CO2-switchable solvents as entrainer in fluid separations - ACS

Jun 29, 2018 - CO2-switchable solvents as entrainer in fluid separations. Boelo Schuur , Mart Nijland , Marek Blahusiak , and Alberto Juan. ACS Sustai...
0 downloads 0 Views 966KB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

CO2-switchable solvents as entrainer in fluid separations Boelo Schuur, Mart Nijland, Marek Blahusiak, and Alberto Juan ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b01771 • Publication Date (Web): 29 Jun 2018 Downloaded from http://pubs.acs.org on July 9, 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 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 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.

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 24 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

CO2-switchable solvents as entrainer in fluid separations Boelo Schuur†,*, Mart Nijland†, Marek Blahušiak†, Alberto Juan‡,¥ †

University of Twente, Sustainable Process Technology Group, Meander building 221, PO Box 217, 7500AE Enschede, The Netherlands, e: [email protected]



University of Twente, Biomolecular Nanotechnology Group, MESA+ institute, Carré building 4326, PO Box 217, 7500AE Enschede, The Netherlands ¥

University of Twente, Molecular Nanofabrication Group, MESA + institute, Carre building 4326, PO Box 217, 7500 AE Enschede, The Netherlands

Abstract CO2-switchable solvents, typically neutral solvents that switch with CO2 into ionic species, were investigated for use as entrainer in fluid separations such as extractive distillation. Their switchable nature was investigated, which may facilitate liquid-liquid extraction or extractive distillation as ionic liquid (IL), whereas during regeneration their decarboxylation into the amine form prevents temperature shoot-up. Studied elements included a property screening, detailed mechanistic and kinetic studies on the switching of 2-ethylhexylamine and N,N-benzyl methylamine. Decarboxylation of a 50 vol% switchable solvent, 25 vol% heptane and 25 vol% toluene mixture at 1.00 104 Pa showed a 40 % CO2 release before reaching the operational pressure, and total decarboxylation took over two hours. An effective increase of the relative volatility of heptane/toluene was found, showing that indeed, CO2-switchable solvents can be applied for extractive distillation. However, low pressure in combination with elevated temperature will lead to quick decarboxylation, limiting the operational window of this class of solvents. Their use in low temperature application such as C4-distillations or liquid-liquid extraction appears more suited.

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

Keywords CO2-Switchable solvents, extractive distillation, liquid-liquid extraction, 2-ethylhexyl amine, benzylmethylamine, entrainer

Introduction Where distillation has been the working horse for separations in the chemical industry for ages, and energy costs in industry amount to about 40-50% of the total costs,1 during recent years the awareness grew that we should become more energy efficient to reduce CO2emissions significantly,2-3 and research has been focused on both replacing traditional separations with e.g. affinity separations,2-4 and on replacing traditional petrochemical production routes with bio-based routes.5-7 For many separation challenges in biorefineries, distillation is not technically feasible,1 and solvent-based approaches may provide sustainable solutions. For a solvent-based separation process to be sustainable, it is essential that a good energy efficiency can be reached and that the solvent is environmentally benign. A class of solvents that has often been claimed as green solvent class is that of the ionic liquids (ILs).8 Especially the negligible volatility is mentioned often as beneficial feature,9 because this eliminates losses through the air, such as with volatile organic solvents. This is indeed true, but this also limits their distillability. The boiling point of mixtures depends on their composition, and when an IL is to be regenerated to high purity, the boiling point will approach that of the IL itself, and since they have a negligible vapor pressure, this will result in a sharp temperature rise, and full regeneration through distillation is simply not possible. Alternatively, IL regeneration by stripping can be applied when the solutes are volatile, but then also large amounts of strip gas are needed for deep regeneration.

ACS Paragon Plus Environment

Page 2 of 24

Page 3 of 24 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

Ideally, an IL could be distillable in the solvent regeneration stage, while in the primary separation stage this is not necessary or even undesirable. A group of solvents that is also known as distillable ILs are the CO2-switchable solvents.10 CO2-switchable solvents switch their behavior upon reaction with CO2,11 affecting the polarity of the solvent. Since 2005, a range of papers has appeared on the use of such solvents, mostly for the extraction of lipids from various sorts of biomass, e.g. soy bean 12 and microalgae,13-14, but also as a reversible protecting group,15 and to switch the ionic strength in aqueous solutions.16 Using their IL character, CO2-switched amines may potentially be applied in e.g. the separation of aliphatics and aromatics by extractive distillation or liquid extraction. The non-volatile character, that normally is a drawback for thermal IL regeneration, can be lost by backswitching, enhancing their distillability. A conceptual process flow diagram of such an extractive distillation process is displayed in Figure 1. In this diagram, after the primary separation, i.e. the extractive distillation, a recovery column is placed, and the CO2 escaping from both the extractive distillation column (EDC) and the recovery column (RC) is sent with the recovered solvent to the carbamate regeneration column (CRC) to regenerate the solvent in its carbamate form. The CO2-switchable solvents might thus be applied in a much wider range of fluid separations than the extraction of natural products. We here report a study on the applicability of this highly interesting class of solvents for other fluid separations than the known lipid extractions, and in our studies, we have investigated on the physical properties of a series of switched amines. Among the most interesting ones from property point of view (viscosity/physical state of switched IL form and boiling point of unswitched amine form) are 2-ethylhexyl amine (2EHA), dibutylamine (DBA) and benzylmethylamine (BMA). For these amines we have studied their application in extractive distillation of heptane-toluene

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

mixtures and for 2EHA and BMA also in detail the mechanism of switching and the rates of switching at temperatures up to 75oC using 1H-NMR and 13C-NMR spectroscopy.

Figure 1. Conceptual process scheme for extractive distillation with carbamate solvent including three columns, EDC = extractive distillation column, RC = recovery column, and CRC = carbamate regeneration column.

Materials and methods Materials All chemicals used were commercially obtained at Sigma-Aldrich, Cambridge Isotope labs (deuterated solvents), or Praxair (CO2). Vapor-liquid equilibrium measurements

ACS Paragon Plus Environment

Page 4 of 24

Page 5 of 24 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

Vapor-liquid equilibria (VLE) were measured in an ebulliometer from Fischer, type VLE 602 at an operational pressure of 104 Pa. After setting the pressure, the temperature was measured and compositional analysis of both the liquid phase and the condensed vapor was performed using GC. Prior to analyzing mixtures, the equipment was validated by measuring pure component boiling points of heptane and toluene at the operational pressure of 104 Pa. Samples of the liquid phase were also analyzed on the solvent composition using quantitative 13C-NMR. In experiments with a solvent, the solvent was applied in a solvent-tofeed ratio S/F = 1 (mass based).

Variable temperature 1H-NMR studies Variable temperature 1H-NMR experiments were performed on a Bruker Avance II 600 MHz provided with an external BCU-Xtreme cooling unit to study on the stability of the ammonium carbamates and on the mechanism of the back-switching from the ammonium carbamates to the amine forms of the solvents. Typically, an amine was bubbled extensively with CO2 (up to 4 h) before mixing it with toluene-d8 and starting the measurement over temperatures ranging from -30 oC to 75 oC for BMA, and from 30 oC to 75 oC for 2-EHA.

Quantitative 13C-NMR measurements Quantitative 13C-NMR measurements were performed on the same machine as the variable temperature 1H-NMR studies using inverse gate coupling mode and a d1 of 2 seconds. After taking the samples from the ebulliometer, they were directly dissolved in toluene-d8 and the NMR tube closed so that no CO2 could escape during the measurement. Results and discussion

Solvent pre-screening

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

Page 6 of 24

The use of CO2-switchable solvents as entrainer in extractive distillation processes is conditional to several solvent properties. Key properties that have been reported for a range of solvents include the boiling point of the amine form, the technical feasibility of switching and the physical appearance of both amine form and the switched form. Ideally, for extractive distillation, both the amine form and the ammonium carbamate form appear as liquid, and the boiling point of the amine form should still be significantly (preferably close to or more than 50ºC) above the boiling points of the mixture to separate. For the separation of heptane and toluene, this means ideally a solvent with a natural boiling temperature close to or above 160 ºC is preferred. Based on literature on CO2-switchable solvents a list of possible solvents with proven switchability is presented in Table 1. In this table, only single amines are listed, as CO2-switchable solvents with two components (e.g. guanidines with alcohols11) were regarded as too complicated. Furthermore, tertiary amines requiring water in the switching chemistry were also not considered. From Table 1, the top three solvents were selected as they appeared most appropriate with high enough boiling points of the amine form, no solidification of the ammonium carbamate form and still reasonable viscosities (exact values were not measured). However, DBA did show gelation upon switching, which might hinder process ability. Due to the gelation, it was decided to limit extensive NMR-studies to the other two promising candidates, i.e. 2EHA and BMA.

Table 1. Property evaluation of CO2 switchable solvents with proven switchability Solvent (amine

Tb /

Amine

Ammonium

form)

ºC

appea-

carbamate

rance

appearance

Comment

ACS Paragon Plus Environment

Ref.

Page 7 of 24 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 25 ºC at 25 ºC Benzylmethylamine 184-

L

L

All criteria met

17

(BMA)

186

2-ethylhexylamine

169

L

L

All criteria met

18

Dibutylamine (DBA) 159

L

Gel

Might be suitable, Tb amine

18

(2EHA)

form suitable, gelation upon switching might hinder application N-ethyl-N-

108

L

L

butylamine (EBA)

Unsuitable for this

17-18

application, Tb amine form too low

N-ethyl-N-

79

L

L

propylamine (EPA

Unsuitable for this

17

application, Tb amine form too low

Dipropylamine

105-

(DPA)

110

L

L

Unsuitable for this

17-18

application, Tb amine form too low

N-ethyl-

36-37

L

L

methylamine

Unsuitable for this

17

application, Tb amine form too low

Diethylamine

55

L

L

Unsuitable for this application, Tb amine form

ACS Paragon Plus Environment

17

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 24

too low N-methyl-

61-63

L

L

propylamine

Unsuitable for this

17

application, Tb amine form too low

Dimethylamine

7

G

L

Unsuitable for this

17

application, Tb amine form too low n-hexylamine

131

L

S

Unsuitable due to

17

solidification upon switching N-tert-butyl-

98

L

S

isopropylamine

Unsuitable due to

17

solidification upon switching

1,3-dibutyl-

49-51

L

S

methylamine

Unsuitable due to

17

solidification upon switching

Allylamine

53

L

S

Unsuitable due to

17

solidification upon switching Piperidine

106

L

S

Unsuitable due to

17

solidification upon switching Pyrrolidone

245

L

S

Unsuitable due to

ACS Paragon Plus Environment

17

Page 9 of 24 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

solidification upon switching Benzylamine

185

L

S

Unsuitable due to

17

solidification upon switching

Solvent switching characteristics When the switchable solvents are applied in extraction,12-13 typically the difference in behavior of the systems between the neutral state and the switched state is studied,13 and the polarity before and after switching is measured, e.g. using the shift in wavelength of a light absorbance maximum of a photochromic dye.13 Also the conditions for switching have been reported, and switching from the ammonium carbamate form back to the amine form is preferred at elevated temperature,13 which consistent with CO2 capture and release by amine solvents and sorbents, that also require an elevated temperature for release.19-20 For potential application in extractive distillations, this might not be ideal, and in order to evaluate the applicability in such separation processes, it is essential to obtain a good insight in both the rate of switching and the mechanism of switching in an environment that is representative for extractive distillation conditions. Therefore, mixtures containing ammonium carbamates of 2EHA and of BMA and deuterated toluene were analyzed by 1HNMR in variable temperature conditions. The measurement results for the CO2-switched 2EHA are displayed in Figure 2. At 30 ºC, characteristic signals appear at 9.1 ppm, 5.2 ppm, 3.2 ppm, and 2.8 ppm. The signals around 7 ppm and 2.1 ppm belong to toluene. The signals

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

below 2 ppm belong to the hydrocarbon tails of 2EHA and its CO2-switched form, and are not displayed here because they are not relevant for studying the mechanism. 1H-NMR, 13CNMR and 2D heteronuclear single quantum coherence (HSQC) spectra of the amine form are presented in the supplementary information, as well as superimposed 1H-NMR spectra before and after CO2 bubbling. Upon increasing the temperature, different 1H-NMR spectra were recorded in the range of 30 – 70 ºC. Afterwards the sample was allowed to cool down to 30 ºC. Both spectra recorded at 30 ºC, initial and after variable temperature experiments, were identical, indicating that the process in the NMR-tube during variable temperature experiments was perfectly reversible. The reversibility was possible due to the use of a sealed NMR-tube, keeping all CO2 that may have liberated from the liquid phase in the tube. Furthermore, it can be seen that the signals at 2.8 ppm, 3.2 ppm and 5.2 ppm maintain their position in the spectrum, but do broaden at increasing temperature. The signal at 9.1 ppm shifts down to 5.3 ppm. Similar signal changes were observed for the 1H-NMR spectra of 2ethoxyethylamine by Kortunov et al.,21 who studied in situ 1H-NMR while CO2-switching the amine. In analogy to the mechanism that Kortunov et al. proposed for 2-ethoxyethylamine, we propose the switching mechanism for 2EHA as displayed in Scheme 1.

ACS Paragon Plus Environment

Page 10 of 24

Page 11 of 24 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 2. Stacked 1H-NMR spectra of 125 mg CO2-switched 2EHA in 0.55 mL toluene-d8 at varying temperature. Upon uptake of CO2 by the amine, a zwitterionic intermediate is formed, which converts to carbamic acid and the acidic proton of carbamic acid can be transferred to a free amine, thereby generating ammonium carbamate. Alternatively, the carbamate can be formed by proton transfer from the zwitterionic intermediate.

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

O

O

+ CO2 R

RNH2

N H2

R

O

N H

OH

carbamic acid

zwitterionic intermediate

primary amine

Page 12 of 24

+ RNH2 O 3

5 R=

6

2 4

1

7 8

R

N H

O

+

RNH3

ammonium carbamate

Scheme 1. Proposed mechanism of switching for 2EHA. The 5.2 ppm signal, assigned to the RNH-COOH and RNH-COO- is relatively sharp at 30 ºC, indicating that the equilibrium at that temperature favors the ammonium carbamate form, whereas the broadening at higher temperature is so strong that the signal is not visible any more above 65 ºC, indicating fast exchange between multiple species, i.e. the zwitterionic intermediate, the carbamic acid and the ammonium carbamate. A second indication for this fast proton exchange at higher temperatures is the broadening of the 3.2 ppm signal for the protons on C1. The signal shifting from 9.1 ppm to 5.3 ppm is an averaged contribution of the RNH3+, RNH2 and RNHCOOH protons. The signal starts relatively sharp at 9.1 ppm, then broadens up to 65 ºC, and then sharpens again. Sharpening of the signal is explained by less exchange, i.e. one of the forms is becoming the more stable form, and equilibrium shifts strongly towards that form. From Figure S2 in the supplementary information, it can be seen that the free amine RNH2 protons show a signal at 0.53 ppm, indicating that this species alone is in the closed environment of the NMR tube not the stable species. In Figure S2 it can be seen as well that CO2-switched 2EHA does not show a signal at 0.53ppm. Most likely, due to the increasing partial pressure of CO2, a total shift towards the free amine is not

ACS Paragon Plus Environment

Page 13 of 24

happening, but an equilibrium has established between the carbamic acid and the amine with fast proton exchange compared to the NMR timescale. In order to validate that indeed a relatively high partial pressure of CO2 is the cause of the limited decarboxylation towards the amine form, two types of experiments were done. In the first type, the total concentration of CO2-switched 2EHA was reduced to 75 mg and 25 mg, respectively in 0.55 mL toluene, and the same temperature program was run. These results are displayed in Figure 3. Secondly, a set of two experiments were done at 70 ºC and in which the characteristic signal for the RNH3+, RNH2 and RNHCOOH protons was followed in time. In the first experiment, the NMR-tube remained sealed, but in the second experiment the tube was taken out of the NMR spectrometer every 20 minutes to flush with N2 for one minute. These results are displayed in Figure 4.

80 70

T / oC

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

60 50 40 30 10

9

8

7

6

5

4

3

2

1

0

chemical shift / ppm Figure 3. Shifting of the signal for the RNH3+, RNH2 and RNHCOOH protons with temperature. Squares represent 125 mg CO2-switched 2EHA in 0.55 mL toluene-d8, circles 75 mg and triangles 25 mg.

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

In Figure 3 it can be seen that at lower CO2-switched 2EHA concentration the change in the chemical shift is much larger, even approaching the chemical shift of the unswitched 2EHA. This is because at the lower concentration of CO2-switched 2EHA, also less CO2 is liberated upon fully back-switching to the neutral amine, and the partial pressure of CO2 in the gas phase in the tube will be less limiting for the back-switching. The second type of experiment, in which N2-sparging was applied with regular intervals to a tube containing the higher 125 mg CO2-switched 2EHA in 0.55 mL toluene-d8, confirms that the partial pressure of CO2 in the tube is important for the back-switching, because in the experiment without N2-sparging the chemical shift is dropping gradually, while in the experiment in which N2-sparging was applied there are stepwise changes visible for the chemical shift. In this N2-sparged tube, the partial pressure of CO2 reached much lower values directly after several purges, allowing the step-wise further back-switching towards the neutral amine.

ACS Paragon Plus Environment

Page 14 of 24

Page 15 of 24 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 4. Chemical Shift versus time for 125 mg CO2-bubbled 2-EHA sparged with N2 (squares), and nonsparged (circles).

Next to the behavior of the CO2-switched primary amine 2EHA, also the switching behavior of the CO2-switched secondary amine BMA was investigated using 1H-NMR variable temperature experiments. Similar to the trend seen for 2EHA, also for BMA upon increasing the temperature from 30 – 75 ºC, there is a strong shift of the signal that at 30 ºC can be found at 8.5 ppm towards 4.0 ppm at 75 ºC. For this sample, to further elucidate on the operational window, spectra were recorded as low as -30 ºC. The stacked 1H-NMR spectra from -30 ºC to 75 ºC are displayed in Figure 5, and the proposed reaction scheme in Scheme 2.

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 5. Stacked 1H-NMR spectra of 100 mg CO2-switched BMA in 0.55 mL toluene-d8 at varying temperature. At 30 ºC the characteristic switching signal representing protons from R1NCOOHR2, R1NH2+R2, and R1NHR2 can be found at 8.6 ppm, and it is a relatively sharp signal. Upon cooling down to -30 ºC this signal shifts to 10.8 ppm and becomes much less pronounced, this is typical for acidic proton signals. When going up in temperature to 75 ºC, an opposite shift is observed to 4.0 ppm, and also here the signal is broadening. The broadening of the signal at higher temperatures is due to faster reactions and species exchange between carbamic acid and the secondary amine. The characteristic signals for the benzylic protons adjacent to N and the CH3-protons appear at 4.5 and 3.5 ppm, respectively, as well as at 3.0 and 2.0 ppm, respectively. Also these signals broaden at higher temperature, indicating the fast species exchange.

Scheme 2. Proposed switching mechanism for BMA. Based on the NMR studies, it is clear that at higher temperature carbamic acid (non-ionic form) is present in significant amounts, but when the CO2 is not actively removed from the gas phase, reverting back to the amine form is a slow process hindered by thermodynamic

ACS Paragon Plus Environment

Page 16 of 24

Page 17 of 24 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

equilibrium. It may be interesting to study these CO2-switchable solvents in extractive distillation applications, and therefore we have executed vapor-liquid equilibrium measurements in which CO2-switched 2EHA and CO2-switched BMA were applied as entrainer.

Vapor-liquid equilibrium measurements In order to measure whether the selected solvents have an effect on the VLE behavior of binary mixtures, the VLE of n-heptane – toluene mixtures were measured in an ebulliometer. Both 2EHA and BMA were applied in their amine form, i.e. without being pretreated with CO2, as well as in their ammonium carbamate form after bubbling extensively with CO2. In a typical experiment done with 2EHA, samples were analyzed with quantitative 13C-NMR to follow the rate of back-switching in the ebulliometer. To minimize errors due to slow relaxation of carbon spins, the quantification was done for carbons 1,2, and 8 (see Figure S4 in the electronic supplementary information for the numbering). The results are shown in Figure 6. Because of the time between sampling and finalization of the 13C-NMR experiments, these results have to be interpreted with caution, as the percentage of switched amines at the moment of sampling was likely higher than displayed. Nevertheless, the graph shows that while reducing the pressure from 105 Pa down to 104 Pa, already a significant percentage of the solvent switched back to the amine state. These results indicate that in order to maintain the switched form longer, a significant partial CO2-pressure is desired. To compare the relative volatility in the ternary systems with the binary mixture VLE, compositions of n-heptane and toluene were expressed as pseudo-binary mixture in which the solvent was disregarded. The VLE results are depicted in Figure 7, the estimated error in

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

the composition of ∆x = ∆y = 0.02 is due to analysis error (0.017), as determined over 196 analyses, and the error in weighing.

percentage CO2-loaded 2EHA

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 18 of 24

100 80 60 40 20 0 0

20

40

60

80

100

120

time / min Figure 6. Percentage of the switched solvent that is still in CO2-switched state during the 13CNMR measurement. At t = 0 there are two data points, one for the initial mixture, one after the operational temperature and pressure of 104 Pa had been achieved. Symbols, squares = C1, circles = C2, triangles C8.

ACS Paragon Plus Environment

Page 19 of 24 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. Vapor liquid equilibrium diagram for n-heptane – toluene at 100 mbar. The line has been simulated in ASPEN Plus using NRTL-LIT property package. The experimental data points are all pseudo-binary points (heptane + toluene = 1, solvent not taken into account), and have all been obtained using a S/F ratio of 1. Symbols correspond to experiments with 2EHA (open squares), CO2switched 2EHA (closed squares), BMA (open triangle), CO2-switched BMA (closed triangles). The results for the CO2-switched solvents are operational vapor liquid correspondence data, because due to escaping CO2, the composition and temperature changed slowly in time and the system was not completely at thermodynamic equilibrium.

It can be seen in Figure 7 that the line representing the VLE of binary n-heptane – toluene mixtures at 100 mbar shows a tangent pinch at high molar fractions of heptane, i.e. the relative volatility α approaches unity at compositions approaching x = y = 1. When using

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

2EHA (open squares) and the CO2-switched 2EHA (closed squares) as solvent in S/F = 1 there is an improvement visible of the relative volatility, and at high molar fractions of heptane the improvement is best visible. However, the improvement in the relative volatility is with these solvents only marginal. Samples from heptane/toluene mixtures with CO2-switched 2EHA appeared hazy, indicating that there is no perfect miscibility under the operational conditions. This likely affects the effectivity of the solvent. The use of BMA (open triangle) improves the relative volatility of n-heptane over toluene more significant, and this improvement is even more pronounced for the CO2-switched BMA (closed triangles). Samples from these mixtures were always clear, showing perfect miscibility of this system. Because of the labor intensity of VLE measurements, and the results with 2EHA indicating that the region x > 0.7 is the most important region, with BMA and CO2-switched BMA only measurements were done starting at x = 0.8. The relative volatility reaches α = 2.3 with BMA and α = 2.7 with CO2-switched BMA. Although these values for the relative volatility are promising and showing that indeed, CO2switchable solvents can be applied as entrainer in extractive distillation, the data series for CO2-switched BMA represents only one experiment, and the reported relative volatility is an operational relative volatility, and not a full thermodynamic equilibrium, because the temperature and composition kept changing with time. This is due to the mode of operation of the ebulliometer, in which the pressure is set, and the temperature monitored. Because of back-switching of CO2-switched BMA during the experiment, CO2 was released resulting in an increasing pressure in the system. This increasing pressure then triggered the pressure control to activate the pump. As a result, over time the solvent was changing from (almost) pure CO2-switched BMA towards BMA. At the same time, upon activation of the pump, also vapors of n-heptane and toluene were pumped out of the system, and because n-heptane is

ACS Paragon Plus Environment

Page 20 of 24

Page 21 of 24 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

more volatile, the fraction n-heptane reduced during the experiment. A similar effect was observed for the three experiments with CO2-switched 2EHA, starting with x = 0.8, x = 0.5 and x = 0.2, respectively. From these results it appears that CO2-switching of BMA can increase the relative volatility in the n-heptane – toluene system as compared to nonswitched BMA, and application of such switchable solvents in affinity separations such as extractive distillation and liquid-liquid extraction can open new windows of opportunities for sustainable separations. To longer maintain the switched form, and to push the equilibrium towards the carbamate, however, lower temperature is desired and for application as entrainer in extractive distillation, fractionations of lower boiling C4-fractions appear more beneficial. Also for liquid-liquid extraction the lower temperature may be applied, which means that yes, CO2-switchable solvents have applicability in fluid separations wider than thus reported, but back-switching rates narrow the operational window to low temperature applications. Follow-up studies should focus on further optimizing the molecular structure of the CO2-switchable solvents in relation to the eventual application to reach higher relative volatilities and become competitive with current solvents that are less benign such as NMP and sulfolane. Supporting information Additional supporting information is provided, including 2D-HSQC analysis and NMR spectra of pure compounds. Acknowledgements We acknowledge prof. Jurriaan Huskens and dr. Wim Brilman for fruitful discussions. Present address

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

Marek Blahusiak is currently working at the Slovak University of Technology in Bratislava, Slovakia. Email: [email protected] References 1. Kiss, A. A.; Lange, J. P.; Schuur, B.; Brilman, D. W. F.; van der Ham, A. G. J.; Kersten, S. R. A., Separation technology–Making a difference in biorefineries. Biomass Bioenergy 2016, 95, 296-309. 2. Sadiq, M. M.; Suzuki, K.; Hill, M. R., Towards energy efficient separations with metal organic frameworks. Chem. Commun. 2018, 54 (23), 2825-2837. 3. Sholl, D. S.; Lively, R. P., Seven chemical separations to change the world. Nature 2016, 532 (7600), 435-437. 4. Blahušiak, M.; Kiss, A. A.; Babic, K.; Kersten, S. R. A.; Bargeman, G.; Schuur, B., Insights into the selection and design of fluid separation processes. Sep. Purif. Technol. 2018, 194, 301-318. 5. Alonso, D. M.; Bond, J. Q.; Dumesic, J. A., Catalytic conversion of biomass to biofuels. Green Chem. 2010, 12 (9), 1493-1513. 6. Van Putten, R. J.; Van Der Waal, J. C.; De Jong, E.; Rasrendra, C. B.; Heeres, H. J.; De Vries, J. G., Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 2013, 113 (3), 1499-1597. 7. Sun, Z.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K., Bright Side of Lignin Depolymerization: Toward New Platform Chemicals. Chem. Rev. 2018, 118 (2), 614-678. 8. Ventura, S. P. M.; e Silva, F. A.; Quental, M. V.; Mondal, D.; Freire, M. G.; Coutinho, J. A. P., Ionic-Liquid-Mediated Extraction and Separation Processes for Bioactive Compounds: Past, Present, and Future Trends. Chem. Rev. 2017, 117 (10), 6984-7052. 9. Ramdin, M.; De Loos, T. W.; Vlugt, T. J. H., State-of-the-art of CO2 capture with ionic liquids. Ind. Eng. Chem. Res. 2012, 51 (24), 8149-8177. 10. Chowdhury, S. A.; Vijayaraghavan, R.; MacFarlane, D. R., Distillable ionic liquid extraction of tannins from plant materials. Green Chem. 2010, 12 (6), 1023-1028. 11. Jessop, P. G.; Heldebrant, D. J.; Li, X.; Eckertt, C. A.; Liotta, C. L., Green chemistry: Reversible nonpolar-to-polar solvent. Nature 2005, 436 (7054), 1102. 12. Phan, L.; Brown, H.; White, J.; Hodgson, A.; Jessop, P. G., Soybean oil extraction and separation using switchable or expanded solvents. Green Chem. 2009, 11 (1), 53-59. 13. Du, Y.; Schuur, B.; Samorì, C.; Tagliavini, E.; Brilman, D. W. F., Secondary amines as switchable solvents for lipid extraction from non-broken microalgae. Biores. Technol. 2013, 149, 253-260. 14. Du, Y.; Schuur, B.; Brilman, D. W. F., Maximizing Lipid Yield in Neochloris oleoabundans Algae Extraction by Stressing and Using Multiple Extraction Stages with N-Ethylbutylamine as Switchable Solvent. Ind. Eng. Chem. Res. 2017, 56 (28), 8073-8080. 15. Peeters, A.; Ameloot, R.; De Vos, D. E., Carbon dioxide as a reversible amine-protecting agent in selective Michael additions and acylations. Green Chem. 2013, 15 (6), 1550-1557. 16. Mercer, S. M.; Jessop, P. G., "Switchable water": Aqueous solutions of switchable ionic strength. ChemSusChem 2010, 3 (4), 467-470. 17. Phan, L.; Andreatta, J. R.; Horvey, L. K.; Edie, C. F.; Luco, A.-L.; Mirchandani, A.; Darensbourg, D. J.; Jessop, P. G., Switchable-Polarity Solvents Prepared with a Single Liquid Component. J. Org. Chem. 2008, 73 (1), 127-132. 18. Tran, K. V. Lipid Extraction From Microalgae Using Switchable Hydrophilicity Solvents. MScthesis, University of Twente, 2013. 19. Brilman, D. W. F.; Veneman, R., Capturing Atmospheric CO2 Using Supported Amine Sorbents. Energy Procedia 2013, 37, 6070-6078. 20. Dutcher, B.; Fan, M.; Russell, A. G., Amine-Based CO2 Capture Technology Development from the Beginning of 2013—A Review. ACS Appl. Mater. Int. 2015, 7 (4), 2137-2148.

ACS Paragon Plus Environment

Page 22 of 24

Page 23 of 24 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

21. Kortunov, P. V.; Siskin, M.; Baugh, L. S.; Calabro, D. C., In Situ Nuclear Magnetic Resonance Mechanistic Studies of Carbon Dioxide Reactions with Liquid Amines in Non-aqueous Systems: Evidence for the Formation of Carbamic Acids and Zwitterionic Species. Energy & Fuels 2015, 29 (9), 5940-5966.

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

TOC

Synopsis The use of CO2-switchable solvents as new sustainable solvent class in fluid separations has been explored in this study.

ACS Paragon Plus Environment

Page 24 of 24