Next Article in Journal
Carotenoid-Enriched Nanoemulsions and γ-Rays Synergistically Induce Cell Death in a Novel Radioresistant Osteosarcoma Cell Line
Next Article in Special Issue
Evaluation of Chokeberry/Carboxymethylcellulose Hydrogels with the Addition of Disaccharides: DART-TOF/MS and HPLC-DAD Analysis
Previous Article in Journal
BAC-FISH Based Physical Map of Endangered Catfish Clarias magur for Chromosome Cataloguing and Gene Isolation through Positional Cloning
Previous Article in Special Issue
Calix[4]arene Polyamine Triazoles: Synthesis, Aggregation and DNA Binding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Poly(dithiophosphate)s, a New Class of Phosphorus- and Sulfur-Containing Functional Polymers by a Catalyst-Free Facile Reaction between Diols and Phosphorus Pentasulfide

1
Polymer Chemistry and Physics Research Group, Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences, Magyar tudósok krt. 2, H-1117 Budapest, Hungary
2
Functional Nanoparticles Research Group, Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences, Magyar tudósok krt. 2, H-1117 Budapest, Hungary
3
Soft Matters Group, Department of Physical Chemistry and Materials Science, Faculty of Chemical Technology and Biotechnology, Budapest University of Technology and Economics, Műegyetem rkp. 3, H-1111 Budapest, Hungary
4
Plasma Chemistry Research Group, Institute of Materials and Environmental Chemistry, Research Centre for Natural Sciences, Magyar tudósok krt. 2, H-1117 Budapest, Hungary
5
Chemical Crystallography Research Laboratory, Research Centre for Natural Sciences, Magyar tudósok krt. 2, H-1117 Budapest, Hungary
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15963; https://doi.org/10.3390/ijms232415963
Submission received: 18 November 2022 / Revised: 6 December 2022 / Accepted: 13 December 2022 / Published: 15 December 2022
(This article belongs to the Special Issue Advances in Multi-Tasked Macromolecules)

Abstract

:
Novel poly(dithiophosphate)s (PDTPs) were successfully synthesized under mild conditions without any additive in the presence of THF or toluene diluents at 60 °C by a direct, catalyst-free reaction between the abundant phosphorus pentasulfide (P4S10) and glycols such as ethylene glycol (EG), 1,6-hexanediol (HD) and poly(ethylene glycol) (PEG). GPC, FTIR, 1H and 31P NMR analyses proved the formation of macromolecules with dithiophosphate coupling groups having P=S and P-SH pendant functionalities. Surprisingly, the ring-opening of THF by the P-SH group and its pendant incorporation as a branching point occur during polymerization. This process is absent with toluene, providing conditions to obtain linear chains. 31P NMR measurements indicate long-time partial hydrolysis and esterification, resulting in the formation of a thiophosphoric acid moiety and branching points. Copolymerization, i.e., using mixtures of EG or HD with PEG, results in polymers with broadly varying viscoelastic properties. TGA shows the lower thermal stability of PDTPs than that of PEG due to the relatively low thermal stability of the P-O-C moieties. The low Tgs of these polymers, from −4 to −50 °C, and a lack of PEG crystallites were found by DSC. This polymerization process and the resulting novel PDTPs enable various new routes for polymer synthesis and application possibilities.

1. Introduction

Macromolecules containing phosphorus and/or sulfur, either in the main chain or pendant, have been intensively investigated worldwide in recent years. This is mainly due to their broad application possibilities, ranging from flame retardancy and environment protection to various high-value-added specialty fields, such as biomaterials, drug delivery, sensors, optical, battery and other energy-related areas, etc. (see, e.g., [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19] and the references therein). Although the reaction of phosphorus pentasulfide (P4S10), a widely available commercial material, with alcohols has been known for a long time [20] to result in the formation of dialkyl dithiophosphates (DTPs), i.e., O,O-diesters of dithiophosphoric acid [21] (Scheme 1), surprisingly, systematic investigations on the utilization of this process for polymer synthesis have not been reported yet, according to the best of our knowledge. Low molecular weight compounds of this kind, especially zinc dialkyl dithiophosphate salts [22,23,24], were recently applied to the preparation of tribological additives as lubricants. Furthermore, the resulting dialkyl dithiophosphates can also be used for several other purposes, e.g., flotation additive [25], wastewater treatment [26] or reagent in organic reactions [27]. The DTPs possess a strong acidic group, and moreover, their anions are highly efficient metal ion complexing agents due to the two sulfur atoms bound to the phosphorus atom [28]. Thus, it can be presumed that the versatility of these functional groups may enable broad application possibilities, especially when they are involved in macromolecular structures.
The synthesis of polymers with dialkyl dithiophosphate moieties can be achieved by different strategies. One of the possibilities is the coupling of dialkyl dithiophosphate complexes of metal ions by using bridging ligands, e.g., 4,4′-bipyridine [29]. The other strategy is the reaction of P4S10 with diols. This is analogous to the coupling of poly(ethyleneimine) chains by using P4S10 [9]. However, a multibranched non-acidic polymer was formed with thiophosphoric acid triamide groups in this case. Surprisingly, only a single publication can be found in the literature on poly(dithiophosphate)s. The reactions of diethyldithiophosphate and P4S10 with diols in bulk at high temperatures (110–140 and 120–180 °C, respectively) were studied by Pudovik et al. [30], who reported on the formation of cyclic dithiophosphate diester products. These authors also claimed the formation of macromolecules, i.e., poly(dithiophosphate)s (PDTPs), but without detailed structural investigations. However, most likely because of the applied high temperatures, this simple method, that is, using the abundant reagents, i.e., diols and P4S10, has not become a known polymerization technique, in contrast to the reaction of phosphoric acid with diols, which were thoroughly investigated by Penczek and his coworkers [31,32].
Herein, we present the synthesis and characterization of polymers obtained by the reaction of P4S10 and various α,ω-diols, such as poly(ethylene glycol) with Mn = 400 g/mol (PEG400), ethylene glycol (EG) and 1,6-hexanediol (HDO) under mild conditions. The property tunability by using different diols simultaneously in the polymerization process, i.e., by copolymerization, is also explored. These systematic investigations demonstrate the applicability of this polymerization technique to obtain a unique class of novel macromolecular structures with tailored properties.

2. Results and Discussion

Preliminary experiments indicated that phosphorus pentasulfide (P4S10) successfully reacts with stoichiometric amounts of diols without the addition of any kind of catalyst, i.e., metals or metal compounds, in tetrahydrofuran (THF) at 60 °C, and this process results in poly(dithiophosphate)s (PDTPs), as shown in Scheme 2. Subsequently, a systematic series of experiments were designed to explore the reaction of P4S10 with various diols, such as poly(ethylene glycol) (PEG400), ethylene glycol (EG), 1,6-hexanediol (HD) and the mixtures of pairs of these diols by using stoichiometric amounts of the reactants, as presented in Table 1. In order to decrease the viscosity of the reaction medium and to provide a solvent for the reactants and product, THF was added to the reaction mixture. Visual observation indicated that P4S10 is insoluble in the mixture of the diols and THF, and it appears as a solid phase at the beginning of the reaction. However, this solid material disappears in the course of the reaction, indicating the formation of a compound soluble in the liquid phase. After the workup process (evaporation of the solvent and drying in vacuo, see Experimental), the dried products were obtained as colorless viscous liquids, with yields higher than 95% in all cases.
First, the reaction of P4S10 was investigated with PEG400 as the diol (P1 sample in Table 1). Gel permeation chromatography (GPC) measurements showed that the main fraction of the product consisted of macromolecules with a higher hydrodynamic radius, i.e., higher molecular weight, than those of the PEG400 reagent (Figure 1). This indicates the coupling of PEG400 molecules. FTIR measurements were also performed on the obtained polymer P1 (Figure 2). A comparison of the IR absorption spectra of PEG400 and P1 implies that the intensity of the broad peak at 3450 cm−1, belonging to the O-H stretching in PEG400, decreased considerably in P1, indicating the reaction of the chain end –OH groups of PEG400 during the polymerization. The absorption peaks in the spectrum of P1 at 797 and 667 cm−1 are also in accordance with the structure forming in the reaction displayed in Scheme 2 because the latter peak belongs to the P=S stretching while the former can be assumed to be the band of the S-H in-plane scissoring [33].
The chemical structure of the product was also investigated by 1H NMR spectroscopy (Figure 3A). The appearance of a broad signal at 5.78 ppm, belonging to the exchangeable acidic protons (and water impurity), and that at 4.3 ppm, assigned to the -CH2-O-P sequences, proves the occurrence of the reaction between PEG400 and P4S10.
Interestingly, an unexpected side reaction occurs when using THF as a diluent during the polymerization process between P4S10 and diols. The ring-opening of THF molecules due to the nucleophilic attack of the P-SH group in the resulting polymers (Scheme 3, Figure S1) leads to the incorporation of such pendant molecular units in the polymer. The reaction between the poly(dithiophosphate) and THF is also proved by carrying out the polymerization process between P4S10 and PEG400 in toluene. In this case, the signals between 1.5 and 2.0 ppm in the 1H NMR spectrum of the sample are absent (Figure S2). Because new -CH2-OH groups appear in the reaction between THF and PDTP, this can lead to the formation of branching points, as depicted in Scheme 3. This is also indicated by the appearance of signals at 4.09 and 4.19 ppm in the 1H NMR spectrum (Figure 3A). From the integration of the peaks in the 1H NMR spectrum, it can be obtained that the ratio of the incorporated THF:PEG400 units is roughly 1:6. This unexpected finding indicates the unique opportunity of a broad range of post-polymerization modifications of poly(dithiophosphate)s by ring-opening reactions similar to the one observed with THF in this case.
The presence of phosphorus in the polymer product was proved by 31P NMR spectroscopy. The desired product (Scheme 2) is formed in high conversion, as proved by the 31P NMR signal at 88 ppm [34]. However, additional coproduct formation is also observed. By comparing the 31P spectra recorded after 2 days (Figure 3B) and 2.5 months (Figure 3C) of polymerization, considerable differences can be observed. The signals at 88 and 81 ppm, assigned to the (RO)2P(S)SH units, disappeared during the 2.5 months. The signal at 96 ppm, which can be assigned to the (RO)2P(S)(SCH2-) structure, does not change during the storage, and the signals at 67 and 64 ppm are also present in both spectra. These two signals can be assigned to the phosphorus atom in (RO)2P(S)(OR) at 67 ppm and in (RO)2P(S)OH at 64 ppm. The absence of any signals around 0 ppm indicates that the product does not contain totally hydrolyzed phosphoric acid units; only thiolated phosphorus centers exist in these polymers. These observations can be explained by assuming two reactions, the hydrolysis reaction of the phosphorus-SH moiety with traces of water and the esterification with the unreacted terminal hydroxyl groups, resulting in the formation of additional branching points (Scheme 4). These assumptions are in accordance with previous investigations on the hydrolysis of zinc dialkyl dithiophosphates [35]. The occurrence of esterification is also confirmed by the determination of the 1H NMR signal areas of CH2OP and other CH2O protons, which increases with time, considering the spectra recorded after two days and two and a half months after the polymerization.
Differential scanning calorimetric measurements show that while a melting peak appears at 2 °C on the curve of PEG400, indicating the crystallization of the PEG oligomers at low temperature, melting is absent on the curve of the P1 polymeric product and only a glass transition can be observed at −47 °C (Figure 4). This considerable difference in the thermal behavior indicates the hindrance of the appropriate orientation of the poly(ethylene glycol) segments required for crystallization due to the presence of the coupling (di)thiophosphate units. Thermogravimetric analysis (TGA) curves (Figure 5) also show the difference between P1 and PEG400 since P1 starts to degrade under an inert atmosphere at ca. 100 °C lower temperature than PEG400, indicating the thermal lability of the P-O-C bond. While the solid residue in the case of PEG400 is negligible at high temperatures, it is 9.5% in the case of the P1 poly(dithiophosphate) sample due to the formation of the resulting nonvolatile inorganic material.
The acidic character of P1 in aqueous conditions was also studied. It was titrated with NaOH solutions. The first step of the two-step titration curve indicates the presence of strong acidic groups, with pKa = 3.5. This means that the P1 polymer behaves as a strong acidic polyelectrolyte in water (Figure S3).
Copolymerization experiments were performed with the simultaneous addition of PEG400 and ethylene glycol (Table 1). GPC measurements showed that there was no considerable difference between the hydrodynamic volume (and the molecular weight) of the formed copolymers P2–P4 (Figure 1). However, when only ethylene glycol was used as a diol (P5), a significantly lower hydrodynamic volume was observed for this product, indicating its lower molecular weight. From the comparison of the 1H NMR spectra of the P1–P4 samples, it can be concluded that the ratio of the integrals of the CH2-O-P signals (at 4.3 ppm) related to the CH2-O-C signals (at 3.5–3.9 ppm) increases with increasing EG content in the feed (Figure S4). This indicates that higher EG content in the feed results in lower P-P average distances in the polymer chains. This is indirect evidence of the ethylene glycol incorporation in the polymer chain. In the case of the copolymers P2–P4, the 31P NMR spectra also prove the partial hydrolysis and esterification side reactions (Figure S5). DSC measurements show slightly increasing Tg with increasing EG unit content in these copolymers. However, a considerable difference was only observed in the case of the high EG-containing copolymers, P4 and P5 (Table 1). Similarly, a slight decrease in the temperature of the first degradation step and an increase in the remaining solid residue ratio was observed in the TGA curves with increasing EG content (Figure 5). The latter observation is in accordance with the shortening of the average P-P distance, i.e., with the increment of the P-content in a given mass of polymers, with increasing EG content in the feed.
The mechanical properties of the copolymers were determined by oscillatory rheological investigations. Based on the observed results shown in Figure 6, it can be concluded that the P1 sample has higher loss moduli than PEG400 due to its poly((di)thiophosphate) structure. The moduli data indicate that the mechanical properties of the obtained polymers can be varied within wide limits by changing the PEG400:EG ratio in the feed. The P1, P2 and P3 samples have nearly the same moduli values, and in accordance with their viscous liquid form, these materials possess considerably lower storage than loss moduli, while the storage modulus for the P2 sample cannot be detected over a wide frequency range, suggesting a completely liquid behavior (Figure 6). While the storage modulus shows frequency independence at low frequencies and nearly linear frequency dependence at high frequencies on a log–log scale, the loss modulus depends linearly on the frequency in the whole applied frequency range, with a slope of 1, typical for viscoelastic liquids. The fact that the storage modulus point series do not intersect with the loss moduli indicates that the relaxation times are shorter than the reciprocally applied high-frequency limit (2 ms), in accordance with the moderate molecular weights and branching structures of the obtained polymer chains. In the case of the physical network type copolymers P4 and P5, only the former could be measured by the applied oscillation rheometry setup. The P4 sample, which has high EG content, i.e., low average P-P distance, has much higher moduli values than the copolymers with higher PEG400 content. Moreover, in the case of the P4 sample, the storage and loss moduli are in the same order of magnitude and both are nearly frequency-independent, indicating the appearance of much stronger secondary intermacromolecular forces in this copolymer composition. The point series are similar to the “imperfect network” case with crosslinks (in this case, the H-bonds) and dangling chains [36].
Copolymers were also synthesized with PEG400 and 1,6-hexanediol comonomers. As displayed in Figure 1, the peaks of the GPC curves of the P6, P7 and P8 copolymers have nearly the same elution volume as P1. Considerable molecular weight decrement can be observed in the case of P9 (Figure 1), similar to P5. Based on the 31P NMR spectra (Figure S6), the hydrolysis and esterification side reactions were also detected in the case of the polymers with 1,6-hexanediol monomeric units. The 1H NMR spectra show the signals of the -CH2- protons adjacent to the unreacted -OH groups (-CH2-OH) in the 3.4–3.85 ppm region and to the reacted ones (-CH2-O-P) between 3.9 and 4.5 ppm (Figure 7). In the case of P9, the conversion of the –OH groups (roughly 69%) can be determined by the comparison of the integrated area of these two signal groups. This relatively low conversion value means that there is a rather high number of unreacted –OH groups, which results in a reasonably high concentration of intermolecular H-bonds, explaining the observed physical network formation at room temperature. Similar physical network consistency can also be found in the case of the P8 copolymer. However, a higher increment in the PEG400 addition in the feed, similar to the EG-containing copolymers P2–P5, results in polymers as viscous liquids at room temperature. The oscillatory rheology response of these copolymers (P6 and P7) shows that the moduli are in the same region and the frequency response is similar to that of the P1–P3 samples. On the DSC curves, a glass transition can be observed at around −45 °C, independent of the composition (Table 1).

3. Materials and Methods

3.1. Materials

P4S10, poly(ethylene glycol) (Mn = 400 g/mol) (PEG400) and 1,6-hexanediol (HDO) were purchased from Merck (Sigma Aldrich, Darmstadt, Germany), while ethylene glycol (EG) was purchased from Molar Chemicals Ltd. (Halásztelek, Hungary) and tetrahydrofuran from VWR (Debrecen, Hungary). All the chemicals were used as received.

3.2. Polymer Syntheses

P4S10 (200–400 mg), a selected diol or mixture of diols and tetrahydrofuran were measured into a vial in stoichiometric amounts (i.e., 1 mole P4S10 to 4 mole diol). The molar ratio of the different diols was systematically varied (Table 1). The reaction mixtures were diluted with tetrahydrofuran to obtain 1 g/mL reactant/THF ratios and stirred at 60 °C for 6 h. Afterwards, the mixture was diluted with ca. 3 mL tetrahydrofuran and filtered using a 0.45 µm syringe filter, followed by the removal of the solvent under reduced pressure. The dried products were obtained after vacuum drying at 60 °C until constant weight. The reaction in toluene was carried out the same way.

3.3. Characterization Methods

1H and 31P NMR spectra were recorded on a Varian 300 MHz spectrometer at 30 °C in CDCl3.
The gel permeation chromatography (GPC) system was equipped with Waters Styragel HR1 and HR4 columns, a Waters 515 HPLC pump, a Waters 717 autosampler, a Jetstream column thermostat and an Agilent 1260 Infinity refractive index detector. Tetrahydrofuran was used as an eluent, with a 0.3 mL/min flow rate at 35 °C. Calibration was made with the polystyrene standards of narrow molecular weight distribution (from PSS Polymer Standards Services GmbH, Mainz, Germany).
Alkalimetric titration curves were recorded using a VWR MD8000L pH meter. For the titrations, 20–40 mg of the polymer was dissolved in 50 mL of distilled water. The titrant, a 0.013 M NaOH solution, was added by a syringe pump at a 10 mL/h rate.
Differential scanning calorimetry (DSC) measurements were performed on Mettler Toledo DSC821e equipment in the −80 to +100 °C temperature range, with a 10 °C/min heating rate in pierced 40 μL alumina crucibles under a nitrogen atmosphere with an 80 mL/min flow rate. The second heating curves were evaluated, and the inflection points of the glass transitions were regarded as glass transition temperatures.
TGA curves were recorded on a Setaram LabsysEvo (Lyon, France) TG-DSC system in a flowing (90 mL/min) nitrogen atmosphere in the 25 to 700 °C temperature range with a heating rate of 10 °C/min.
Oscillatory rheology measurements, applicable for determining the mechanical behavior of samples with a liquid consistency, were performed on an Anton Paar Physica MCR 301 rheometer at 25 °C with a cone-plate geometry probe (diameter: 25 mm; cone angle: 1°; sample gap: 0.054 mm) applying 1% strain in the 0.5–500 1/s angular frequency range.

4. Conclusions

An unexplored process, namely, the reaction between phosphorus pentasulfide (P4S10) and diols such as ethylene glycol (EG), 1,6-hexanediol (HD) and poly(ethylene glycol) (PEG), was investigated for the synthesis of a novel class of functional phosphorus- and sulfur-containing polymers, poly(dithiophosphate)s (PDTPs) (see Scheme 2), under mild conditions, that is, without any catalyst in the presence of THF or toluene diluents at a low temperature, i.e., at 60 °C, in contrast to the only reported case at 120–180 °C [30]. It was found that this simple process successfully results in PDTPs with higher than 95% yields. The structure of the resulting polymers was revealed by FTIR, 1H and 31P NMR spectroscopies and GPC measurements. Interestingly, the formation of branching points was detected in the case using THF as a diluent, which is attributed to the ring-opening of this molecule by the pendant sulfide group in the polymer chain. In contrast, this process is absent when toluene is used. By reacting P4S10 with mixtures of PEG and EG or HD, PDTP copolymers were obtained. The prepared PDTPs possess low glass transition temperatures (Tgs) in the range of −4 to −50 °C, indicating the formation of elastic chains. It is important to note that there is no crystalline fraction of PEG in the PDTPs. This can be attributed to the prevention of the necessary orientation and packing of the PEG chains in PDTPs. It has to be noted that this behavior might be useful in such areas as ion-conducting membranes for batteries. The thermogravimetric analyses show that the investigated PDTPs have lower thermal stability than PEG, which is attributed to the lower stability of the P-O-C bonds than the C-O-C ether bonds in PEG. Testing the mechanical properties of the investigated PDTPs by oscillatory rheology measurements has indicated that the moduli of these polymers can be varied in a broad range, depending on the composition, i.e., on the ratio of the applied diol comonomers. In sum, it can be concluded that the reaction of P4S10 with diols under mild conditions enables the synthesis of a variety of PDTPs with interesting properties. These findings are expected to open new routes for polymer synthesis and various application possibilities, such as lithium or hydrogen ion-conducting polymer electrolytes for batteries and fuel cells, flame retardants, pH-responsive polymers, etc., with the involvement of the novel poly(dithiophosphate)s.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232415963/s1.

Author Contributions

Conceptualization, Á.S., G.S., L.T., B.G., L.B., B.I. and E.K.; Methodology, Á.S., G.S., L.T., B.G., L.B., B.I. and E.K.; Validation, Á.S., G.S., L.T., B.G., L.B., B.I. and E.K.; Investigation, Á.S., G.S., L.T., B.G., L.B., B.I. and E.K.; Data Curation, Á.S., G.S., L.T., B.G., L.B., B.I. and E.K.; Writing—Original Draft Preparation, Á.S., G.S., B.I. and E.K.; Writing—Review and Editing, Á.Sz., G.S., L.T., B.G., L.B., B.I. and E.K; Supervision, Á.S., B.I. and E.K.; Project Administration, Á.S. and B.I.; Funding Acquisition, B.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by project no. VEKOP-2.3.2-16-2017-00013, supported by the European Union and the State of Hungary, co-financed by the European Regional Development Fund, and the National Research, Development and Innovation Office, Hungary (K135946, PD128612 and K124544).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Additional data on polymer analyses can be obtained by request from the corresponding authors.

Acknowledgments

The authors thank Beatrix Jázmin Sóvári and Bálint Imre for their technical assistance in the GPC and FTIR measurements, respectively. The research was performed within project no. VEKOP-2.3.2-16-2017-00013, supported by the European Union and the State of Hungary, co-financed by the European Regional Development Fund. The authors also acknowledge the support of the National Research, Development and Innovation Office, Hungary (K135946, PD128612 and K124544). B. Gyarmati acknowledges the János Bolyai Research Scholarship of the Hungarian Academy of Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

DSCdifferential scanning calorimetry
DTPdialkyl dithiophosphate
EGethylene glycol
FTIRFourier-transform infrared spectroscopy
GPCgel permeation chromatography
HD1,6-hexanediol
Mnnumber average molecular weight
Mpeakpeak molecular weight
NaOHsodium hydroxide
NMRnuclear magnetic resonance spectroscopy
P4S10phosphorus pentasulfide
PDTPpoly(dithiophosphate)
PEGpoly(ethylene glycol)
PEG400poly(ethylene glycol) with 400 g/mol number average molecular weight
Tgglass transition temperature
TGAthermogravimetric analysis
THFtetrahydrofuran

References

  1. Huo, S.; Song, P.; Yu, B.; Ran, S.; Chevali, V.S.; Liu, L.; Fang, Z.; Wang, H. Phosphorus-containing flame retardant epoxy thermosets: Recent advances and future perspectives. Prog. Polym. Sci. 2021, 114, 101366. [Google Scholar] [CrossRef]
  2. Liu, Y.; Zhang, P. Review of Phosphorus-Based Polymers for Mineral Scale and Corrosion Control in Oilfield. Polymers 2022, 14, 2673. [Google Scholar] [CrossRef] [PubMed]
  3. Todorova, Z.; Tumurbaatar, O.; Todorova, J.; Ugrinova, I.; Koseva, N. Phosphorus-containing star-shaped polymer conjugates for biomedical applications. Eur. Polym. J. 2021, 142, 110151. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Chen, X.; Xu, J.; Zhang, Q.; Gao, L.; Wang, Z.; Qu, L.; Wang, K.; Li, Y.; Cai, Z.; et al. Cross-Linked Polyphosphazene Nanospheres Boosting Long-Lived Organic Room-Temperature Phosphorescence. J. Am. Chem. Soc. 2022, 144, 6107–6117. [Google Scholar] [CrossRef] [PubMed]
  5. Allcock, H.R. Metallocene and other transition metal derivatives of phosphazene oligomers and polymers: A retrospective summary. Polymer 2022, 249, 124761. [Google Scholar] [CrossRef]
  6. Mu, X.; Li, X.; Liao, C.; Yu, H.; Jin, Y.; Yu, B.; Han, L.; Chen, L.; Kan, Y.; Song, L.; et al. Phosphorus-fixed stable interfacial nonflammable gel polymer electrolyte for safe flexible lithium-ion batteries. Adv. Funct. Mater. 2022, 32, 2203006. [Google Scholar] [CrossRef]
  7. Kracíková, L.; Ziółkowska, N.; Androvič, L.; Klimánková, I.; Červený, D.; Vít, M.; Pompach, P.; Konefał, R.; Janoušková, O.; Hrubý, M.; et al. Phosphorus-Containing Polymeric Zwitterion: A Pioneering Bioresponsive Probe for 31P-Magnetic Resonance Imaging. Macromol. Biosci. 2022, 22, 2100523. [Google Scholar] [CrossRef]
  8. Georgieva, R.; Tsevi, R.; Kossev, K.; Kusheva, R.; Balgjiska, M.; Petrova, R.; Tenchova, V.; Gitsov, I.; Troev, K. Immobilization of Aminothiols on Poly(oxyalkylene phosphates). Formation of Poly(oxyethylene phosphates)/Cysteamine Complexes and Their Radioprotective Efficiency. J. Med. Chem. 2002, 45, 5797–5801. [Google Scholar] [CrossRef] [PubMed]
  9. Sun, J.; Sun, G.; Zhao, X.; Zhao, H.; Yang, Z.; Yan, L.; Jiang, X.; Cui, Y. Efficient removal of Pb(II) and Cr(VI) from acidic wastewater using porous thiophosphoryl polyethyleneimine. New J. Chem. 2021, 45, 16196–16204. [Google Scholar] [CrossRef]
  10. Daglar, O.; Alkan, B.; Gunay, U.S.; Hizal, G.; Tunca, U.; Durmaz, H. Ultrafast synthesis of phosphorus-containing polythioethers in the presence of TBD. Eur. Polym. J. 2022, 162, 110931. [Google Scholar] [CrossRef]
  11. Zhang, X.H.; Theato, P. (Eds.) Sulfur-Containing Polymers: From Synthesis to Functional Materials; John Wiley-VCH: Weinheim, Germany, 2021. [Google Scholar]
  12. Lee, T.; Dirlam, P.T.; Njardarson, J.T.; Glass, R.S.; Pyun, J. Polymerizations with elemental sulfur: From petroleum refining to polymeric materials. J. Am. Chem. Soc. 2021, 144, 5–22. [Google Scholar] [CrossRef] [PubMed]
  13. Geven, M.; d’Arcy, R.; Turhan, Z.Y.; El-Mohtadi, F.; Alshamsan, A.; Tirelli, N. Sulfur-based oxidation-responsive polymers. Chemistry, (chemically selective) responsiveness and biomedical applications. Eur. Polym. J. 2021, 149, 110387. [Google Scholar] [CrossRef]
  14. Mutlu, H.; Döpping, D.A.; Huber, B.; Theato, P. Elemental Sulfur Mediated Novel Multicomponent Redox Polycondensation for the Synthesis of Alternating Copolymers Based on 2,4-Thiophene/Arene Repeating Units. Macromol. Rapid Commun. 2021, 42, 2000695. [Google Scholar] [CrossRef] [PubMed]
  15. Ziegenbalg, N.; Elbinger, L.; Schubert, U.S.; Brendel, J.C. Polymers from S-vinyl monomers: Reactivities and properties. Polym. Chem. 2022, 13, 5019–5041. [Google Scholar] [CrossRef]
  16. Scheiger, J.M.; Hoffmann, M.; Falkenstein, P.; Wang, Z.; Rutschmann, M.; Scheiger, V.W.; Grimm, A.; Urbschat, K.; Sengpiel, T.; Matysik, J.; et al. Inverse Vulcanization of Norbornenylsilanes: Soluble Polymers with Controllable Molecular Properties via Siloxane Bonds. Angew. Chem. Int. Ed. 2022, 61, e202114896. [Google Scholar] [CrossRef] [PubMed]
  17. Ceneviva, L.V.S.; Mierzati, M.; Miyahara, Y.; Nomura, C.T.; Taguchi, S.; Abe, H.; Tsuge, T. Poly(3-mercapto-2-methylpropionate), a Novel α-Methylated Bio-Polythioester with Rubber-like Elasticity, and Its Copolymer with 3-hydroxybutyrate: Biosynthesis and Characterization. Bioengineering 2022, 9, 228. [Google Scholar] [CrossRef]
  18. Li, C.; Liu, J.; Hong, Y.; Lin, R.; Liu, Z.; Chen, M.; Lam, J.W.Y.; Ning, G.-H.; Zheng, X.; Qin, A.; et al. Click Synthesis Enabled Sulfur Atom Strategy for Polymerization-Enhanced and Two-Photon Photosensitization. Angew. Chem. Int. Ed. 2022, 61, e202202005. [Google Scholar] [CrossRef]
  19. Abbasi, A.; Nasef, M.M.; Yahya, W.Z.N.; Moniruzzaman, M.; Ghumman, A.S.M. Preparation and characterization of sulfur-vinylbenzyl chloride polymer under optimized reaction conditions using inverse vulcanization. Eur. Polym. J. 2021, 143, 110202. [Google Scholar] [CrossRef]
  20. Carius, L. Die Ersetzung des Sauerstoffs durch Schwefel in organischen Verbindungen. J. Prakt. Chem. 1860, 79, 375–377. [Google Scholar] [CrossRef]
  21. Mastin, T.W.; Norman, G.R.; Weilmuenster, E.A. Chemistry of the Aliphatic Estersof Thiophosphoric acids. I. J. Am. Chem. Soc. 1945, 67, 1662–1664. [Google Scholar] [CrossRef]
  22. Li, H.; Zhang, Y.; Li, C.; Zhou, Z.; Nie, X.; Chen, Y.; Cao, H.; Liu, B.; Zhang, N.; Said, Z.; et al. Extreme pressure and antiwear additives for lubricant: Academic insights and perspectives. Int. J. Adv. Manuf. Technol. 2022, 120, 1–27. [Google Scholar] [CrossRef]
  23. Spikes, H. The history and mechanisms of ZDDP. Tribol. Lett. 2004, 17, 469–489. [Google Scholar] [CrossRef]
  24. Dixena, R.K.; Sayanna, E.; Badoni, R.P. A study on tribological behaviours of ZDDP in polymer thickened lubricating greases. Lubr. Sci. 2016, 28, 177–186. [Google Scholar] [CrossRef]
  25. Zou, S.; Lin, Q.; Wang, S.; Ma, X.; Zhong, H. A novel surfactant O,O’-bis(2-butoxyethyl) ammonium dithiophosphate: Synthesis, selective flotation and adsorption mechanism towards galena. Miner. Eng. 2022, 179, 107466. [Google Scholar] [CrossRef]
  26. Xu, Y.; Xie, Z.G.; Xue, L. Chelation of heavy metals by potassium butyl dithiophosphate. J. Environ. Chem. 2011, 23, 778–783. [Google Scholar] [CrossRef] [PubMed]
  27. Ozturk, T.; Ertas, E.; Mert, O. A Berzelius Reagent, Phosphorus Decasulfide (P4S10), in Organic Syntheses. Chem. Rev. 2010, 110, 3419–3478. [Google Scholar] [CrossRef] [PubMed]
  28. Sharma, D.; Firdoos, T.; Radha, A.; Kumar, S.; Shakya, S.; Jassal, A.K.; Pandey, S.K. Synthesis, experimental and theoretical analyses of bis(2-ethylphenyl)phosphorodithioates of nickel(II). J. Mol. Struct. 2022, 1263, 133166. [Google Scholar] [CrossRef]
  29. Cabras, V.; Pilloni, M.; Scano, A.; Lai, R.; Aragoni, M.C.; Coles, S.J.; Ennas, G. Mechanochemical Reactivity of Square-Planar Nickel Complexes and Pyridyl-Based Spacers for the Solid-State Preparation of Coordination Polymers: The Case of Nickel Diethyldithiophosphate and 4,4′-Bipyridine. Eur. J. Inorg. Chem. 2017, 13, 1908–1914. [Google Scholar] [CrossRef] [Green Version]
  30. Pudovik, A.N.; Cherkasov, R.A.; Kondrat’eva, R.M. Polyperesterification of diethyl ester of dithiophosphoric acid with glycols. Vysokomol. Soyed. 1967, 9, 1118–1122. [Google Scholar]
  31. Penczek, S.; Pretula, J.; Kubisa, P.; Kaluzynski, K.; Szymanski, R. Reactions of H3PO4 forming polymers. Apparently simple reactions leading to sophisticated structures and applications. Prog. Polym. Sci. 2015, 45, 44–70. [Google Scholar] [CrossRef]
  32. Pretula, J.; Kaluzynski, K.; Wisniewski, B.; Szymanski, R.; Loontjens, T.; Penczek, S. Formation of Poly(ethylene phosphates) in Polycondensation of H3PO4 with Ethylene Glycol. Kinetic and Mechanistic Study. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 830–843. [Google Scholar] [CrossRef]
  33. Nyquist, R.A. Vibrational spectroscopic study of organophosphorus compounds: P-S-H and P-H groups. Spectrochim. Acta 1969, 25, 47–66. [Google Scholar] [CrossRef]
  34. Carter, J.M.; Brown, E.M.; Irish, E.E.; Bowden, N.B. Characterization of Dialkyldithiophosphates as Slow Hydrogen Sulfide Releasing Chemicals and Their Effect on the Growth of Maize. J. Agric. Food Chem. 2019, 67, 11883–11892. [Google Scholar] [CrossRef] [PubMed]
  35. Burn, A.J.; Gosney, I.; Warrens, C.P.; Wastle, J.P. Phosphorus-31 NMR Investigation of the Heterogeneous Hydrolytic Decomposition of Zinc(lI) Bis(O,O-Dialkyl Dithiophosphate) Lubricant Additives. J. Chem. Soc. Perkin Trans. 1995, 2, 265–268. [Google Scholar] [CrossRef]
  36. Rubinstein, M.; Colby, R.H. Polymer Physics, 1st ed.; Oxford University Press: Oxford, UK, 2003; p. 293. [Google Scholar]
Scheme 1. The reaction of phosphorus pentasulfide (P4S10) with an alcohol.
Scheme 1. The reaction of phosphorus pentasulfide (P4S10) with an alcohol.
Ijms 23 15963 sch001
Scheme 2. The formation of poly(dithiophosphate) in the reaction of a diol with phosphorus pentasulfide.
Scheme 2. The formation of poly(dithiophosphate) in the reaction of a diol with phosphorus pentasulfide.
Ijms 23 15963 sch002
Figure 1. The gel permeation (GPC) chromatograms of PEG400 and P1–P9 PDTP samples.
Figure 1. The gel permeation (GPC) chromatograms of PEG400 and P1–P9 PDTP samples.
Ijms 23 15963 g001
Figure 2. The FTIR spectra of PEG400 and the P1 PDTP sample.
Figure 2. The FTIR spectra of PEG400 and the P1 PDTP sample.
Ijms 23 15963 g002
Figure 3. The 1H NMR (A) and the 31P NMR (B,C) spectra of the P1 PDTP sample recorded two days (B) and two and a half months (C) after the polymerization (solvent: CDCl3).
Figure 3. The 1H NMR (A) and the 31P NMR (B,C) spectra of the P1 PDTP sample recorded two days (B) and two and a half months (C) after the polymerization (solvent: CDCl3).
Ijms 23 15963 g003
Scheme 3. Branching point formation via the ring-opening reaction between the poly(dithiophosphate) and the tetrahydrofuran solvent molecule.
Scheme 3. Branching point formation via the ring-opening reaction between the poly(dithiophosphate) and the tetrahydrofuran solvent molecule.
Ijms 23 15963 sch003
Scheme 4. The hydrolytic and esterification reactions of the poly(dithiophosphate), the latter also leading to branching.
Scheme 4. The hydrolytic and esterification reactions of the poly(dithiophosphate), the latter also leading to branching.
Ijms 23 15963 sch004
Figure 4. The differential scanning calorimetry (DSC) curves of PEG400 and the P1 PDTP sample.
Figure 4. The differential scanning calorimetry (DSC) curves of PEG400 and the P1 PDTP sample.
Ijms 23 15963 g004
Figure 5. The thermogravimetric analysis (TGA) curves with the corresponding solid residue values of PEG400 and the P1–P4 PDTP samples.
Figure 5. The thermogravimetric analysis (TGA) curves with the corresponding solid residue values of PEG400 and the P1–P4 PDTP samples.
Ijms 23 15963 g005
Figure 6. The storage (filled symbols) and loss (open symbols) moduli values of PEG400, P1–P4, P6 and P7 PDTP samples, obtained by oscillation rheometry.
Figure 6. The storage (filled symbols) and loss (open symbols) moduli values of PEG400, P1–P4, P6 and P7 PDTP samples, obtained by oscillation rheometry.
Ijms 23 15963 g006
Figure 7. The 1H NMR spectrum of the P9 PDTP copolymer sample.
Figure 7. The 1H NMR spectrum of the P9 PDTP copolymer sample.
Ijms 23 15963 g007
Table 1. The diol ratios in the syntheses of poly(dithiophosphate)s, the glass transition temperature (Tg) values and the peak molecular weights (Mpeak) of the resulting polymers (reaction conditions: 1 g/mL reactant/THF ratio, 60 °C, 6 h).
Table 1. The diol ratios in the syntheses of poly(dithiophosphate)s, the glass transition temperature (Tg) values and the peak molecular weights (Mpeak) of the resulting polymers (reaction conditions: 1 g/mL reactant/THF ratio, 60 °C, 6 h).
SampleDiol Ratios in the Feed (Mol%)Tg (°C) *Mpeak (g/mol) **
PEG400EGHD
P110000−471700
P275250−441700
P350500−421600
P425750−342200
P501000−4400
P675025−471400
P750050−461500
P825075−431600
P900100−451000
* Determined by differential scanning calorimetry (DSC). ** Determined by gel permeation chromatography (GPC).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Szabó, Á.; Szarka, G.; Trif, L.; Gyarmati, B.; Bereczki, L.; Iván, B.; Kovács, E. Poly(dithiophosphate)s, a New Class of Phosphorus- and Sulfur-Containing Functional Polymers by a Catalyst-Free Facile Reaction between Diols and Phosphorus Pentasulfide. Int. J. Mol. Sci. 2022, 23, 15963. https://doi.org/10.3390/ijms232415963

AMA Style

Szabó Á, Szarka G, Trif L, Gyarmati B, Bereczki L, Iván B, Kovács E. Poly(dithiophosphate)s, a New Class of Phosphorus- and Sulfur-Containing Functional Polymers by a Catalyst-Free Facile Reaction between Diols and Phosphorus Pentasulfide. International Journal of Molecular Sciences. 2022; 23(24):15963. https://doi.org/10.3390/ijms232415963

Chicago/Turabian Style

Szabó, Ákos, Györgyi Szarka, László Trif, Benjámin Gyarmati, Laura Bereczki, Béla Iván, and Ervin Kovács. 2022. "Poly(dithiophosphate)s, a New Class of Phosphorus- and Sulfur-Containing Functional Polymers by a Catalyst-Free Facile Reaction between Diols and Phosphorus Pentasulfide" International Journal of Molecular Sciences 23, no. 24: 15963. https://doi.org/10.3390/ijms232415963

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop