Oguzhan
Aslanturk
a,
Gokhan
Sagdic
a,
Emrah
Cakmakci
b,
Hakan
Durmaz
*a and
Ufuk Saim
Gunay
*a
aDepartment of Chemistry, Istanbul Technical University, 34469 Istanbul, Türkiye. E-mail: durmazh@itu.edu.tr; gunayuf@itu.edu.tr
bDepartment of Chemistry, Marmara University, 34722 Istanbul, Türkiye
First published on 13th September 2024
The Michael reaction, a cornerstone in organic chemistry, continues to revolutionize the field with its unparalleled versatility in forming carbon–carbon, carbon–oxygen, carbon–nitrogen, and carbon–sulfur bonds, paving the way for groundbreaking advancements in complex molecule and macromolecule construction. In this study, imide-yne reaction was employed at the macromolecular level for the first time to prepare linear poly(imide ester)s. A wide range of bisimides and dipropiolates were reacted through imide-yne click polymerization in the presence of 1,4-diazabicyclo[2.2.2] octane (DABCO) at room temperature. The polymerizations proceed in an anti-Markovnikov fashion, yielding the E-isomer as the major product. Polymers were obtained in high yields and their molecular weights were found to be in the range of 5.64–12.67 kDa. The remaining unreacted double bonds in the linear polymers were found to undergo further functionalization with thiols using a strong organocatalyst 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD), which was also supported by a model study. Post-polymerization modification study prompted us to prepare imide-yne monomers that can react with dithiols to synthesize poly(imide thioether)s through nucleophilic thiol–ene click reaction using TBD as the catalyst. The obtained polymers displayed a wide range of glass transition temperatures and thermal stability. Thus, it can be said that the proposed method enables the synthesis of new polyimide-based structures with tailorable thermal properties. It is believed that the proposed strategy will make a significant contribution to expanding the versatility of active alkyne chemistry at the macromolecular level.
Michael addition reactions are well-known reactions and have found tremendous applications in both synthetic organic chemistry and polymer science.26–30 This reaction is highly compatible with several nucleophiles, reaction media, and conditions. Moreover, the reaction mostly proceeds at ambient temperature and affords the final products with high yields and/or efficiencies, emphasizing operation simplicity and energy-saving. All these features put Michael reactions in a special place in “click” chemistry, a term coined by Sharpless and coworkers.31 In Michael reactions, amine, thiol, and oxygen-based nucleophiles are mostly studied compounds to react with electron-deficient double and triple bonds.32 Among them, the amine-based Michael reaction, namely the aza-Michael reaction, has some privileges such as the reaction taking place mostly without the use of any additives or catalysts.27,33–35 Similarly, the addition of imide to a reactive double or a triple bond is a typical aza-Michael addition reaction and has been employed successfully in organic chemistry to construct various nitrogen-based heterocyclic compounds.36–39 Unlike amines, a catalyst is required to activate imide for the Michael addition reaction. Although the participation of imide in a Michael addition reaction was first examined in the late 1940s,36 this chemistry has rarely been tested at the macromolecular level. Park et al., prepared poly(imide-aramid-sulfone)s by reacting divinyl sulfone and bis(4-(vinylsulfonyl)phenyl)terephthalamide with pyromellitic diimide in the presence of tetrabutylammonium hydroxide; moderate molecular weight polymers in good yields were obtained in these polymerizations.40
Although it has not been studied as much as reactive double bonds, the chemistry of activated alkynes in Michael addition reactions has been rejuvenated in the last two decades and has attracted much attention from every discipline of chemistry since the reactive triple bond has been shown to undergo rapid and efficient reactions with the nucleophiles. This chemistry, also known as nucleophile-yne or X-yne, has become particularly important in polymer and material science as it enables ultra-fast polymer synthesis, functionalization, gelation, network preparation, etc.32,41,42 In this regard, thiol–yne,43–50 amino-yne,33,51–57 hydroxyl-yne,30,58–61 phospho-yne,62 and finally imine-yne63 reactions have been proven to be robust and versatile strategies for the preparation of the above-mentioned polymer-based structures. As emphasized above, imides can also be added to the activated alkynes via aza-Michael reactions and this chemistry has been used several times in organic studies.64–68 For instance, in one of the early studies, the Trost group performed an α-addition to the alkynoates using phthalimide and sulfonamide in the presence of phosphine catalysts. The reactions proceeded at high temperatures and the target products were obtained in good to high yields.64 Later, Mola et al., reported the preparations of lactams, imides, and nucleosides via nucleophilic addition to activated triple bonds and 1,4-diazabicyclo[2.2.2]octane (DABCO) and 4-(dimethylamino)-pyridine (DMAP) were found to be the best catalysts for the synthesis of imides.66 Shahraki et al. reported the N-vinylation of heterocyclic compounds, including phthalimide, using ethyl and methyl esters of acetylenedicarboxylic acid using an equimolar amount of pyridine over 24 hours.67 Petit et al., used tosylacetylene as a protecting group for various N–H group-bearing compounds, including imides, in the presence of a base or a catalyst, and observed an excellent stereochemistry control depending on the base or catalyst.68 Despite all these nice organic examples, the imide-Michael reaction involving activated alkynes has not been extensively implemented as an alternative polymer synthesis method. Herein we offer a new and versatile method to add to the toolkit of X-yne click polymerization reactions and expand the library of activated alkyne-based polymer synthesis. In this current study, we demonstrate for the first time that imide-yne click polymerization could be a new synthetic method and energy-efficient approach for the preparation of a series of poly(imide ester)s using various bisimides and dipropiolates as monomers at room temperature, and in the presence of DABCO as a catalyst (Scheme 1).
Imide-yne monomers 2m and 3m were prepared using the same procedure described for 1m, and their experimental procedures and characterization data can be found in the ESI.†
Next, the optimum conditions for the proposed imide-yne polymerization were examined. For this purpose, bisimide 1, chosen for its high solubility, and dipropiolate a, chosen for its simplicity and easy monitoring by 1H NMR, were used as model monomers, and DABCO was used as the catalyst. Also, monomer concentrations were set to 0.25 M, and the mole ratios of bisimide:dipropiolate:DABCO was set to be 1:1:0.1. The results for the obtained polymer, namely P1a, are located in Table 1. Initially, we kinetically screened the effect of solvent on polymerization. Since monomer 1 and the other imide monomers have apparent solubility issues, we could have tested only NMP, DMSO, and DMAc together with DMF as solvents (entries 1–4, Table 1). As seen in this table, all solvents afforded polymer within 1 h, and a slight increase in molecular weights was observed over time. Among them DMF (entry 1, Table 1) and DMAc (entry 4, Table 1) gave little higher molecular weights; the former produced Mw = 8.3 kDa in 1 h, increased to 10.2 in 4 h and 10.8 in 24 h, and the latter gave Mw = 8.04 kDa in 1 h, increased to 11.8 in 4 h and 12.6 in 24 h. The other two solvents (i.e., NMP and DMSO) produced little lower molecular weights compared to these solvents (entries 2 and 3, Table 1). According to the results obtained from the solvent trials, we decided to use DMAc as the polymerization solvent, and since no distinct change was observed in the molecular weights after 4 h, the polymerization duration was fixed as 4 h for other experiments. Temperature effect on polymerization was next examined; however, increasing the reaction temperature to 40 °C and 80 °C did not improve molecular weights (entries 5 and 6, Table 1). As expected, lowering monomer concentrations to 0.1 M yielded a low molecular weight polymer (entry 7, Table 1); however, increasing monomer concentrations to 0.5 M gave an insoluble polymer (entry 8, Table 1). The same insoluble polymer was also observed when the catalyst concentration was increased to 0.25 equivalent (per monomer) while keeping the monomer concentrations at 0.25 M (entry 9, Table 1). It is worth mentioning here that either increasing the amount of dipropiolate or DABCO leads to a black precipitate appearing in a gel-like form, which might be attributed to a competitive enyne reaction that can take place in the presence of dipropiolate and DABCO.74 It should also be noted here that organocatalysts other than DABCO, such as TEA, TBD, DBU, and DMAP, were also tested in the optimum conditions but none of them yielded polymer. As such, we did not change the catalyst (i.e., DABCO) or its equivalent (0.1 equivalent per monomer) or increase the monomer concentrations, and kept the aforementioned optimum conditions for the rest of the studies.
Entry | Solvent | M w (kDa)b/Đb | |||||
---|---|---|---|---|---|---|---|
1 h | 2 h | 4 h | 6 h | 8 h | 24 h | ||
a Reaction conditions: 0.3 mmol of bisimide (1), 0.3 mmol of dipropiolate (a) and 0.03 mmol of DABCO in 1.2 mL of solvent at room temperature. b Determined by GPC calibrated based on linear PS standards in THF. c Reaction was carried out at 40 °C. d Reaction was carried out at 80 °C. e Reaction was carried out using monomer concentrations of 0.1 M. f Reaction was carried out using monomer concentrations of 0.5 M and upon the addition of DABCO, an insoluble polymer was obtained. g 0.25 equivalent of DABCO was utilized and an insoluble polymer was obtained. | |||||||
1 | DMF | 8.3/2.47 | 9.6/2.36 | 10.2/2.37 | 10.5/2.39 | 10.6/2.40 | 10.8/2.00 |
2 | NMP | 6.92/2.25 | 9.02/2.33 | 9.66/2.44 | 10.3/2.43 | 10.1/2.39 | 10.53/1.85 |
3 | DMSO | 3.6/1.71 | 4.16/1.81 | 6.4/2.16 | 7.94/2.40 | 8.96/2.55 | 9.17/2.44 |
4 | DMAc | 8.04/2.28 | 10.04/2.52 | 11.8/2.48 | 12.3/2.59 | 12.4/2.77 | 12.6/2.13 |
5c | DMAc | 10.57/2.41 | |||||
6d | DMAc | 9.19/2.06 | |||||
7e | DMAc | 7.54/1.69 | |||||
8f | DMAc | Insoluble | |||||
9g | DMAc | Insoluble |
Before performing other reactions to create a polymer library, the obtained polymer, P1a, was analyzed by NMR (Fig. 1). From the 1H NMR spectra, it can be seen that the alkyne proton of propiolate unit at δ 2.96 ppm disappeared and the integral ratio of the N–H proton of the imide structure reduced to a large extent. Importantly, the characteristic N-vinyl proton signals after the imide-yne reaction were detected as doublets and found to resonate at around δ 8.5–5.5 ppm with some of them overlapping with the aromatic proton signals. Here, NCHCH for the E- and the Z-isomers were found to resonate at δ 6.75–6.72 and 5.70–5.53 ppm, respectively. Aromatic protons were also detected between δ 8.16–7.59 ppm. When the integral ratios of the vinyl protons were compared, the ratio for the E-isomer was found to be 91% for P1a. The structure of P1a was further validated by 13C NMR analysis (Fig. S55†). As seen in this figure, propiolate carbons resonated at around δ 75 ppm completely disappeared and new vinylic carbons stemming from imide-yne addition were detected at δ 160.71 and 100.38 ppm. NMR results show that the expected product (i.e., P1a) was successfully prepared and the imide-yne polymerization reaction proceeded smoothly.
Fig. 1 Overlaid 1H NMR spectra of bisimide 1 (top) in d6-DMSO (500 MHz), dipropiolate a (middle) in CDCl3 (500 MHz), and P1a (bottom) in d6-DMSO (500 MHz). |
To diversify the poly(imide ester)s, different bisimide structures along with different dipropiolates given in Scheme 1 were reacted under the optimum conditions, and the obtained results are collected in Table 2. Unfortunately, bisimides 2 and 3 were found to have limited solubility in DMAc, so they were not utilized directly in imide-yne click polymerizations; however, these two compounds were used in the preparation of imide-yne monomers (see below). Bisimide 1 was combined with all dipropiolates (entries 1–5, Table 2), and for dipropiolate b, the reaction time was reduced to 1 h due to solubility problems after that time (entry 2, Table 2). Low to moderate molecular weight polymers ranging from Mw = 6.87–12.67 kDa were obtained from these polymerizations in good to high yields. Also, bisimides 4 and 5 that were synthesized by thiol-maleimide reactions were found to yield lower molecular weight polymers (entries 6 and 7, Table 2) and were less reactive towards imide-yne click reaction which could be due to the limited resonance stabilization of the imide anion. Moreover, bisimide 6 synthesized by amino-maleimide reaction yielded THF insoluble polymer so GPC data is not available for this polymer (entry 8, Table 2).
Entry | Bismide | Dipro-piolate | Polymer | M wb (kDa) | Đ | T gc (°C) | Isolated yieldd (%) |
---|---|---|---|---|---|---|---|
a Reaction conditions: 0.3 mmol of bisimide, 0.3 mmol of dipropiolate and 0.03 mmol of DABCO in 1.2 mL of DMAc at room temperature for 4 h. b Determined by GPC calibrated based on linear PS standards in THF. c Determined by DSC from the second heating cycle. d Gravimetric yield obtained after first precipitation. e Polymerization was carried out for 1 h. f The resulting polymer was observed to be insoluble in THF and therefore GPC data could not be obtained. | |||||||
1 | 1 | a | P1a | 12.43 | 2.48 | 199 | 90 |
2e | 1 | b | P1b | 12.67 | 1.96 | 193 | 82 |
3 | 1 | c | P1c | 7.82 | 1.95 | 121 | 72 |
4 | 1 | d | P1d | 12.65 | 1.68 | 147 | 84 |
5 | 1 | e | P1e | 6.87 | 1.88 | 162 | 80 |
6 | 4 | a | P4a | 6.55 | 2.96 | 61 | 62 |
7 | 5 | a | P5a | 5.64 | 2.83 | 130 | 70 |
8f | 6 | a | P6a | — | — | 205 | 94 |
The poly(imide ester)s were also characterized via FT-IR spectroscopy (Fig. S60, S67, S74, S81, S89, S95, S101, and S106†). The FT-IR spectra of the poly(imide ester)s displayed the absence of the vibration bands for propiolate triple bonds (–CC–) at around 2100 cm−1 and the presence of the characteristic imide carbonyl stretching vibration bands at around 1775 cm−1 and 1710 cm−1. Besides, a new band appeared at around 1640 cm−1 which was attributed to the newly formed double bonds (–N–CHCH–CO).
The Tg values of the synthesized poly(imide ester)s are collected in Table 2. It can be seen that the polyimides synthesized via imide-yne click polymerization displayed a wide range of Tg values ranging from 61 °C to 205 °C. This shows that the Tg values; i.e., the processability and flexibility of the resulting polyimides can be tuned, highlighting the modularity of the proposed method. Among the polyimides based on 1, P1a and P1b displayed the highest Tg values. P1a and P1b contain short chains between their propiolate end groups, respectively. When these short chains were replaced with a relatively long flexible hexanediol the Tg value decreased (P1d). When the propiolate was changed to a more flexible disulfide bond-containing one (P1c), the molecular weight was found to be lower and the Tg value decreased significantly. P1e which has rigid DOPO units, displayed a lower Tg value than expected which could be ascribed to its lower molecular weight compared to P1a.
P1a, P4a, and P5a polymers are produced from the reaction of the same dipropiolate but using different bisimides. When the Tg values of the homologous P1a, P4a, and P5a are compared, the Tg values were found to decrease in the following order: P1a > P5a > P4a. Bisimides 4 and 5 have flexible thioether linkages in their structures and therefore they displayed significantly lower Tg values than P1a which is composed of a rigid aromatic bisimide (1). Finally, the highest Tg value was obtained for the polymer P6a which has a rigid piperazine core. The nitrogen groups in these piperazine units create additional dipolar interactions which increase the intramolecular interactions and thus limit the degree of rotation and lead to reduced free volumes.
Since reactive double bonds are still present in the resulting imide-yne polymers, we hypothesized that these bonds can be further functionalized by thiol–ene click chemistry. To test this foresight, we designed a model study as depicted in Scheme 2a. In this experiment, cis-1,2,3,6-tetrahydrophthalimide was reacted with methyl propiolate in the presence of DABCO, resulting in N-vinylated product, which was characterized by 1H NMR. Fig. S47† indicates the full disappearance of alkyne signal δ 2.98 ppm and the appearance of vinyl signals between δ 7.68–5.65 ppm, full agreement with the integral ratios confirms the expected product. This product was then reacted with 1-propanethiol in CHCl3 at room temperature. It is known that nucleophilic thiol–ene reactions require a suitable catalyst to drive the reaction. It has been shown previously by our group47–49 and others50,75,76 that TBD is a robust organocatalyst due to its stronger basicity and nucleophilicity and can easily provoke a thiol compound to a reactive double bond. Here, we implemented this catalyst in the second step for further functionalization, and the reaction proceeded for 5 min at room temperature. After purification, the modified compound was characterized by NMR. Full disappearance of vinyl protons (Fig. S50†) and carbons (Fig. S51†) and the appearance of acetal signal (NCHS) at δ 5.48 ppm indicate that nucleophilic thiol–ene click reaction occurred quantitatively.
Scheme 2 Model imide-yne reaction and the following nucleophilic thiol–ene click reaction (a) and post-polymerization modification of P1a with 1-propanethiol (b). |
The results obtained from model experiments indicate that the main chain reactive double bonds of an imide-yne polymer could be readily functionalized in the same way described in the model reaction. To examine this, a post-polymerization modification reaction was employed on P1a and the reactive double bonds on the main chain were further reacted with 1.5 equiv. of 1-propanethiol (per alkene unit) in the presence of 0.25 equiv. of TBD for 15 min (Scheme 2b). According to the 1H NMR spectrum of modified P1a (Fig. 2), the addition was found to be quantitative. N-Vinyl proton peaks were diminished and a new peak emerged at δ 5.56 ppm corresponding to the acetal proton. Also, the expected signals arising from 1-propanethiol were detected at δ 3.30 (SCH2), partially overlapping with the H2O peak at 3.33 ppm as well as at δ 1.52 (SCH2CH2) and 0.85 (SCH2CH2CH3) ppm. Finally, aromatic protons appeared as singlets at δ 8.04, 7.85 & 7.70 ppm. 13C NMR spectrum also displayed the full disappearance of main chain double bond carbons and the formation of corresponding acetal carbon at around δ 50 ppm after the reaction (Fig. S125†). Collectively, the NMR results indicate a successful post-modification reaction. It should also be noted that modified P1a exhibited a lower Tg value, decreasing from 199 °C to 118 °C (see Fig. S126 and S128†). This decrease can be attributed to the increased flexibility and free volume after modification.
It is worth mentioning here that we found that we could only add the imide groups to the propiolate's carbon once, using DABCO as the catalyst. We also attempted a second imide addition but failed; however, we were able to achieve a second addition using thiols in the presence of TBD. Following this observation, we decided that the selective mono addition of imides to the triple bonds could be utilized for the preparation of imide-yne monomers and these monomers can undergo further thiol-Michael addition reactions to produce poly(imide thioether)s. In our previous publications, we demonstrated that when activated alkynes are reacted with dithiol compounds, the outcomes of the reactions are typical polythioethers that proceed via double thiol-Michael addition.47–49 We also indicated in these studies that dialkyl acetylenedicarboxylates are more reactive and offer better results than dipropiolates in terms of molecular weight and yield, as the alkyne unit is connected to two carbonyl groups, making the alkyne more electron-deficient and therefore more reactive towards thiols. As such, in this study, when preparing imide-yne monomers, we purposely reacted the imide compounds with DMAD. Scheme 3 depicts the synthesis of imide-yne monomers and the following poly(imide thioether)s. In the first step, bisimide compounds 1–3 were reacted with excess DMAD in CHCl3 at room temperature using DABCO as the catalyst. All three monomers were isolated by precipitation in hexane. It should be noted here that bisimide monomers 2 and 3 are not highly soluble in organic solvents and therefore are not used in imide-yne polymerizations as mentioned previously. The same solubility issue is also valid in the preparation of monomers using CHCl3 as the reaction solvent. Moreover, when compounds 2 and 3 were mixed with CHCl3, a suspension was formed, and we observed that this situation did not change when DMAD and DABCO were added to the reaction medium. However, a clear solution started to form when stirring was continued, indicating the expected monomer formation.
The obtained monomers could be easily characterized by NMR, Fig. 3a shows the 1H NMR spectrum of representative monomer 3m. As seen in this figure, both E- (δ 7.18 ppm) and Z-isomer (δ 6.72 ppm) products are formed simultaneously, with the E-isomer in excess (∼90%). Next, nucleophilic thiol–ene step-growth polymerization was performed using 3m and HDT. The polymerization proceeded for 5 min at room temperature in CHCl3 and low molecular weight polymer (Mw = 7.44 kDa) was obtained in high yield, almost the same results were found for the other two monomers (Table 3). Notably, increasing polymerization duration did not improve molecular weight or yield (data not shown). Fig. 3b displays the corresponding 1H NMR spectrum of polymer (P3m). As can be seen in this figure, vinyl signals completely disappeared and new peaks regarding the polymeric structure could be detected. Here, NCH protons appear at δ 5.36 ppm, SCH protons at δ 4.21 ppm, and SCH2 protons at δ 2.50 ppm, respectively, confirming poly(imide thioether) formation.
Entry | Polymer | M wb (kDa) | Đ | T gc (°C) | Isolated yieldd (%) |
---|---|---|---|---|---|
a Reaction conditions: 1 mmol of imide-yne monomer, 1 mmol of HDT, and 0.25 mmol of TBD in 1 mL of CHCl3 at room temperature for 5 minutes. b Determined by GPC calibrated based on linear PS standards in THF. c Determined by DSC from the second heating cycle. d Gravimetric yield obtained after first precipitation. | |||||
1 | P1m | 8.24 | 1.80 | 97 | 81 |
2 | P2m | 8.73 | 1.81 | 98 | 88 |
3 | P3m | 7.44 | 1.93 | 91 | 85 |
The Tg values of the synthesized poly(imide thioether)s were found to be similar (Table 3) and no significant change was observed regarding their structures. As noted above for polymers P1c, P4a, and P5a which contain sulfur linkages, the poly(imide thioether)s exhibited lower Tg values compared to poly(imide ester)s.
Traditional polyimides are thermally stable and display exceptional thermal properties. The TGA thermograms of all polyimides synthesized herein are given in the ESI† and the results are collected in Table 4. All poly(imide ester)s based on bisimide 1, displayed similar TGA spectra. The 10% weight loss temperatures (T10) were above 310 °C for all samples (P1a, P1b, P1c, P1d, and P1e). The presence of the disulfide linkages led to a remarkable decrease in T10 temperatures (P1c). This early decomposition in P1c is also reflected in the Tmax temperatures. While the Tmax values were over 400 °C in P1a, P1b, P1d, and P1e, in P1c, the Tmax value was shifted to lower temperatures and besides, instead of one Tmax temperature, two maximum weight loss temperatures were observed for P1c. Nevertheless, the early decomposition positively affected the char yield of P1c, and a high char yield of 15.8% was obtained at 750 °C under a nitrogen atmosphere. As expected, the highest char yield (21.4%) was obtained for the P-containing poly(imide ester), i.e., P1e, among the poly(imide ester)s synthesized by using the bismide 1. P4a, which contains thioether bonds-containing bisimide, displayed lower T10 and Tmax temperatures but a significant amount of char similar to P1c compared to other poly(imide ester)s based on bisimide 1. Again, in this case, the presence of thiol groups led to early degradation but in turn, produced a remarkable char yield. P5a displayed a higher char yield compared to P4a due to the presence of the aromatic groups in its structure. P6a displayed the lowest T10 temperature among the studied poly(imide ester)s which could be ascribed to the presence of unreacted monomers and solvent. Nevertheless, P6a produced a high amount of char yield due to the additional thermal stability brought by piperazine rings.
Entry | T 10a (°C) | T maxb (°C) | Charc (%) |
---|---|---|---|
a T 10 is the 10% weight loss temperature. b T max values were obtained from the derivative weight loss curves. c Char value at 750 °C under a nitrogen atmosphere. | |||
P1a | 399 | 457 | 15.2 |
P1b | 392 | 442 | 10.5 |
P1c | 311 | 324/387 | 15.8 |
P1d | 367 | 434 | 14.8 |
P1e | 391 | 426 | 21.4 |
P4a | 314 | 361 | 17.6 |
P5a | 258 | 384 | 25.4 |
P6a | 163 | 273/423 | 24.0 |
P1m | 195 | 353/405 | 7.4 |
P2m | 295 | 346/408 | 11.5 |
P3m | 304 | 344/407 | 8.8 |
Moreover, poly(imide thioether)s displayed relatively lower T10 and Tmax temperatures due to the flexible and thermally weak thioether linkages and displayed two-step degradation profiles. The first maximum weight loss temperature was attributed to the decomposition of the thiol linkages while the second maximum weight loss temperature was ascribed to the degradation of the imide groups. The Tmax temperatures of the synthesized poly(imide thioether)s were found to be similar regardless of their structures but P2m produced relatively higher char yield compared to P1m and P3m, which could be ascribed to the presence of thermally stable benzophenone units in P2m.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4py00918e |
This journal is © The Royal Society of Chemistry 2024 |