DOI:
10.1039/D5SC05783C
(Edge Article)
Chem. Sci., 2025, Advance Article
Lewis-acid induced mechanochemical degradation of polyvinylidene fluoride: transformation into valuable products
Received
31st July 2025
, Accepted 5th September 2025
First published on 9th September 2025
Abstract
Polyvinylidene fluoride (–[CH2CF2]n–, PVDF) waste poses significant environmental challenges due to its recalcitrant nature and widespread use. This study addresses the end-of-life management of PVDF by introducing a novel, sustainable mechanochemical approach for its valorisation. We investigated the degradation of PVDF into value-added materials using ball milling with anhydrous AlCl3 to achieve a quantitative mineralisation producing AlF3 and halide-functionalised graphite, along with gaseous products (HCl and CH4). Mechanistic key steps involve Lewis-acid catalysed C–F bond activation, dehydrofluorination and aromatisation. This approach provides an effective solution for PVDF waste management while offering a promising route for the production of high-value materials from polymer waste streams. Our findings contribute to sustainable practices in polymer recycling and resource recovery, respond to pressing environmental concerns associated with fluoropolymer disposal, and demonstrate the potential to convert polymer wastes into useful products.
Introduction
Polyvinylidene fluoride (–[CH2CF2]n–, PVDF) belongs to the class of perfluoro and polyfluoro alkyl substances (PFAS) and has been used extensively in chemical processing equipment, electronic components, water treatment systems and architectural coatings.1,2 PVDF acts as a vital binder in lithium-ion batteries to secure the cohesion between active materials and current collectors.3 While PVDF's exceptional chemical stability enhances battery performance, the increasing demand for these batteries translates to a growing challenge in waste management. Unlike many polymers, PVDF is highly resistant to degradation due to the inherent strength of its C–F bonds.4–6 Research efforts focused on developing efficient PVDF recycling methods are important for the sustainable development of lithium-ion battery technology, as they address the environmental impact of this key component in battery electrodes. Complementing these efforts, research into the controlled degradation of PVDF offers promising alternative strategies for responsible end-of-life management, potentially opening new avenues for material recovery and reuse.
Studies by Hori and co-workers demonstrated a complete mineralisation of PVDF in supercritical water at 380 °C using a 5.8-fold molar excess of oxygen.7 Interestingly, complete mineralisation was also achieved in subcritical water at 300 °C using a 31-fold molar excess of hydrogen peroxide relative to the fluorine content of PVDF.8 In both cases a carbon rich residue was formed which was not characterised further. Using molten NaOH, Yanagihara and Katoh accomplished the degradation of various PFAS including PVDF to NaF and carbonates.9 In addition, Baron and co-workers reported the degradation of PVDF performing mechanochemistry, in which NaOH was used as co-milling agent to form NaF.10 For the latter, water production causes strong agglomeration, which hampers an efficient fluoride/hydroxide exchange and again only unidentified carbon rich compounds and organic phases were generated.
In recent years, mechanochemistry11 has attracted significant interest because of its alignment with green chemistry principles by enabling solvent-free reactions, reducing waste, improving energy efficiency and developing new synthetic routes. This approach minimises environmental impact and maximises atom economy in chemical processes.12,13 For instance, Gouverneur and co-workers established a mechanochemical fluorination approach, using CaF2 to form fluorochemicals without prior production of HF.14 A review article by Aydonat et al. examines how mechanochemistry can enhance polymer sustainability.15 The paper discusses recent advances in mechanophores and applications to various polymers, contributing to efforts in sustainable materials development and plastic waste management.
Dehydrofluorination reactions using solid Lewis acidic Al compounds were reported in the literature. Aluminium chlorofluoride (ACF, AlClxF3−x; x: 0.05–0.3) and tempered γ-Al2O3 at 700 °C have shown reactivity towards fluorinated molecules.16–20
In this work, we report on a unique procedure for the quantitative mineralisation of PVDF by ball milling in the presence of AlCl3 to form AlF3 and, remarkably, a fluorinated and chlorinated graphite. Additionally, valuable gases such as HCl and CH4 were generated during the mechanochemical process (Scheme 1).
 |
| Scheme 1 Reaction of PVDF and AlCl3 forming functionalised graphite, methane, aluminium fluoride and hydrogen chloride. | |
Results and discussion
Mechanochemical degradation of PVDF was conducted in a Fritsch Premium Line 7 planetary mill from Fritsch GmbH, Germany. Anhydrous AlCl3 (267 mg, 2 mmol) and PVDF (192 mg, 3 mmol) were placed in a 45 mL ZrO2 jar equipped with a gassing lid and five ZrO2 balls (each 2.5 mg, 10 mm in diameter). After milling at 800 rpm for 7 h, an insoluble black powder was obtained. Common protic and aprotic solvents (H2O, EtOH, iPrOH, DMSO, CH2Cl2, CHCl3, CCl4, CH3CN, acetone, C6H6) did not dissolve the black powder and impeded subsequent extraction attempts. Consequently, the bulk material was analysed as such; it is referred to as PMPX (planetary milled powder, X means the milling time in hours, PMPXm relates to the milled mixture using a PVDF membrane).
Gas phase analytics
The gaseous contents of the jars after milling were vacuum transferred into a C6D6 solution in JYoung NMR tubes for further characterisation. In the 1H NMR spectrum (Fig. 1a) a distinct signal at 0.10 ppm and a low intensity signal at 4.51 ppm were assigned to CH4 and H2, respectively. Furthermore, gas chromatography (SI Fig. 1) and OFCEAS (Optical Feedback Cavity Enhanced Absorption Spectroscopy) confirm the presence of CH4 after milling (Fig. 1b). In addition, OFCEAS experiments revealed the formation of HCl and trace amounts of CO and CO2, most likely originating from the polymer's oxidation during milling by residual air in the jar. Note that imperfect jar and gassing lid seals may allow some air ingress.
 |
| Fig. 1 (a) 1H NMR spectrum (300 MHz, C6D6) from the gas phase after milling AlCl3 and PVDF for 7 hours (PMP7). (b) Results of the Optical Feedback Cavity Enhanced Absorption Spectroscopy (OFCEAS) for the gaseous contents of PMP4 and PMP7 using H2 as matrix gas. | |
X-ray powder diffraction, XRD
XRD patterns were collected from four samples, each consisting of AlCl3 and PVDF (using powdery PVDF or a PVDF membrane) mixtures that were milled for varying durations. The powders were measured without any further processing. Fig. 2a shows the influence of the milling time on the crystallinity. After 2 h of milling, weak reflections attributed to α-AlF3 and β-AlF3 were observable in the diffractogram for PMP2. After 4 h of milling the reflections for β-AlF3 diminished, while those for the thermodynamically stable α-AlF3 phase21 increased significantly, suggesting enhanced crystallinity of the α-AlF3 phase in PMP4. Notably, the diffraction pattern of the sample milled for 7 h displayed no distinct reflections. This result suggests a substantial amorphisation of the sample PMP7, caused by mechanical impact and increased defect concentrations due to prolonged milling.22,23 When AlCl3 and a PVDF membrane (PMP7m) was milled for 7 h, intense reflections attributed to α-AlF3 and minor reflections for AlF3·3H2O could be observed. Note that Scholz and colleagues reported on the mechanochemical generation of nanocrystalline aluminium hydroxide fluoride samples AlFx(OH)3−x·nH2O (x = 1.5) exhibiting pyrochlore structure by milling aluminium hydroxide and hydrous aluminium fluoride.24
 |
| Fig. 2 (a) X-ray powder diffractograms (Mo Kα source λ = 0.7107 Å) from milling approaches of AlCl3 and PVDF at various milling times. PMPX (planetary milled powder, X means the milling time in hours). For PMP7m PVDF membrane was used. The reference reflections for AlF3·3H2O, α- and β-AlF3 are depicted in grey. (b) Elemental mapping of PMP7 using STEM and EDX with a scale bar of 500 nm. | |
Scanning and transmission electron microscopy, S/TEM
S/TEM measurements were conducted to gain insight into the material. For this, vacuum was applied to PMP7 to remove HCl and CH4 from the material. Energy dispersive X-ray spectroscopy (EDX) suggests an atomic ratio of 1 to 3.1 for Al and F, which is in accordance with the formation of AlF3. Furthermore, EDX analysis revealed a slight excess of fluorine and chlorine atoms when accounting for the stoichiometry of AlF3. This is in accordance with the presence of a fluorinated and chlorinated material. Furthermore, the elemental mapping in Fig. 2b shows a homogenous distribution of Al, Cl, F and C atoms over the whole particles, which demonstrates that there is no phase separation of AlF3 and the graphitic material.
Multinuclear MAS NMR spectroscopy
The 27Al MAS NMR spectrum of PMP7 displays only a signal at −16 ppm for [AlF6] entities (SI Fig. 15).25 The linewidth is 400–500 Hz less than the ones observed for other amorphous doped aluminium fluorides,26,27 suggesting less amorphicity.
In the 19F MAS NMR spectrum (Fig. 3a) for PMP7 one can detect the signal for amorphous AlF3 moieties at −167 ppm.25 In addition, several signals were observed in the characteristic region for CFy (y = 1, 2 or 3) moieties.28 The positions of these signals are consistent with those typically found in commercially available fluorinated graphite (also shown in Fig. 3a for comparison), which indicates the generation of functionalised graphite through the mechanochemical treatment. However, there are subtle differences. The signal at −130 ppm is more intense for PMP7, while the signal at −115 ppm observed in the commercial fluorinated graphite sample is missing for PMP7. The 19F MAS NMR spectrum of PMP4 shows a simplified profile, dominated by a prominent signal attributed to α-AlF3 at −170 ppm. Smaller, less intense signals are observed at −130 and −187 ppm. The comparison of the spectra suggests that longer milling times lead to an increased amount of CFy (with y = 1, 2, 3) moieties. Note that the signal for PVDF at −95 ppm is not observed in the 19F MAS NMR for PMP4 and PMP7. The 19F MAS NMR spectrum of PMP7m is compatible with the presence of amorphous as well as α-AlF3 (SI Fig. 17).
 |
| Fig. 3 (a) 19F MAS NMR ( rot = 25 kHz) spectra of PMP4, PMP7, fluorinated graphite (F-graphite, >61 wt% F) and PVDF as reference; asterisks * represent spinning sidebands. am. stands for amorphous. (b) 19F–13C CP (top) and 13C{1H} MAS NMR ( rot = 10 kHz) spectra of PMP7, F-graphite and PVDF. | |
The 13C{1H} MAS NMR spectrum of PMP7 depicted in Fig. 3b reveals a signal at 0 ppm at a characteristic chemical shift for CH4.29 The signal disappeared when vacuum was applied to the sample implying that the methane molecules were weakly bound to or physically trapped within the graphitic network. In addition, a broad signal centered at 130 ppm indicates the presence of sp2 carbon atoms. The 19F–13C CP MAS NMR spectrum depicted in Fig. 3a (top) reveals that some of these sp2 carbon atoms have a weak interaction with fluorine atoms within the graphitic material.30 Notably, a second signal at 160 ppm can be assigned to sp2 C–F moieties.31 Comparison with the 13C{1H} MAS NMR spectrum of fluorinated graphite revealed significant differences. The spectrum of fluorinated graphite only displayed signals at 88 and 110 ppm, corresponding to CF and CF2 functionalities,32 respectively. In contrast, the 13C{1H} MAS NMR spectrum of PMP7 is dominated by the broad sp2 C resonance at 130 ppm. The absence of signals at 88 and 110 ppm shows that PMP7 has significantly fewer CF and CF2 moieties than fluorinated graphite. This observation aligns with the suggested composition of PMP7, where sp2 C atoms are the primary component, with only a limited presence of C–F bonds. The 13C{1H} MAS NMR spectrum of PMP7m shows mainly a broad signal at −130 ppm (SI Fig. 18).
Attenuated total reflection infrared spectroscopy, ATR-IR
ATR IR spectra (Fig. 4a) were measured to investigate the mechanochemical influence of ball milling on PVDF powder and PVDF membrane with and without the presence of AlCl3. In the IR spectra of PMP4, PMP7 and PMP7m a band at 640 cm−1 was observed and is assigned to Al–F vibrational bands.33 The broader linewidth of the Al–F band in PMP7 suggests an increased structural disorder or potential structural changes.25–27 In addition, the IR spectra of PMP4 and PMP7m show similarities to the spectrum of pure α-AlF3 (SI Fig. 23), with additional bands characteristic of C–F stretching vibrations at 1400–1000 cm−1.34,35 An extension of the milling time to 7 h of the PVDF powder results in intensified bands in this region. For PMP7 a distinct vibrational mode can be found at 805 cm−1 in a characteristic region for C–Cl bands suggesting that the graphite is also partially chlorinated during the milling process.36 Interestingly, the IR spectra of PMP7m and PMP4 exhibit notable similarities, while PMP7m and PMP7 show significant divergence. This suggests that when the PVDF membrane is used in the milling process, the resulting material PMP7m resembles PMP4 rather than PMP7. Additives of the real life PVDF sample may cause the difference of the degradation process. The formation of a graphitic structure by dehydrofluorination is indicated by a C
C band at 1590 cm−1.37 For comparison, after 7 h of milling neat PVDF, the IR data revealed a significant broadening of the C–F stretching bands between 1000 and 1400 cm−1, but no C
C bands that are characteristic for dehydrofluorination of PVDF were visible.34,35 Note that pyrolysis of PVDF at elevated temperatures has been reported to result in dehydrofluorination.38
 |
| Fig. 4 (a) ATR IR spectra (diamond) of PMP samples after 7 and 4 h of milling, PMP7 sample using PVDF membrane (PMP7m), 7 h milled PVDF and neat PVDF. (b) Raman spectra recorded at 532 nm for PMP7, fluorinated graphite (>61 wt% F) and pristine graphite. | |
Raman spectroscopy
To gain a deeper understanding of the graphitic material, Raman spectra were recorded and compared with spectra of pristine graphite and fluorinated graphite (Fig. 4b). Graphite exhibits distinctive Raman features: the G-band (1582 cm−1), associated with the in-plane vibrations of sp2 carbons and the D-band (1354 cm−1), indicating defects or disorder.39 These bands were observed in all samples. However, the position and shape of these bands varied significantly between the samples. The milled bulk material PMP7 also shows similar D- and G-bands, but they are much broader than those in graphite. A broadening of D- and G-bands is typically associated with a high and variable degree of functionalisation and disorder in the graphitic material. This is also apparent from the Raman spectrum of commercial fluorinated graphite. The blue shift in the G-band (to 1590 cm−1) and the concomitant red shift (to 1339 cm−1) in the D-band for PMP7 are consistent with literature reports for functionalised graphite and are associated often with graphite amorphisation.40 Such shifts have been observed previously in oxidised or fluorinated graphite/graphene as well.41,42 In the case of PMP7, we attribute the functionalisation with fluorine and chlorine atoms as the primary reason for an altered electronic structure and bonding environment, which directly affects the C
C bond vibrational frequencies and contributes to the G- and D-band broadening and frequency shifts.
X-ray photoelectron spectroscopy, XPS
Raman spectroscopy indicated a distortion of the graphitic material, prompting further investigation. To elucidate the chemical state of PMP7, X-ray photoelectron spectra were measured, which is a surface-sensitive technique for near-surface layers. F 1s and C 1s XPS spectra of PVDF and PMP7 are depicted in Fig. 5. The F 1s XPS spectrum of PVDF predominantly exhibits a peak at 687.6 eV, characteristic for CF2 groups,43 along with a less intense peak at 688.3 eV attributed to CF3 groups. The CF3 groups are likely originating from terminal CF3 moieties within PVDF.44 In contrast, the F 1s XPS spectrum of PMP7 reveals a broad peak centered at 687.1 eV, likely resulting from the overlap of signals arising from CFx and Al–F species.43,45 This interpretation is further supported by the significantly broadened full width at half maximum (FWHM) of 3.2 eV, compared to the typical FWHM of 1.8 eV observed for narrower peaks in PVDF. The C 1s XPS spectrum of PVDF exhibits peaks corresponding to CH2 in the neighbourhood of CF2 at binding energies of 286.2 and 290.7 eV, respectively.43 Additionally, small peaks attributed to CHF and CF3 moieties were observed at 287.9 and 293.0 eV, respectively.43 After a 7 hour milling process of PVDF in the presence of AlCl3, a decrease in the CF2 signal intensity was noted in the C 1s spectrum of PMP7. Concomitantly, peaks at 286.4, 289.4, and 291.1 eV associated with CCl, CCl2, and CCl3 moieties were observed,43 although these peaks could also be assigned to CH2, CF, and CF2 entities, respectively. The presence of chlorinated and fluorinated functionalities in the graphitic structure is expected to result in a higher binding energy shift of the sp2 C peak in XPS spectra. A clear identification of a distinct sp2 C peak is not possible, because the expected π to π* transition peaks46 at around 291–293 eV cannot be clearly identified due to the CF2 and CCl3 peaks in this region for PMP7. From the C 1s XPS analysis of PMP7, we can assume an approximate functionalisation of 14% by fluorine and chlorine atoms within the graphitic material (SI Table 2). Overall, the XPS data are in accordance with a defluorination of PVDF by the milling process.
 |
| Fig. 5 XPS high-resolution spectra of C 1s and F 1s of neat PVDF and PMP7. adv. C means adventitious carbon. | |
Proposed mechanism
The proposed PVDF degradation mechanism (Scheme 2a) involves a Lewis-acidic interaction of AlCl3 at a C–F bond of the –[CH2CF2]n– moiety producing a carbenium-like ion and an aluminate by fluorine–aluminium bond formation.17,25,26,47,48 In particular, the use of anhydrous AlCl3 as a Lewis-acidic co-milling agent is crucial for the reaction; hydrous AlCl3·6H2O or aluminium hydroxide/oxide proved to be ineffective. Subsequently, a chloride ion from the [AlCl3F]− attacks the C–H bond, leading to the release of HCl gas by dehydrochlorination and eventually the formation of AlF3. Alternatively, the carbenium intermediate could undergo chlorination by the [AlCl3F]− anion followed by Lewis-acid supported HF elimination (dehydrofluorination), leading to a partially chlorinated olefinic species. On the other hand, initial dehydrofluorination of fluorinated alkyl moieties in PVDF or derivatives can occur by release of HF molecules,19,49,50 which in turn can react further with AlClxFy (x = 3 − y, y ≠ 3) to generate AlF3 and HCl. Note that incomplete PVDF degradation was observed when less than two equivalents of AlCl3 were used. This is evidenced by the ATR-IR spectrum of the material obtained by a reaction of 1 eq. AlCl3 and 3 eq. PVDF (PVDF eq. are given with respect to the monomer block), where characteristic C–F vibrational modes remained detectable (SI Fig. 22). However, the resulting dehydrohalogenated product obtained with 2 eq. AlCl3 (a polyolefinic species) then undergoes C–C coupling and formation reactions to yield halogenated aromatic rings releasing again HF or HCl.51 To further elucidate the presence of intermediate radicals in the latter transformations, azobisisobutyronitrile (AIBN) was added to a mixture of AlCl3 and PVDF. Note that AIBN can be used for trapping alkenyl radicals.52 Milling of this mixture for 7 h yielded again an insoluble black powder. The gaseous products evolved were vacuum transferred into a C6D6 solution for subsequent gas and liquid phase analyses. NMR spectroscopic investigations and GC/MS data show the presence of 1,3,5-trifluorobenzene and chlorinated olefinic compounds (SI Fig. 2–8). In addition, MALDI-MS data revealed that the Mw of the material is 1000 m/z less when compared to Mw of PMP7 (2000 m/z; SI Fig. 20–21). This confirms that the addition of radicals hampers the formation of the graphitic network. High temperatures (up to 1000 K)53–55 during milling and AlCl3 as co-milling agent can facilitate the aromatic ring formation. Once aromatic rings are formed, a Lewis-acid-mediated Scholl reaction, involving H2 release, leads to the formation of fluorine- and chlorine-functionalised graphitic entities.56,57 The H2 evolution observed by the corresponding signals in the 1H NMR spectrum of the gaseous content from PMP7 depicted in Fig. 1a suggests that a reducing H2 atmosphere activates the graphitic material,58,59 promoting the formation of CH4, which adsorbs on the graphitic layers.
 |
| Scheme 2 (a) Proposed reaction mechanism for the degradation of PVDF with AlCl3 using ball milling. (b) Model reaction of Scholl reaction using anthracene and 0.1 eq. AlCl3 forming graphite and methane. (c) Raman spectra of anthracene and product after milling anthracene and 0.1 eq. AlCl3 recorded at 532 nm. (d) 1H NMR spectrum (300 MHz, C6D6) from the gas phase after milling 0.1 eq. AlCl3 and anthracene. | |
Importantly, the occurrence of the Scholl reaction and CH4 formation by H2 reduction steps are supported by the independent milling reaction of anthracene with a catalytic amount of 0.1 eq. AlCl3 (Scheme 2b), which produced graphitic entities (identified by Raman spectroscopy, Scheme 2c) and CH4 (detected by 1H NMR spectroscopy, Scheme 2d).
Conclusions
In conclusion, this study highlights the potential of mechanochemistry to effectively transform fluorinated polymer waste such as PVDF into valuable products. The mechanochemical approach paves the way for further advancements in polymer degradation and demonstrates the transformative potential of mechanochemistry in opening up chemical synthesis and waste management towards a circular economy.60
Through ball milling with anhydrous AlCl3, a quantitative mineralisation of PVDF was achieved, yielding high-value materials such as AlF3 and functionalised graphite alongside gaseous HCl and CH4. AlF3 is used in various industries, including the production of aluminium, glass and ceramics.61 Functionalised graphite exhibits promising properties for applications in energy storage, catalysis and advanced materials.62–64 Note that a study by Ding et al. demonstrated that a AlF3 coated graphite anode exhibits a higher initial discharge capacity and improved rate performance.65 Traditional synthesis of fluorinated graphite typically involves reacting graphite with F2 or HF gas streams under high temperature and pressure conditions.30,66 In addition, the solvent-free approach in this study aligns with green chemistry principles.12,13
Whereas in general reaction mechanisms during mechanochemical treatment of PFAS remain largely unclear,67 the proposed mechanism for the PVDF degradation in this study by a Lewis-acid may contribute to an understanding of fundamental processes involved, and guide the development of efficient and sustainable degradation strategies. Model reactions have shown that Scholl reactions are a key-step for the generation of graphitic materials in polymer degradation. Thus, dehydrofluorination of PVDF is followed by the Scholl reaction that generates intermediate H2, which in turn implies that a partial oxidation of the carbon network occurs. H2 can react further to produce CH4. The presented strategy is, however, limited to fluoropolymers which can undergo dehydrofluorination.
Mechanochemical fluorinated polymer degradation is different from reported reductive methods. Thus, Crimmin and co-workers used a strong Mg(I) reducing agent to defluorinate polytetrafluoroethylene.68 The group of Qu and Kang developed a defluorination strategy of PFAS using a carbazole based photo reductant KQGZ (10,13-diphenyl-9H-dibenzo[a,c]carbazole) at low temperature.69
Methods
Experimental procedures, characterisation methods and data for the control reactions can be found in the SI.
General information and materials
All samples were prepared in an MBraun glovebox filled with argon or in JYoung NMR tubes using conventional Schlenk techniques. Anhydrous aluminium trichloride (purity 99.99%) was obtained from ABCR and used as received. Polyvinylidene fluoride powder was purchased from Apollo Scientific. The PVDF membrane (ROTI®Fluoro pore size 0.2 μm) was purchased from Carl Roth. Benzene-d6 was bought from Eurisotop and distilled over Solvona®. The reference materials graphite, fluorinated graphite (>61 wt% fluorine) and anthracene were obtained from Sigma Aldrich and used as received. Azobisisobutyronitrile (purity 98%) was obtained from Sigma Aldrich.
Author contributions
M. Bui conceived the project, designed and carried out the experiments and wrote the manuscript. C. Heinekamp performed the S/TEM imaging and EDX analysis. E. Fuhry recorded the Raman spectra. S. Weidner performed the MALDI-TOF MS experiments. J. Radnik carried out the XPS measurements. MAS NMR experiments were measured by K. Scheurell. M. Ahrens, K. Balasubramanian, F. Emmerling and T. Braun supervised the project.
Conflicts of interest
There are no conflicts to declare.
Data availability
Details of the experimental procedures and characterization of the complexes can be found in the SI.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5sc05783c.
Acknowledgements
The authors acknowledge Mehmet Bayat for the OFCEAS measurements (BAM Division 1.4) and Jakob Bölle for experimental support. This work was financially supported by the German Research Foundation (DFG) through the CRC 1349 “Fluorine Specific Interactions” (project number 387284271).
Notes and references
- A. Ghobadi Moghadam and A. Hemmati, Sci. Rep., 2023, 13, 8076 CrossRef CAS.
- W. Xia and Z. Zhang, IET Nanodielectr., 2018, 1, 17–31 CrossRef.
- X. Zhong, J. Han, L. Chen, W. Liu, F. Jiao, H. Zhu and W. Qin, Appl. Surf. Sci., 2021, 553, 149564 CrossRef CAS.
- T. Ahrens, J. Kohlmann, M. Ahrens and T. Braun, Chem. Rev., 2015, 115, 931–972 CrossRef CAS.
- H. Amii and K. Uneyama, Chem. Rev., 2009, 109, 2119–2183 CrossRef CAS PubMed.
- D. O'Hagan, Chem. Soc. Rev., 2008, 37, 308–319 RSC.
- H. Hori, T. Sakamoto, K. Ohmura, H. Yoshikawa, T. Seita, T. Fujita and Y. Morizawa, Ind. Eng. Chem. Res., 2014, 53, 6934–6940 CrossRef CAS.
- H. Hori, H. Tanaka, K. Watanabe, T. Tsuge, T. Sakamoto, A. Manseri and B. Ameduri, Ind. Eng. Chem. Res., 2015, 54, 8650–8658 CrossRef CAS.
- N. Yanagihara and T. Katoh, Green Chem., 2022, 24, 6255–6263 RSC.
- Q. Zhang, J. Lu, F. Saito and M. Baron, J. Appl. Polym. Sci., 2001, 81, 2249–2252 CrossRef CAS.
- A. A. L. Michalchuk, E. V. Boldyreva, A. M. Belenguer, F. Emmerling and V. V. Boldyrev, Front. Chem., 2021, 9, 685789 CrossRef CAS PubMed.
- K. J. Ardila-Fierro and J. G. Hernández, ChemSusChem, 2021, 14, 2145–2162 CrossRef CAS PubMed.
- J. Alić, M.-C. Schlegel, F. Emmerling and T. Stolar, Angew. Chem., Int. Ed., 2024, e202414745 Search PubMed.
- C. Patel, E. André-Joyaux, J. A. Leitch, X. M. de Irujo-Labalde, F. Ibba, J. Struijs, M. A. Ellwanger, R. Paton, D. L. Browne, G. Pupo, S. Aldridge, M. A. Hayward and V. Gouverneur, Science, 2023, 381, 302–306 CAS.
- S. Aydonat, A. H. Hergesell, C. L. Seitzinger, R. Lennarz, G. Chang, C. Sievers, J. Meisner, I. Vollmer and R. Göstl, Polym. J., 2024, 56, 249–268 CrossRef CAS.
- C. P. Marshall, G. Scholz, T. Braun and E. Kemnitz, Cat. Sci. Technol., 2020, 10, 391–402 CAS.
- C. G. Krespan and V. A. Petrov, Chem. Rev., 1996, 96, 3269–3302 CAS.
- G. Meißner, K. Kretschmar, T. Braun and E. Kemnitz, Angew. Chem., Int. Ed., 2017, 56, 16338–16341 CrossRef.
- G. Meißner, D. Dirican, C. Jäger, T. Braun and E. Kemnitz, Cat. Sci. Technol., 2017, 7, 3348–3354 RSC.
- B. Calvo, C. P. Marshall, T. Krahl, J. Kröhnert, A. Trunschke, G. Scholz, T. Braun and E. Kemnitz, Dalton Trans., 2018, 47, 16461–16473 RSC.
- C. L. Bailey, S. Mukhopadhyay, A. Wander, B. G. Searle and N. M. Harrison, J. Phys. Chem. C, 2009, 113, 4976–4983 CrossRef CAS.
- T. Einfalt, O. Planinšek and K. Hrovat, Acta Pharm., 2013, 63, 305–334 CAS.
- M. H. Enayati and F. A. Mohamed, Int. Mater. Rev., 2014, 59, 394–416 CAS.
- G. Scholz, S. Brehme, M. Balski, R. König and E. Kemnitz, Solid State Sci., 2010, 12, 1500–1506 CAS.
- T. Krahl and E. Kemnitz, Cat. Sci. Technol., 2017, 7, 773–796 CAS.
- M. Bui, K. F. Hoffmann, T. Braun, S. Riedel, C. Heinekamp, K. Scheurell, G. Scholz, T. M. Stawski and F. Emmerling, ChemCatChem, 2023, 15, e202300350 CrossRef CAS.
- T. Krahl and E. Kemnitz, J. Fluorine Chem., 2006, 127, 663–678 CrossRef CAS.
- E. W. Hagaman, D. K. Murray and G. D. Del Cul, Energy Fuels, 1998, 12, 399–408 CrossRef CAS.
- T. Koskela, M. Ylihautala, J. Vaara and J. Jokisaari, Chem. Phys. Lett., 1996, 261, 425–430 CrossRef CAS.
- K. Guérin, J. P. Pinheiro, M. Dubois, Z. Fawal, F. Masin, R. Yazami and A. Hamwi, Chem. Mater., 2004, 16, 1786–1792 CrossRef.
- R. S. Matthews, Org. Magn. Reson., 1982, 18, 226–230 CrossRef CAS.
- J. Giraudet, M. Dubois, K. Guérin, J. P. Pinheiro, A. Hamwi, W. E. E. Stone, P. Pirotte and F. Masin, J. Solid State Chem., 2005, 178, 1262–1268 CrossRef CAS.
- U. Gross, S. Rüdiger, E. Kemnitz, K.-W. Brzezinka, S. Mukhopadhyay, C. Bailey, A. Wander and N. Harrison, J. Phys. Chem. A, 2007, 111, 5813–5819 CAS.
- W. Zhang, M. Dubois, K. Guérin, P. Bonnet, H. Kharbache, F. Masin, A. P. Kharitonov and A. Hamwi, Phys. Chem. Chem. Phys., 2010, 12, 1388–1398 RSC.
- Y. Bormashenko, R. Pogreb, O. Stanevsky and E. Bormashenko, Polym. Test., 2004, 23, 791–796 CrossRef CAS.
- J. C. L. Chow, P. K. Ummat and W. R. Datars, Phys. Rev. B Condens. Matter, 1999, 271, 165–172 CrossRef CAS.
- P. E. Fanning and M. A. Vannice, Carbon, 1993, 31, 721–730 CrossRef CAS.
- M. L. O'Shea, C. Morterra and M. J. D. Low, Mater. Chem. Phys., 1990, 26, 193–209 CrossRef.
- A. C. Ferrari, Solid State Commun., 2007, 143, 47–57 CrossRef CAS.
- A. C. Ferrari and J. Robertson, Phys. Rev. B:Condens. Matter Mater. Phys., 2000, 61, 14095–14107 CrossRef CAS.
- K. N. Kudin, B. Ozbas, H. C. Schniepp, R. K. Prud'homme, I. A. Aksay and R. Car, Nano Lett., 2008, 8, 36–41 CrossRef.
- V. Gupta, T. Nakajima and B. Žemva, J. Fluorine Chem., 2001, 110, 145–151 CrossRef CAS.
- J. F. Watts, Surf. Interface Anal., 1993, 20, 267 CrossRef.
- F. Boschet, T. Ono and B. Ameduri, Macromol. Rapid Commun., 2012, 33, 302–308 CrossRef CAS PubMed.
- A. Hess, E. Kemnitz, A. Lippitz, W. E. S. Unger and D. H. Menz, J. Catal., 1994, 148, 270–280 CrossRef.
- T. R. Gengenbach, G. H. Major, M. R. Linford and C. D. Easton, J. Vac. Sci. Technol. A, 2021, 39, 013204 CrossRef CAS.
- W. Gu, M. R. Haneline, C. Douvris and O. V. Ozerov, J. Am. Chem. Soc., 2009, 131, 11203–11212 CrossRef CAS PubMed.
- J. Wang, Y. Ogawa and N. Shibata, Sci. Rep., 2019, 9, 19113 CrossRef.
- C. Heinekamp, A. G. Buzanich, M. Ahrens, T. Braun and F. Emmerling, Chem. Commun., 2023, 59, 11224–11227 RSC.
- S. E. S. Farley, D. Mulryan, F. Rekhroukh, A. Phanopoulos and M. R. Crimmin, Angew. Chem., Int. Ed., 2024, 63, e202317550 CrossRef CAS.
- G. Montaudo, C. Puglisi, E. Scamporrino and D. Vitalini, J. Polym. Sci., Part A: Polym. Chem., 1986, 24, 301–316 CrossRef CAS.
- A. K. Morri, Y. Thummala, S. Ghosh and V. R. Doddi, ChemistrySelect, 2021, 6, 2387–2393 CrossRef CAS.
- H. Kulla, S. Haferkamp, I. Akhmetova, M. Röllig, C. Maierhofer, K. Rademann and F. Emmerling, Angew. Chem., Int. Ed., 2018, 57, 5930–5933 CrossRef CAS.
- K. S. McKissic, J. T. Caruso, R. G. Blair and J. Mack, Green Chem., 2014, 16, 1628–1632 RSC.
- S. You, M. W. Chen, D. D. Dlott and K. S. Suslick, Nat. Commun., 2015, 6, 6581 CrossRef CAS.
- S. Grätz, D. Beyer, V. Tkachova, S. Hellmann, R. Berger, X. Feng and L. Borchardt, Chem. Commun., 2018, 54, 5307–5310 RSC.
- R. Scholl and C. Seer, Liebigs Ann. Chem., 1912, 394, 111–177 CrossRef.
- Z. J. Pan and R. T. Yang, J. Catal., 1990, 123, 206–214 CrossRef CAS.
- G. Bomchil, A. Hüller, T. Rayment, S. J. Roser, M. V. Smalley, R. K. Thomas, J. W. White, A. D. Buckingham, R. Mason, E. W. J. Mitchell and J. W. White, Philos. Trans. R. Soc. B, 1980, 290, 537–552 CAS.
- Note that during the process the following publications were published on a mechnaochemical degradation of PVDF. In these cases, however, the fate of the carbon is unclear:
(a) L. Yang, Z. Chen, C. A. Goult, T. Schlatzer, R. S. Paton and V. Gouverneur, Nature, 2025, 640, 100–106 CrossRef CAS PubMed;
(b) M. Hattori, D. Saha, M. Z. Bacho and N. Shibata, Nat. Chem., 2025 DOI:10.1038/s41557-025-01855-3.
- T. Nakajima, B. Žemva and A. Tressaud, Advanced inorganic fluorides: synthesis, characterization and applications, Elsevier, 2000 Search PubMed.
- H. Groult, T. Nakajima, L. Perrigaud, Y. Ohzawa, H. Yashiro, S. Komaba and N. Kumagai, J. Fluorine Chem., 2005, 126, 1111–1116 CrossRef CAS.
- F. Yin, J.-H. Liu, Y. Zhang, M.-N. Liu, L.-Y. Wang, Z.-C. Yu, W.-H. Yang, J. Zhang and Y.-Z. Long, Adv. Funct. Mater., 2024, 34, 2406417 CrossRef CAS.
- K. Henry, M. Colin, G. Chambery, B. Vigolo, S. Cahen, C. Hérold, V. Nesvizhevsky, S. Le Floch, V. Pischedda, S. Chen and M. Dubois, Dalton Trans., 2024, 53, 9473–9481 RSC.
- F. Ding, W. Xu, D. Choi, W. Wang, X. Li, M. H. Engelhard, X. Chen, Z. Yang and J.-G. Zhang, J. Mater. Chem., 2012, 22, 12745–12751 RSC.
- W. Feng, P. Long, Y. Feng and Y. Li, Adv. Sci., 2016, 3, 1500413 CrossRef.
- S. Verma, T. Lee, E. Sahle-Demessie, M. Ateia and M. N. Nadagouda, Chem. Eng. J. Adv., 2022, 13, 1–11 Search PubMed.
- D. J. Sheldon, J. M. Parr and M. R. Crimmin, J. Am. Chem. Soc., 2023, 145, 10486–10490 CrossRef CAS.
- H. Zhang, J.-X. Chen, J.-P. Qu and Y.-B. Kang, Nature, 2024, 635, 610–617 CrossRef.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.