Kamila
Błażek
a,
Hynek
Beneš
b,
Zuzana
Walterová
b,
Sabina
Abbrent
b,
Arantxa
Eceiza
c,
Tamara
Calvo-Correas
c and
Janusz
Datta
*a
aGdansk University of Technology, Faculty of Chemistry, Department of Polymers Technology, 11/12 Gabriela Narutowicza Street, 80-233 Gdansk, Poland. E-mail: janusz.datta@pg.edu.pl
bInstitute of Macromolecular Chemistry, CAS, Heyrovského nám. 2, Praque 162 06, Czech Republic
cMaterials+Technologies’ Research Group (GMT), Department of Chemical and Environmental Engineering, Polytechnic School, University of the Basque Country, Pza Europa 1, Donostia-San Sebastian 20018, Spain
First published on 1st February 2021
Bio-based cyclic carbonates are of significant research interest as monomers for non-isocyanate polyurethane (NIPU) synthesis. This research describes the synthesis of a series of five-membered bis(cyclic carbonate)s using bio-based polyether polyols (PO3G) with different molecular weights (250, 650 and 1000 g mol−1) and carbon dioxide as green feedstocks. The utilization of CO2 as a source of carbon in the chemical reaction is in agreement with the sustainable chemical industry. Furthermore, in order to support the green and sustainable polymer chemistry approach, the syntheses were attempted under solvent-free conditions. The implemented synthetic methods are focused on the design of processes and final products that minimize negative environmental impact. Detailed chemical structure analysis of synthesized products was performed using a combination of spectroscopy techniques (ATR-FTIR as well as 1D and 2D NMR techniques), mass spectrometry (MALDI-TOF MS) and chromatography analysis (SEC). The formation of the main product with two terminal cyclic carbonates was confirmed and the formed side products were also identified, characterized and quantified. Finally, as a proof of concept, the synthesized bis(cyclic carbonate)s were successfully used for the preparation of NIPU thermosets. Chemical and mechanical properties of the produced materials suggest their high potential for future applications, e.g. as sound absorbing materials.
Nowadays, the synthesis of cyclic carbonate(s) from an appropriate epoxide and carbon dioxide (CO2) is an active field of research.7 This is mainly due to the necessity of designing new technologies able to mitigate the environmental impact of CO2, which as a greenhouse gas, has a significant impact on an enhanced greenhouse effect. Due to the fact that CO2 is chemically inert, abundant, non-flammable, non-toxic and highly attractive one-carbon (C1) building block, its conversion into valuable chemicals is attracting attention from a wide academic community, industry and society.8 For these reasons, the capture, storage and utilization of CO2 are serious challenges to overcome and move toward sustainable development.9 Another issue is the adoption of rules and practices of green production, ranging from the substitution of hazardous components with more sustainable alternatives for the design and development of completely novel, green polymers.10 Synthesis of new, biomass-derived NIPUs can be accomplished using novel five-membered bis(cyclic carbonate)s prepared by chemical fixation of CO2 from bis-epoxides derived from bio-based feedstocks, such as vegetable oils, bio-based 1,4-butanediol, resorcinol, trimethylolpropane,11 isosorbide,12 sorbitol,13 vanillin,14 glycerol, pentaerythritol,15 furfuryl alcohol,16 gallic acid,17 tannic acid18 and polytrimethylene ether glycols.9,19
While DGEBA (Bisphenol A diglycidyl ether), BDO (1,4-butanediol), PPG (poly(propylene glycol)), PEG (poly(ethylene glycol)) diglycidyl ethers are commercially available products, diglycidyl ethers of other compounds, especially obtained from bio-based resources, require special preparation strategies.20,21 Currently, two main methods for the syntheses of glycidyl ethers with terminal epoxy groups through the reaction with an epihalohydrin are known.22 The first concerns the use of Lewis acids and consists of the halohydrin intermediate process and the dehydrochlorination process. In this type of reaction, halohydrin ether as a side-product may be formed and undesired polymerization reaction can proceed. The second method involves the use of phase-transfer catalysts (PTC), sodium hydroxide solution and organic solvents. Although, by-products formed by this method require special procedures for disposal, which is unfavourable from an industrial point of view.22
In this research study, we have explored the potential of bio-based polyether polyols as precursors for the synthesis of fully bio-sourced NIPUs. First, a series of bis(cyclic carbonate)s with different molecular weights (250, 650 and 1000 g mol−1) was synthesized using a two-step process involving polyol epoxidation and subsequent chemical fixation of CO2. Detailed molecular structure characterization of the prepared intermediates and final products was carried out using ATR-FTIR, 1D- and 2D-NMR, and SEC and MALDI-TOF mass spectrometry. To the best of our knowledge and despite their great potential as monomers for NIPUs, the synthesis of cyclic carbonates has never been reported in such a detailed way. As the last step, the synthesized bis(cyclic carbonate)s were used for the preparation of NIPU thermosets by polyaddition with polyfunctional bio-based amines. The produced crosslinked polyurethane materials were characterized using FTIR, DMTA, TGA and the NIPU network mechanical and swelling properties were determined.
Fig. 1 Scheme of the reaction between polyether polyol and epichlorohydrin (I step) and cycloaddition of CO2 into diglycidyl ether (II step). |
(1) |
(2) |
(3) |
The chemical structures of the synthesized diglycidyl ethers were first analysed by FTIR (Fig. 2). The FTIR spectra have confirmed the successful epoxidation reaction revealing the formation of oxirane ring moieties (epoxy groups) as two new bands at 908 cm−1 (νC–O of the oxirane group) and 841 cm−1 (νC–O–C of the oxirane group), and the consumption of polyols’ hydroxyl groups – indicated by the almost complete disappearance of the stretching OH band (ν-OH) at 3590–3310 cm−1. The epoxy band intensity was proportional to the epoxy group content and progressively increased (ED1000 < ED650 < ED250) as the molecular weight of the used polyols decreased. All epoxidized samples further contained typical asymmetric and symmetric stretching vibrations of C–H groups (ν-CH2– and ν-CH3) at 2930 and 2850 cm−1, and the stretching vibrations of ether functional groups (ν-C–O–C–) at 1100 cm−1.
In order to identify the formed products, including side-products MALDI TOF mass spectrometry analysis of the synthesized diglycidyl ethers was conducted. Initially, MALDI TOF mass spectra of the starting materials, i.e., bio-based polyols (PO3G250, PO3G650 and PO3G1000, Fig. S1 in the ESI†), showed a broad distribution of molecular ions (ionized by Na+) with a trimethylene ether repeating unit (Δm/z = 58 Da) and with peak positions revealing a linear structure with OH and H end-groups. After epoxidation, the MALDI-TOF mass spectra of all three synthesized diglycidyl ethers (ED250, ED650, and ED1000, Fig. 3A–C) proved the formation of the structures bearing oxirane end-groups. Owing to the high purity of the original PO3G1000 polyol, the MALDI TOF mass spectrum of ED1000 was smooth and showed three well-distinguished distributions of molecular ions (sodium adducts). The peaks of all three distributions were characterized by a mass increment of 58 Da corresponding to the mass of the repeating unit in poly(trimethylene ether) glycols (PO3G).
The most intensive distribution (marked with black asterisks) corresponded to the targeted diglycidyl ether 2, i.e. the linear PO3G structure with two terminal oxirane rings (Fig. 3D).
Besides this main product, two by-product species were identified to have formed during the epoxidation. These by-products were composed of linear PO3G repeating units (Δm/z of 58 Da) with one (the full red circle marked distribution) or two (the blue cross marked distribution) glycidyl ether terminal groups bearing a chloromethyl side-chain. The possible reaction mechanism of this side reaction involves the initial formation of alkoxy anion 4 and its subsequent reaction with an epoxy group of epichlorohydrin thus producing the chloride-containing alkoxy anion 5 and final elimination of the chloride anion to form glycidyl ether 6 and 7 bearing one or two chloromethyl side-chains (Fig. 4).
Moreover, the MALDI TOF mass spectra of ED250 and ED650 also indicated the formation of additional by-products (Fig. 3A and B) assigned to the OH-bearing structures 8, 9 and 10 (Fig. 3D), which were probably formed by the hydrolysis of the oxirane rings (the violet number sign and the green square marked distributions) and the elimination of the chloride anion (the grey triangle marked distribution) as shown in Fig. 4. MALDI TOF mass spectrometry further revealed the presence of chlorohydrin derivatives 1 and 11 in ED650 (Fig. 3B) formed due to incomplete dehydrochlorination.
The extent of these side reactions as well as the yield of epoxidation (the epoxy conversion) can be assessed from the 13C NMR results (Fig. 5). Normally, 1H NMR spectra would be used for such analysis.
Fig. 5 13C NMR spectra of the synthesized diglycidyl ethers (ED250, ED650 and ED1000), peak denotation and end group characterization and quantification. |
However, in this case, the low abundance and similarity of different end-groups resulted in unresolved, overlapping peaks and 13C NMR had to be used instead. Here, the dominant methylene signals 3 (67.7 ppm) and 4 (30.3 ppm) originate from the trimethylene ether repeating unit of the polymer backbone (also see the 13C NMR spectra of the original PO3Gs in Fig. S2 in the ESI†), while the methylene signals 1 (58.5 ppm) and 2 (33.4 ppm) represent CH2 carbons next to the unreacted OH end-groups. Furthermore, the presence of structures identified by MALDI TOF mass spectrometry (Fig. 3) was verified and quantified by 13C NMR spectroscopy in the form of different existing end-groups (Fig. 5). Firstly, three different glycidyl ether (epoxy) end-groups were identified confirming the formation of the main product 2 (the black asterisk signals) and the by-products with chloromethyl (6, 7; the red circle signals) and hydroxymethyl (10; the blue triangle signals) side-chains.31 Moreover, ED250 and ED650 samples have also been shown to contain the chlorohydrin derivatives 1 and 11 (the magenta trefoil signals), and the OH-bearing structures 8 and 9 (the green square signals).
Since all the 13C NMR signals corresponding to the end-groups were identified, end-group quantification was performed using the sum integral of one non-overlapping signal of each species (the underlined signals in Fig. 5). The sum of three different glycidyl ether end-groups then corresponds to the degree of epoxy conversion, estimated to be 88%, 72% and 29% for ED1000, ED650 and ED250, respectively. This is in very good agreement with the MALDI TOF results confirming the high efficiency and purity of the ED1000 sample.
For further NIPU preparation via the polyaddition of bis(cyclic carbonate)s with polyamines, it is necessary to convert bisepoxides (prepared in step I) into the bis(cyclocarbonate)s (prepared in step II) in high yield and selectivity by using an environmentally friendly catalyst.33 The TBAB catalyst was previously employed as a green and efficient catalyst for the cycloaddition of CO2 with mono-epoxides34 and aromatic di-epoxides.33 Therefore, it was also adopted for the conversion of 2 into 3 with a complete epoxy group consumption during CO2 cycloaddition, as no residual epoxide peaks remained in the 13C NMR spectra of all synthesized DC samples (the absence of methylene and methine signals of the oxirane ring at ca. 43.9 ppm and 50.8 ppm, respectively, Fig. 6). Moreover, the characteristic carbon signals of cyclic carbonate ring at 155.8 ppm (CO), 76.1 ppm (CH) and 66.6 ppm (CH2)35 confirmed the formation of 3 with the yield of 28% (DC250), 37% (DC650) and 49% (DC1000) respectively. 1H–13C HSQC spectra of the synthesized bis(cyclic carbonate)s (Fig. S3 in the ESI†) helped us to distinguish individual protons and carbons connected to the chlorine atom and OH groups. Fig. 6 shows determination and quantifications of the identified end-groups giving the total content of cyclic carbonate groups (the black asterisk, the red circle and the blue square signals) of 82% (DC250), 56% (DC650) and 74% (DC1000). The methylene signals 1 (58.5 ppm) and 2 (33.4 ppm) representing CH2 carbons next to unreacted OH end-groups indicated the residual amount of unreacted PO3G chains. Additionally, the DC250 and DC650 samples contained certain amounts of chlorohydrin derivatives (1 and 11) and the OH-bearing species (8 and 9), already formed as by-products during the previous step I.
Fig. 6 13C NMR spectra of the synthesized bis(cyclic carbonate)s (DC250, DC650 and DC1000), peak denotation and end group characterization and quantification. |
The structure of the main product 3 and the formation of by-products during CO2 cycloaddition were further investigated by MALDI-TOF mass spectrometry (Fig. 7). The MALDI-TOF mass spectra of all synthesized bis(cyclic carbonate)s distinguished five distributions of molecular ions with a mass increment of 58 Da corresponding to the trimethylene ether repeating unit with different end-groups (Fig. 7A–C). The most intensive distribution (the black asterisk marked signals) was assigned to the linear PO3G molecules bearing two terminal cyclic carbonates, which proved the successful formation of the main product – bis(cyclic carbonate)s 3 (Fig. 7D). Two other distributions marked with full red circles and blue crosses were also assigned to bis(cyclic carbonates) species 12 and 13, which contained one and two chloromethyl side chains in terminal groups, respectively. They were formed via CO2 cycloaddition from their epoxidized analogues 6 and 7. It was thus found that the presence of a chlorine atom in the side-chain group did not suppress the reactivity of diglycidyl ethers towards CO2 cycloaddition. The side reactions proceeded to a very low extent in the case of DC1000, while intensive MALDI TOF signals of by-products were observed for DC250 and DC650. In particular, mainly mono(cyclic carbonate) structures bearing only one cyclic carbonate end-group (14–19) were detected. The incomplete CO2 cycloaddition together with the hydrolysis of the oxirane ring leading to the formation of 8, 18 and 19 species were thus confirmed to be the most important side reactions during step II.
Finally, the molecular weight average (Mn) and dispersity (Đ) of the synthesized bis(cyclic carbonate)s were determined using MALDI TOF mass spectrometry and SEC (Table 1 and Fig. S4 in the ESI†). The theoretical Mn of the synthesized bis(cyclic carbonate)s was calculated from the Mn values of the initial bio-based poly(trimethylene glycol)s (PO3G250, PO3G650 and PO3G1000) determined using the 13C NMR (Fig. S2 in the ESI†). The calculation was performed based on the assumption of full conversion and occurrence of no side reactions.
Sample | M n (theor.) [g mol−1] | M n (SEC) [g mol−1] | Đ (SEC) | M n (MALDI-TOF) [g mol−1] | Đ (MALDI-TOF) |
---|---|---|---|---|---|
a PS calibration. | |||||
DC250 | 460 | 540 | 1.4 | 710 | 1.3 |
DC650 | 900 | 800 | 2.0 | 790 | 1.5 |
DC1000 | 1300 | 1200 | 2.2 | 1350 | 1.5 |
Both experimentally determined Mn values correlate well with each other and are in good agreement with the theoretical Mn. The difference in molecular weight confirms the occurrence of side reactions as confirmed by FTIR, NMR and MALDI TOF above. The difference between the theoretical and the determined Mn weights were higher in the case of DC250 than DC650 and DC1000, which agrees well with the above-mentioned findings that the side reactions proceed to a higher extent during DC250 synthesis. Also, the relatively high dispersity of the synthesized bis(cyclic carbonate)s is caused by the high dispersity of the original bio-based polyether polyols (see Table S1 in the ESI†).
Fig. 8 Polyaddition of the synthesized bis(cyclic carbonate)s 3 and bio-based amine hardener Priamine 1071. |
FTIR analysis was used to confirm the final structure of crosslinked NIPUs (Fig. 9). It was found that during the polyaddition, the bis(cyclic carbonate)s 3 were quantitatively converted into NIPUs 16 as indicated by the almost complete disappearance of the typical carbonyl group vibration of 3 at about 1790 cm−1. The final conversion of the cyclic carbonate groups was estimated to be 94% (NIPU250), 100% (NIPU650) and 86% (NIPU1000). New peaks appeared at around 3300 cm−1 and 1700 cm−1 corresponding to the O–H (hydroxyl group)/N–H (urethane linkage) vibrations and the CO vibrations of the urethane linkages, respectively.36,37 The formation of urethane linkages was further confirmed by the presence of CO bending (1719 cm−1), N–H bending (1533 cm−1) and C–N stretching (1370 cm−1) bands, and the C–O asymmetric stretching of N–CO–O and C–O–C linkages at about 1245 cm−1.38 The strong absorption band at 1098 cm−1 is related to the ether (C–O–C) vibrations of the PO3G part of NIPUs. The bond assigned to the asymmetric stretching vibrations of N–CO–O and C–O–C of urethane groups were observed at 1245 cm−1. Moreover, the additional peak formation at around 1700 cm−1 for NIPU650 and the carbonyl peak splitting for NIPU1000 can be related to the formation of H-bonded ordered domains,39 as the NIPU structure provides a vast number of H-bonding groups, especially OH groups in the vicinity of urethane linkages.40
Fig. 9 The FTIR spectra of NIPU250, NIPU650 and NIPU1000 non-isocyanate polyurethanes synthesized by the ring-opening polyaddition of the DC250, DC650 and DC1000 bis(cyclic carbonate)s, respectively. |
Fig. 10 shows the temperature dependences of the storage (E′) and loss (E′′) moduli and the loss factor (tanδ) of the NIPUs. All NIPUs present a rubbery plateau at a temperature higher than 50 °C confirming their chemically crosslinked structure. The values of storage modulus in the rubbery region (E′R) are indirectly proportional to the average molar mass between crosslinks (Mc) corresponding to the network with the highest density in the case of NIPU250 (Table 2). These results agree well with the swelling degree of NIPUs (SR values in Table 2). While the NIPU1000 network swelled to a high degree due to the presence of long polyether chains between crosslinks, the NIPU250 sample swelled considerably less, confirming the formation of the densest chemically crosslinked network.41 Moreover, the high gel content (GC) of NIPU250 and NIPU650 provide evidence of optimal crosslinking in these materials.42,43 In contrast, NIPU1000 exhibits a low content of gel fraction, which indicates the formation of a large amount of “free species” – structures not chemically bonded to the polyurethane network.44
Sample | Dynamic-mechanical properties | Swelling properties | ||||
---|---|---|---|---|---|---|
T α [°C] | tanδmax | E′R [MPa] | M c [g mol−1] | SR [%] | GC [%] | |
T α is the main (alpha) transition temperature; tanδmax is maximum at tanδ curve; E′R is the storage modulus at the rubbery plateau at Tα + 100 °C; Mc is the molecular weight of the section between the crosslinks; SR is the swelling index; GC is the gel content. | ||||||
NIPU250 | −4.4/24.2 | 0.73 | 8.26 | 1174 | 334 | 97 |
NIPU650 | −8.6/13.5 | 0.78 | 3.05 | 3192 | 372 | 87 |
NIPU1000 | −27.4 | 0.99 | 1.26 | 7737 | 85 | 59 |
During the synthesis of NIPU1000, free (macro)cyclic structures (not covalently incorporated into the polyurethane network) were probably formed as a consequence of the relatively high molecular weight of the DC1000 monomer.45
The tanδ curves of all NIPUs demonstrated a clear maximum (Fig. 10), which corresponds to the main transition temperature (Tα, Table 2). The main transition of the crosslinked NIPUs is connected to the free volume available for network segment relaxation.24,41
The decreasing length of poly(trimethylene oxide) chain in NIPUs shifts the Tα values to higher temperatures due to reduced mobility of the chain segments as a consequence of the increased crosslink density.27 In the case of NIPU250 and NIPU650, the second maximum on tanδ curves appears at 24.2 °C (NIPU250) and 13.5 °C (NIPU650), which suggests a partial (micro)phase separation as a consequence of low miscibility between the long non-polar aliphatic chains of the amine hardener and more polar short chains of the poly(trimethylene oxide) polyol.
Moreover, due to the presence of long pendant chains (fatty acid chains of the amine hardener), all NIPU materials exhibit a broad main transition region and high tanδmax values (Table 2), which demonstrates their high ability to absorb acoustic energy. Generally, tanδ > 0.3 over a broad temperature range is required for efficient vibration damping materials.46–48
As the last step, thermal degradation of NIPUs was investigated using TGA under a nitrogen atmosphere (Table 3 and Fig. S6 in the ESI†). All prepared NIPUs were thermally stable up to ca. 200 °C with the initial decomposition temperature (T5%) of 265, 267 and 286 °C for NIPU250, NIPU650 and NIPU1000, respectively. The thermal degradation of NIPU250 and NIPU650 was shown to proceed in two main steps, typical for these types of polyurethane materials.49 The first weight loss takes place between 250 and 360 °C (Tmax1 at 319 and 323 °C for NIPU250 and NIPU650, respectively) and corresponds to the degradation of the urethane linkages, while the second weight loss is attributed to the decomposition of soft segments – poly(trimethylene oxide) chains (Tmax2 around 460 °C).20,50 In contrast, NIPU1000 decomposes in one broad step with Tmax1 = 393 °C, which reflects the diversity of the decomposed structures24 which is in line with the phase-mixed and homogeneous morphology without segregation into hard/soft domains indicated by DMTA.
Sample | T 5% [°C] | T 50% [°C] | T 90% [°C] | T max1 [°C] | T max2 [°C] | Solid residues at 650 °C [%] |
---|---|---|---|---|---|---|
T 5%, T50%, T90% is a temperature of 5%, 50% and 90% mass loss, respectively; Tmax1 and Tmax2 are the temperatures of the maximum rate of mass loss during the first and the second degradation step. | ||||||
NIPU250 | 265 | 413 | 458 | 319 | 444 | 0.1 |
NIPU650 | 267 | 383 | 456 | 323 | 444 | 1.4 |
NIPU1000 | 286 | 385 | 443 | 393 | — | 2.0 |
The TGA results also clearly reveal that NIPU1000 exhibits the highest initial degradation temperature.51 Taking into account the best thermal stability of DC1000 among the synthesized bis(cyclic carbonate)s (Fig. S7 in the ESI†), the high purity of the synthesized DC1000 is most probably the key factor improving the thermal stability of the final NIPU1000.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0py01576h |
This journal is © The Royal Society of Chemistry 2021 |