High-Temperature Batch and Continuous-Flow Transesterification of

Jan 15, 2018 - Biobased glycerol and its acetal derivatives (GAs: glycerol acetals) fit in the current list of top biomass-derived platform chemicals ...
1 downloads 14 Views 952KB Size
Subscriber access provided by READING UNIV

Article

High-Temperature Batch and Continuous-Flow Transesterification of Alkyl and Enol Esters with Glycerol and its Acetal Derivatives Roberto Calmanti, Manuele Galvan, Emanuele Amadio, Alvise Perosa, and Maurizio Selva ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b04297 • Publication Date (Web): 15 Jan 2018 Downloaded from http://pubs.acs.org on January 15, 2018

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 35 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

High-Temperature Batch and Continuous-Flow Transesterification of Alkyl and Enol Esters with Glycerol and its Acetal Derivatives Roberto Calmanti, Manuele Galvan, Emanuele Amadio, Alvise Perosa, Maurizio Selva* Dipartimento di Scienze Molecolari e Nanosistemi dell’Università Ca’ Foscari Venezia, Via Torino 155, 30172 – Venezia Mestre (Italy) * Corresponding Author: Prof. Maurizio Selva. Web: http://www.unive.it/persone/selva. e-mail: [email protected]

KEYWORDS Transesterification, alkyl esters, isopropenyl acetate, glycerol, glycerol acetals, high-temperature (HT) reactions

ABSTRACT A new procedure for the transesterification of alkyl acetates and formates with model glycerol acetals (GAs: Solketal and glycerol formal) was explored in the absence of any catalysts at 180275 °C. Highly selective transformations occurred in both batch and continuous-flow (CF) modes: particularly, the enol derivative isopropenyl acetate (iPAc) was the best performing reactant by

ACS Paragon Plus Environment

1

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

Page 2 of 35

which quantitative acetylation reactions were achieved with yields on GAs acetates > 95%. An excess acylating agent was necessary (2-20 molar equivs.), but the unconverted ester was fully recovered and could be reused. The reaction plausibly involved multiple mechanisms where either the electrophilic and the nucleophilic activation of reagents took place through both traces of acetic acid (formed in situ by the hydrolysis of esters) and the autoprotolysis of GAs. iPAc confirmed a superior performance than other esters also for the high-temperature conversion of glycerol: in this case, although acylation and acetalization processes were simultaneously possible, conditions were optimised to achieve the exhaustive transesterification of glycerol to triacetin, in both batch and CF modes. Triacetin was isolated in 99% yield.

INTRODUCTION Bio-based glycerol and its acetal derivatives (GAs: glycerol acetals) fit in the current list of top biomass-derived platform chemicals which are used as building blocks for higher value-added chemical products and materials.1,2,3,4,5,6 Among the procedures for the chemical conversion of glycerol and GAs, an emerging field is related to high-temperature (HT) transformations carried out in the absence of any catalyst. The study of HT-driven routes was originally aimed at improving biodiesel production not only through the transesterification of natural oils, but also by the simultaneous conversion of the co-product glycerol into glycerol dicarbonate (GDC) or triacetin (TA), respectively.7,8,9,10,11 To this end, supercritical fluid-based (SCF) reactions with nontoxic acyl acceptors such as dimethyl carbonate (DMC) or methyl acetate were used (Scheme 1). Scheme 1. Transesterification of natural triglycerides with DMC or methyl acetate.

ACS Paragon Plus Environment

2

Page 3 of 35 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

At 250-380 °C and 150-200 bar, such SCF-technologies proved highly tolerant towards common impurities (FFA and water) present in the reactant oils and they allowed extremely fast kinetics and yields >90% with simplified downstream processing in which high-standard glycerol-free biodiesel as blends of FAMEs and GDC or TA were obtained without further purification.12,13,14,15,16,17 Moreover, energy costs and capital investments of plants running on SCF-technologies (SCF-plants) could be efficiently mitigated by integrating the SCF-based reactions within biorefinery units able to recover or exchange waste/excess heat.18 Comparative studies and simulations demonstrated that the total energy consumption and the output potential environmental impact per mass of product obtained from SCF-plants for the biodiesel production could be even lower than that of conventional base-catalyzed transesterification processes.19,20,21 HT-reactions were also successfully investigated by us for the upgrading of model glycerol acetals, specifically glycerol formal and solketal, and glycerol by using dialkyl carbonates.22,23,24 For example, at 275-300 °C and 20-40 bar, the continuous-flow (CF) reaction of DMC with GAs yielded mono-transcarbonated products with a 98% selectivity at complete conversion; while, the same CF-process with glycerol was optimized to obtain glycerol carbonate in a 83-92% yield. Other Authors recently described also a batch reaction between supercritical DMC and glycerol to isolate glycerol carbonate in a 98% yield after only 15 min.25

ACS Paragon Plus Environment

3

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

Page 4 of 35

This scenario prompted us to further investigate HT-reactions with the new objective of implementing the conversion of glycerol and GAs into the corresponding ester derivatives. Notwithstanding the many catalytic procedures described with carboxylic acids and anhydrides (mostly AcOH and Ac2O),26,27,28,29,30,31,32,33 the catalyst-free esterification of glycerol has seldom been reported: poor conversions (35-49%) and moderate product yields (70%, triacetin) were achieved with both AcOH and Ac2O, respectively.34,35,36 Issues related to the corrosivity of reactants, the legal restrictions to the use of Ac2O, and the formation of co-product water severely limited large-scale applications of such methods either in the presence or in the absence of catalysts.37,38 Water not only altered the esterification equilibrium, but also contributed in the deactivation of catalysts if used.26-33 A literature survey indicated that the transesterification reactions of esters with both glycerol and GAs were more promising and safer processes. The acetylation of glycerol with alkyl (methyl, ethyl) acetates was reported over several homogeneous and heterogeneous acid and base catalysts,39,40,41,42,43,44,45,46 while the transesterification of esters with GAs was almost exclusively finalized at the kinetic resolution of racemic GAs esters by lipase-based biocatalysts.47,48,49,50 The analysis also proved that the HT-based reactions of esters with glycerol and its acetals still represented a largely unexplored area. In the present work, a systematic inspection of such processes has been carried out by using model reactants including formate, acetate, and lactate esters, isopropenyl acetate (iPAc: an active acetylating agent,51,52,53,54), glycerol, and its acetals glycerol formal and solketal. At 200-240 °C and 10-50 bar, in the absence of any catalysts, batch and continuous-flow protocols were achieved for the selective and high-yield synthesis of esters of GAs and triacetin, respectively. Some mechanistic hypotheses were formulated to account for the product distribution and the different reactivity of the esters.

ACS Paragon Plus Environment

4

Page 5 of 35 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

EXPERIMENTAL SECTION Materials and equipment including autoclaves and the continuous-flow (CF) apparatus used in this study are listed and described in the SI section. All the reactions described in the experimental and the results and discussion sections, were repeated twice to ensure reproducibility. Under the same set of conditions (T, p, flow rates) duplicated tests afforded values of conversion and amount of products (determined by GC/MS) which differed by less than 5% from one experiment to another. Commercially available esters used in this study were safe and nontoxic compounds: these included methyl acetate (2a), ethyl acetate (2b), propyl acetate (2c), ethyl formate (2d), S-(-)ethyl lactate (2e) and isopropenyl acetate (2f: iPAc). Reactants 2a-2f were flammable products with hazardous (H) phrases and precautionary statements (P) as H225, P210, and P403 + P235. However, they were low toxic compounds. For example, LD50 (oral, acute) for methyl-, ethyl-, and propyl acetates are 5001 mg/kg, 5620 mg/kg, and 6640 mg/kg. 55 Ethyl lactate is even edible and present in many foods. Overall, such esters were safer than acids or anhydrides as esterification reagents. Batch reactions. Mixtures of acetal 1a or 4a-a’ (4.0 mmol) and each ester 2a-f were prepared by varying the reactant molar ratio (Q= ester:acetal) from 1 to 20, the (excess) ester serving both as a reagent and a solvent. Each homogeneous solution was charged in a flat-bottomed glass reactor which was placed inside a 100-mL stainless steel autoclave. This expedient avoided the contact of reagents with the autoclave walls, thereby ruling out metal catalysis. The autoclave was sealed, degassed via three vacuum-nitrogen cycles, and then electrically heated at the desired temperature (120-220 °C). The reaction was allowed to proceed from 1 to 24 h, during which the reacting mixture was kept under magnetic stirring. The observed autogenous pressure was in the range 2-

ACS Paragon Plus Environment

5

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

Page 6 of 35

20 bar. At the end of the experiment the reactor was rapidly cooled to rt and vented. The reaction mixtures were analysed by GC/FID or GC/MS. The same procedure was used also to run experiments with: i) added AcOH in a 1-20 mol% amount with respect to the reacting acetal; ii) glycerol in place of acetals. In this case, irrespective of the reactants molar ratio (ester:glycerol, from 1 to 20), starting mixtures were biphasic, but turned to homogeneous solutions at the end of the experiments. Continuous-flow (CF) reactions. Homogeneous solutions of acetal 1a or 4a-4a’ and the selected ester (2a-f) were prepared by mixing the reactants in different molar ratios from 5 to 20. Before any reaction, the CF-reactor (a stainless-steel tubular reactor: 0.52 m x 1/4’’ filled with ground-glass Raschig rings) was primed and conditioned by delivering the chosen solution of reactants (10 mL) at room temperature and atmospheric pressure, and at a flow rate of 0.1 mL/min. The reaction was then started by setting the temperature and the pressure at the desired values (200–275 °C and 1-50 bars, respectively). T and p were controlled by a thermostatic oven and a back-pressure regulator. Once an amount of the reacting mixture equalled 5 times the inner volume of the reactor (2.1 mL) was allowed to flow, samples at the outlet of the reactor were taken up at regular intervals of 30 minutes through a Rheodyne valve, and analysed by GC/FID and GC/MS. At the end of the experiment, the oven was set to 100 °C and the reactor was flushed with methanol (100 mL at 0.5 mL/min), cooled to room temperature and vented. A similar procedure was used also for the CF-reactions of glycerol with both methyl and isopropenyl acetates. Diglyme was necessary as a co-solvent due to the poor mutual miscibility of the reactants. Homogeneous solutions were achieved by mixing glycerol, the selected ester (2a or 2f), and diglyme in a 1:1-20:33 molar ratio, respectively. This mixture was used to run experiments described in Table 2.

ACS Paragon Plus Environment

6

Page 7 of 35 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

The product esters included: (2,2-dimethyl-1,3-dioxolan-4-yl)methyl acetate (3a: solketal acetate, (2,2-dimethyl-1,3-dioxolan-4-yl)methyl formate (3b: solketal formate), (2,2-dimethyl1,3-dioxolan-4-yl)methyl 2-hydroxypropanoate (3c: solketal lactate), a 3:2 isomer mixture of 1,3dioxan-5-yl formate and (1,3-dioxolan-4-yl)methyl formate (5a-a’: isomers of glycerol formal formate), a 3:2 isomer mixture of 1,3-dioxan-5-yl acetate and (1,3-dioxolan-4-yl)methyl acetate (5b-b’: isomers of glycerol formal acetate), and triacetin (8) (see below, Schemes 2, 4, and 5 for detailed structures). These compounds were purified and characterized by MS and NMR. All data including also isolated yields are reported in the SI section.

RESULTS AND DISCUSSION Batch HT-transesterification of different esters with solketal (1a). Conditions to begin this study were chosen based on our recent results reported for the transcarbonation of dialkyl carbonates with glycerol acetals.22-24 As described in the experimental section, mixtures of 1a and each of the esters 2a-2f in a 1:20 molar ratio (Q), respectively, were set to react in an autoclave at different temperatures (150 to 220 °C) and times (1 to 10 h). In the absence of catalyst, the transesterification process occurred at T≥180 °C. The structure of the products, i.e. solketal esters 3a, 3b, and 3c (R1=Me, H, and CH3CH(OH), respectively; Scheme 2) was confirmed through their GC/MS and 1H/13C MNR characterization. Compounds 3a and 3b were also isolated in 96 and 60 % yields from the reactions of 1a with 2f and 2d, respectively. Scheme 2. Reactions of solketal with different esters at high temperatures.

ACS Paragon Plus Environment

7

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

Page 8 of 35

Each reactant ester 2a-2f gave remarkably different outcomes. As an example, Figure 1 compares the results achieved at 200 °C for 5 h, under an autogenous pressure of 8 bars.

Figure 1. Batch HT-reaction between solketal and esters 2a-f: conversion of solketal and selectivity towards Solketal esters 3a, 3b, and 3c are shown. Conditions: ester:solketal molar ratio (Q)=20, 200 °C, 5h. The conversion of solketal did not exceed 10% with acetates 2a-c, while it was substantially higher for ethyl lactate, ethyl formate (2e-d) and i-propenyl acetate (2f): 57%, 67% and 100% (blue bars), respectively. Except for ethyl lactate, the transesterification selectivity was always

ACS Paragon Plus Environment

8

Page 9 of 35 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

very high (≥98%, red bars). Only trace amounts of monoacetins, diacetins and triacetin were observed ( 96% by GC/MS and is therefore not shown.

100

60 50

0 22 0 20

pe

m

0 18

40 30 20 10 0

ra

0 16

20

re tu

10

0 14

5

Mo

] C



0 12

la

o ati rR

Conversion

80 70

[%]

90

Te

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

Page 10 of 35

]

[Q

Figure 2. Conversion of solketal as a function of the temperature and the reactants molar ratio (Q) during the batch reaction of solketal and iPAc. All tests were carried out for 5 hours. The selectivity towards product 3a was always > 96%.

The conversion was favoured both by higher temperature (bars of the same color) as well as by higher Q ratio (left to right). Yet, a quantitative reaction took place also with a moderate excess of iPAc (Q = 5, 220 °C: green bar). This behaviour was further substantiated by a separate

ACS Paragon Plus Environment

10

Page 11 of 35 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

experiment, not reported in Figure 2, demonstrating that at the same T (220 °C), the transesterification could be completed even at Q=2, albeit over 10 hours. Whichever the conditions used, vacuum distillation of the final mixtures allowed an almost complete recovery of the unconverted excess reagent: more than 95 wt% of iPAc could be reused as such for further transesterification reactions. GC/MS analyses of the final mixtures also provided some insights to speculate on the reaction mechanism, particularly the detection of traces of acetic acid among the reaction products. The presence of AcOH was plausibly due to a hydrolysis side-reaction of iPAc occurring at high temperature.60 In fact, both commercial Solketal and iPAc had water contents of 1, and 0.1-0.15 mol%, respectively (see Karl-Fisher tests in the SI section for details), and attempts to dehydrate solketal failed because of its highly hydrophilic nature. This suggested that the investigated transesterification could be an acid-catalysed process. To explore this hypothesis, further reactions of Solketal and iPAc in the presence of added AcOH (1, 5 and 20 mol% respect to solketal) were run. Most representative results are summarised in Table 1 (Experiments were followed for 5 hours: details on the trend of conversion vs time are reported on SI, see figure S25). Table 1. The reaction of iPAc with Solketal in the presence of added AcOH.

Entry

Added AcOH (mol%)a

1

noned

2

1

3

5

4

20

Qb

T/tb

Conv. (%)c 57

20

180 °C

68

3h

69 83

a

mol % with respect to Solketal. b Q was the iPAc:Solketal molar ratio; Q, T and t (temperature and time, respectively) were kept constant throughout runs 1-4. c Conversion of Solketal by GC/MS. d Traces of AcOH (98%). The effect still appeared limited even by quadrupling the acid amount (20 mol%, entry 4). A similar poor catalytic performance of AcOH was reported also for the closely related esterification of fatty acids with supercritical methyl acetate.61 This led to conclude that an acidcatalysed pathway was hardly the only active mechanism for reactions of Figures 1-2. An autoprotolysis equilibrium of Solketal could also be invoked.62 Analogous (autoprotolysis) processes have been described for OH groups of both aliphatic alcohols and phenols at elevated temperatures in the absence of any catalyst.63,64,65 Scheme 3 illustrates two mechanistic hypotheses. In the right-hand side red pathway, a high temperature hydrolysis of iPAc forms acetic acid which catalyzes the subsequent transesterification reaction through electrophilic activation. In lefthand side blue pathway, HT-induced autoprotolysis of Solketal generates an ion pair [(ROH2+)(RO-)] whose components act cooperatively: the cation as a proton donor and the anion as a nucleophile. In both cases, the coproducts enol or enolate rapidly tautomerize to acetone. Also, traces of water present in the reactants might play a role in the autoprotolysis equilibrium.66,67

Scheme 3. Transesterification of iPAc with solketal. Right and left: electrophilic and cooperative activation of reactants, respectively.

ACS Paragon Plus Environment

12

Page 13 of 35 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

Similar mechanisms were hypothesized also for esters 2a-e; though, the corresponding acids (AcOH, HCO2H, and CH3CH(OH)CO2H) were not observed. The concentration of these compounds (acids) could be below the detection limit because of the moderate hydrolysis of esters 2a-e which were far less reactive than 2f (see Figure 1). The extension of the HT-transesterification protocol. Glycerol Formal (GLyF) was investigated as another model acetal of glycerol. GlyF was used as a 3:2 commercially available mixture of 5-hydroxy-1,3-dioxane (4a) and (1,3-dioxolan-4-yl)methanol (4a’) respectively. Isomers 4a/4a’ were set to react under conditions similar to those of Figure 2 (T=120-220 °C, 5 h) by using methyl acetate (2a), ethyl formate (2d), and iPAc (2f) in a 20-molar excess (Q=20). The HT-reaction proved feasible for all the reactant esters, and as for Solketal, the best results were achieved with iPAc. For example, at 220 °C and after 5 hours, the reaction of 2a, 2d and 2f with GlyF showed a conversion of 13%, 54% and 96%, respectively, thereby confirming the same reactivity trend previously noticed, i.e. methyl acetate < ethyl formate < iPAc. The selectivity towards the expected formate and acetate derivatives of GlyF was always > 97%, and the structure of products (5a-a’and 5b-b’) was confirmed by GC/MS and 1H/13C MNR analyses (Scheme 4).

ACS Paragon Plus Environment

13

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

Page 14 of 35

Under the same reaction conditions (220 °C, 5h), the reversibility of the reaction of conventional esters (Scheme 4, top) was proved by heating a mixture of glycerol formal acetate (5a-5a’, 4.0 mmol) and MeOH in a 1:20 molar ratio, respectively: GlyF (4a-a’) was achieved with almost quantitative conversion (98%) and full selectivity. From the above mentioned HT-reactions of GlyF with iPAc and 2d, products as mixtures of isomers 5a-a’ and 5b-b’ were isolated in 92% and 45% yields, respectively, in the same 3:2 isomeric ratio of the reagent (4a-4a’). Figure 3 shows some illustrative results of the effect of the temperature on conversion and selectivity of the HT-transesterification of iPAc with GlyF. The conversion gradually improved from 24% up to a quantitative value as the temperature was increased from 120 to 220 °C. In this interval however, the comparison of Figures 2 and 3 showed that GlyF was systematically less reactive than Solketal. The same trend was reported by us in the reaction of GlyF and Solketal with light dialkyl carbonates,22 and similar findings were recently described also by others:68 more in general, the application/implementation of the Hansen approach and the COSMO-RS model indicated that glycerol formal had not only a stronger structuration in the liquid state than Solketal, but formaldehyde-based acetals were less reactive than ketal acetone-based homologues, under acidic (hydrolytic) conditions. Although this offers an interesting basis for discussion, the interpretation of experimental and modelling results is still far from explaining the different behaviour of GAs at a molecular level. Further investigations will be necessary to clarify such aspects. For the sake of completeness, a minor competitive process of transacetalization of GlyF with acetone (coproduced from iPAc) was also observed under the conditions of Figure 3. Traces of Solketal acetate (3a, 97%.

The transesterification of iPAc with Solketal was clearly influenced by the pressure (Figure 4B). At each of the selected temperatures, the conversion was more than doubled when the pressure was increased from 10 to 30 bars: particularly, a quantitative reaction was achieved at 275 °C. Only minor improvements (≤ 8%) were noticed by further rising the pressure up to 50 bars at both 225 and 250 °C. By contrast, the conversion was lower than 15% for reactions run at the same temperatures, but at atmospheric pressure. These results highlighted the fundamental role of phase transitions in the investigated reactions. If the pressure was high enough, despite the high T, most of the reacting mixture was present as a condensed liquid phase in which the contact between iPAc and Solketal was effective for a productive reaction. On the contrary, an abrupt decrease of the conversion occurred below a threshold value of about 30 bars when reactants, especially the more volatile iPAc (bp = 97 °C), preferentially partitioned in the vapor phase inside the reactor. To get further insights into this aspect, the phase diagram of pure iPAc was also predicted by using an extended Antoine equation (for details, see Figure S2 in the SI section).73 Under the assumption that the theoretic liquid-vapor pressure profile of pure iPAc could be an acceptable approximation for the binary mixtures of Figure 4 (that contained a large excess of the ester 2f: Q=20), the change of conversion shown in the range of 10-30 bars of Figure 4B well matched the predicted liquidvapor transition of iPAc (Figure S2). A similar dependence of conversion vs pressure was noticed also in the HT-transcarbonation of glycerol acetals:22 in that case however, the effect of the pressure was even more pronounced because reactions took place in the absence of any (inert) filler inside the CF-reactor, thereby making the contact of reactants more difficult.

ACS Paragon Plus Environment

18

Page 19 of 35 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

Under the optimized conditions for T and p (275 °C and 30 bars), the iPAc:Solketal molar ratio (Q) was scaled down from 20 to 2 to examine the effect of the relative amounts of reactants. Experiments proved that the Q ratio could be substantially decreased: a quantitative process was still observed at Q=5, i.e. fourfold below the initial value. However, a further reduction impacted on the conversion of 1a. This (conversion) did not exceed 82% at Q=2. In all cases, the selectivity towards ester 3a was >98%. Considering the (limited) capacity of the CF-reactor, the results confirmed not only the reliable performance of the process, but also the perspective for large-scale applications and further process intensification, in line with the principles of green engineering.74 The CF-arrangement could operate virtually indefinitely once the high-temperature regime was reached. No catalyst had to be prepared, activated, recovered or disposed of, thereby simplifying both upstream and downstream operations for the delivery of reactants, the recovery of products and the recycle of the unconverted iPAc. Not to mention that iPAc was a cheap reactant available on a large scale, and that the integration of the process in a waste heat recovery system provided by a biorefinery, could efficiently relieve the energy demand of the reaction. In short, if the framework of energy existing within an industrial park could be used, high-temperature reactions could be driven at very competitive, if any, costs with a further valorization of “waste” energy.18,74 The CF-protocol proved efficient also for the reaction of iPAc with glycerol formal; though, in analogy to batch reactions, GlyF was less reactive than Solketal. Under the conditions of Figure 4B (30 bars), conversions of GlyF did not exceed 31% and 76% at 225 and 250 °C, respectively. Only at 275 °C, a substantially quantitative reaction was achieved (conv.: 97%): isomer esters 5a and 5a’ were obtained in a 3:2 ratio and with an overall transesterification selectivity > 96%. The reaction of glycerol with methyl and isopropenyl acetates. Batch and continuous-flow reactions were performed by using mixtures of glycerol and methyl or isopropenyl acetates (2a

ACS Paragon Plus Environment

19

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

Page 20 of 35

and 2f) in a molar ratio (Q=2: glycerol) variable between 1 and 20. Batch (autoclave) tests were run at 120-220 °C, while CF-experiments required higher temperatures up to 300 °C and dyglime as a co-solvent. In the absence of any catalyst, the reactions proceeded with both esters. The whole spectrum of the observed products is shown in Scheme 5.

Scheme 5. The batch reaction of glycerol with esters 2a and 2f.

The structure of products acetals (1a and 3a) and acetins (6, 7 and 8) were assigned by GC/MS and by comparison to authentic samples when available (compounds 1a, 3a, and 8). The reaction conversion and the product distribution achieved under the best conditions found in this work, are described in Table 2. For comparison, the Table also shows some of the best results already reported for the batch transesterification of glycerol with methyl acetate in the presence of either an acid or a base catalyst [3%Y/SBA-3 and CaSn(OH)6, respectively]. At 180 °C, the batch transesterification of methyl acetate with glycerol yielded mono-, di-, and tri-acetin products. As expected, increasing of the Q ratio and of the reaction time favoured the conversion, but it also resulted in the formation of derivatives of multiple transesterification (entries 1-2). The selectivity was elusive also for catalytic processes: although these were faster

ACS Paragon Plus Environment

20

Page 21 of 35

than HT-reactions, not even triacetin 8 was obtained as a sole product despite the use of a large excess of ester 2a (entries 3-4). Table 2. Batch and continuous-flow reactions of glycerol with methyl and isopropenyl acetate

Ent.

Cond.

a

Ester

1

Cat.

a

Q

T/p/t

Conv.

(°C/bar/h)c

(%)d

Products distribution (%) d

Ref.

b

6a-6a’

7a7a’

8

1a

3a

none

1

180/10/5

20

90

10

none

20

180/10/24

84

63

35

2

3%Y/SBA-3

14

110/1/3

97

8

30

62

CaSn(OH)6

10

70/1/0.25

68

77

23

none

1

180/8/5

73

43

12

1

32

12

none

5

180/8/5

73

54

16

1

22

7

none

20

180/8/24

100

2a

none

20

300/50/5

78

2f

none

20

300/50/5

100

2 2a Batch

3 4 5 6

2f

7 8e

Cont.flow

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

9e

41 43

100 65

33

2 100

a

The catalyst (if present) and reactions conditions (batch or continuous-flow) used. b Q = Glycerol:ester molar ratio. c For batch reactions, p was the autogenous pressure in the autoclave reactor (entries 1-2, and 5-7). d Conversion of glycerol and products distribution determined by GC/MS. e Under continuous-flow conditions, a mixture of glycerol, ester and diglyme as a cosolvent in a 1:20:33 molar ratio, respectively, was used. The equilibria involved in the consecutive transesterification steps were clearly perturbed by the increasing amount of MeOH originated as a reaction co-product. In analogy to previously reported studies for catalytic processes,75 the reversible nature of these reactions was confirmed when a solution of triacetin and methanol in a 1:10 molar ratio, respectively, was set to react at 180 °C for 5 hours. A mixture of diacetin, monoacetin, and glycerol in 17, 1, and 1 % amounts, respectively, was achieved.

ACS Paragon Plus Environment

21

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

Page 22 of 35

As in the case of acetals, the HT-conversion of glycerol was higher with iPAc (2f) than with methyl acetate. At 180 °C and moderate Q ratios (1-5), batch reactions of 2f yielded complex mixtures including not only acetins, but also Solketal (1a) and its methyl ester (3a) (entries 6-7). Both 1a and 3a came from a competitive acetalization of glycerol induced by the acetone released during the transesterification with iPAc (Schemes 2 and 5). This was proved by the reaction of a mixture of glycerol and acetone in a 1: 10 molar ratio, respectively: at 180 °C, after 5 hours, a 70% conversion was achieved with full selectivity towards solketal 1a. A larger iPAc excess substantially inhibited the acetalization reaction and, at the same time, the irreversible loss of acetone from 2f allowed not only a quantitative process, but also the exhaustive acetylation of glycerol towards triacetin 8 (entry 8, Q=20). This derivative was isolated in 99% yield. In the continuous-flow mode, mixtures of mono-, di-, and tri-acetin products were still obtained with methyl acetate (entry 8), while the reaction of 2f proved successful for the selective synthesis of triacetin (entry 9). Although preliminary, these data further proved the concept, demonstrating that iPAc was not only effective for the conversion of GAs into the corresponding methyl esters, but also for the straightforward transesterification of glycerol into its fully acetylated derivative triacetin. This outcome also raised two challenging perspectives for the implementation of: i) the CFtransesterification of crude glycerol as received from the biodiesel production. The here presented HT-strategy would avoid costly techniques required for the purification of off-grade glycerol and common drawbacks due to the poisoning of catalysts by the impurities of the crude reagent (salts, soaps, etc.).76 ii) a selective one-pot preparation of solketal esters (3a and its derivatives) by reacting glycerol with both iPAc and its higher enol ester homologues. These subjects are currently under investigation in our lab.

ACS Paragon Plus Environment

22

Page 23 of 35 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

CONCLUSION In conclusion, the here described HT-processes exemplify an archetype of clean reactions for the conversion of glycerol and its acetals to their ester derivatives, by using innocuous reagents and producing minimal, if any, wastes. The overall protocol discloses a new perspective for the upgrading of renewables, though for a genuine sustainable path, the procedure should be conveniently integrated within a biorefinery plant where synthetic operations may take advantage of modern technologies for the recovery of waste/excess heat and the recycle of reactants/solvents.

ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI:--------. Description of batch and continuous-flow apparatus and full spectral characterization (1H, 13C NMR and MS) of compounds (PDF) AUTHOR INFORMATION Corresponding Author Prof. Maurizio Selva. Web: http://www.unive.it/persone/selva. e-mail: [email protected] Author Contributions Maurizio Selva conceived the experiments, and he wrote the paper along with Roberto Calmanti; Roberto Calmanti and Manuele Galvan performed the reactivity experiments; Alvise Perosa and Emanuele Amadio contributed to the interpretation and discussion of data and results.

ACS Paragon Plus Environment

23

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

Page 24 of 35

ABBREVIATIONS CF, continuous-flow; iPAc, isopropenyl acetate; GAs, glycerol acetals; τ, residence time.

REFERENCES

(1) Bozell, J. J.; Petersen, G. R. Technology development for the production of biobased products from biorefinery carbohydrates-the US Department of Energy’s “Top 10” revisited. Green Chem. 2010, 12 (4), 539–554, DOI: 10.1039/B922014C. (2) Farmer, T. J.; Mascal, M. In Introduction to Chemicals from Biomass, 2nd Ed., J. Clark and F. Deswarte, Eds., John Wiley & Sons, Ltd, Chichester, 2015, UK, DOI: 10.1002/9781118714478. (3) Pagliaro, M.; Rossi, M. In The Future of Glycerol, 2nd Ed. Royal Society of Chemistry, 2010, DOI: 10.1039/9781849731089. (4) Len, C.; Luque, R. Continuous flow transformations of glycerol to valuable products: an overview. Sustain. Chem. Process. 2014, 2 (1), 1, DOI: 10.1186/2043-7129-2-1. (5) Tan, H. W.; Abdul Aziz, A. R.; Aroua, M. K. Glycerol production and its applications as a raw material: A review. Renew. Sustain. Energy Rev. 2013, 27, 118–127, DOI: 10.1016/j.rser.2013.06.035. (6) Zhou, C.-. H.; Beltramini, J. N.; Fan, Y.-. X.; Lu, G. Q. Chemoselective catalytic conversion of glycerol as a biorenewable source to valuable commodity chemicals. Chem. Soc. Rev. 2008, 37, 527-549, DOI: 10.1039/B707343G.

ACS Paragon Plus Environment

24

Page 25 of 35 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

(7) Fabbri, D.; Bevoni, V.; Notari, M.; Rivetti, F. Properties of a potential biofuel obtained from soybean oil by transmethylation with dimethyl carbonate. Fuel 2007, 86 (5–6), 690–697, DOI: 10.1016/j.fuel.2006.09.003. (8) Su, E. Z.; Zhang, M. J.; Zhang, J. G.; Gao, J. F.; Wei, D. Z. Lipase-catalyzed irreversible transesterification of vegetable oils for fatty acid methyl esters production with dimethyl carbonate as the acyl acceptor. Biochem. Eng. J. 2007, 36 (2), 167–173, DOI: 10.1016/j.bej.2007.02.012. (9) Saka, S.; Isayama, Y. A new process for catalyst-free production of biodiesel using supercritical methyl acetate. Fuel 2009, 88 (7), 1307–1313, DOI: 10.1016/j.fuel.2008.12.028. (10) Niza, N. M.; Tan, K. T.; Lee, K. T.; Ahmad, Z. Biodiesel production by non-catalytic supercritical methyl acetate: Thermal stability study. Appl. Energy 2013, 101, 198–202, DOI: 10.1016/j.apenergy.2012.03.033. (11) Rathore, V.; Tyagi, S.; Newalkar, B.; Badoni R. P. Glycerin free Synthesis of Jatropha and Pongamia Biodiesel in Supercritical Dimethyl and Diethyl Carbonate. Ind. Eng. Chem. Res., 2014, 53 (26), 10525–10533, DOI: 10.1021/ie5011614. (12) It should be noted that due to the presence of water and FFA in natural oils, transesterification reactions with both DMC and methyl acetate (Scheme 1) also yielded minor amounts of glycerol carbonate and acetic acid, respectively. (13) Tan, K. T.; Lee, K. T. A review on supercritical fluids (SCF) technology in sustainable biodiesel production: Potential and challenges. Renew. Sustain. Energy Rev. 2011, 15 (5), 2452– 2456, DOI: 10.1016/j.rser.2011.02.012.

ACS Paragon Plus Environment

25

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

Page 26 of 35

(14) Marx, S. Glycerol-free biodiesel production through transesterification: A review. Fuel Process. Technol. 2016, 151, 139–147, DOI: 10.1016/j.fuproc.2016.05.033. (15) Tang, Y.; Yan, T.; Shen, B.; Li, H.; Jeje, A. Synthesis of no-glycerol biodiesel through transesterification catalyzed by CaO from different precursors. Can. J. Chem. Eng. 2016, 94 (8), 1466–1471, DOI: 10.1002/cjce.22543. (16) https://www. biofuelsystems.com/biodiesel/specification.htm (last access March 8th, 2017). (17) Casas, A.; Ruiz, J. R.; Ramos, M. J.; Pérez, A. Effects of triacetin on biodiesel quality. Energy and Fuels 2010, 24 (8), 4481–4489, DOI: 10.1021/ef100406b. (18) Ng, R. T. L.; Tay, D. H. S.; Ng, D. K. S. Simultaneous Process Synthesis, Heat and Power Integration in a Sustainable Integrated Biorefinery. Energy & Fuels 2012, 26 (12), 7316–7330, DOI: 10.1021/ef301283c. (19) Marulanda, V. F. Biodiesel production by supercritical methanol transesterification: process simulation and potential environmental impact assessment. J. Clean. Prod. 2012, 33, 109–116, DOI: 10.1016/j.jclepro.2012.04.022. (20) Glisic, S.; Skala, D. The problems in design and detailed analyses of energy consumption for biodiesel synthesis at supercritical conditions. J. Supercrit. Fluids 2009, 49 (2), 293–301, DOI: 10.1016/j.supflu.2008.12.011. (21) Van Kasteren, J. M. N.; Nisworo, A. P. A process model to estimate the cost of industrial scale biodiesel production from waste cooking oil by supercritical transesterification. Resour. Conserv. Recycl. 2007, 50 (4), 442–458, DOI: 10.1016/j.resconrec.2006.07.005. ACS Paragon Plus Environment

26

Page 27 of 35 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

(22) Selva, M.; Guidi, S.; Noè, M. Upgrading of glycerol acetals by thermal catalyst-free transesterification of dialkyl carbonates under continuous-flow conditions. Green Chem. 2015, 17 (2), 1008–1023, DOI: 10.1039/C4GC01750A. (23) Selva, M.; Perosa, A.; Guidi, S.; Cattelan, L. In Encyclopedia of Inorganic and Bioinorganic Chemistry,

Wiley

&

Sons,

Ltd:

Chichester,

UK,

2016,

pp321-333,

DOI:

10.1002/9781119951438.eibc2428. (24) Guidi, S.; Calmanti, R.; Noè, M.; Perosa, A.; Selva, M. Thermal (catalyst-free) transesterification of diols and glycerol with dimethyl carbonate: A flexible reaction for batch and continuous-flow applications. ACS Sustain. Chem. Eng. 2016, 4 (11), 6144–6151, DOI: 10.1021/acssuschemeng.6b01633. (25) Ilham, Z.; Saka, S. Esterification of glycerol from biodiesel production to glycerol carbonate in non-catalytic supercritical dimethyl carbonate. Springerplus 2016, 5 (1), 923, DOI: 10.1186/s40064-016-2643-1. (26) Venkatesha, N. J.; Bhat, Y. S.; Prakash, B. S. J. Volume accessibility of acid sites in modified montmorillonite and triacetin selectivity in acetylation of glycerol. RSC Adv. 2016, 6 (51), 45819–45828, DOI: 10.1039/C6RA05720A. (27) Huang, M. Y.; Han, X. X.; Hung, C. Te; Lin, J. C.; Wu, P. H.; Wu, J. C.; Liu, S. Bin. Heteropolyacid-based ionic liquids as efficient homogeneous catalysts for acetylation of glycerol. J. Catal. 2014, 320 (1), 42–51, DOI: 10.1016/j.jcat.2014.09.022. (28) Fukumura, T.; Toda, T.; Seki, Y.; Kubo, M.; Shibasaki-Kitakawa, N.; Yonemoto, T. Catalytic synthesis of glycerol monoacetate using a continuous expanded bed column reactor ACS Paragon Plus Environment

27

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

Page 28 of 35

packed with cation-exchange resin. Ind Eng Chem Res 2009, 48, 1816–1823, DOI: 10.1021/ie800625g. (29) Dalla Costa, B. O.; Decolatti, H. P.; Legnoverde, M. S.; Querini, C. A. Influence of acidic properties of different solid acid catalysts for glycerol acetylation. Catal. Today 2017, 289, 222– 230, DOI: 10.1016/j.cattod.2016.09.015. (30) Gorji, Y. M.; Ghaziaskar, H. S. Optimization of Solketalacetin Synthesis as a Green Fuel Additive from Ketalization of Monoacetin with Acetone. Ind. Eng. Chem. Res. 2016, 55 (25), 6904–6910, DOI: 10.1021/acs.iecr.6b00929. (31) Dodson, J. R.; Leite, T. d C. M.; S. Pontes, N.; Peres Pinto, B.; Mota, C. J. A. Green Acetylation of Solketal and Glycerol Formal by Heterogeneous Acid Catalysts to Form a Biodiesel Fuel Additive. ChemSusChem 2014, 7 (9), 2728–2734, DOI: 10.1002/cssc.201402070. (32) Perosa, A.; Moraschini, A.; Selva, M.; Noè, M. Synthesis of the fatty esters of solketal and glycerol-formal: Biobased specialty chemicals. Molecules 2016, 21 (2), 170, DOI: 10.3390/molecules21020170. (33) García, E.; Laca, M.; Pe, E.; Garrido, A.; Peinado, J. New Class of Acetal Derived from Glycerin as a Biodiesel Fuel Component New Class of Acetal Derived from Glycerin as a Biodiesel Fuel Component. Energy & Fuels 2008, 22, 4274–4280, DOI: 10.1021/ef800477m. (34) Sun, J.; Tong, X.; Yu, L.; Wan, J. An efficient and sustainable production of triacetin from the acetylation of glycerol using magnetic solid acid catalysts under mild conditions. Catal. Today 2016, 264, 115–122, DOI: 10.1016/j.cattod.2015.07.011.

ACS Paragon Plus Environment

28

Page 29 of 35 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

(35) Rezayat, M.; Ghaziaskar, H. S. Continuous synthesis of glycerol acetates in supercritical carbon dioxide using Amberlyst 15. Green Chem. 2009, 11 (5), 710–715, DOI: 10.1039/B815674C. (36) Ranu, B. C.; Dey, S. S.; Hajra, A. Highly efficient acylation of alcohols, amines and thiols under solvent-free and catalyst-free conditions. Green Chem. 2003, 5, 44–46, DOI: 10.1039/b211238h. (37) Yadav, G. D.; Joshi, A.V. A green route for the acylation of resorcinol with acetic acid. Clean Techn Env. Policy 2002, 4, 157–164, DOI: 10.1007/s10098-002-0148-9. (38) Odell, L. R.; Skopec, J.; McCluskey, A. A “cold synthesis” of heroin and implications in heroin signature analysis. Utility of trifluoroacetic/acetic anhydride in the acetylation of morphine. Forensic Sci. Int. 2006, 164 (2–3), 221–229, DOI: 10.1016/j.forsciint.2006.02.009. (39) Morales, G.; Paniagua, M.; Melero, J. A.; Vicente, G.; Ochoa, C. Sulfonic AcidFunctionalized Catalysts for the Valorization of Glycerol via Transesterification with Methyl Acetate. Ind. Eng. Chem. Res. 2011, 50 (10), 5898–5906, DOI: 10.1021/ie102357c. (40) Meireles, B. A.; Pereira, V. L. P. Synthesis of bio-additives: transesterification of ethyl acetate with glycerol using homogeneous or heterogeneous acid catalysts. J. Braz. Chem. Soc. 2013, 24, 17–57, DOI: 10.1590/S0103-50532013000100004. (41) Khayoon, M. S.; Hameed, B. H. Yttrium-grafted mesostructured SBA-3 catalyst for the transesterification of glycerol with methyl acetate to synthesize fuel oxygenates. Appl. Catal. A Gen. 2013, 460–461, 61–69, DOI: 10.1016/j.apcata.2013.03.045.

ACS Paragon Plus Environment

29

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

Page 30 of 35

(42) Ranu, B. C.; Dutta, P.; Sarkar, A. Highly selective acylation of alcohols and amines by an indium triiodide-catalysed transesterification process. J. Chem. Soc. Perkin Trans. 1 2000, No. 14, 2223–2225, DOI: 10.1039/B002181O. (43) Sandesh, S.; Kristachar, P. K. R.; Manjunathan, P.; Halgeri, A. B.; Shanbhag, G. V. Synthesis of biodiesel and acetins by transesterification reactions using novel CaSn(OH)6 heterogeneous

base

catalyst.

Appl.

Catal.

A

Gen.

2016,

523,

1–11,

DOI:

10.1016/j.apcata.2016.05.006. (44) Sutter, M.; Dayoub, W.; Métay, E.; Raoul, Y.; Lemaire, M. 1-O-Alkyl (di)glycerol ethers synthesis from methyl esters and triglycerides by two pathways: catalytic reductive alkylation and transesterification/reduction. Green Chem. 2013, 15 (3), 786-797, DOI: 10.1039/C3GC36907B. (45) Mallesha, N.; Rao, S. P.; Suhas, R.; Gowda, D. C. A green method for selective acetylation of primary alcohols using ethyl acetate and solid potassium carbonate. J. Chem. Res. 2011, 35 (9), 536–539, DOI: 10.3184/174751911X13155057300536. (46) Kharchafi, G.; Jérôme, F.; Adam, I.; Pouilloux, Y.; Barrault, J. Design of well balanced hydrophilic-lipophilic catalytic surfaces for the direct and selective monoesterification of various polyols. New J. Chem. 2005, 29 (7), 928–934, DOI: 10.1039/B418509A. (47) Vanttinen, E.; Kanerva, L. T. Lipase-catalysed transesterification in the preparation of optically active solketal. J. Chem. Soc. Perkin Trans. 1994, 1 (23), 3459–3463, DOI: 10.1039/P19940003459. (48) Sakai, T.; Kishimoto, T.; Tanaka, Y.; Ema, T.; Utaka, M. Low-temperature method for enhancement of enantioselectivity in the lipase-catalyzed kinetic resolutions of solketal and some ACS Paragon Plus Environment

30

Page 31 of 35 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

chiral alcohols. Tetrahedron Lett. 1998, 39 (43), 7881–7884, DOI: 10.1016/S00404039(98)01747-X. (49) Mine, Y.; Zhang, L.; Fukunaga, K.; Sugimura, Y. Enhancement of enzyme activity and enantioselectivity by cyclopentyl methyl ether in the transesterification catalyzed by Pseudomonas cepacia lipase co-lyophilized with cyclodextrins. Biotechnol. Lett. 2005, 27 (6), 383–388, DOI: 10.1007/s10529-005-1527-1. (50) Boncel, S.; Zniszczoł, A.; Szymańska, K.; Mrowiec-Białoń, J.; Jarzebski, A.; Walczak, K. Z. Alkaline lipase from Pseudomonas fluorescens non-covalently immobilised on pristine versus oxidised multi-wall carbon nanotubes as efficient and recyclable catalytic systems in the synthesis of

Solketal

esters.

Enzyme

Microb.

Technol.

2013,

53

(4),

263–270,

DOI:

10.1016/j.enzmictec.2013.05.003. (51) a) Zeng, T.; Song, G.; Li, C.-J. Separation, recovery and reuse of N-heterocyclic carbene catalysts in transesterification reactions. Chem. Commun. 2009, No. 41, 6249–6251, DOI: 10.1039/B910162D. (52) Pelagalli, R.; Chiarotto, I.; Feroci, M.; Vecchio, S. Isopropenyl acetate, a remarkable, cheap and acylating agent of amines under solvent- and catalyst-free conditions: a systematic investigation. Green Chem. 2012, 14 (8), 2251-2255, DOI: 10.1039/c2gc35485c. (53) Chiarotto, I. Tetraethylammonium hydrogen carbonate: A cheap, efficient, and recyclable catalyst for transesterification reactions under solvent-free conditions. Synth. Commun. 2016, 46 (22), 1840–1847, DOI: 10.1080/00397911.2016.1233343.

ACS Paragon Plus Environment

31

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

Page 32 of 35

(54) Assaf, G.; Checksfield, G.; Critcher, D.; Dunn, P. J.; Field, S.; Harris, L. J.; Howard, R. M.; Scotney, G.; Scott, A.; Mathew, S.; et al. The use of environmental metrics to evaluate green chemistry improvements to the synthesis of (S{,}S)-reboxetine succinate. Green Chem. 2012, 14 (1), 123–129, DOI: 10.1039/C1GC15921F. (55) www.sigmaaldrich.com, last access, Oct. 2017; (see, SDS). (56) March, J.; Smith, M. B.Eds. March’s Advanced Organic Chemistry, Reactions, Mechanisms, and Structure, 6th Ed., J. Wiley & Sons, 2007, DOI: 10.1002/0470084960. (57) Baev, A. K. Specific Intermolecular Interactions of Organic Compounds, Springer-Verlag Berlin Heidelberg, 2012, DOI: 10.1007/978-3-642-21622-0. (58) Vu, T. H. T.; Au, H. T.; Nguyen, T. H. T.; Nguyen, T. T. T.; Do, M. H.; Bui, N. Q.; Essayem, N. Esterification of Lactic Acid by Catalytic Extractive Reaction: An Efficient Way to Produce a Biosolvent Composition. Catal. Letters 2013, 143 (9), 950–956, DOI 10.1007/s10562-013-10774. (59) Chisholm, M. H.; Delbridge, E. E.; Gallucci, J. C. Modeling the catalyst resting state in aryl tin(iv) polymerizations of lactide and estimating the relative rates of transamidation, transesterification and chain transfer. New J. Chem. 2004, 28 (1), 145–152, DOI: 10.1039/B306700A. (60) Noyce, D. S.; Pollack, R. M. Mechanisms for the acid-catalyzed hydrolysis of vinyl acetate and isopropenyl acetate. J. Am. Chem. Soc. 1969, 91 (25), 7158–7163, DOI: 10.1021/ja01053a045.

ACS Paragon Plus Environment

32

Page 33 of 35 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

(61) Campanelli, P.; Banchero, M.; Manna, L. Synthesis of biodiesel from edible, non-edible and waste cooking oils via supercritical methyl acetate transesterification. Fuel 2010, 89 (12), 3675–3682, DOI: 10.1016/j.fuel.2010.07.033. (62) Purich, D. L.; Allison, R. D. In The Handbook of Biochemical Kinetics, Academic Press, 2000, pp75. (63) Vieitez, I.; da Silva, C.; Alckmin, I.; Borges, G. R.; Corazza, F. C.; Oliveira, J. V.; Grompone, M. A.; Jachmanián, I. Continuous catalyst-free methanolysis and ethanolysis of soybean oil under supercritical alcohol/water mixtures. Renew. Energy 2010, 35 (9), 1976–1981, DOI: 10.1016/j.renene.2010.01.027. (64) Takebayashi, Y.; Hotta, H.; Shono, A.; Yoda, S.; Furuya, T.; Otake, K. Noncatalytic orthoselective methylation of phenol in supercritical methanol: The mechanism and acid/base effect. Ind. Eng. Chem. Res. 2008, 47 (3), 704–709, DOI: 10.1021/ie0707321. (65) Horikawa, Y.; Uchino, Y.; Sako, T. Alkylation and Acetal Formation using Supercritical Alcohol without Catalyst. Chem. Lett. 2003, 32 (3), 232-233, DOI: 10.1246/cl.2003.232. (66) Roses, M.; Rafols, C.; Bosch, E. Autoprotolysis in Aqueous Organic Solvent Mixtures. Anal. Chem. 1993, 65, 2294-2299, DOI: 10.1021/ac00065a021. (67) Fonrodona, G.; Rafols, C.; Bosch, E.; Roses, M. Autoprotolysis in aqueous organic solvent mixtures. Water/alcohol binary system. Anal. Chim. Acta 1996, 335, 291-302, DOI: 10.1016/S0003-2670(96)00329-7. (68) Moity, L.; Benazzouz, A.; Molinier, V.; Nardello-Rataj, V.; Elmkaddem, M. K.; de Caro, P.; Thiébaud-Roux, S.; Gerbaud, V.; Marion, P.; Aubry, J.-M. Glycerol acetals and ketals as biobased solvents: positioning in Hansen and COSMO-RS spaces, volatility and stability towards hydrolysis and autoxidation. Green Chem. 2015, 17 (3), 1779–1792, DOI: 10.1039/C4GC02377C. ACS Paragon Plus Environment

33

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

Page 34 of 35

(69) Jas, G.; Kirschning, A. Continuous Flow Techniques in Organic Synthesis. Chem. – A Eur. J. 2003, 9 (23), 5708–5723, DOI: 10.1002/chem.200305212. (70) Wiles, C.; Watts, P. Continuous flow reactors: a perspective. Green Chem. 2012, 14 (1), 38–54, DOI: 10.1039/C1GC16022B. (71) Harriot, P. Chemical Reactor Design, Marcel Dekker Eds, New York, 2003. (72) Kwon, E. E.; Yi, H.; Jeon, Y. J. Boosting the value of biodiesel byproduct by the noncatalytic transesterification of dimethyl carbonate via a continuous flow system under ambient pressure. Chemosphere 2014, 113, 87–92, DOI: 10.1016/j.chemosphere.2014.04.055. (73) Wang, Z.; Wu, B.; Zhu, J.; Chen, K.; Wang, W. Isobaric vapor–liquid–liquid equilibrium for isopropenyl acetate + water + acetic acid at 101 kPa. Fluid Phase Equilib. 2012, 314, 152–155, DOI: 10.1016/j.fluid.2011.10.025. (74) Anastas, P. T.; Zimmerman, J. B. Through the 12 Principles of Green Engineering. Env. Sci. Tech., 2003, 37, 94A-101A, DOI: 10.1021/es032373g. (75) Lopez, D. E.; Goodwin, J. G. Jr.; Bruce, D. A.; Lotero, E. Transesterification of triacetin with methanol on solid acid and base catalysts. Appl. Catal. A: General 2005, 295, 97–105, DOI: 10.1016/j.apcata.2005.07.055. (76) Guidi, S.; Noè, M.; Riello, P.; Perosa, A.; Selva, M. Towards a Rational Design of a Continuous-Flow Method for the Acetalization of Crude Glycerol: Scope and Limitations of Commercial Amberlyst 36 and AlF3 3H2O as Model Catalysts. Molecules 2016, 21, 657, DOI: 10.3390/molecules21050657.

ACS Paragon Plus Environment

34

Page 35 of 35 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

For Table of Contents Use Only TABLE OF CONTENTS (TOC) GRAPHIC

SYNOPSIS: A sustainable approach for the valorization of glycerol and its acetal through the hightemperature transesterification with alkyl and enol esters

ACS Paragon Plus Environment

35