Chemistry & Engineering

RESULTS AND DISCUSSION Catalyst Screening. Numerous metal catalysts were screened for production of MAA,

163 164

with manganese (II) oxalate and ruthenium (I) dicarbonyl propionate identified as the most

165

efficient at selective decarboxylation of a 1.0 M solution of IA at 250 oC for 3 h. Other catalysts

166

screened included simple metal salts of chromium, cobalt, copper, iron, manganese,

167

molybdenum, nickel, and zinc, but they provided lower yields of MAA (data provided in Table

168

S1) and were thus not explored further. Optimization using manganese (II) oxalate. Once the most effective catalyst was

169 170

identified, we modified reaction conditions in an effort to maximize yield. Lower reaction

171

temperatures, pressures and catalyst loadings were of interest, along short reaction times and

172

enhanced selectivity. Another objective was to increase the concentration of IA for process

173

economic and engineering reasons. A significant consequence of low starting material

174

concentration is that the majority of the reaction vessel is occupied by water, thereby resulting in

175

high internal pressure due to the high vapor pressure of water (see Figure S1). Table 2 depicts initial attempts at optimization using manganese (II) oxalate at 5.0 mol

176 177

%, which was half of the amount used during screening studies. In addition, the concentration of

178

IA was increased to 5.5 M. A headspace pressure of 400 psig nitrogen (at room temperature)

179

was added in an effort to keep MAA in the liquid phase. We hypothesized that MAA in the

180

vapor phase might condense in the upper reaches of the reactor and undergo polymerization. As seen in Table 2, optimization reactions conducted at lower temperatures (200 – 225

181 182

o

183

MAA in excess of 7.2%. In addition, carbon monoxide was detected among the headspace gases

184

(Figure 6a), along with nitrogen and carbon dioxide. Carbon dioxide is the expected gaseous

C) and with shorter reaction times (1.5 – 2.0 h) than during catalyst screening failed to give

9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

185

coproduct from decarboxylation of IA. We suspected that carbon monoxide arose from

186

decarbonylation of oxalic acid via dissociation of manganese (II) oxalate under our conditions,

187

thereby leading to decarbonylation of oxalic acid to carbon monoxide and other species. This

188

was confirmed by conducting an experiment in the presence of 10 equivalents of oxalic acid for

189

every equivalent of manganese (II) oxalate with otherwise similar conditions. The headspace

190

gases from this experiment displayed a spike in carbon monoxide (Figure 6c), thus indicating

191

that carbon monoxide was evolving from decomposition of oxalic acid.

192

Reaction temperature affected ∆P and yield of MAA. As seen in Table 2, ∆P and yield of

193

MAA increased with increasing temperature at constant reaction time (1.5 h) and with longer

194

reaction times at constant temperature (200 oC). For example, entries 2-5 represent

195

decarboxylations conducted at progressively higher temperatures (200 – 250 oC) and otherwise

196

similar reaction conditions. Both ∆P and yield of MAA increased as the temperature was

197

increased from 200 to 250 oC.

198

In an effort to stabilize manganese (II) oxalate, an equimolar amount of PPh3 (5.0 mol %)

199

was added to the reaction mixture (entries 6-8, Table 2). The rationale behind addition of

200

phosphines originated with Crooks, et al., who noted that ruthenium carbonyl carboxylate

201

polymers form more stable ruthenium dimers capped by phosphines upon addition of

202

tributylphospshine and pyridine.22 He proposed that these dimers were the active catalyst species

203

in the decarboxylation of organic acids. We hypothesized that phosphines would form stable

204

dimeric species with manganese upon addition to our system. Addition of PPh3 enhanced the

205

yield of MAA. For instance, comparison of decarboxylation at 212 oC for 1.5 h with (entry 6)

206

and without (entry 3) PPh3 (5.0 mol %) in the presence of manganese (II) oxalate (5.0 mol %)

207

revealed higher ∆P (196 versus 135 psig) and yield of MAA (5.3 versus 2.4%) with PPh3.

10 ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33

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 Sustainable Chemistry & Engineering

208

Furthermore, higher ∆P (326 versus 221 psig) and yield of MAA (13.0 versus 7.0%) was noted

209

with PPh3 (entry 7) at 225 oC than without added PPh3 but otherwise similar reaction conditions

210

(entry 4). Lastly, PPh3 alone (entry 8) demonstrated greater efficacy at decarboxylation of IA

211

than did manganese (II) oxalate (entry 4) at similar reaction conditions (225 oC and 1.5 h), as

212

indicated by higher ∆P (282 versus 221 psig) and yield of MAA (14.0 versus 7.0%). The control

213

experiment conducted in the absence of catalyst at 225 oC for 1.5 h (entry 1) gave MAA in 2.6%

214

yield, which was lower than catalyzed reactions conducted utilizing otherwise similar conditions

215

(entries 4, 5, 7, and 8). However, suppression of carbon monoxide was unsuccessful with PPh3

216

(Figure 6b).

217

Lower partial vapor pressures (pH2O; 191 – 499 psig) were achieved in Table 2 than

218

during catalyst screening experiments due to lower temperatures investigated as well as higher

219

concentrations of IA in water. For comparison, a pH2O of 540 psig was calculated according to

220

Raoult’s Law with a reaction temperature of 250 oC and a concentration of IA of 1.0 M (Table

221

S1). Increasing the concentration of IA to 5.0 M with otherwise similar conditions resulted in a

222

pH2O of 499 psig (entry 5, Table 2). Further reductions in reaction temperature afforded

223

progressively lower pH2O at constant IA concentration (Table 2).

224

Optimization with ruthenium catalysis. Due to the relatively low yield of MAA given

225

by manganese (II) oxalate along with evolution of carbon monoxide, ruthenium catalysis was

226

explored. Ruthenium carbonyl carboxylates were of interest because of our previous work on

227

tandem isomerization-decarboxylation of oleic acid.23-25 When using catalytic ruthenium (I)

228

dicarbonyl propionate in the present system, IA was first isomerized to mesaconic acid, which

229

was in turn decarboxylated to MAA (Figure 2). Isomerization was confirmed by a control

230

experiment subjecting dibutyl itaconate to the reaction conditions, thereby inhibiting

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

231

decarboxylation via protection of the carboxylic acid moieties as butyl esters. When the reaction

232

was performed at 225 °C for only 15 minutes, dibutyl mesaconate was present in 44% with

233

dibutyl itaconate constituting most of the remaining product (see Figures S2 and S3). As

234

expected, a negligible amount of citraconic dibutyl ester was detected, as it is the cis isomer of

235

mesaconate and is therefore less thermodynamically favored. Isomerization was anticipated

236

because extensive isomerization of double bond location and geometry was observed in our

237

previous studies using oleic acid as the substrate.23-25 Under the same conditions with

238

manganese (II) oxalate, less than 5% isomerization to mesaconic dibutyl ester was observed.

239

Catalytic ruthenium on carbon (Ru/C) was previously investigated for decarboxylation of IA, but

240

the authors concluded that the catalyst was inferior to Pd/C and Pt/C due to higher propylene

241

production via unwanted decarboxylation of MAA.17 However, we did not observe propylene

242

during the course of our study (Figure 7).

243

Depicted in Table 3 is catalytic decarboxylation of a 5.5 M solution of IA in water using

244

ruthenium dicarbonyl propionate (0.1 mol % Ru) at 200 – 225 oC and 0.5 – 2.0 h with a

245

headspace pressure of 400 psig nitrogen. Yields of MAA ranged from 8.2 – 33.8%, which were

246

higher than those observed in Table 2 despite the lower catalyst loading, generally shorter

247

reaction times and lower reaction temperatures explored in Table 3. Similar to manganese-

248

catalyzed decarboxylations, the yield of MAA increased when reaction time was increased from

249

0.5 to 1.0 h at constant temperature (212 oC), as seen by comparison of entries 1 and 2.

250

However, a further increase in reaction time to 2.0 h (entry 3) at otherwise similar conditions did

251

not provide higher yield relative to a reaction time of 1.0 h. . In addition, ∆P increased with

252

increasing reaction time at constant temperature (entries 1 – 3, Table 3), despite yield plateauing

253

at 1.5 h. These results indicate that headspace pressure does not necessarily correlate with yield

12 ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

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 Sustainable Chemistry & Engineering

254

of MAA. Entries 7 – 10 (Table 3) were conducted in the presence of 0.5 mol % PPh3 in an

255

effort to stabilize the ruthenium catalyst. A higher yield of MAA was achieved at 200 oC and 1.5

256

h with 0.5 mol % PPh3 added to the system (14.9%; entry 7) versus without (8.6%; entry 4) and

257

otherwise similar conditions. In addition, a higher ∆P was observed without PPh3 addition (120

258

versus 116 psig) despite the lower yield, thus providing further evidence that headspace pressure

259

does not correlate to yield of MAA. When investigated at progressively higher temperatures in

260

the presence of PPh3 (entries 7 – 10), the isolated yield of MAA continued to increase, reaching a

261

maximum of 33.8% at 225 oC and 1.5 h (entry 10). This compares favorably to decarboxylations

262

performed in the absence of PPh3, where a yield of only 14.9% was observed at otherwise similar

263

conditions (entry 6). The addition of PPh3 had a positive effect on the isolated yield in these

264

reactions, and it is possible that other additives could further increase the isolated yield. Lastly,

265

as seen in Figure 7, carbon monoxide was not observed with ruthenium catalysis, thereby

266

indicating that carbon monoxide was from oxalate decomposition.

267

Comparison of manganese-catalyzed (Table 2) to ruthenium-catalyzed decarboxylations

268

revealed several important differences. Ruthenium-catalyzed reactions provided MAA in higher

269

yield (33.8% versus 14.0%) with lower catalyst concentration (0.1 versus 5.0 mol %), as seen by

270

comparison of entry 10 in Table 3 to entry 10 in Table 2. Both ruthenium and manganese-

271

catalyzed reactions were facilitated by addition of PPh3. However, the percentage of PPh3

272

(relative to IA) was an order of magnitude lower for ruthenium-catalyzed reactions. A further

273

demonstration of the efficacy of ruthenium relative to manganese is comparison of experiments

274

conducted at similar reaction temperature and time. For instance, manganese-catalyzed

275

decarboxylations conducted at 200 oC for 1.5 h (entry 3, Table 2) to the corresponding ruthenium

13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

276

experiment (entry 4, Table 3) revealed a six-fold difference in yield of MAA (8.6 versus 1.5%)

277

despite the lower catalyst loading for ruthenium.

278

Lowering vapor pressure with co-solvents. Lower pressures are desirable from

279

engineering and economic standpoints due to safety and cost considerations. Lower pressure is

280

achieved primarily by lower reaction temperatures yielding lower pressures according to Gay-

281

Lussac’s Law and/or by replacement of a fraction of water with an appropriate co-solvent

282

possessing lower vapor pressure according to Raoult’s Law. Replacement of a portion of the

283

solvent (water) with tetraglyme to form a mixture (1:1 mole ratio) of water and tetraglyme

284

facilitated considerably lower vapor pressures relative to water at constant volume, as seen in

285

Figure 5. Tetraglyme was chosen because it is chemically inert to our process and has a lower

286

vapor pressure than water. Note that the data collected for Figure 5 was conducted in the

287

absence of IA and catalyst.

288

Application of a tetraglyme/water solvent system (1:1 mole ratio) to decarboxylation of

289

IA in the presence of catalytic ruthenium carbonyl propionate yielded a considerably lower ∆P

290

(133 versus 277 psig) while reducing the yield of MAA. This is seen by comparison of the first

291

and third entries of Table 4. While yields were typically lower in water/tetraglyme mixtures than

292

in water alone, the water/tetraglyme mixtures preserved considerably more of the starting acids,

293

which would facilitate recycling to produce additional MAA. Replacement of all of the water

294

with tetraglyme (entry 2, Table 4) resulted in further yield reduction (3.9%). We speculate that

295

this was due to anhydride formation, as anhydrides are anticipated to be less reactive toward

296

decarboxylation than carboxylic acids. Furthermore, anhydride formation in the absence of

297

water is probably thermodynamically favored under our conditions. For instance, the

298

equilibrium constants determined in a previous study for diacid ↔ anhydride + water in aqueous

14 ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

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 Sustainable Chemistry & Engineering

299

solution at 20 oC ranged from 3.2 – 25 for dimethylmaleic acid and other maleic acid

300

derivatives.27 As a consequence, water is a necessary component in our reaction medium, as it

301

mitigates formation of anhydrides in situ. Lastly, as anticipated from the data presented in

302

Figure 5, replacement of water with tetraglyme afforded lower pH20, as seen by comparison of the

303

first entry in Table 4 (100% water; pH20 of 313 psig) with the other entries (pH20 ≤ 209 psig).

304 305

CONCLUSIONS

306

Ruthenium (I) dicarbonyl propionate was identified as the most efficient catalyst at selective

307

decarboxylation of IA to afford sustainable MAA with high selectivity. Also investigated was

308

manganese (II) oxalate (5 mol %), which was identified as providing the highest yield of MAA

309

after screening over 20 simple transition metal catalysts. However, accumulation of carbon

310

monoxide during the course of manganese-catalyzed decarboxylation suggested that

311

decomposition of oxalate under our conditions was problematic. This was confirmed with a

312

control experiment that demonstrated elevated carbon monoxide levels when oxalic acid was

313

added to the system. Attempts at stabilization of the catalyst with triphenylphosphine were

314

unsuccessful, but it exhibited greater catalytic efficacy (14.0% yield) than the manganese catalyst

315

(4.8% yield) at 5 mol % with otherwise similar reaction conditions (1.5 h and 225 oC). Evolution

316

of carbon monoxide was not observed during ruthenium-catalyzed decarboxylation.

317

Additionally, lower catalyst loadings (< 1 mol %), lower reaction temperatures and pressures and

318

higher yields were achieved utilizing ruthenium catalysis, thereby improving reaction parameters

319

relative to manganese-catalyzed decarboxylations. High substrate concentration (5.5 M), low

320

reaction temperature (200 – 225 oC), low reaction pressure (≤ 425 psig), low ruthenium catalyst

321

load (0.1 mol %), high selectivity (> 90%), and avoidance of propylene production (excessive

15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

322

decarboxylation) represent advances over previous contributions in this area. Additionally,

323

ruthenium catalyzed yields increased further with the addition of PPh3, and could potentially be

324

further enhanced with the appropriate additive. Furthermore, we achieved direct decarboxylation

325

of IA as opposed to its conjugate base, thereby avoiding basification and acidification steps. We

326

also report that utilization of co-solvents such as tetraglyme lowers the vapor pressure of water

327

within the reaction vessel by > 100 psig while minimizing decomposition of starting acids. We

328

continue to investigate additives, reaction conditions and solvent conditions that will lead to

329

increased yield, conversion and selectivity as well as stabilization of products within the reactor

330

system once they have formed.

331 332

Supporting Information

333

Table S1 depicts screening results of catalytic decarboxylation using various catalysts; vapor

334

pressure of steam as a function of temperature is shown in Figure S1; Figures S2 and S3 display

335

GC-MS and NMR data of isomerized dibutyl itaconate.

336 337

AUTHOR INFORMATION

338

Corresponding Author

339

*B. R. Moser. Fax: 309-681-6534. Tel: 309-681-6511. E-mail: [email protected].

340 341

Author contributions

342

All authors contributed to conceptual design of experiments as well as manuscript preparation.

343

All authors have given approval to the final version of the manuscript.

344

16 ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

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 Sustainable Chemistry & Engineering

345

Notes

346

The authors declare no competing financial interest.

347 348

Disclaimer

349

Mention of trade names or commercial products in this publication is solely for the purpose of

350

providing specific information and does not imply recommendation or endorsement by the U.S.

351

Department of Agriculture. USDA is an equal opportunity provider and employer.

352 353

ACKNOWLEDGEMENTS

354 355

Benetria Banks, Daniel Knetzer and Erin Walter are acknowledged for excellent technical

356

assistance.

357 358

Funding

359 360

This work was part of the in-house research of the Agricultural Research Service of the United

361

States Department of Agriculture.

362 363

REFERENCES

364

(1)

Collins, T. Toward sustainable chemistry. Science 2001, 291, 48–49.

365

(2)

Posada, J. A.; Patel, A. D.; Roes, A.; Blok, K.; Faaij, A. P.; Patel, M. K. Potential of

366

bioethanol as a chemical building block for biorefineries: preliminary sustainability

367

assessment of 12 bioethanol-based products. Bioresour. Technol. 2013, 135, 490−499.

17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

368

(3)

Zhou, C. H.; Beltramini, J. N.; Fan, Y. X.; Lu, G. Q. Chemoselective catalytic conversion

369

of glycerol as a biorenewable source to valuable commodity chemicals. Chem. Soc Rev.

370

2008, 37, 527–549.

371

(4)

fuels from biomass. Angew. Chem. Int. Ed. 2007, 46, 7184–7201.

372 373

(5)

(6)

Choudhary, T. V.; Phillips, C. B. Renewable fuels via catalytic hydrodeoxygenation. Appl. Catal. A: Gen. 2011, 397, 1–12.

376 377

Corma, A.; Iborra, S.; Velty, A. Chemical routes for the transformation of biomass into chemicals. Chem. Rev. 2007, 107, 2411–2502.

374 375

Huber, G. W.; Corma, A. Synergies between bio- and oil refineries for the production of

(7)

Kozliak, E.; Kubatova, A.; Artemyeva, A. A.; Nagel, E.; Zhang, C.; Rajappagowda, R.

378

B.; Smirnova, A. L. Thermal liquefaction of lignin to aromatics: efficiency, selectivity,

379

and product analysis. ACS Sustainable Chem. Eng. 2016, 4, 5106–5122.

380

(8)

Gen. 2001, 221, 367–377.

381 382

(9)

(10)

(11)

389

Wilczynski, R.; Juliette, J. J. Methacrylic Acid and Derivatives. In: Kirk-Othmer Encyclopedia of Chemical Technology, 5th Ed., Vol. 16, Wiley: 2006; p 227.

387 388

van Haveren, J.; Scott, E. L.; and Sanders, J. Bulk chemicals from biomass. Biofuels Bioprod. Bioref. 2008, 2, 41–57.

385 386

Witcoff, H. A.; Reuben, B. G.; Plotkin, J. S. Industrial Organic Chemicals, 2nd edition. Wiley-Interscience: 2014; pp 1–662.

383 384

Nagai, K. New developments in the production of methyl methacrylate. Appl. Catal. A:

(12)

Johnson, D. W.; Eastham, G. R.; Poliakoff, M. Method of producing acrylic and methacrylic acid. U.S. Patent 8,933,179 B2. 2015.

18 ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

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

390

ACS Sustainable Chemistry & Engineering

(13)

ethylene and catalyst systems for use therein. WO Patent 09619434. 1996.

391 392

(14)

Dubois, J. L. Procede de fabrication d’un methylacrylate de methyl derive de la biomasse. WO Patent 2010058119. 2010.

393 394

Tooze, R. P.; Eastham, G. R.; Whiston, K.; Wang, X. L. Process for the carbonylation of

(15)

Carlsson, M.; Habenicht, C.; Kam, L. C.; Antal Jr.; M. J.; Bian, N.; Cunningham, R. J.;

395

Jones Jr., M. Study of sequential conversion of citric to itaconic to methacrylic acid in

396

near-critical and supercritical water. Ind. Eng. Chem. Res. 1994, 33, 1989–1996.

397

(16)

Johnson, D. W.; Eastham, G. R.; Poliakoff, M.; Huddle, T. A. A process for the

398

production of methacrylic acid and its derivatives and polymers produced therefrom. WO

399

Patent 2012/069813. 2012.

400

(17)

Le Notre, J.; Witte-van Dijk, C. M.; van Haveren, J.; Scott, E. L.; Sanders, J. P. M.

401

Synthesis of bio-based methacrylic acid by decarboxylation of itaconic acid and citric

402

acid catalyzed by solid transition-metal catalysts. ChemSusChem 2014, 7, 2712–2720.

403

(18)

itaconic acid reactions in real time. J. Phys. Chem. A 2001, 105, 10839–10845.

404 405

Li, J.; Brill, T. B. Spectroscopy of hydrothermal solutions 18: pH-dependent kinetics of

(19)

Kuo, T. M.; Kurtzman, C. P.; Levinson, W. E. Production of itaconic acid by Pseudozyma antarctica. U.S. Patent 7,479,381. 2009.

406 407

(20)

Batti, M. A. Process for the production of itaconic acid. U.S. Patent 3,162,582. 1964.

408

(21)

Tsai, Y. C.; Huang, M. C.; Lin, S. F.; Su, Y. C. Method for the production of itaconic acid using Aspergillus terreus solid state fermentation. U.S. Patent 6,171,831. 2001.

409 410

(22)

Crooks, G. R.; Johnson, B. F. G.; Lewis, J.; Williams, I. G.; Gamlen, G. Chemistry of

411

polynuclear compounds. Part XVII. Some carboxylate complexes of ruthenium and

412

osmium carbonyls. J. Chem. Soc. A. 1969, 276 –2766.

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

413

(23)

Murray, R. E.; Doll, K. M.; Liu, Z. Process for isomerization and decarboxylation of

414

unsaturated organic compounds with a metal catalyst or catalyst precursor. U.S. Patent

415

Application 2014/0275592 (September 18, 2014).

416

(24)

converting alkenoic fatty acids into alkenes. ACS Catal. 2014, 4, 3517–3520.

417 418

Murray, R. E.; Walter, E. L.; Doll, K. M. Tandem isomerization-decarboxylation for

(25)

Moser, B. R.; Knothe, G.; Walter, E. L.; Murray, R. E.; Dunn, R. O.; Doll, K. M.

419

Analysis and properties of the decarboxylation products of oleic acid by catalytic

420

triruthenium dodecacarbonyl. Energy Fuels 2016, 30, 7443–7451.

421

(26)

Hill: 1997; pp 2-48 – 2-75.

422 423

Perry, R. H.; Green, D. W. Perry’s Chemical Engineers’ Handbook, 7th Ed, McGraw-

(27)

Eberson, L.; Welinder, H. Studies on cyclic anhydrides. III. Equilibrium constants for the

424

acid-anhydride equilibrium in aqueous solutions of certain vicinal diacids. J. Am. Chem.

425

Soc. 1971, 93, 5821–5826.

20 ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

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

426

427 428

ACS Sustainable Chemistry & Engineering

Table 1. Comparison of existing bio-based routes to MAA with the current study ChemSusChem WO 2012/069813 2014, 7, 2717-2720 Feedstock Monosodium salt of IA Monosodium salt of IA Reactor High water diluent High water diluent Low [IA] Low [IA] Catalysts Heterogeneous Pt, Pd Group I and II bases and Ru [Catalyst] (mol %) 2.5 30 – 100 [Starting Acid] 0.15 M 0.3 – 3.0 Temp (oC) 200 – 250 255 – 265 Pressure (psi) 551 450 - 3000 Time 1h 150 seconds Solvent Supercritical water Supercritical water Reaction mode Batch (unvented) Continuous (vented) Yield of MAA Up to 51% (isolated) Up to 63% (HPLC) Selectivity1 Up to 84% Up to 99% 2 Propylene detected? Yes Not determined 1 Selectivity is defined as [MAA / (MAA + crotonic acid) × 100]. 2

Ind. Eng. Chem. Res. 1994, 33, 1989-1996 Citric acid or IA High water diluent Low [IA] NaOH 0 – 50 0.01 – 0.50 230 – 400 4000 - 5000 1 – 250 seconds Supercritical water Continuous (vented) Up to 75% (HPLC) Up to 100% Yes

Current study IA Low water diluent High [IA] Mn(II) oxalate [Ru(CO)2(CH3CH2COO)]n 0.1 – 10 1.0 – 10.0 190 – 250 1 - 425 1–3h Subcritical water Batch (unvented) Up to 40% (HPLC) Up to 95% No

Propylene is indicative of over-reaction (double decarboxylation of IA or triple for citric) and is therefore undesirable.

429 430 431 432 433

21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 22 of 33

434

Table 2. Results of catalytic decarboxylation of 5.5 M IA with Mn (II) oxalate and triphenylphosphine (PPh3) with 400 ± 5

435

psig N2 headspace pressure Entry 1 2 3 4 5 6 7 8

Mn (II) oxalate PPh3 Time Temp CA MAA ∆P (mol %) (h) (°C) (mol %) (area %) (isolated)1 (psig) 1.5 225 69 0 0 0.8 2.6 1.5 200 67 5.0 0 0.8 1.5 1.5 212 135 5.0 0 0.8 2.4 1.5 225 250 5.0 0 0.7 4.8 1.5 250 342 5.0 0 0.9 5.7 1.5 212 196 5.0 5.0 2.1 5.3 1.5 225 326 5.0 5.0 2.4 13.0 1.5 225 282 0 5.0 2.9 14.0 Yield of MAA is reported as percent of theoretical after distillation.

IA+CCA+MCA (area %)2 69.5 56.4 58.5 41.4 6.2 64.6 33.4 19.4

Sel3 76.5 65.2 75.0 87.2 86.4 71.6 84.4 82.8

pH2O4 (psig) 312 191 244 312 499 243 311 311

436

1

437

2

IA = itaconic acid; CCA = citraconic acid; MSA = mesaconic acid. Values were determined by HPLC.

438

3

Selectivity is [MAA/(MAA + CA) × 100].

439

4

Partial vapor pressure of water (pH2O, psig) was calculated from mole fraction of water and vapor pressure data from reference

440

26.

441 442 443 444 445 22 ACS Paragon Plus Environment

Page 23 of 33

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

446

ACS Sustainable Chemistry & Engineering

Table 3. Results of catalytic decarboxylation of 5.5 M IA with 0.1 mol % [Ru(CO)2(CH3CH2COO)]n1 Entry

Time Temp CA mol % ∆P (h) (°C) (psig) PPh3 (area %) 1 0.5 212 75 0.0 0 2 1.0 212 128 0.0 0 3 2.0 212 201 0.0 0 4 1.5 200 120 0.0 0 5 1.5 212 184 0.0 0 6 1.5 225 277 0.0 0 7 1.5 200 116 0.5 0 8 1.5 212 206 0.5 0 9 1.0 225 251 0.5 0 10 1.5 225 289 0.5 0 All headspace pressures are 400 ± 5 psi.

MAA (isolated)2 8.2 13.6 13.7 8.6 17.0 14.9 14.9 21.6 25.4 33.8

IA+CCA+MCA (area %)3 92.5 57.8 27.7 74.9 34.4 3.8 46.0 25.2 12.3 6.2

Sel4 100 100 100 100 100 100 100 100 100 100

pH2O5 (psig) 245 245 245 191 245 313 190 245 313 313

447

1

448

2

Yield of MAA is reported as percent of theoretical after distillation.

449

3

IA = itaconic acid; CCA = citraconic acid; MSA = mesaconic acid. Values were determined by HPLC.

450

4

Selectivity was calculated according to the same method as in Table 2.

451

5

Partial vapor pressure of water was calculated according to the same method in Table 2.

452 453 454 455

23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Page 24 of 33

456

Table 4. Results of catalytic decarboxylation of 5.5 M IA using 0.1 mol % [Ru(CO)2(CH3CH2COO)]n with varied solvent

457

compositions Entry 1 2 3 4 5

IA+CCA+MCA mole ratio mole ratio ∆P CA MAA 1 (area %)2 TG : water IA: water (psi) (area %) (isolated) 0:1 1:10 277 0.9 14.9 16.8 1:0 N/A 108 3.7 3.9 67.6 3.0 12.6 75.1 1:1 0.72:1 133 1:3 1:2 139 3.8 7.6 82.6 1:4 1:2.5 153 3.3 9.2 49.9 Yield of MAA is reported as percent of theoretical after distillation.

Sel3 94.3 51.3 80.8 66.7 73.6

pH2O4 (psig) 313 0 101 188 209

458

1

459

2

IA = itaconic acid; CCA = citraconic acid; MSA = mesaconic acid. Values were determined by HPLC.

460

3

Selectivity was calculated according to the same method as in Table 2.

461

4

Partial vapor pressure of water was calculated according to the same method in Table 2.

24 ACS Paragon Plus Environment

Page 25 of 33

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 Sustainable Chemistry & Engineering

O

HCN

HO

CN

H2SO4

O

NH2H2SO4

CH3OH

O O Methyl methacrylate

H2O

O OH Methacrylic acid

Figure 1. Production of methacrylic acid from acetone by the conventional petrochemical acetone-cyanohydrin route.

25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Itaconic Acid

Mesaconic Acid

O

Citraconic Acid

O OH

HO

O

OH

HO

O

H2O

CO2

Citramalic Acid

Crotonic Acid O

O OH

HO

HO

OH

HO

O

H2O

O

OH

O

CO2 2-Hydroxyisobutyric Acid OH HO

H2O

O

H2O

Methacrylic Acid O OH

Figure 2. Bio-based route to methacrylic acid from itaconic acid. Note that mesaconic and citramalic acids are intermediates along the way to methacrylic acid. Also note that 2hydroxyisobutyric acid is the hydrated form of methacrylic acid.

26 ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33

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 Sustainable Chemistry & Engineering

O HO

O OH

O conjugate acid

NaOH HO

O

O O-Na+

O conjugate base

catalyst

O-Na+

HCl

OH + NaCl methacrylic acid

CO2

Figure 3. Production of methacrylic acid through decarboxylation of monosodium itaconate, the conjugate base of itaconic acid. In this example, the sodium salt is depicted, but others such as potassium, ammonium, and others may be utilized.

27 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Simple Fermentation Sugars

O

O OH

HO

H2O, catalyst OH

O Itaconic acid

Methacrylic acid CO2 CO2

O HO

O

OH O OH

OH

Citric acid

Figure 4. Bio-based methacrylic acid from simple sugars and citric acid via an itaconic acid intermediate.

28 ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

600 Water:Glyme, 1:1 500

Pressure, psig

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 Sustainable Chemistry & Engineering

Water Only Glyme only

400

300

200

100

0 100

120

140

160

180

200

220

240

260

Temperature, °C

Figure 5. Comparison of pressure (psig) as a function of temperature (Celsius) at constant volume among water, a 1:1 (mole ratio) mixture of water and tetraglyme and tetraglyme, demonstrating that vapor pressure can be decreased with prudent choice of an appropriate low vapor pressure organic co-solvent.

29 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

Figure 6. Micro GC chromatogram of samples of headspace gas after decarboxylation of IA using 5.0 mol % manganese (II) oxalate with (a) and without (b) 5.0 mol % PPh3 showing evolution of both carbon monoxide (0.47 min) and carbon dioxide (2.2 min). Chromatogram c) depicts the results of an experiment conducted in the presence of excess oxalic acid. The signal at 0.38 min corresponds to nitrogen.

30 ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33

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 Sustainable Chemistry & Engineering

Figure 7. Representative micro GC chromatogram of a sample of headspace gas collected after decarboxylation of IA at 200 oC for 2.0 h utilizing 0.1 mol % [Ru(CO)2(CH3CH2COO)]n. Signals corresponding to nitrogen (N2) and carbon dioxide (CO2) are at 0.38 and 2.2 minutes, respectively. If present, propylene would appear at 1.0 minutes.

31 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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

For Table of Contents Use Only Bio-Based Methacrylic Acid via Selective Catalytic Decarboxylation of Itaconic Acid James C. Lansing, Rex E. Murray and Bryan R. Moser

Methacrylic acid is synthesized from renewable itaconic acid using transition metal catalysts and subcritical water, which represents a sustainable alternative to the hazardous petrochemical route.

32 ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

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 Sustainable Chemistry & Engineering

Simple bio-based sugars can be converted to itaconic acid, which in turn can be catalytically transformed into methacrylic acid, a commodity monomer with numerous practical applications, including as components in protective eyewear and car headlights. 223x87mm (96 x 96 DPI)

ACS Paragon Plus Environment