Next Article in Journal
Self-Assembled Bimetallic Aluminum-Salen Catalyst for the Cyclic Carbonates Synthesis
Next Article in Special Issue
Side-Arm Assisted Anilido-Imine Based Rare-Earth Metal Complexes for Isoprene Stereoselective Polymerization
Previous Article in Journal
Pharmacological Efficacy of Ginseng against Respiratory Tract Infections
Previous Article in Special Issue
High-Solids, Solvent-Free Modification of Engineered Polysaccharides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate)

Polymer Physics Laboratory, Institute of Science and Engineering, Kakuma Campus, Kanazawa University, Kanazawa 920-1192, Ishikawa, Japan
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(13), 4096; https://doi.org/10.3390/molecules26134096
Submission received: 13 June 2021 / Revised: 2 July 2021 / Accepted: 2 July 2021 / Published: 5 July 2021

Abstract

:
We report that lithium salts in lithium-ion batteries effectively modify the physical properties of poly (methyl methacrylate) (PMMA). The glass transition temperature (Tg) is an indicator of the heat resistance of amorphous polymers. The anionic species of the salts strongly affected the glass transition behavior of PMMA. We focused on the additive effects of various lithium salts, such as LiCF3SO3, LiCOOCF3, LiClO4, and LiBr, on the Tg of PMMA. The large anions of the former three salts caused them to form macroscopic aggregates that acted as fillers in the PMMA matrix and to combine the PMMA domains, increasing Tg. On the other hand, LiBr salts dispersed microscopically in the PMMA matrix at the molecular scale, leading to the linking of the PMMA chains. Thus, the addition of LiBr to PMMA increased Tg as well as the relaxation time in the range of glass to rubber transition.

Graphical Abstract

1. Introduction

Transparent plastics, in particular lightweight and highly processable poly(methyl methacrylate) (PMMA), have replaced various inorganic glassy materials in many optical components, such as optical films for displays, optical lenses, optical fibers, and touch panels. PMMA, in particular, has the best transparency among the existing transparent plastics. The ester groups in PMMA make it possible to modify the physical properties of acrylic resins by copolymerization with other monomers or by modifying them with other chemical species [1,2,3,4,5].
Ionomers, which are ionic polymers with a small amount of ionic groups introduced into polymer chains, are also effective in the modification of PMMA using polar groups. Interestingly, these ionic groups aggregate in the host polymer matrix to form microphase-separated structures, which directly affect their molecular mobility. The introduction of a small amount of ionic groups into polymers, such as ionomers, has been attempted since the 1950s [6]; accordingly, a detailed characterization of their molecular aggregation states has been strongly desired. The formation of such ionic aggregates is considered to have a direct effect on the physical and mechanical properties of ionomers. However, since ionomers are complicated multi-body systems [7], it is difficult to elucidate the relationship between the molecular interactions in them and their mechanical properties.
We recently developed a simple method for modifying PMMA by adding salts. This method was inspired by the solid electrolyte system of lithium-ion batteries. Poly(ethylene oxide) (PEO), which is a rubbery polymer at room temperature, has been conventionally used as a solid electrolyte material [8,9]. It is well known that the addition of lithium ions increases the glass transition temperature (Tg) of PEO, and this phenomenon has been recognized as a problem for batteries because the increase in Tg inhibits the conduction of ions. This seems to be owing to the ion-dipole interaction between the oxygen atoms in PEO and the lithium cations, leading to the suppression of the segmental motion of PEO chains [10,11]. Furthermore, the electrolytes based on PMMA have also been reported [12,13,14,15].
In light of the above, we investigated the additive effects of lithium salts on the glass transition behavior of PMMA, based on the intermolecular interaction between the ester groups of the PMMA side chains [16] and the added lithium salts [17,18,19]. To date, we have succeeded in reducing the birefringence [17] and improving the heat resistance [18] of transparent plastics by adding lithium salts to PMMA and utilizing the intermolecular interaction between the polar groups of PMMA and the added salts. The objective of this work was to clarify the relationship between the salt aggregation structure and the molecular mobility of PMMA chains in PMMA/salt-doped systems. For this purpose, we investigated the additive effects of lithium salts with various anions on the dynamic mechanical spectra of the hosting PMMA matrix.

2. Results and Discussion

2.1. Effects of the Anion Characteristics of Salts

We examined the dynamic mechanical spectra of the PMMA sheets doped with several lithium salts with large anions: LiCF3SO3, LiCOOCF3, and LiClO4. These lithium salts have been used as typical fillers for lithium electrolytes. The resulting dynamic mechanical properties of these samples for temperatures in the 30–200 °C range and salt content of 7 mol% are shown in Figure 1. In addition, the Tg values of the PMMA matrix, estimated from the peak temperature of the E” curve, are summarized in Table 1. The Tg values measured at 10 Hz were different from previous differential scanning calorimetry (DSC) data [18,19] but showed similar behavior as DSC. This difference in values is because the DSC data results from the heat capacity changes under the glass transition, whereas the dynamic mechanical data are associated with the mechanical relaxation due to molecular motion and depend on the mechanical frequency. These data indicate that doping with lithium salts increases the Tg values of the PMMA matrix. It is likely that the obvious increase in Tg is caused by the molecular interaction between the lithium cations and ester groups in PMMA. This is because the molecular interactions between the cations and ester groups in the present PMMA/salt samples were confirmed by the ATR spectra for wavenumbers in the 1650–1780 cm−1 range, as shown in Figure S2.
We decomposed the corresponding peaks of the ATR spectra and estimated the concentration of PMMA/salt aggregates based on their peak areas to better determine the effects of the anionic species of the added salts on the Tg of PMMA. The deconvoluted ATR spectra of the different PMMA/salt samples are shown in Figure 2, fitted by Lorentzian functions [20].
In Figure 2a, the stretching vibration modes of ClO4 are separated, yielding the peaks at 623 cm−1 and 635 cm−1, corresponding to the free anions and bound or contact ClO4 anions [21,22], respectively. Figure 2b shows the peak deconvolution of the ATR spectra of the PMMA/LiCOOCF3 sample with the added salt at 7 mol%. Here, we removed the pristine PMMA spectrum from the ATR spectrum of its doped spectrum because the bands ascribed to the anions in the 1700 cm−1 band overlapped with the carbonyl symmetric stretching vibration mode of pristine PMMA at 1724 cm−1. According to Regis et al. [23], the asymmetric stretching vibration band of COO is composed of four modes, at 1695 cm−1 (free ions), 1736 cm−1 (triplet anions), 1710 cm−1 (ion pairs), and 1729 cm−1 (aggregates). Figure 2c shows the peak decomposed ATR spectra of PMMA/LiCF3SO3 with 7 mol% of the added salt. In addition, we removed the PMMA spectra from the ATR spectra of the doped samples because the PMMA peaks for wavenumbers in the 1020–1080 cm−1 range overlapped. The symmetric stretching vibration mode of SO3 is composed of three modes at 1032 cm−1 (free ions), 1040 cm−1 (ion pairs), and 1045 cm−1 (triplet anions) [24]. The contributions of the different anionic species (ClO4, CF3SO3, and COOCF3) were quantitatively estimated from the peak areas of the decomposed spectra (Table 2).
Here, we considered the fraction of non-free anions as the aggregation degree. Solubility parameters were evaluated by estimating the cohesive energies of these salts using the atomic cluster contribution method. As shown in the Supporting Information, the fraction of non-free anions is related to the solubility parameter, strongly suggesting that the fraction of non-free anions can be considered the level of the aggregation state of anions in the doped PMMA samples. The aggregation degrees for the different PMMA matrices were in the following order: LiCF3SO3 > LiCOOCF3 > LiClO4.
As shown in Figure 3, the reciprocal Tg was proportional to the aggregation degree, according to the following empirical relationship:
1 T g = 1 ϕ T g 0
where T g 0 is the glass transition temperature of pristine PMMA, and φ is the aggregation degree of the salt anions, estimated from the fraction of non-free anions. The inverse proportionality of Tg implies the homogeneous dispersion of the added salt in the host PMMA matrix. It is likely that the molecular interaction between the polar groups of PMMA and the anion aggregates suppresses the overall segmental motion of PMMA. The effects of the cohesive force between the anions of the added salt are shown in the Supporting Information. It is interesting to note, however, that a chemical crosslinking process reduces the Tg of PMMA, as demonstrated by Pásztor et al. [25].

2.2. Effects of Anion Sizes of the Salts

We investigated the effects of the anion sizes of the salts on the formation of the salt aggregates by comparing the doping effects of LiCF3SO3 and LiBr on the Tg of the PMMA matrix. As shown in Figure 8, there are also no significant peaks derived from the salt in the XRD pattern of PMMA doped with LiBr. Figure 4 shows the temperature dependence of the dynamic mechanical spectra, tanδ, for the PMMA/LiCF3SO3 and PMMA/LiBr samples for [Li]/[C=O] in the 0–7 mol% range. The Tg values of PMMA/LiCF3SO3 monotonously shifted toward higher temperatures with increasing salt concentrations. This behavior was also observed for the LiCOOCF3 and LiClO4 doped samples. On the other hand, the doping of LiBr broadened the glass-transition peak as the concentration of LiBr increased [18,22].
Figure 5 shows the rheological behavior, from rubbery to flow regions for both LiCF3SO3 and LiBr-doped samples. The averaged relaxation times, which were estimated from the crossover points of G’ and G” master curves, were 0.40 s for pristine PMMA, 0.63 s for PMMA/LiCF3SO3, and 180 s for PMMA/LiBr [18,22]. Furthermore, as shown in Figure 5b, the master curve of PMMA can be successfully superimposed on the curve of PMMA/LiCF3SO3 by shifting to a higher modulus on the longitudinal axis under a slight horizontal shift on the frequency axis. The horizontal shift results from the ratio of average relaxation times. Thus, the addition of the conventional LiCF3SO3 salt shifted the overall segmental motion of the PMMA chains toward higher frequencies, based on a simple thermo-rheological process. Note that the longitudinal shift in PMMA/LiCF3SO3 indicates an overall increase in the modulus, which can be considered as a filler effect of the salt aggregates homogeneously dispersed in the PMMA matrix. This is because the cohesive force of the salts is superior to the intermolecular interactions with the PMMA chains.
On the other hand, the relaxation time spectra of PMMA/LiBr expanded considerably over time. The terminal zone in the G’ and G” spectra was prolonged and suppressed the melt-flow behavior of the PMMA chains. Importantly, the G’ and G” curves of pristine PMMA could not be superimposed on the curves of PMMA/LiBr. This implies that the LiBr salts effectively expanded the melt-flow regions of PMMA. Therefore, the prolonged terminal zone in PMMA/LiBr suggests that the PMMA chains were associated with the Br anions, and the associated PMMA chains exhibited an increase in molecular weight and/or gelation, as seen in ionomers. In addition, it was confirmed using a conventional methanol-based immersion test that the association in PMMA/LiBr is owing to physical (rather than chemical) linkages. On the other hand, and as described above, the LiCF3SO3 salts aggregate in the PMMA matrix where the salt aggregates suppress the overall segmental mobility of the PMMA chains. The structural images are shown in Figure 6.
Figure 7 compares the ATR spectra of pristine PMMA and PMMA/LiBr with 7 mol% salt sheets, with and without rinsing with methanol.
The positions of the absorption bands in the ATR spectra and their assignments are summarized in Table S2 in the Supporting Information. The ATR spectra were normalized to a peak height at 1435 cm−1, assigned to the symmetric stretching vibration mode of OCH3 [26]. Figure 7 [27,28] shows that only the intensity of the 1724 cm−1 band ascribed to carbonyl groups decreased after adding LiBr but increased owing to the methanol rinsing, indicating that methanol washing weakened the association between carbonyl groups and the salt anions. Interestingly, a broad peak appeared in the 1500–600 cm−1 range for PMMA/LiBr, as shown in Figure 7. According to a classical coupled vibration model between C=O and lithium, the vibration mode of the coupling interaction is expected in the 1400–1500 cm−1 range. Details of the calculations are provided in the Supporting Information. This result led us to conclude that strong interactions comparable to covalent bonds, such as single and double bonds, are formed between oxygen atoms and lithium cations. This may be ascribed to the ionic association of PMMA.

3. Materials and Methods

3.1. Materials and Sample Preparation

PMMA pellets (Acrypet VH; Mw = 120,000, molecular weight distribution index of Mw/Mw = 2.2) were commercial grades. LiCF3SO3 (purity ≥ 98.0%; Kanto Chemical Industry Co., Ltd., Tokyo, Japan), LiCOOCF3 (purity 97%; Alfa Aesar, Massachusetts, America), LiClO4·3H2O (purity 97%; Nacalai Tesque, Inc., Kyoto, Japan), and LiBr (purity ≥ 98.0%; Tokyo Chemical Industry Co., Ltd., Tokyo, Japan) were purchased and used without any additional purification. The salt content was estimated as the molar ratio of lithium ions (Li) to PMMA carbonyl groups (C=O). The PMMA/salt films were cast using a mixture of dichloromethane and methanol at a weight ratio of 9:1 for 1 h. The salt concentration was fixed at 0.07 molar ratio of lithium cations with respect to the PMMA ester groups for each sample, except for the measurements of viscoelastic properties. These LiClO4, LiCOOCF3, and LiCF3SO3 salts used in this work were relatively soluble with PMMA; on the other hand, LiBr was less soluble. In order to clarify the additive effects of LiCF3SO3 and LiBr, we used the maximum concentration of LiBr, which was 7 mol%.
After drying in a fume hood, the samples were dried at 160 °C for 30 h to evaporate the residual solvents. Finally, the PMMA/salt sheets with a thickness of approximately 300 µm were obtained by compression molding at 200 °C and 30 MPa for 10 min and were quenched at 25 °C for 3 min. The ionic conductivity values of these sheets were too low to be precisely estimated.

3.2. Characterization

X-ray diffraction (XRD) measurements of the prepared samples were performed using a SmartLab diffractometer (Rigaku Co., Ltd., Tokyo, Japan) to estimate the miscibility of the added salts in the samples. The X-ray radiation source was Cu Kα, and an X-ray tube voltage of 40 kV and a current of 30 mA were used. The diffraction angle ranged from 5° to 80°. The scan step and duration were 0.1° and 1 s, respectively.
The XRD patterns are shown in Figure 8. The absence of crystallization in the added salts suggests that the salts were uniformly dispersed in the PMMA matrix.
Ultraviolet-visible (UV-vis) absorption spectra were obtained using V650 (JASCO Co., Ltd., Tokyo, Japan) for wavelengths in the 400–700 nm range. The UV-vis spectra of the prepared samples showed that the transmittance values of the samples with salt concentrations of 7 mol% were above 80%, except for PMMA/LiBr (Figure S1).

3.3. Measurements

Dynamic mechanical analysis was performed using Rheogel-E4000 (UBM Co., Ltd., Kyoto, Japan). Rectangular specimens with the 5 mm width and 20 mm length were cut out from the sheets with a thickness of approximately 300 μm. The temperature dependences of the tensile storage modulus (E’) and loss modulus (E”) were measured for temperatures in the 30–200 °C range with an initial chuck distance of 10 mm. The frequency and the heating rate were 10 Hz and 2 °C/min, respectively. The frequency dependences of the oscillatory shear moduli G’ and G” were measured using a cone-and-plate rheometer in the temperature range of 200–240 °C using Rheogel-E4000 (UBM Co., Ltd., Kyoto, Japan). Attenuated total reflection (ATR) measurements were performed using Spectrum 100 (PerkinElmer Co., Ltd., Kanagawa, Japan) with a KRS-5 ATR prism for evaluating the infrared spectra under a nitrogen flow. The accumulation count was 16, and the resolution was 4 cm−1.

4. Conclusions

LiCF3SO3, LiCOOCF3, and LiClO4, which are conventionally used as lithium electrolytes, and an alkali halide, LiBr, were incorporated into PMMA. In the case of LiCF3SO3, LiCOOCF3, and LiClO4, the aggregates were homogeneously dispersed in the PMMA matrix, and these salts thermo-rheologically shifted the segmental mobility of PMMA toward lower temperatures. In addition, the Li salts effectively acted as fillers in the PMMA matrix, suggesting that they can be used as stiffening agents. On the other hand, the alkali halide molecules acted as physical cross-linkages, leading to the ionic association of the PMMA chains and resulting in the expansion of the terminal relaxation time spectra of PMMA into the flow region. The addition of these lithium salts enables controlling the melt flow behavior of PMMA.

Supplementary Materials

The following are available online. Figure S1: UV-Vis spectra of PMMA/salt samples with [Li]/[C=O] = 0.07 and PMMA/LiBr with [Li]/[C=O] = 0.01. PMMA/LiBr with 1 mol% of the salt concentration showed low transmittance below 30%. Figure S2: ATR spectra of PMMA/salt samples in the range of the stretching vibration mode of carbonyl groups with [Li]/[C=O] = 0.07. Figure S3: Relationship between the Tg of the PMMA/salt samples and cohesive energies of salts, for [Li]/[C=O] = 0.07. Figure S4: Three masses are connected by a Hookean spring to each other. This system describes coupled motion along one dimension. Table S1: Functional groups and molar attraction constants of the salts. Table S2: Wavenumbers and assignments of infrared absorption bands of PMMA.

Author Contributions

Conseptualization, A.I.; methodology, A.I.; data curation, A.I.; formal analysis, A.I. and K.-h.N.; writing-original draft preparation, A.I.; writing-review and editing, A.I. and K.-h.N.; supervision, K.-h.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by JSPS KAKENHI Grant Number 21K14680.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in supplementary material.

Acknowledgments

I would like to thank M. Yamaguchi for useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Xu, J.; He, J.; Fan, D.; Tang, W.; Yang, Y. Thermal decomposition of dithioesters and its effect on RAFT polymerization. Macromolecules 2006, 39, 3753–3759. [Google Scholar] [CrossRef]
  2. Ma, J.; Lu, M.; Cao, C.; Zhang, H. Synthesis and characterization of PMMA/SiO2 organic-inorganic hybrid materials via RAFT-mediated miniemulsion polymerization. Polym. Compos. 2013, 34, 626–633. [Google Scholar] [CrossRef]
  3. Tan, W.L.; Bakar, M.A.; Bakar, N.H.H.A. Effect of anion of lithium salt on the property of lithium salt-epoxidized natural rubber polymer electrolytes. Ionics 2013, 19, 601–613. [Google Scholar] [CrossRef]
  4. Hayashi, M.; Kimura, T.; Oba, Y.; Takasu, A. One-Pot Synthesis of Dual supramolecular associative PMMA-based copolymers and the precise thermal property tuning. Macromol. Chem. Phys. 2021, 222, 2000302. [Google Scholar] [CrossRef]
  5. Ali, U.; Karim, K.J.B.A.; Buang, N.A. A Review of the properties and applications of poly (methyl methacrylate) (PMMA). Polym. Rev. 2015, 55, 678–705. [Google Scholar] [CrossRef]
  6. Longworth, R.; Vaughan, D.J. Physical Structure of Ionomers. Nature 1968, 218, 85–87. [Google Scholar] [CrossRef]
  7. Kim, J.; Jackman, R.J.; Eisenberg, A. Filler and percolation behavior of ionic aggregates in styrene-sodium methacrylate ionomers. Macromolecules 1994, 27, 2789–2803. [Google Scholar] [CrossRef]
  8. Fenton, D.E.; Parker, J.M.; Wright, P.V. Complexes of alkali metal ions with poly (ethylene oxide). Polymer 1973, 14, 589. [Google Scholar] [CrossRef]
  9. Armand, M. Polymer solid electrolytes-an overview. Solid State Ionics 1983, 9–10, 745–754. [Google Scholar] [CrossRef]
  10. Tominaga, Y.; Izumi, Y.; Kwak, G.H.; Asai, S.; Sumita, M. Effect of supercritical carbon dioxide processing on ionic association and conduction in a crystalline poly(ethylene oxide)-LiCF3SO3 complex. Macromolecules 2003, 36, 8766–8772. [Google Scholar] [CrossRef]
  11. Tominaga, Y.; Asai, S.; Sumita, M. Relation between ionic conductivity and solubility of CO2 in pressurized solid polymer electrolytes. Macromolecules 2007, 40, 3348–3354. [Google Scholar] [CrossRef]
  12. Agnihotry, S.A.; Pradeep, P.; Sekhon, S.S. PMMA based gel electrolyte for EC smart windows. Electrochim. Acta 1999, 44, 3121–3126. [Google Scholar] [CrossRef]
  13. Shukla, N.; Thakur, A.K. Role of salt concentration on conductivity optimization and structural phase separation in a solid polymer electrolyte based on PMMA-LiClO4. Ionics 2009, 15, 357–367. [Google Scholar] [CrossRef]
  14. TianKhoon, L.; Ataollahi, N.; Hassan, N.H.; Ahmad, A. Studies of porous solid polymeric electrolytes based on poly (vinylidene fluoride) and poly (methyl methacrylate) grafted natural rubber for applications in electrochemical devices. J. Solid State Electrochem. 2016, 20, 203–213. [Google Scholar] [CrossRef]
  15. Kurapati, S.; Gunturi, S.S.; Nadella, K.J.; Erothu, H. Novel solid polymer electrolyte based on PMMA:CH3COOLi effect of salt concentration on optical and conductivity studies. Polym. Bull. 2019, 76, 5463–5481. [Google Scholar] [CrossRef]
  16. Ito, A.; Ayerdurai, V.; Miyagawa, A.; Matsumoto, A.; Okada, H.; Courtoux, A.; Yamaguchi, M. Effects of residual solvent on glass transition temperature of poly(methyl methacrylate). Nihon Reoroji Gakkaishi 2018, 46, 117–121. [Google Scholar] [CrossRef] [Green Version]
  17. Ito, A.; Maeno, R.; Yamaguchi, M. Control of optical and mechanical properties of poly(methyl methacrylate) by introducing lithium salt. Opt. Mater. 2018, 83, 152–156. [Google Scholar] [CrossRef]
  18. Ito, A.; Phulkerd, P.; Ayerdurai, V.; Soga, M.; Courtoux, A.; Miyagawa, A.; Yamaguchi, M. Enhancement of the glass transition temperature of poly(methyl methacrylate) by salt. Polym. J. 2018, 50, 857–863. [Google Scholar] [CrossRef]
  19. Miyagawa, A.; Ayerdurai, V.; Nobukawa, S.; Yamaguchi, M. Viscoelastic properties of poly(methyl methacrylate) with high glass transition temperature by lithium salt addition. J. Polym. Sci. Part. B Polym. Phys. 2016, 54, 2388–2394. [Google Scholar] [CrossRef]
  20. Stancik, A.L.; Brauns, E.B. A simple asymmetric lineshape for fitting infrared absorption spectra. Vib. Spectrosc. 2008, 47, 66–69. [Google Scholar] [CrossRef]
  21. Wieczorek, W.; Lipka, P.; Żukowska, G.; Wyciślik, H. Ionic interactions in polymeric electrolytes based on low molecular weight poly(ethylene glycol)s. J. Phys. Chem. B 1998, 102, 6968–6974. [Google Scholar] [CrossRef]
  22. Ito, A. Effect of Addition of Lithium Salts on Properties of Poly(Methyl Methacrylate). Ph.D. Thesis, Japan Advanced Institute of Science and Technology (JAIST), Nomi, Japan, 2019. [Google Scholar]
  23. Regis, A.; Corset, J. IR study of solutions of lithium trifluoroacetate in non-aqueous solvents. Structure of ion pairs and triple ions. Chem. Phys. Lett. 1975, 32, 462–465. [Google Scholar] [CrossRef]
  24. Frech, R.; Huang, W. Polymer conformation and ionic association in complexes of lithium, sodium and potassium triflate with poly (ethylene oxide) oligomers. Solid State Ionics 1994, 72, 103–107. [Google Scholar] [CrossRef]
  25. Pásztor, S.; Becsei, B.; Szarka, G.; Thomann, Y.; Thomann, R.; Mühlhaupt, R.; Iván, B. The scissors effect in action: The fox-flory relationship between the glass transition temperature of crosslinked poly(methyl methacrylate) and Mc in nanophase separated poly(methyl methacrylate)-l-polyisobutylene conetworks. Materials 2020, 13, 4822. [Google Scholar] [CrossRef]
  26. Willis, H.A.; Zichy, V.J.I.; Hendra, P.J. The laser-Raman and infra-red spectra of poly(methyl methacrylate). Polymer 1969, 10, 737–746. [Google Scholar] [CrossRef]
  27. Brinkhuis, R.H.G.; Schouten, A.J. Thin-film behavior of poly(methyl methacrylates). 2. An FT-IR study of Langmuir-Blodgett films of isotactic PMMA. Macromolecules 1991, 24, 1496–1504. [Google Scholar] [CrossRef]
  28. Havriliak, S.; Roman, N. The infra-red absorption characteristics of syndiotactic poly(methyl methacrylate) from 1050 cm−1 to 1300 cm−1. Polymer 1966, 7, 387–400. [Google Scholar] [CrossRef]
Figure 1. The dynamic mechanical properties of the prepared PMMA/salt samples with [Li]/[C=O] = 0.07.
Figure 1. The dynamic mechanical properties of the prepared PMMA/salt samples with [Li]/[C=O] = 0.07.
Molecules 26 04096 g001
Figure 2. ATR spectra of the PMMA samples doped with (a) LiClO4, (b) LiCOOCF3, and (c) LiCF3SO3. Their concentrations of lithium salts are [Li]/[C=O] = 0.07. The black lines denote the experimental spectra and the dashed gray lines denote the computational ones.
Figure 2. ATR spectra of the PMMA samples doped with (a) LiClO4, (b) LiCOOCF3, and (c) LiCF3SO3. Their concentrations of lithium salts are [Li]/[C=O] = 0.07. The black lines denote the experimental spectra and the dashed gray lines denote the computational ones.
Molecules 26 04096 g002
Figure 3. The inverse dependence of Tg on the aggregation degree for the PMMA/salt samples with [Li]/[C=O] = 0.07.
Figure 3. The inverse dependence of Tg on the aggregation degree for the PMMA/salt samples with [Li]/[C=O] = 0.07.
Molecules 26 04096 g003
Figure 4. Temperature dependences of dynamic mechanical spectra for (a) PMMA/LiCF3SO3 and (b) PMMA/LiBr.
Figure 4. Temperature dependences of dynamic mechanical spectra for (a) PMMA/LiCF3SO3 and (b) PMMA/LiBr.
Molecules 26 04096 g004
Figure 5. (a) Storage and loss moduli of PMMA/LiCF3SO3 and PMMA/LiBr with [Li]/[C=O] = 0.01 plotted against the frequency reduced to 200 °C. (b) The longitudinal shift of PMMA/LiCF3SO3 toward PMMA.
Figure 5. (a) Storage and loss moduli of PMMA/LiCF3SO3 and PMMA/LiBr with [Li]/[C=O] = 0.01 plotted against the frequency reduced to 200 °C. (b) The longitudinal shift of PMMA/LiCF3SO3 toward PMMA.
Molecules 26 04096 g005
Figure 6. A schematic of the models of (a) PMMA/LiCF3SO3 and (b) PMMA/LiBr with conventional lithium salts and lithium halide.
Figure 6. A schematic of the models of (a) PMMA/LiCF3SO3 and (b) PMMA/LiBr with conventional lithium salts and lithium halide.
Molecules 26 04096 g006
Figure 7. ATR spectra normalized to the peak height at the wavenumber of 1435 cm−1, for PMMA (black), PMMA/LiBr after rinsing with methanol (blue), and PMMA/LiBr (red), for wavenumbers in the 1000–1800 cm−1 range. Their concentrations of LiBr are [Li]/[C=O] = 0.07.
Figure 7. ATR spectra normalized to the peak height at the wavenumber of 1435 cm−1, for PMMA (black), PMMA/LiBr after rinsing with methanol (blue), and PMMA/LiBr (red), for wavenumbers in the 1000–1800 cm−1 range. Their concentrations of LiBr are [Li]/[C=O] = 0.07.
Molecules 26 04096 g007
Figure 8. XRD patterns of PMMA/salt samples with [Li]/[C=O] = 0.07.
Figure 8. XRD patterns of PMMA/salt samples with [Li]/[C=O] = 0.07.
Molecules 26 04096 g008
Table 1. The values of Tg for the prepared PMMA/salt samples with [Li]/[C=O] = 0.07.
Table 1. The values of Tg for the prepared PMMA/salt samples with [Li]/[C=O] = 0.07.
Sample CodeTg/°C
PMMA127
PMMA/LiClO4126
PMMA/LiCOOCF3133
PMMA/LiCF3SO3137
PMMA/LiBr133
Table 2. The contribution ratios of the structures of anionic species (ClO4, CF3SO3, and COOCF3) calculated using the area ratios of the ATR spectra peaks [22]. Their concentrations of lithium salts are [Li]/[C=O] = 0.07.
Table 2. The contribution ratios of the structures of anionic species (ClO4, CF3SO3, and COOCF3) calculated using the area ratios of the ATR spectra peaks [22]. Their concentrations of lithium salts are [Li]/[C=O] = 0.07.
Sample CodePeak Area/%
PMMA/LiClO4PMMA/LiCOOCF3PMMA/LiCF3SO3
Free72.344.117.3
Ion pair27.740.842.7
Triple ion5.740.0
Aggregates9.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ito, A.; Nitta, K.-h. Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate). Molecules 2021, 26, 4096. https://doi.org/10.3390/molecules26134096

AMA Style

Ito A, Nitta K-h. Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate). Molecules. 2021; 26(13):4096. https://doi.org/10.3390/molecules26134096

Chicago/Turabian Style

Ito, Asae, and Koh-hei Nitta. 2021. "Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate)" Molecules 26, no. 13: 4096. https://doi.org/10.3390/molecules26134096

Article Metrics

Back to TopTop