Sustainable Energy & Fuels

22 downloads 0 Views 856KB Size Report
In this study, we reported on the propylene carbonate (PC) production mediated by lanthanum oxide (La-O)/tetrabutyl ammonium bromide (TBAB) binary mixture ...
View Article Online View Journal

Sustainable Energy & Fuels Interdisciplinary research for the development of sustainable energy technologies

Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: A. K. Qaroush, F. Alsoubani, A. Al-Khateeb, E. Nabih, I. Israa Al-Ramahi, M. Khanfar, K. I. Assaf and A. F. Eftaiha, Sustainable Energy Fuels, 2018, DOI: 10.1039/C8SE00092A. Volume 1 Number 1 2017 Pages 1–100 Volume 1 | Number x | 2017

Sustainable Energy & Fuels Interdisciplinary research for the development of sustainable energy technologies rsc.li/sustainable-energy

This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication.

Sustainable Energy & Fuels

Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available. You can find more information about Accepted Manuscripts in the author guidelines.

Pages 0000–0000

Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the ethical guidelines, outlined in our author and reviewer resource centre, still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains.

rsc.li/sustainable-energy

Page 1 of 9

Please do notEnergy adjust & margins Sustainable Fuels View Article Online

DOI: 10.1039/C8SE00092A

ARTICLE An Efficient Atom-Economical Chemoselective CO2 Cycloaddition using Lanthanum Oxide/Tetrabutyl Ammonium Bromide Received 00th January 20xx, Accepted 00th January 20xx

a*

b

b

b

Abdussalam K. Qaroush, Fatima A. Alsoubani, Ala'a M. Al-Khateeb, Enas Nabih, Esraa Alb c d b Ramahi, Mohammad F. Khanfar, Khaleel I. Assaf, Ala'a F. Eftaiha*

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

DOI: 10.1039/x0xx00000x www.rsc.org/

The cycloaddition reaction between carbon dioxide (CO2) and epoxides is an attractive example for the synthesis of valueadded chemicals exploiting CO2 greenhouse emissions. In this study, we reported on the propylene carbonate (PC) production mediated by lanthanum oxide (La-O)/tetrabutyl ammonium bromide (TBAB) binary mixture. The effect of different reaction parameters including catalyst/co-catalyst loading, the amount of substrate, CO2 pressure, temperature, and reaction time were investigated. The binary catalytic system showed high catalytic conversion and selectivity. Catalyst reusability experiments indicated that the binary system was recyclable over five times without losing its activity/selectivity. In addition, a plausible reaction mechanism based on spectroscopic and kinetic measurements were presented, which was supported by density functional theory (DFT) calculations. 15

Introduction Carbon dioxide (CO2) is considered as a cheap, abundant, non-toxic, non-flammable, and chemically stable C1 chemical feedstock.1,2 It has a great potential to act as a green solvent.3 However, its increased atmospheric accumulation is the main cause to the overall global warming phenomenon,4,5 which is considered nowadays as an unsolved problem. In order to resolve this issue, it is essential to reduce the growing amounts of CO2 resulting from the anthropogenic effect.6 The mitigation policy of CO2 built-up in the atmosphere lies in three main subcategories, viz., carbon capture and sequestration (CCS), carbon capture and recycling (CCR), and carbon capture and utilisation (CCU).7–10 Catalytic conversion of CO2 into useful chemicals is of potential importance to tackle environmental problems and supporting sustainable development.11 Cyclic carbonates12–14 are one of the interesting commercially available products, that can be synthesized from CO2, due to its 100% atom economical synthesis. One of these interesting carbonates is propylene carbonate (PC), a safer alternative for hazardous dipolar aprotic solvents like hexamethylphosphoramide (HMPA), together with its implementation as an accelerator for curing silicate-systems used as foundry sand binders.12 PC has been the main focus of several

a.

Department of Chemistry, Faculty of Science, The University of Jordan, Amman 11942, Jordan. Department of Chemistry, The Hashemite University, P.O. Box 150459, Zarqa 13115, Jordan. c. Pharmaceutical and Chemical Engineering Department, German Jordanian University, P.O. Box 35247 Amman 11180, Jordan. d. Department of Life Sciences and Chemistry, Jacobs University Bremen, Campus Ring 1, 28759 Bremen, Germany. † Footnotes rela)ng to the )tle and/or authors should appear here. Electronic Supplementary Information (ESI) available: [details of any supplementary information available should be included here]. See DOI: 10.1039/x0xx00000x b.

researchers due to being a green carbonylating agent, ease of 16 preparation with almost quantitative yields, its implementation as an organic solvent in the electrolytes for high-energy density 17 batteries, and electrolytic capacitors. Therefore, it is an important aspect to catalyse the production of the title molecules under mild conditions with low catalyst loading and high 18 efficiency/selectivity. Cyclic carbonates have been synthesized through the use of different homogeneous or heterogeneous catalysts, that promote the cycloaddition of CO2 with epoxides, e.g. Salen-metal 19 20 compounds, metalloporphyrins, metal-organic frameworks 21–23 24 25,26 (MOFs), zeolites, , mesoporous silica, polyoxometallates 27 16,28 29–33 (POMs), smectite, ionic liquids, and most importantly 34 metallic oxides due to its commercial availability. Heterogeneous catalysis, among others, offers many advantages compared to homogenous catalysed processes, viz., extra catalyst stability together with the ease of recovery/reusability, minimal cost for 35,36 product separation and, solvent-free processes. In this respect, metallic oxides have been studied extensively as heterogeneous catalysts for the cycloaddition reaction of CO2 and epoxides due to 37 their acid/base character. Yano et al. first reported on the use of a commercial magnesium oxide in N,N-dimethylformamide (DMF) as a solvent with low conversions, however, no data were given 38 regarding the selectivity. Moreover, Arai and coworkers utilised different metal oxides (M = Mg, Ca, Al, Zn, Zr, La, and Ce) at higher pressures, up to 80 bars in the same polar solvent. Among these catalysts, La-O was superior in terms of yields. Although MgO, CeO2 and ZnO showed higher selectivity with moderate yields and conversions, the previous figures of merit (conversion, selectivity, and yield were 72.6%, 74.5%, 54.1%, respectively), ranked La-O as the best candidate. A solvent-free, commercial based La-O catalyst was superior among other zirconia doped oxide heterogeneous counter parts. It showed that the propylene oxide (PO) conversion

J. Name., 2013, 00, 1-3 | 1

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

Sustainable Energy & Fuels

Please do notEnergy adjust & margins Sustainable Fuels

Page 2 of 9 View Article Online

DOI: 10.1039/C8SE00092A

Journal Name

Results and discussion 1.

1.0

Transmittance (%)

of ca. 86% and PC yield and selectivity of 55% and 64%, 39 respectively. This makes La-O an interesting case study for further investigations. Tetrabutylammonium bromide (TBAB) played an important role 40–43 in the production of cyclic carbonates from epoxides and CO2. Its role is represented in the nucleophilic attack of the bromide anion on the electrophilic carbon of PO and the formation of bromo-alkoxide intermediate, which attacks CO2 to form the anionic carbonate that rearranges into PC and subsequent regenerates the TBAB once again for another reaction cycle. Caló et 44 al. attributed the reactivity of TBAB to the bulkiness of the tetrahedral cation, where the separation of ions becomes more obvious as less electrostatic attraction is feasible. As a consequence, more bromide will be available for the nucleophilic attack. To our knowledge, the only metallic oxide used with TBAB as a catalyst combination for the production of PC was introduced by Eddaoudi and coworkers,45 where they used MOF based Y2O3/TBAB binary mixture. It was reported that the use Y2O3 resulted in much lower yields compared to the MOF based catalyst. Herein, the synergistic effect between the best performing metal oxide reported in the literature for catalysing the production of PC in the presence of the TBAB as a co-catalyst was investigated. In comparison with similar systems reported in the affiliated literature, our catalyst combination, La-O/TBAB, showed high selectivity (> 99%), lower catalyst loading and pressure applied, together with an additional emphasis on the in situ formation of lanthanum carbonate (La2CO3)46 as the active surface rather than La-O, which was common for the cycloaddition reaction of CO2 and PO. In this work, the influence of different experimental parameters and conditions, on the cycloaddition reaction, has been investigated, namely, the effect of catalyst loading, co-catalyst concentration, [PO], CO2 pressure, temperature, and reaction time. Furthermore, the binary catalyst was recycled five times without losing its activity. Density functional theory (DFT) calculations were also used to unveil the mechanism of the cycloaddition reaction.

0.9

0.8

1466 cm-1 1371 cm-1 1650 1500 1350 1200

0.7

0.6

3609 cm-1 3700 3600 3500 3400

4000 3500 3000 2500 2000 1500 1000

500

-1

Wavenumver (cm )

Figure 1. ATR-FTIR spectra of La-O before (black) and after (red) the cycloaddition reaction.

A further peak was observed at 3609 cm−1, which can be ascribed to OH groups bound to the catalyst, viz., the surface was comprised of both LaCO3OH and La-O. In addition, the absence of −1 peaks within the region of 1620-1780 cm indicated that there was 47 no bulk carbonate. Upon performing the cycloaddition reactions, a correspondent spectrum between the sample as received and the reused (washed and dried) La-based catalyst was observed with a −1 higher intensity within the region 1350-1620 cm , namely, the −1 −1 peak centred at 1466 cm was red-shifted to 1464 cm . This was anticipated by a further modification of the surface under the reaction conditions and fortified by the kinetic measurements with a second-order reaction with respect to CO2 (vide infra). Notably, the surface is modified through La-O rather than La-OH as having −1 the same peak centred at 3609 cm . Thermal gravimetric analysis (TGA) of the material under nitrogen showed a mass loss of about 2% when the sample was heated up to 300 °C which might be attributed to water elimination. Three major steps were recorded upon further heating at ∼332, 487 and 662 °C, with an overall weight loss of about 10%. The first prominent mass loss was consistent with that trace obtained from 48 La(OH)3 as reported by Füglein and Walter. The other two weight losses in the TGA pattern, corresponded to dioxycarbonate 49,50 (La2O2CO3) and the oxide species (La-O), respectively. It was noteworthy that the release of CO2 between 600 and 800 °C (Figure S1, electronic supplementary information (ESI)) was verified using a coupled TG-FTIR.

Catalyst Characterization

La-O is well-known to react with both CO2 and H2O, (key components of the atmosphere). Therefore, we analysed the material before submitting the catalyst for testing. A simple Brauner-Emmet-Teller (BET) analysis revealed that the material was 2 non-porous with a surface area of 7.9 m /g. The material was further subjected to extra analysis using Attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy, which revealed carbonated surface as shown by the prominent −1 peaks at 1466 and 1371 cm , which can be ascribed to the symmetric vibrations of the monodentate carbonate group (Figure 47 1).

2.

Optimisation of Reaction Conditions

In the current work, the cycloaddition process using the lanthanumbased catalyst (in the absence of the co-catalyst) showed insignificant conversion (6.6%), poor selectivity (50%) and low percent yield (< 1%) at the investigated arbitrary conditions (Table 1, run 1). The use of the co-catalyst exhibited a mild conversion with respect to PO and gave a yield of about 13.6% PC (Table 1, run 1 2), further peaks were observed via H NMR for tributylamine (TBA) till a ratio of cat/co-cat of 1:2 by (wt./wt.) as shown in (Figure S2, ESI). When a ratio of 1:4 was used, the peaks of TBAB were prominent with no degradation losses. That was a clear indicator 1 that TBAB was not decomposed by the catalyst as verified using H NMR in the absence of PO (Figure S3, ESI). This was not the case when PO was presented, in accordance with previous observations 51–53 by North and coworkers, where TBA was formed and detected via gas chromatography–mass spectrometry (GC-MS). In agreement

2 | J. Name., 2012, 00, 1-3

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

ARTICLE

Page 3 of 9

Please do notEnergy adjust & margins Sustainable Fuels View Article Online

DOI: 10.1039/C8SE00092A

Sustainable Energy & Fuels

ARTICLE

% Yield 99% as evidenced by H NMR. Therefore, this catalyst loading was chosen as an optimal loading for further experiments. b.

The effect of [PO]

PO was selected as a model substrate to optimise the cycloaddition reaction conditions. Slight changes were observed under different PO concentrations, (Figure 2B, 1-7 mL, yielded 50-55%). Upon increasing the amount up to three-folds (10 mL of PO), a yield of 61% was obtained. The reason behind the poor performance might be attributed to the fixed number of the Lanthanum-Lewis acid active sites found in the catalyst. c.

40

30

The effect of CO2 pressure

The effect of pressure was also studied over a range of 1-40 bar of CO2 (Figure 2C). Applying 1 bar gave the lowest yield (17%) among different trials. Increasing the pressure from 5 to 10 bars yielded 30 to 52%, respectively. When higher pressures were applied (20 and 40 bars), lower yields of 43 and 36% were obtained, respectively.

B. 55

45

0.10 0.15 0.20 0.25 0.30

1

Catalyst Loading (g) 50

The Effect of Catalyst Loading

65

A.

% Yield of PC

CO2 (bar) 10 10 10 0

50

C.

40 30 20

50

3 5 7 10 PO Volume (mL)

D.

35 20 5

51 5 10 20 Pco2 (bar)

40

85 120 130 152 183 o Temperature ( C)

Figure 2. Optimisation of the reaction conditions upon varying: A. Catalyst loading. B. Substrate (PO) volume. C. The pressure of CO2 and D. Temperature on the overall % yield of PC while fixing the other parameters. For A, the PO volume = 3 mL, ܲ஼ைమ = 10 bar, Temp. = 152 °C, mass of TBAB = 0.2 g and a reaction time of 2 h. For B and C, the mass of La-O, was 0.2 g, while keeping the other variables same as mentioned in A. In D, the optimised reaction conditions in terms of epoxide volume (3 mL), CO2 pressure (10 bar), masses of cat. and co-cat (0.2 g:0.2 g La-O/TBAB) and a reaction time of 2 h were adopted.

e.

The Effect of Time

The effect of time was also another crucial factor to be investigated. A semi-logarithmic relationship was observed with a reaction time ranging from 1 to 24 h with an achieved selectivity over 99%. Ranging from 1 to 6 h resulted in yield from 27 to 79%. Almost quantitative yields were obtained after 9 h, and a full conversion was obtained after 24 h (Figure 3). Turnover numbers (TONs) and turnover frequencies (TOFs) of the investigated binary catalytic system are shown in Table S1, ESI.

J. Name., 2013, 00, 1-3 | 3

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

a.

PO (mL) 3 3 0 3

The effect of Temperature

% Yield of PC

Run Cat (g) Co-cat (g) 1 0.2 0 2 0 0.2 3 0.2 0.2 4 0.2 0.2 N.R.: no Reaction

d.

The effect of temperature was a crucial factor to deal with upon studying this reaction at different temperatures. The stability of the TBAB co-catalyst increased as shown by the TGA. It revealed that co-catalyst decomposition temperature raised from 112 to 126 °C upon mixing with basic catalyst (Figure S4, ESI). Moreover, the optimal selectivity and yields for the PC production were considered. Starting at 85 °C, modest yields were obtained (ca. 7%). Increasing the temperature improved the reaction yield, where a maximum value of 54% was reached at 183 °C, however, a very dark brown colour was observed indicating the decomposition of TBAB. At 152 °C, the reaction maintained the activity and selectivity of the binary system La-O/TBAB. % Yield of PC

Table 1. Control experiments to study the impact of cat (run 1), co-cat (run 2), CO2 (run 3) and PO (run 4) on the cycloaddition reaction at 152 °C for 2 h.

This might be explained by the competition between CO2 and PO to react with the Lewis acid, lanthanum-based active sites. This makes the use of 10 bar of CO2 as the optimal parameter to be used for the binary catalytic system of La-O/TBAB.

% Yield of PC

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

45

with other studies reported in the literature, TBAB alone under relevant conditions (120 °C, 20 bar CO2, 6 h, and 7 mL PO) gave a yield of 16%. Individually, no reaction took place in the absence of PO or CO2, no oligomers, glycols or cyclic ethers were observed as 1 verified by H NMR (Table 1, run 3 & 4). The catalyst combination La-O/TBAB showed a clear synergistic effect with a yield of 45% (152 °C, 10 bar CO2, 2h, and 3 mL PO), which was consistent with 51 the finding of North and collagues, as well as Eddaoudi and co45 38 workers. Regarding the reaction medium, Bhanage et al. reported on the use of toluene together with a solvent-free cycloaddition of epoxides with CO2 using different heterogeneous metal oxide-based catalysts. They showed that once toluene was utilised, it gave much lower conversions but higher selectivity when ethylene carbonate was formed (conversion was improved from 41% to 82%, and selectivity was decreased from 78% to 64%, for toluene, and solvent-free trials, respectively), no justification was given for such behaviour. The former observations were taken into consideration in our studies by using a lower boiling point solvent, namely dichloromethane (DCM), for the ease of product separation purposes.

Please do notEnergy adjust & margins Sustainable Fuels

Page 4 of 9 View Article Online

DOI: 10.1039/C8SE00092A

ARTICLE

Journal Name

90

% Yield of PC

80 70 60 50 40 30 20 0

3

6

9

12

15

18

21

24

Reaction time (h) Figure 3. Reaction profile of PO (3 mL) and CO2 (10 bar) using 1:1 by weight La-O/TBAB at 152 °C as a function of time.

Cycloaddition of Different Substrates

Different substrates were reported ranging from aromatic, cycloaliphatic to sustainable epoxides (1-6, Figure 4). At the optimised conditions, cyclic carbonate esters (7-12) were produced in descending order as follows: styrene oxide > 1,2-epoxybutane > epibromohydrin > PO > cyclohexene oxide with yields of 94, 90, 85, 83 and 71%, respectively. It was anticipated that increasing the bulkiness of the substrate will prohibit the access to the active site and thus resulting in a lower reaction yield. This was confirmed while no reaction was observed for Limonene oxide (6).

(7-12)

100 O

% Yield of CC

80

O

O

60

(7)

No Reaction

40 20

St y

re ne

1, ox 2id Ep e ox yb Ep u t ib an ro e m oh Pr y dr op in yle Cy ne clo ox he id xe e ne ox Lim id on e en e ox id e

0

(10)

(8)

(11)

(9)

(12)

Figure 4. Percentage yield of cyclic carbonate esters (7-12) using different epoxides (1-6) under the optimized reaction conditions. Epoxide (3 mL) and CO2 (10 bar) using 0.2 g : 0.2 g La-O/TBAB at 152 °C and a reaction time of 6 h.

4.

vacuum oven. Epoxide (3 mL) and CO2 (10 bar) using 1:1 by weight LaO/TBAB at 152 °C and a reaction time of 6 h.

Catalyst Reusability

The system showed a reasonable stability and reusability (Figure 5) after five cycles without losing its activity and selectivity. The performance mainly attributed to the synergistic effect rather than individual contributions of the binary system components. The increased reaction yield after the third catalytic cycle might be attributed to measurements uncertainty and/or carbonation of the La-O catalyst, which created new active sites for the cycloaddition reaction (vide infra). 60 50

% Yield of PC

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

(1-6)

40 30 20 10 0 1

2

3

4

5

Number of catalytic cycles Figure 5. Catalyst reusability of the binary system La-O/TBAB (under the optimised reaction conditions). Drying of the catalyst was carried out in a

5.

Reaction Kinetics

The integrated rate law of the cycloaddition reaction was obtained by varying the concentration/pressure of one reactant as a function of time while keeping excess quantities of the other reactant together with the cat/co-cat binary system at a constant temperature. The results of the kinetic study obey the following rate equation: Rate = k [PO] [CO2]2 (1) It was found that the reaction was pseudo first-order with respect to PO and second-order with respect to CO2 as confirmed from the straight lines obtained by plotting ln [PO] and the reciprocal of CO2 pressure against time (Figure S5, ESI). The first-order dependence of the substrate was consistent when a bimetallic aluminium-salen complex combined with TBAB was exploited for the cycloaddition of CO2 to the epoxides.52,54 The chemisorption of CO2 by the strong Brönsted basic sites (O2−) within La-O might be the main reason for

4 | J. Name ., 2012, 00 , 1-3

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

3.

100

Page 5 of 9

Please do notEnergy adjust & margins Sustainable Fuels View Article Online

DOI: 10.1039/C8SE00092A

ARTICLE 46,47,55

the second-order kinetics of CO2 in which one mole reacts with the surface and the other mole react to form PC accordingly.

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

6.

Reaction Mechanism

Based on the demonstrated spectroscopic and kinetic data, Scheme 1 shows a plausible catalytic cycle of the investigated cycloaddition reaction between CO2 and PC exploiting the La-O/TBAB binary system. Upon feeding the autoclave with CO2, the surface of the basic La-O catalyst (A, Scheme 1) reacts with CO2 to form the carbonated species (B, Scheme 1) as verified by ATR-FTIR spectroscopy (Figure 1). Meanwhile, an epoxide molecule coordinated to Lewis acidic sites of lanthanum (C, Scheme 1) which tend to orient the molecule to be ring-opened by a TBAB bromide ion to give the alkoxide ion (D, Scheme 1). Instantaneously, as shown in the 1H NMR (Figure S2, ESI), TBAB was decomposed in the presence of PO to give tributylamine (TBA, F, Scheme 1) and a presumable formation of butyl bromide which was not shown

earlier. We suggested that the butyl bromide decomposed at the investigated conditions due to its low flash point (10 °C), which was the presumable reason behind the brownish colour of the solution upon the reaction completion. Unfortunately, we could not verify this assumption via a suitable spectroscopic technique. The formed tertiary amine reacts with the CO2 to form a carbamate activated 52 species (G, Scheme 1) which tends to coordinate to the active species (D, Scheme 1), that and rearranged upon an intramolecular nucleophilic attack of the coordinated alkoxide ion to the carbamate ion eliminating. Afterwards, a TBA molecule was eliminated as a leaving group, which can be reintroduced once again within the catalytic cycle. Upon the formation of the coordinated ionic carbonate (E), another intermolecular rearrangement formed the organic carbonate, PC, as a target molecule to regenerate the stabilized carbonated species (B) to start the catalytic cycle once again.

Scheme 1. A plausible reaction mechanism for the formation of PC catalyzed by the binary system La-O/TBAB as suggested by ATR-FTIR spectroscopy, kinetic data and DFT calculations.

7.

Density Functional Theory (DFT) Calculations

The reaction mechanism was further verified by using quantumchemical calculations, employing the DFT method (B3LYP/aug-ccPVDZ). For the sake of simplicity and less expensive calculations, we chose to use a simplified cluster of lanthanum oxide, namely, La-O56 in order to understand the energetic transformations during the cycloaddition process. All calculations were performed in the gasphase. The energy diagram and the DFT-optimised structures for the catalysed reaction are shown in Figure 6. Firstly, CO2 was sorbed by La-O to form the carbonate species (B, Figure 6) which was stabilised through multiple dipole-dipole interactions. Later, the epoxide molecule coordinated to the lanthanum centre (C, Figure 6), which led to an overall stabilising in the energy of −31 kcal −1 mol . The ring-opening reaction of the epoxide by tetramethylammonium bromide (TMAB) (for simplicity, it was chosen instead of TBAB) was found to introduce additional stabilisation throughout ion-ion interactions, in which the

ammonium cation is in close proximity to the anionic carbonate anionic, whereas the bromide anion was accessible to attack the epoxide (D, Figure 6). The next step involved the intermolecular rearrangement and elimination of trimethylamine (E, Figure 6). The ring-closure step involved an intramolecular nucleophilic attack of the O atom towards the carbon bearing the bromide (E, Figure 6), excluding the latter as a good leaving group to produce the PC, while the bromide anion bound to the lanthanum. The transition state for this step is associated with an energy barrier (65 kcal −1 mol ). The ring-closing step was further studied as a function of C--O bond distance (Figure S6, ESI). The most stable structure is obtained once the PC is formed with a bond distance of ~1.4 Å, while shorter and longer distances were less stable. An energy barrier was obtained at 2.2 Å, which resembled the transition state.

J. Name ., 2013, 00 , 1-3 | 5

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

Sustainable Energy & Fuels

Please do notEnergy adjust & margins Sustainable Fuels

Page 6 of 9 View Article Online

DOI: 10.1039/C8SE00092A

Journal Name

Figure 6. Free energy profile of the minimum energy path for the formation of PC catalyzed by the binary system La-O/TMAB (it was chosen instead of TBAB for simplicity) calculated at the B3LYP/aug-cc-PVDZ level of theory.

8.

A Comparative Study

binary catalyst of La-O/TBAB (Run 1, Table 2) showed a lower catalyst loading, better selectivity and yields, lower periods of cycloaddition together with the lower pressure of CO2 applied without the formation of oligomeric materials as shown by Bhanage et al. (Run 2, Table 2). Moreover, the same research group utilised different metal oxides (M = Mg, Ca, Al, Zn, Zr, La, and Ce) at higher pressures (ܲ஼ைమ = 80 bar) in DMF as a reaction medium. Among these catalysts, La-O was superior in terms of yield and selectivity. 39 Adeleye et al., reported on using commercial La-O with yield and selectivity of 55% and 64%, respectively. On the other hand, the use of main group oxides such magnesium oxide (MgO) gave 41% of PC 37 in the presence of DMF as a solvent as reported by Yano et al. 57 Kaneda and colleagues, reported on the use of a mixed Mg-Al oxide and yielded 88% of PC and PO conversion of 96%. Moreover, the activity of the Mg-Al mixed oxide was greatly larger than that of a physical mixture of MgO and Al2O3. Eddaoudi et al.,45 prepared a yttrium-based as a heterogeneous catalyst which showed a good performance once combined with TBAB. Upon the use of Y2O3 as a control sample, it gave much lower yields than the corresponding yttrium-based MOF, viz., gea-MOF-1, when accompanied with tetrabutylammonium bromide (TBAB) as a cocatalyst, the authors attributed the lower reactivity of the lanthanide-based oxide due to the higher accessibility of the Lewis acid Y-sites, which is a considered as a pre-requisite for the epoxide activation step.

Among other catalysts, the binary catalytic system showed a reasonable activity with extra selectivity towards PC formation. Our Table 2. Different heterogeneous catalysts reported in the literature together with our binary catalyst La-O/TBAB.

Run 1 2 3 4 5 6 7

Catalyst a La-O/TBAB b La-O c La-O MgOd e Mg-Al mixed oxide g Y2O3/TBAB h gea-MOF-1, TBAB

Yield (%) 95 54.1 55 41 88 29 88

Selectivity (%) >99 74.5 64 n.a. f n.a. n.a. n.a.

a

Reference Current work 38 39 37 57 45 45 b

Reaction conditions: PO used: 3 ml; CO2: 10 bar; catalyst: 0.2 g; (DCM: 5 ml); temperature: 152 °C; time: 9 h. Reaction conditions PO used: 57 mmol; CO2: 8 MPa (80 bar); catalyst: 1 g (DMF: 4 ml); temperature: 150 °C; time: 15 h. Co-oligomer of PO and PC was formed. c Reaction conditions: Supercritical conditions, catalyst loading, 10% (w/w); reaction temperature, 170 °C; CO2: 70 bar; reaction time; 20 h; stirring speed; 350 rpm. d Reaction conditions: MgO (0.4 g, calcined for 3 h), CO2 pressure 20 kg cm2 (19.6 bar); PO used (10 mmol); DMF (3 ml); 135 °C; 12 h. e Reaction conditions: Mg-Al mixed oxide (Mg/Al) = 5, calcined at 400 °C) 0.5 g; epoxide 4 mmol, DMF: 3 mL; CO2: 5 atm; 100 °C; and 24 h. f Conversion of PO was 96%. g Reaction conditions: PO used (100 mmol); Y2O3 (0.15mmol); TBAB (0.15mmol); 120 °C; 20 bar, 6h. h Reaction conditions: PO used (100 mmol); gea-MOF-1 (60 mg, corresponding to 0.15 mmol yttrium), TBAB (0.15mmol); 120 °C; 20 bar, 6h. n.a.: not available.

Conclusions The utilisation of a highly selective catalyst (La-O/TBAB) for the cycloaddition reaction was reported. Upon addition of TBAB as a cocatalyst, the conversion was enhanced dramatically due to the presumable nucleophilic attack of the bromide ion while epoxides are bound to the La-active sites. All substrates were active towards CO2 cycloaddition except sterically bulk substrates due to presumable inaccessibility of the epoxide to La-atom. The reaction mechanism was found in accordance with both DFT calculations and kinetics. The La-O/TBAB binary catalyst showed modest reusability performance over five cycles.

Experimental 1.

Chemicals

Unless otherwise stated, all chemicals were used without further purification. La-O was purchased from Sigma-Aldrich and used as received without any activation. The morphology and structural properties of lanthanum oxide were reported previously by Adeleye et al.39 CO2 (99.95%, Food Grade) was purchased from Advanced Technical Gases Co. (Amman, Jordan). Propylene oxide, limonene oxide, styrene oxide, epibromohydrin, 1,2-epoxybutane, and CDCl3 were purchased from Sigma-Aldrich, dichloromethane and tetrabutylammonium bromide were purchased from TEDIA and Fluka, respectively.

6 | J. Name ., 2012, 00 , 1-3

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

ARTICLE

Page 7 of 9

Please do notEnergy adjust & margins Sustainable Fuels View Article Online

DOI: 10.1039/C8SE00092A

Sustainable Energy & Fuels

Conflicts of interest

Instruments

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

1

Solution H nuclear magnetic resonance (NMR) spectra were 1 collected at room temperature using (AVANCE-III 400 MHz ( H: 400 MHz), FT-NMR NanoBay spectrometer (Bruker, Switzerland) in CDCl3. Ex situ attenuated total reflectance-Fourier transform infrared (ATR-FTIR) spectra were recorded on a Bruker Vertex 70FT-IR spectrometer at room temperature coupled with a Vertex PtATR accessory. Thermogravimetric analysis (TGA) for the TBAB and La-O/TBAB (1:4) were performed using a NETZSCH STA 409 PC LUXX instrument (heating rate 10 K/min). TGA for the La-O material was carried using a NETZSCH TG 209F1 Libra with a heating rate of 10 K/min from 22-900 °C using Al2O3 crucibles. Gas chromatography (GC) was performed using helium as a carrier gas with a hewlett packrard (HP 6890) using a nonpolar capillary column (Length: 15 m, diameter: 0.53 µm, film thickness: 1.5 µm). GC oven was raised from 40 till 140 °C, with a heating rate of 10 K/min. An FID detector temperature of 260 °C was used. Brauner Emmet Teller (BET) was analysed using a Quantachrome Autosorb Automated Gas Sorption System, Quantachrome Corporation, the material was activated at 60 °C (in vacuuo, 3 h), and the software used was Autosorb for ® Windows , version 1.20. 3.

Cycloaddition Reaction

The cycloaddition reaction was carried out in a 300-mL stainless steel autoclave equipped with a magnetic stirring bar in a thermostatted oil bath maintained at 152 °C, as optimised reaction conditions. Upon loading the catalyst, the cocatalyst and epoxide, together with DCM were added. The reactor was left to pressurize to the appropriate amount of CO2 and was heated to the desired temperature. After fulfilling the reaction time, the reactor was cooled, and the gases were vented off, followed by evaporating the DCM. The reaction mixture was filtered through a filter paper to remove the catalyst. Yields and selectivity were determined using a 1 representative sample using either H NMR or GC. For reproducibility purposes, each trial for the cycloaddition reaction was repeated two times and the average value was given in the corresponding tables. 4.

Kinetic Measurements

A simplified kinetic study was carried out in a 100-mL stainless autoclave that was heated in a thermo-coupled heating mantle. The 1 reaction order of PO and CO2 was evaluated using H NMR and GC while keeping the La-O/TBAB constant according to the optimised conditions. 5.

Density Functional Theory (DFT) Calculations 58

All calculations were performed using the Gaussian 09 program. The B3LYP function was chosen to optimise the geometries and calculate the energies of the reactants, intermediates, transition states, and products. The aug-cc-PVDZ basis set was applied to the hydrogen (H), carbon (C), oxygen (O), Nitrogen (N), and Bromine 59 (Br) atoms, while LANL2DZ ECP basis set was used to describe the Lanthanum (La) atom. The harmonic vibration frequency calculations were also evaluated to verify the local minima points of the potential energy surface with all the positive frequencies.

There are no conflicts to declare.

Acknowledgements AFE acknowledges the Deanship of Scientific Research at the Hashemite University for the financial support. The effort of Brett Urash (University of California Santa Barbara, UCSB) in discussion and selection of the lanthanum oxide cluster for DFT calculations is highly appreciated.

References 1 A. Goeppert, M. Czaun, G. K. Surya Prakash and G. A. Olah, Energy Env. Sci, 2012, 5, 7833–7853. 2 A. Otto, T. Grube, S. Schiebahn and D. Stolten, Energy Env. Sci, 2015, 8, 3283–3297. 3 A. Galia and G. Filardo, in Carbon Dioxide as Chemical Feedstock, Wiley-VCH Verlag GmbH & Co. KGaA, 2010, pp. 15–31. 4 S. J. Davis, K. Caldeira and H. D. Matthews, Science, 2010, 329, 1330–1333. 5 P. T. Brown and K. Caldeira, Nature, 2017, 552, 45. 6 G. P. Peters, R. M. Andrew, T. Boden, J. G. Canadell, P. Ciais, C. Le Quere, G. Marland, M. R. Raupach and C. Wilson, Nat. Clim Change, 2013, 3, 4–6. 7 P. Markewitz, W. Kuckshinrichs, W. Leitner, J. Linssen, P. Zapp, R. Bongartz, A. Schreiber and T. E. Muller, Energy Env. Sci, 2012, 5, 7281–7305. 8 S. D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A. G. Russell, Q. Wei and M. Fan, RSC Adv, 2013, 3, 22739–22773. 9 M. E. Boot-Handford, J. C. Abanades, E. J. Anthony, M. J. Blunt, S. Brandani, N. Mac Dowell, J. R. Fernandez, M.-C. Ferrari, R. Gross, J. P. Hallett, R. S. Haszeldine, P. Heptonstall, A. Lyngfelt, Z. Makuch, E. Mangano, R. T. J. Porter, M. Pourkashanian, G. T. Rochelle, N. Shah, J. G. Yao and P. S. Fennell, Energy Env. Sci, 2014, 7, 130–189. 10 A. K. Qaroush, H. S. Alshamaly, S. S. Alazzeh, R. H. Abeskhron, K. I. Assaf and A. F. Eftaiha, Chem Sci, 2018, 9, 1088–1100. 11 Q.-W. Song, Z.-H. Zhou and L.-N. He, Green Chem, 2017, 19, 3707–3728. 12 A.-A. G. Shaikh and S. Sivaram, Chem. Rev., 1996, 96, 951–976. 13 J. H. Clements, Ind. Eng. Chem. Res., 2003, 42, 663–674. 14 J. W. Comerford, I. D. V. Ingram, M. North and X. Wu, Green Chem., 2015, 17, 1966–1987. 15 A. K. Qaroush, A. S. Al-Hamayda, Y. K. Khashman, S. I. Vagin, C. Troll and B. Rieger, Catal Sci Technol, 2013, 3, 2221–2226. 16 Y. Du, F. Cai, D.-L. Kong and L.-N. He, Green Chem, 2005, 7, 518– 523. 17 JEFFSOL Alkylene Carbonates ®, Huntsman, 2001. 18 J. Qin, P. Wang, Q. Li, Y. Zhang, D. Yuan and Y. Yao, Chem Commun, 2014, 50, 10952–10955. 19 A. Decortes, A. M. Castilla and A. W. Kleij, Angew. Chem. Int. Ed., 2010, 49, 9822–9837. 20 D. Bai, S. Duan, L. Hai and H. Jing, ChemCatChem, 2012, 4, 1752– 1758. 21 M. H. Beyzavi, C. J. Stephenson, Y. Liu, O. Karagiaridi, J. T. Hupp and O. K. Farha, Front. Energy Res., 2015, 2, 63. 22 J. Liang, Y.-Q. Xie, X.-S. Wang, Q. Wang, T.-T. Liu, Y.-B. Huang and R. Cao, Chem Commun, 2018, 54, 342–345.

J. Name., 2013, 00, 1-3 | 7

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

2.

ARTICLE

Please do notEnergy adjust & margins Sustainable Fuels

Page 8 of 9 View Article Online

DOI: 10.1039/C8SE00092A

Journal Name

23 J. Zhu, P. M. Usov, W. Xu, P. J. Celis-Salazar, S. Lin, M. C. Kessinger, C. Landaverde-Alvarado, M. Cai, A. M. May, C. Slebodnick, D. Zhu, S. D. Senanayake and A. J. Morris, J. Am. Chem. Soc., 2018, 140, 993–1003. 24 M. Tu and R. J. Davis, J. Catal., 2001, 199, 85–91. 25 M. Liu, K. Gao, L. Liang, J. Sun, L. Sheng and M. Arai, Catal Sci Technol, 2016, 6, 6406–6416. 26 M. Liu, X. Lu, L. Shi, F. Wang and J. Sun, ChemSusChem, 10, 1110–1119. 27 H. Yasuda, L.-N. He, T. Sakakura and C. Hu, J. Catal., 2005, 233, 119–122. 28 J. Peng and Y. Deng, New J Chem, 2001, 25, 639–641. 29 Y. Xie, Z. Zhang, T. Jiang, J. He, B. Han, T. Wu and K. Ding, Angew. Chem. Int. Ed., 2007, 46, 7255–7258. 30 Y. Xiong, H. Wang, R. Wang, Y. Yan, B. Zheng and Y. Wang, Chem Commun, 2010, 46, 3399–3401. 31 Y. Xiong, J. Liu, Y. Wang, H. Wang and R. Wang, Angew. Chem. Int. Ed., 2012, 51, 9114–9118. 32 M. Liu, L. Liang, X. Li, X. Gao and J. Sun, Green Chem, 2016, 18, 2851–2863. 33 M. Liu, J. Lan, L. Liang, J. Sun and M. Arai, J. Catal., 2017, 347, 138–147. 34 S. Fujita, M. Arai and B. M. Bhanage, in Transformation and Utilization of Carbon Dioxide, eds. B. M. Bhanage and M. Arai, Springer Berlin Heidelberg, Berlin, Heidelberg, 2014, pp. 39–53. 35 O. Deutschmann, H. Knözinger, K. Kochloefl and T. Turek, in Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH & Co. KGaA, 2009. 36 V. B. Saptal and B. M. Bhanage, Curr. Opin. Green Sustain. Chem., 2017, 3, 1–10. 37 T. Yano, H. Matsui, T. Koike, H. Ishiguro, H. Fujihara, M. Yoshihara and T. Maeshima, Chem Commun, 1997, 1129–1130. 38 B. M. Bhanage, S. Fujita, Y. Ikushima and M. Arai, Appl. Catal. Gen., 2001, 219, 259–266. 39 A. I. Adeleye, D. Patel, D. Niyogi and B. Saha, Ind. Eng. Chem. Res., 2014, 53, 18647–18657. 40 W. J. Kruper and D. D. Dellar, J. Org. Chem., 1995, 60, 725–727. 41 J. Sun, S. Fujita, B. M. Bhanage and M. Arai, Catal. Today, 2004, 93–95, 383–388. 42 M. North and R. Pasquale, Angew. Chem. Int. Ed., 2009, 48, 2946–2948. 43 S. Senthilkumar, M. S. Maru, R. S. Somani, H. C. Bajaj and S. Neogi, Dalton Trans, 2018, 47, 418–428. 44 V. Caló, A. Nacci, A. Monopoli and A. Fanizzi, Org. Lett., 2002, 4, 2561–2563. 45 V. Guillerm, WeselińskiŁukasz J., Y. Belmabkhout, A. J. Cairns, V. D’Elia, WojtasŁukasz, K. Adil and M. Eddaoudi, Nat Chem, 2014, 6, 673–680. 46 O. V. Manoilova, S. G. Podkolzin, B. Tope, J. Lercher, E. E. Stangland, J.-M. Goupil and B. M. Weckhuysen, J. Phys. Chem. B, 2004, 108, 15770–15781. 47 B. Klingenberg and M. A. Vannice, Chem. Mater., 1996, 8, 2755– 2768. 48 E. Füglein and D. Walter, J. Therm. Anal. Calorim., 2012, 110, 199–202. 49 M. Hoang, A. E. Hughes, J. F. Mathews and K. C. Pratt, J. Catal., 1997, 171, 313–319. 50 P. Jeevanandam, Y. Koltypin, O. Palchik and A. Gedanken, J Mater Chem, 2001, 11, 869–873. 51 M. North, P. Villuendas and C. Young, Tetrahedron Lett., 2012, 53, 2736–2740.

52 M. North and R. Pasquale, Angew. Chem. Int. Ed., 2009, 48, 2946–2948. 53 A. Decortes, A. M. Castilla and A. W. Kleij, Angew. Chem. Int. Ed., 2010, 49, 9822–9837. 54 W. Clegg, R. W. Harrington, M. North and R. Pasquale, Chem. – Eur. J., 2010, 16, 6828–6843. 55 C. Chu, Y. Zhao, S. Li and Y. Sun, Phys Chem Chem Phys, 2016, 18, 16509–16517. 56 C. Chu, Y. Zhao, S. Li and Y. Sun, J. Phys. Chem. C, 2016, 120, 2737–2746. 57 K. Yamaguchi, K. Ebitani, T. Yoshida, H. Yoshida and K. Kaneda, J. Am. Chem. Soc., 1999, 121, 4526–4527. 58 Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J.R., & Fox, D. J., Gaussian 09, Gaussian Inc. 59 P. Jeffrey Hay and Willard R. Wadt, J. Chem. Phys., 1985, 82, 299.

8 | J. Name., 2012, 00, 1-3

This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Sustainable Energy & Fuels Accepted Manuscript

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

ARTICLE

Sustainable Energy & Fuels Accepted Manuscript

Published on 06 April 2018. Downloaded on 06/04/2018 16:06:15.

Page 9 of 9 Sustainable Energy & Fuels DOI: 10.1039/C8SE00092A

View Article Online