DOI:
10.1039/D5PY00580A
(Paper)
Polym. Chem., 2025, Advance Article
A continuous flow based irreversible polycondensation enables synthesis of polycarbonate diols beyond batch limitations†
Received
10th June 2025
, Accepted 11th July 2025
First published on 14th July 2025
Abstract
Aliphatic polycarbonate polyols have emerged as valued precursors for high-performance polyurethanes due to their superior hydrolytic and thermal stability. These polyols are industrially produced using conventional batch reactors which suffer from limited control over polycondensation and low efficiency of production. In this work we report a continuous-flow platform for the continuous production of polycarbonate diols. This process is based on the irreversible polycondensation of short-chain diols with diphenyl carbonate in the presence of a methanesulfonic acid catalyst. This system enables quasi-first-order kinetics with high yield in short residence time, affording polycarbonate diols with a molecular weight range of Mn = 1200–2600 g mol−1 and a dispersity range of 1.9–2.3. We extend the continuous process for the production of polyurethanes by utilizing the crude polycarbonate diols directly for polyurethane synthesis, enabling an integrated and purification-free process. This work establishes a unified flow platform for the continuous production of step-growth polymers with high precision, providing a promising solution to the longstanding limitations of batch production.
Introduction
Polyols, as key precursors in polyurethane production, are generally classified into four main types: polyether, polyester, polycarbonate, and polyamide polyols. Among them, aliphatic polycarbonate polyols stand out due to the presence of carbonate linkages with delocalized electron clouds formed by three oxygen atoms, which impart excellent hydrolytic stability, thermal resistance, and mechanical strength.1 These properties are markedly superior to those of conventional polyether and polyester polyols, making polycarbonate-based systems highly attractive for high-performance and durable polyurethane applications.2–10
Besides ring-opening polymerization methods11–17 and coupling of CO2 and diols,18–21 step-growth polycondensation of diols with carbonate monomers, including dimethyl carbonate (DMC), diethyl carbonate (DEC), diphenyl carbonate (DPC) and carbonyldiimidazole, is a major approach for the synthesis of polycarbonates. Among them, DMC and DEC are widely employed for the polycondensation with primary diols efficiently affording polycarbonate catalyzed by metal or organo-based catalysts.22–28 In contrast, DPC and carbonyldiimidazole, benefiting from the low nucleophilicity of its by-product, are more suitable for the polycondensation with less reactive secondary diols (e.g., isosorbide),29,30 heterogeneous polycondensation,31,32 and high-molecular-weight and structurally well-defined polycarbonate materials.33–36 However, traditional synthesis of polycarbonate polyols is predominantly carried out via batch polycondensation, which suffers from prolonged reaction times and broad molecular weight distributions. This may negatively impact the mechanical performance, durability, and long-term service stability of the final polyurethane products. Therefore, the development of new synthetic strategies that enable efficient and controllable preparation of polycarbonate polyols is essential for advancing the performance of polyurethane materials.
Microfluidic reactors are powerful platforms for polymer synthesis through microscale effects to enhance mass and heat transfer. They have been widely employed in chain-growth polymerizations, including ring-opening polymerization, free-radical polymerization, and ring-opening metathesis polymerization.37–43 However, step-growth polymerizations remain challenging to implement in microflow systems. This difficulty arises from the in situ generated small-molecule byproducts, which are not readily removed from the confined reaction environment, thereby hindering the forward progression of equilibrium-controlled reactions. Thus, the suppression of the reverse reaction becomes a critical requirement for enabling continuous-flow polycondensation. Achieving this demands a significant kinetic disparity between the forward and reverse reactions.42 Although reversible step-growth polymerizations remain elusive under continuous-flow conditions, a limited number of examples involving irreversible polycondensation reactions have been demonstrated.44–53
Oshimura et al. reported an irreversible polycondensation of diols with DPC catalyzed by dilithium tetra-tert-butylzincate. Production of aliphatic polycarbonates (Mn ≥ 10 kg mol−1) can be achieved under mild conditions (25–70 °C).54 This irreversible polycondensation approach significantly enhanced the structural stability and controllability of the polymers, laying a solid foundation for the continuous industrial production of polycarbonate polyols. We envision that the use of this approach could open new avenues for the production of polycarbonate diols by step-growth polymerization using a continuous microfluidic approach.
In this work, we report for the first time a continuous synthesis method for aliphatic polycarbonate polyols using a series of aliphatic diols and DPC in a microflow reactor system catalyzed by organic acids. Due to the lack of nucleophilicity of the generated phenol, we directly utilize the polycarbonate diols for the production of industrially relevant polyurethanes by reacting them with polyisocyanate precursors. Polyurethanes are ideal synthetic targets as they represent a versatile class of functional polymer materials, extensively applied as elastomers, adhesives, coatings, foams etc. By leveraging the enhanced mass and heat transfer in the microflow system, this approach offers a highly efficient and controllable pathway for the synthesis of polycarbonate polyols, providing new insights into their industrial production and application (Scheme 1).
 |
| Scheme 1 Schematic representation of the synthesis of aliphatic polycarbonate polyol and its corresponding polyurethane using continuous flow chemistry. | |
Results and discussion
To investigate the catalytic irreversible polycondensation between diols and diphenyl carbonate (DPC), we initially screened several commercially available catalysts, including DBU (1,8-diazabicyclo[5.4.0]undec-7-ene), NaOH, TBD (1,5,7-triazabicyclo[4.4.0]dec-5-ene), sulfamic acid (ASA), and methanesulfonic acid (MSA). The initial attempt was made using 1,6-hexanediol (HDL) and DPC at 140 °C with an equimolar feed ratio of [HDL]0/[DPC]0 = 1
:
1 and a catalyst loading of 3 mol%. After 4 hours, a portion of the reaction mixture was withdrawn and analyzed by 1H NMR spectroscopy. The results revealed that base-catalyzed systems exhibited relatively low NMR yield (approximately 10–46 mol%, Table 1, entries 1–3) compared to acid catalysts. This is likely attributed to the stronger acidity of the in situ generated phenol relative to HDL, which leads to neutralization of the base catalysts and subsequently suppresses their catalytic activity.1 In contrast, the ASA-catalyzed reaction afforded a yield of 16 mol% (Table 1, entry 4), whereas the MSA-catalyzed system afforded a significantly higher yield of 90 mol%, corresponding to a number-average molecular weight (Mn, NMR) of approximately 1300 g mol−1 (Table 1, entry 5). Upon extending the reaction time to 4–7 hours, the yield further increased to 94–97 mol% (Fig. S1†), with Mn, NMR reaching 1800 g mol−1, confirming the high catalytic efficiency of MSA in promoting the irreversible polycondensation.
Table 1 Polycondensation of 1,6-hexanediol (HDL) and DPC catalyzed by different catalystsa
Having confirmed the catalytic activity of MSA, we next investigated the polymerization kinetics of the reaction in a batch reactor. The monomer conversion increased progressively with time evolution, under the defined conditions (140 °C, [HDL]0/[DPC]0 = 1
:
1, catalyst loading = 3 mol%). The kinetic plot exhibited a linear correlation, indicating that the MSA catalyzed reaction followed apparent first-order kinetics (Fig. 1, kobs = 0.025 min−1). As the temperature increased, kobs rose accordingly. At 160 °C, a maximum kobs of 0.050 min−1 was observed, suggesting that elevated temperatures significantly promote the polycondensation process (Fig. 1a). The kobs values obtained at different reaction temperatures were further used to construct an Arrhenius plot, from which the apparent activation energy (ΔEa) for the MSA-catalyzed reaction was calculated. The resulting value of ΔEa = 18.9 kcal mol−1 indicates a relatively moderate activation barrier that is compatible with the experimental conditions used (120–160 °C) for the MSA-catalyzed polycondensation process (Fig. 1b).
 |
| Fig. 1 (a) Catalytic kinetic processes of methanesulfonic acid catalysis at different temperatures (120–160 °C); (b) Arrhenius plot of MSA catalyzed polycondensation reaction. Experimental conditions: [HDL]0/[DPC]0 = 1 : 1, catalyst loading = 3 mol%. | |
To gain mechanistic insight into the MSA-catalyzed coupling reaction between aliphatic diols and diphenyl carbonate (DPC), density functional theory (DFT) simulation was performed. Coulembier and Taton revealed a dual activation mechanism of the ROP of cyclic ester using benzonic acid as a catalyst.55 Therefore, we postulate that MSA serves as a bifunctional catalyst to facilitate the alcoholysis of DPC through both the hydrogen bond donor and acceptor (Fig. S2†). Overall, the reaction is thermodynamically favorable (ΔG ≈ −18.01 kcal mol−1), where the rate-determining step is a nucleophilic attack through TS1 (ΔG‡ = 22.18 kcal·mol−1), which is in good agreement with the experimentally determined apparent activation energy ΔEa (18.9 kcal mol−1) (Fig. 2a).
 |
| Fig. 2 (a) The DFT simulation of initial coupling reaction between propanol and DPC catalyzed by MSA, leading to the generation of a tetrahedral intermediate (Int2). (b) The transition state for phenol generation with a high energy barrier for the reverse reaction. To reduce computational cost, n-propanol was employed as a model alcohol. Geometry optimizations were calculated at the ωB97XD/6-31+G(d) level, and single-point energies were refined at the ωB97XD/6-31+G(d′,p′) level. | |
DPC is activated by protonated MSA via hydrogen bonding, while the alcohol is simultaneously activated through interaction with MSA's sp2 oxygen, establishing a dual activation mechanism. Phenol elimination to form Int4 is both kinetically and thermodynamically favorable, with a low forward barrier (9.75 kcal mol−1) and a high reverse barrier (29.2 kcal mol−1), due to phenol's low nucleophilicity (Fig. 2b). This irreversibility supports the potential for non-equilibrium polycondensation in continuous-flow irreversible carbonate synthesis.
Based on the results of the batch polycondensation process, we next studied the polycondensation using HDL and DPC in a continuous-flow reactor. According to the Carothers equation,56 to obtain a polycarbonate polyol with an Mn of approximately 2000 g mol−1, the molar ratio of [HDL]0/[DPC]0 was set at 1
:
0.9, introducing a slight stoichiometric imbalance to limit the degree of polymerization. The initial investigation of the effect of the flow rate on polymerization performance was performed with a catalyst loading = 3 mol%, T = 160 °C, using a reactor with an inner diameter of 1 mm (d = 1 mm) and a fixed residence time of 4 h. The influence of the flow rate on continuous-flow irreversible polycondensation was systematically evaluated. As the flow rate increased from 0.004 to 0.033 mL min−1, the monomer conversion gradually increases from 61 to 96 mol%, accompanied by a notable growth in Mn, NMR from 200 to 2000 g mol−1. The increase in molecular weight with higher flow rates may be attributed to reduced polymer chain entanglement under rapid flow conditions (Fig. 3a). This enhancement is more accurately attributed not to polymer chain disentanglement, which is negligible at such low flow velocities, but rather to the hydrodynamic behavior characteristic of laminar flow regimes. Under these conditions, residence time distribution (RTD), axial dispersion, and parabolic flow profiles can significantly affect the spatial homogeneity of reaction progress along the channel. Higher flow rates can mitigate these effects by reducing axial dispersion and suppressing RTD broadening, leading to more uniform reaction conditions and thus higher conversions and narrower molecular weight distributions.57,58
 |
| Fig. 3 (a) NMR yield and Mn, NMR for the polycondensation of HDL and DPC using MSA at 160 °C utilizing continuous-flow reactors with different flow rates for a 1 mm tube diameter (d = 1 mm); (b) the influence on polymerization performance using continuous-flow reactors with different tube diameters; (c) kinetic plots for the polycondensation of HDL and DPC catalyzed by MSA at 160 °C using continuous-flow reactors (d = 1 mm). | |
The inner diameter of the tubular reactor is a critical design parameter in continuous-flow systems, as it directly influences fluid dynamics, mass and heat transfer efficiency, reaction rate, and system pressure drop. To evaluate the effect of tube diameter on the polymerization performance, experiments were conducted at a fixed flow rate of 0.026 mL min−1. When the reactor diameter was set to 0.5 mm, the NMR yield reached 82 mol%, with a number-average molecular weight (Mn, NMR) of approximately 400 g mol−1. Increasing the diameter to 0.8 mm resulted in a slight drop in yield to 79 mol%, with Mn, NMR around 500 g mol−1. Interestingly, further expansion of the tube diameter to 2.0 mm led to a marked increase in yield (94 mol%) and molecular weight (Mn, NMR ≈ 1400 g mol−1). Notably, the highest yield (97 mol%) and molecular weight (Mn, NMR ≈ 2600 g mol−1) were achieved at a tube diameter of 1.0 mm (Fig. 3b). This suggests that a 1.0 mm diameter provides an optimal balance between mass transfer and intrinsic reaction kinetics. Under these conditions, the polymerization is likely dominated by the inherent catalytic kinetics rather than mass transfer limitations, resulting in maximum yield and chain growth. The polymerization followed pseudo-first-order kinetics with respect to monomer consumption, as evidenced by the linear evolution of ln[1/(1 − p)] versus time (Fig. 3c).
To investigate the feasibility of synthesizing polycarbonate polyols in a continuous-flow reactor, we evaluated a series of diols with varying chain lengths. Diols including 1,7-heptanediol (1,7-HD), 1,8-octanediol (1,8-OD), 1,9-nonanediol (1,9-ND), 1,10-decanediol (1,10-DD), and 1,4-cyclohexanedimethanol (CHDM) were selected for continuous flow polycondensation under the given conditions. The corresponding polycarbonate polyols were obtained, with number-average molar masses ranging from 1400 to 2600 g mol−1 (ĐM = 1.9–2.1), showing unimodal distributions (Fig. 4a and b). Thermal analysis revealed distinct melting temperatures (Tm) and crystallization behaviors depending on the diol structure. For linear aliphatic diols (from 1,6-hexanediol to 1,9-nonanediol), a clear odd–even effect on melting temperature was observed (23.8–42.4 °C)59–61 (Fig. 4c). The melting points were generally lower than those of higher-molecular-weight analogues, which can be attributed to chain-end disruption and the presence of smaller crystalline domains in low-molecular-weight polyols.62 All long-chain polycarbonate polyols exhibited crystallization temperatures (Tc), which increased with the length of the aliphatic spacer (−12.3–33 °C) (Fig. 4d). In contrast, the polycarbonate polyol derived from 1,4-cyclohexanedimethanol showed no detectable melting transition, likely due to the inability of the rigid cycloaliphatic moieties to pack into an ordered crystalline lattice. Notably, the glass transition temperature (Tg) of the obtained polycarbonate polyol was not observed. This could have resulted from the low molecular weight of the polyols and their high chain flexibility, thereby leading to a broad and weak thermal transition.63,64
 |
| Fig. 4 (a) and (b) The SEC curves of various polycarbonate polyols using a continuous flow reactor. (c) The 2nd heating DSC curves of obtained polycarbonates. (d) The cooling DSC curves of the obtained polycarbonates. | |
Based on the results of the synthesis of polycarbonate diols in micro-flow reactors and concerning the much lower reactivity of phenol with isocyanates compared to aliphatic alcohols, we directly employed the crude mixture of polyols and the by-product phenol for polyurethane synthesis under micro-flow conditions. We selected the polyaddition between poly(hexamethylene carbonate) diol and 4,4′-methylene-bis(cyclohexyl isocyanate) (HMDI) as a model reaction. Given that 3 mol% MSA enables nearly complete conversion in under 1 second, we reduced the catalyst loading to 0.1 mol% to investigate the polymerization kinetics. The reaction was monitored in situ using FTIR-ATR with a flow-through cell by following the band associated with carbamate group formation at ∼1600 cm−1. The results show that the inner diameter of the tubular reactor strongly influences the reaction rate. A diameter of 0.8 mm yielded an observed rate constant of kobs = 0.14 min−1, significantly higher than those of 1 mm (kobs = 0.09 min−1) and 2 mm (kobs = 0.07 min−1). This enhancement is likely due to increased shear and improved diffusion in narrower channels, which facilitate more efficient mixing between the polycarbonate diol and HMDI (Fig. 5a–d).
 |
| Fig. 5 (a) FTIR spectroscopy characterized polycondensation between polycarbonate and HMDI with time evolution using a continues-flow reactor (d = 0.8 mm); (b) the diameter of the continuous-flow reactor is 1 mm; (c) d = 2 mm; experimental conditions for all kinetic studies: [poly(hexamethylene carbonate) diol]0/[HMDI]0/[MSA]0 = 1 : 1.1 : 0.001 at 25 °C; (d) semilogarithmic plot of polycondensation; (e) the 2nd heating (10 °C min−1 under Ar) DSC curve of the obtained polycarbonate-based polyurethane; and (f) the SEC curve of polycarbonate-based polyurethane using a continuous flow reactor. | |
Polyurethane pressure-sensitive adhesives (PU-PSAs) offer tunable adhesion properties by tailoring the match between the polyol and diisocyanate components, enabling their application across diverse fields including electronic encapsulation, medical dressings, high-end labeling, and removable tapes.65–68 Taking 1,6-hexadiol based polycarbonate as an example due to its commercial and industrial availability, we synthesized a polyurethane prepolymer using a micro-flow reactor and subsequently crosslinked them with trimethylolpropane (TMP) to obtain PU-PSA networks. The polyurethane adhesives were coated using a precision coating machine to form thin films with low viscosity, and then cured at 80 °C in a convection oven for 24 hours. The resulting PU-PSA samples were characterized for their adhesive performance using 180° peel strength testing on stainless steel in accordance with ISO 29862:2024 (Scheme 2). The results revealed a non-linear correlation between the peel strength and degrees of crosslinking (Fig. 6a). At a low molar ratio of 0.11, the adhesive showed limited cohesion strength, yielding a modest peel force. As the crosslinker ratio increased to 0.22, the peel strength reached a maximum (∼0.054 N cm−1), suggesting the formation of an optimal crosslinked network that balanced interfacial adhesion and cohesive integrity. However, further increasing the crosslinker content led to a noticeable decline in peel strength. These results indicate that a moderate crosslinking level is critical for achieving high peel resistance in polyurethane adhesives based on polycarbonate polyols (Fig. 6b–f).
 |
| Scheme 2 Schematic illustration of the preparation of polyurethane adhesive samples and the 180° tape peel test. | |
 |
| Fig. 6 (a) The comparison of the adhesive strengths of polyurethane with different ratios of TMP; (b) [polycarbonate diol]0/[HMDI]0/[TMP]0 = 1 : 1.5 : 0.11; (c) [polycarbonate diol]0/[HMDI]0/[TMP]0 = 1 : 1.5 : 0.16; (d) [polycarbonate diol]0/[HMDI]0/[TMP]0 = 1 : 1.5 : 0.22; (e) [polycarbonate diol]0/[HMDI]0/[TMP]0 = 1 : 1.5 : 0.27; and (f) [polycarbonate diol]0/[HMDI]0/[TMP]0 = 1 : 1.5 : 0.33. | |
Conclusion
Aliphatic polycarbonate polyols are promising and valuable building blocks for high-performance polyurethanes due to their superior hydrolytic and thermal stability. However, conventional batch step-growth polymerizations suffer from limited control over the polymer structure and low efficiency of production. We developed a continuous-flow strategy for the synthesis of polycarbonate diols via the MSA-catalyzed irreversible polycondensation of diols and diphenyl carbonate. Mechanistic and kinetic results revealed that MSA operates through dual hydrogen-bond activation, enabling a quasi-first-order reaction with high NMR yield in short residence time. The resulting polycarbonate diols exhibited controlled molecular weights and could be directly applied to polyurethane synthesis in flow. This study establishes a practical, scalable, and continuous platform for the efficient production of functional polyurethanes.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the ESI.†
Acknowledgements
The authors are thankful for the financial support from the National Key Research & Development Program of China (2021YFC2101900), the National Natural Science Foundation of China (No. U24A20492) and Nanjing Municipal Science and Technology Bureau, and also appreciate the high performance center of Nanjing Tech University for supporting the computational resources.
References
- M. Scharfenberg, S. Hofmann, J. Preis, J. Hilf and H. Frey, Rigid Hyperbranched Polycarbonate Polyols from CO2 and Cyclohexene-Based Epoxides, Macromolecules, 2017, 50(16), 6088–6097 CrossRef CAS
. - R. Zhu, Y. Wang, Z. Zhang, D. Ma and X. Wang, Synthesis of polycarbonate urethane elastomers and effects of the chemical structures on their thermal, mechanical and biocompatibility properties, Heliyon, 2016, 2(6), e00125 CrossRef PubMed
. - Y.-H. Wu, C.-C. Wang and C.-Y. Chen, Effect of the cyclic structure content on aliphatic polycarbonate-based polyurethane, Polym. J., 2021, 53(6), 695–702 CrossRef CAS
. - M. Song, X. Yang and G. Wang, Preparation of polycarbonate diols (PCDLs) from dimethyl carbonate (DMC) and diols catalyzed by KNO3/γ-Al2O3, RSC Adv., 2018, 8(61), 35014–35022 RSC
. - M. Rogulska, Polycarbonate-based thermoplastic polyurethane elastomers modified by DMPA, Polym. Bull., 2019, 76(9), 4719–4733 CrossRef CAS
. - C. Wongsamut, R. Suwanpreedee and H. Manuspiya, Thermoplastic polyurethane-based polycarbonate diol hot melt adhesives: The effect of hard-soft segment ratio on adhesion properties, Int. J. Adhes. Adhes., 2020, 102, 102677 CrossRef CAS
. - S. M. Cakić, M. Špírková, I. S. Ristić, J. K. B-Simendić, M. M-Cincović and R. Poręba, The waterborne polyurethane dispersions based on polycarbonate diol: Effect of ionic content, Mater. Chem. Phys., 2013, 138(1), 277–285 CrossRef
. - M. Fuensanta, J. A. Jofre-Reche, F. Rodríguez-Llansola, V. Costa, J. I. Iglesias and J. M. Martín-Martínez, Structural characterization of polyurethane ureas and waterborne polyurethane urea dispersions made with mixtures of polyester polyol and polycarbonate diol, Prog. Org. Coat., 2017, 112, 141–152 CrossRef CAS
. - W. Chunarrom and H. Manuspiya, The dielectric and polarization behavior of polyurethane-based polycarbonate diols with different content levels of fluorinated hard segments, Polym. Chem., 2021, 12(8), 1136–1146 RSC
. - F. Campana, G. Brufani, F. Mauriello, R. Luque and L. Vaccaro, Green polyurethanes from bio-based building blocks: Recent advances and applications, Green Synth. Catal., 2024 DOI:10.1016/j.gresc.2024.08.001
. - G. A. Bhat and D. J. Darensbourg, Progress in the catalytic reactions of CO2 and epoxides to selectively provide cyclic or polymeric carbonates, Green Chem., 2022, 24(13), 5007–5034 RSC
. - J. Huang, J. C. Worch, A. P. Dove and O. Coulembier, Update and Challenges in Carbon Dioxide-Based Polycarbonate Synthesis, ChemSusChem, 2020, 13(3), 469–487 CrossRef CAS PubMed
. - G.-W. Yang, R. Xie, Y.-Y. Zhang, C.-K. Xu and G.-P. Wu, Evolution of Copolymers of Epoxides and CO2: Catalysts, Monomers, Architectures, and Applications, Chem. Rev., 2024, 124(21), 12305–12380 CrossRef CAS PubMed
. - A. J. Plajer and C. K. Williams, Heterocycle/Heteroallene Ring-Opening Copolymerization: Selective Catalysis Delivering Alternating Copolymers, Angew. Chem., Int. Ed., 2022, 61(1), e202104495 CrossRef CAS PubMed
. - C. Zhang, X. Geng, X. Zhang, Y. Gnanou and X. Feng, Alkyl borane-mediated metal-free ring-opening (co)polymerizations of oxygenated monomers, Prog. Polym. Sci., 2023, 136, 101644 CrossRef CAS
. - W. Yu, E. Maynard, V. Chiaradia, M. C. Arno and A. P. Dove, Aliphatic Polycarbonates from Cyclic Carbonate Monomers and Their Application as Biomaterials, Chem. Rev., 2021, 121(18), 10865–10907 CrossRef CAS PubMed
. - B. Grignard, S. Gennen, C. Jérôme, A. W. Kleij and C. Detrembleur, Advances in the use of CO2 as a renewable feedstock for the synthesis of polymers, Chem. Soc. Rev., 2019, 48(16), 4466–4514 RSC
. - Y. Chai, Q. Chen, C. Huang, Q. Zheng, M. North and H. Xie, Introducing the reversible chemistry of CO2 with diols mediated by organic superbases into polycarbonate synthesis, Green Chem., 2020, 22(15), 4871–4877 RSC
. - Z. Chen, N. Hadjichristidis, X. Feng and Y. Gnanou, Cs2CO3-promoted polycondensation of CO2 with diols and dihalides for the synthesis of miscellaneous polycarbonates, Polym. Chem., 2016, 7(30), 4944–4952 RSC
. - M. Tamura, Y. Nakagawa and K. Tomishige, Direct CO2 Transformation to Aliphatic Polycarbonates, Asian J. Org. Chem., 2022, 11(11), e202200445 CrossRef CAS
. - Y. Gu, K. Matsuda, A. Nakayama, M. Tamura, Y. Nakagawa and K. Tomishige, Direct Synthesis of Alternating Polycarbonates from CO2 and Diols by Using a Catalyst System of CeO2 and 2-Furonitrile, ACS Sustainable Chem. Eng., 2019, 7(6), 6304–6315 CrossRef CAS
. - L. Meabe, N. Lago, L. Rubatat, C. Li, A. J. Müller, H. Sardon, M. Armand and D. Mecerreyes, Polycondensation as a Versatile Synthetic Route to Aliphatic Polycarbonates for Solid Polymer Electrolytes, Electrochim. Acta, 2017, 237, 259–266 CrossRef CAS
. - J. Sun and D. Kuckling, Synthesis of high-molecular-weight aliphatic polycarbonates by organo-catalysis, Polym. Chem., 2016, 7(8), 1642–1649 RSC
. - W. Fang, Z. Zhang, Z. Yang, Y. Zhang, F. Xu, C. Li, H. An, T. Song, Y. Luo and S. Zhang, One-pot synthesis of bio-based polycarbonates from dimethyl carbonate and isosorbide under metal-free condition, Green Chem., 2020, 22(14), 4550–4560 RSC
. - B. Jiang and C. M. Thomas, Efficient synthesis of camphor-based polycarbonates: a direct route to recyclable polymers, Catal. Sci. Technol., 2023, 13(13), 3910–3915 RSC
. - G. Fiorani, A. Perosa and M. Selva, Dimethyl carbonate: a versatile reagent for a sustainable valorization of renewables, Green Chem., 2018, 20(2), 288–322 RSC
. - H. Yao, Y. Zuo, J. Zhang, H. Yang and G. Zhang, Organocatalytic Condensation Polymerization for Synthesis of High Performance Aliphatic Polycarbonate, Macromol. Rapid Commun., 2025, 2401044 CrossRef CAS PubMed
. - J. H. Park, J. Y. Jeon, J. J. Lee, Y. Jang, J. K. Varghese and B. Y. Lee, Preparation of High-Molecular-Weight Aliphatic Polycarbonates by Condensation Polymerization of Diols and Dimethyl Carbonate, Macromolecules, 2013, 46(9), 3301–3308 CrossRef CAS
. - C. Ma, F. Xu, W. Cheng, X. Tan, Q. Su and S. Zhang, Tailoring Molecular Weight of Bioderived Polycarbonates via Bifunctional Ionic Liquids Catalysts under Metal-Free Conditions, ACS Sustainable Chem. Eng., 2018, 6(2), 2684–2693 CrossRef CAS
. - H. Wang, X. Zhang, Z. Zhang, Y. Zhang, Z. Yang, W. Zhang, C. Li and F. Xu, Quantum-Chemically Guided Insights on Ionic Liquid Catalysts for Biobased Polycarbonate Synthesis—Mechanistic Exploration and Cation-Functionalized Optimization, ACS Sustainable Chem. Eng., 2025, 13(12), 4857–4872 CrossRef CAS
. - R. Chong, F. Qian, Z. Sun, M. Wei, Z. Zhang, J. Zhang, M. He, Q. Chen and J. Qian, Efficient Transesterification Preparation of Polycarbonate Diol with High Molecular Weight Using TiO2/SiO2 Solid Acid Catalysts, ChemistrySelect, 2023, 8(18), e202204646 CrossRef CAS
. - X.-L. Shen, Z.-Q. Wang, Q.-Y. Wang, S.-Y. Liu and G.-Y. Wang, Synthesis of Poly(isosorbide carbonate) via Melt Polycondensation Catalyzed by Ca/SBA-15 Solid Base, Chin. J. Polym. Sci., 2018, 36(9), 1027–1035 CrossRef CAS
. - M. Zheng, W. Peng, Y.-Y. Zhang, X. Zhang and W. Chen, Synthesis, Characterization, and Application of High-Molecular-Weight Polyethylene-like Polycarbonates: Toward Sustainable and Recyclable Fibers, Macromolecules, 2024, 57(15), 7020–7030 CrossRef CAS
. - Z.-W. Xu, J.-L. Pan, J. Xu, W.-M. Cheng, Z.-L. Li and C. Cheng, Recyclable, Reprocessable, Anticorrosive, and Hydrolysis-Resistant Long-Chain Aliphatic Polycarbonates as Polyethylene-like Materials, ACS Appl. Polym. Mater., 2024, 6(2), 1551–1562 CrossRef CAS
. - J. V. Olsson, D. Hult, S. García-Gallego and M. Malkoch, Fluoride-promoted carbonylation polymerization: a facile step-growth technique to polycarbonates, Chem. Sci., 2017, 8(7), 4853–4857 RSC
. - E. H. Choi, J. Lee, S. U. Son and C. Song, Biomass-derived furanic polycarbonates: Mild synthesis and control of the glass transition temperature, J. Polym. Sci., Part A: Polym. Chem., 2019, 57(17), 1796–1800 CrossRef CAS
. - J.-i. Yoshida, Basics of flow microreactor synthesis, Springer, 2015 Search PubMed
. - F. Bally, C. A. Serra, V. Hessel and G. Hadziioannou, Homogeneous polymerization: benefits brought by microprocess technologies to the synthesis and production of polymers, Macromol. React. Eng., 2010, 4(9–10), 543–561 CrossRef CAS
. - T. Junkers, Precise macromolecular engineering via continuous-flow synthesis techniques, J. Flow Chem., 2017, 7(3–4), 106–110 CrossRef CAS
. - M. H. Reis, F. A. Leibfarth and L. M. Pitet, Polymerizations in Continuous Flow: Recent Advances in the Synthesis of Diverse Polymeric Materials, ACS Macro Lett., 2020, 9(1), 123–133 CrossRef CAS PubMed
. - X. Hu, N. Zhu, Z. Fang and K. Guo, Continuous flow ring-opening polymerizations, React. Chem. Eng., 2017, 2(1), 20–26 RSC
. - Y. Su, Y. Song and L. Xiang, Continuous-Flow Microreactors for Polymer Synthesis: Engineering Principles and Applications, Top. Curr. Chem., 2018, 376(6), 44 CrossRef PubMed
. - C. Tonhauser, A. Natalello, H. Löwe and H. Frey, Microflow technology in polymer synthesis, Macromolecules, 2012, 45(24), 9551–9570 CrossRef CAS
. - D. Kessler, H. Löwe and P. Theato, Synthesis of Defined Poly(silsesquioxane)s: Fast Polycondensation of Trialkoxysilanes in a Continuous-Flow Microreactor, Macromol. Chem. Phys., 2009, 210(10), 807–813 CrossRef CAS
. - P. Wang, K. Wang, J. Zhang and G. Luo, Preparation of poly(p-phenylene terephthalamide) in a microstructured chemical system, RSC Adv., 2015, 5(79), 64055–64064 RSC
. - H. Seyler, D. J. Jones, A. B. Holmes and W. W. H. Wong, Continuous flow synthesis of conjugated polymers, Chem. Commun., 2012, 48(10), 1598–1600 RSC
. - G. Pirotte, S. Agarkar, B. Xu, J. Zhang, L. Lutsen, D. Vanderzande, H. Yan, P. Pollet, J. R. Reynolds, W. Maes and S. R. Marder, Molecular weight tuning of low bandgap polymers by continuous flow chemistry: increasing the applicability of PffBT4T for organic photovoltaics, J. Mater. Chem. A, 2017, 5(34), 18166–18175 RSC
. - G. Pirotte, J. Kesters, P. Verstappen, S. Govaerts, J. Manca, L. Lutsen, D. Vanderzande and W. Maes, Continuous Flow Polymer Synthesis toward Reproducible Large-Scale Production for Efficient Bulk Heterojunction Organic Solar Cells, ChemSusChem, 2015, 8(19), 3228–3233 CrossRef CAS PubMed
. - R. Singh, K. Veeramani, R. Bajpai and A. Kumar, High-Throughput Template-Free Continuous Flow Synthesis of Polyaniline Nanofibers, Ind. Eng. Chem. Res., 2019, 58(15), 5864–5872 CrossRef CAS
. - X. Lopez de Pariza, T. Erdmann, P. L. Arrechea, L. Perez, C. Dausse, N. H. Park, J. L. Hedrick and H. Sardon, Synthesis of Tailored Segmented Polyurethanes Utilizing Continuous-Flow Reactors and Real-Time Process Monitoring, Chem. Mater., 2021, 33(20), 7986–7993 CrossRef CAS
. - F. Siragusa, L. Crane, P. Stiernet, T. Habets, B. Grignard, J.-C. M. Monbaliu and C. Detrembleur, Continuous Flow Synthesis of Functional Isocyanate-Free Poly(oxazolidone)s by Step-Growth Polymerization, ACS Macro Lett., 2024, 13(5), 644–650 CrossRef CAS PubMed
. - Y. Feng, X. Li, T. Ma, Y. Li, D. Ji, H. Qin, Z. Fang, W. He and K. Guo, Preparation of chemically recyclable bio-based semi-aromatic polyamides using continuous flow technology under mild conditions, Green Chem., 2024, 26(9), 5556–5563 RSC
. - J. L. Rapp, M. A. Borden, V. Bhat, A. Sarabia and F. A. Leibfarth, Continuous Polymer Synthesis and Manufacturing of Polyurethane Elastomers Enabled by Automation, ACS Polym. Au, 2024, 4(2), 120–127 CrossRef CAS PubMed
. - M. Oshimura, T. Hirata, T. Hirano and K. Ute, Synthesis of aliphatic polycarbonates by irreversible polycondensation catalyzed by dilithium tetra-tert-butylzincate, Polymer, 2017, 131, 50–55 CrossRef CAS
. - C. Jehanno, L. Mezzasalma, H. Sardon, F. Ruipérez, O. Coulembier and D. Taton, Benzoic Acid as an Efficient Organocatalyst for the Statistical Ring-Opening Copolymerization of ε-Caprolactone and L-Lactide: A Computational Investigation, Macromolecules, 2019, 52(23), 9238–9247 CrossRef CAS
. - W. H. Carothers, Polymers and polyfunctionality, Trans. Faraday Soc., 1936, 32(0), 39–49 RSC
. - D. J. Walsh, D. A. Schinski, R. A. Schneider and D. Guironnet, General route to design polymer molecular weight distributions through flow chemistry, Nat. Commun., 2020, 11(1), 3094 CrossRef CAS PubMed
. - M. H. Reis, T. P. Varner and F. A. Leibfarth, The Influence of Residence Time Distribution on Continuous-Flow Polymerization, Macromolecules, 2019, 52(9), 3551–3557 CrossRef CAS
. - R. A. Pérez-Camargo, G. Liu, L. Meabe, Y. Zhao, H. Sardon, A. J. Müller and D. Wang, Using Successive Self-Nucleation and Annealing to Detect the Solid–Solid Transitions in Poly(hexamethylene carbonate) and Poly(octamethylene carbonate), Macromolecules, 2021, 54(20), 9670–9680 CrossRef
. - I. Flores, R. A. Pérez-Camargo, E. Gabirondo, M. R. Caputo, G. Liu, D. Wang, H. Sardon and A. J. Müller, Unexpected Structural Properties in the Saturation Region of the Odd–Even Effects in Aliphatic Polyethers: Influence of Crystallization Conditions, Macromolecules, 2022, 55(2), 584–594 CrossRef CAS
. - J. Huang, R. Yan, Y. Ni, N. Shi, Z. Li, C. Ma and K. Guo, Cyclic Polycarbonates by N-Heterocyclic Carbene-Mediated Ring-Expansion Polymerization and Their Selective Depolymerization to Monomers, ACS Sustainable Chem. Eng., 2022, 10(46), 15007–15016 CrossRef CAS
. - M. Rubinstein and R. H. Colby, Polymer physics, Oxford university press, 2003 Search PubMed
. - V. N. Novikov and E. A. Rössler, Correlation between glass transition temperature and molecular mass in non-polymeric and polymer glass formers, Polymer, 2013, 54(26), 6987–6991 CrossRef CAS
. - L.-P. Blanchard, J. Hesse and S. L. Malhotra, Effect of molecular weight on glass transition by differential scanning calorimetry, Can. J. Chem., 1974, 52(18), 3170–3175 CrossRef CAS
. - S. Mapari, S. Mestry and S. T. Mhaske, Developments in pressure-sensitive adhesives: a review, Polym. Bull., 2021, 78(7), 4075–4108 CrossRef CAS
. - L. Zeng, L. Yang, J. Liu, S. Lu, L. Ai, Y. Dong, Z. Ye and P. Liu, Preparation and Properties of Flame-Retardant Polyurethane Pressure Sensitive Adhesive and Its Application, J. Compos. Sci., 2023, 7(2), 85 CrossRef CAS
. - D. M. Fitzgerald, Y. L. Colson and M. W. Grinstaff, Synthetic pressure sensitive adhesives for biomedical applications, Prog. Polym. Sci., 2023, 142, 101692 CrossRef CAS PubMed
. - Z. Czech and R. Hinterwaldner, Pressure-sensitive adhesives based on polyurethanes, in Technology of Pressure-Sensitive Adhesives and Products, CRC Press, 2008, pp. 11–1–11–21 Search PubMed
.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.