Next Article in Journal
Biquaternionic Dirac Equation Predicts Zero Mass for Majorana Fermions
Next Article in Special Issue
Asymmetric Cellulosic Membranes: Current and Future Aspects
Previous Article in Journal
Durrmeyer-Type Generalization of Parametric Bernstein Operators
Previous Article in Special Issue
Surface Modifications of Nanofillers for Carbon Dioxide Separation Nanocomposite Membrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Specific Structure and Properties of Composite Membranes Based on the Torlon® (Polyamide-imide)/Layered Perovskite Oxide

1
Institute of Chemistry, Saint Petersburg State University, Universitetskiy pr. 26, Saint Petersburg 198504, Russia
2
Institute of Macromolecular Compounds, Russian Academy of Sciences, Bolshoy pr. 31, Saint Petersburg 199004, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2020, 12(7), 1142; https://doi.org/10.3390/sym12071142
Submission received: 31 May 2020 / Revised: 15 June 2020 / Accepted: 1 July 2020 / Published: 8 July 2020
(This article belongs to the Special Issue Asymmetric Membranes)

Abstract

:
The use of perovskite-type layered oxide K2La2Ti3O10 (Per) as a modifier of the Torlon® polyamide-imide (PAI) membrane has led to the formation of an specific structure of a dense nonsymmetrical film, namely, a thin perovskite-enriched layer (3–5 μm) combined with the polymer matrix (~30 μm). The PAI/Per membrane structure was studied by SEM in combination with energy dispersive microanalysis of the elemental composition which illustrated different compositions of top and bottom surfaces of the perovskite-containing membranes. Measurement of water and alcohol contact angles and calculation of surface tension revealed hydrophilization of the membrane surface enriched with perovskite. The transport properties of the nonsymmetrical PAI/Per membranes were studied in the pervaporation of ethanol‒ethyl acetate mixture. The inclusion of 2 wt.% Per in the PAI gives a membrane with a high separation factor and increased total flux.

Graphical Abstract

1. Introduction

Progressive development worldwide depends on technologies that provide energy conservation, as well as minimizing waste and having zero emissions. The concept of an ecological future pushes into the foreground the use of membrane technologies as a favorable substitute for traditional separation processes [1,2,3,4].
The key element determining the effectiveness of membrane processes is the membrane. Most processes are carried out at elevated temperatures, therefore the use of membranes based on polymers of a heteroaromatic structure with enhanced thermal and chemical resistance is preferable [5,6,7,8]. We chose Torlon® polyamide-imide (PAI), a thermoplastic that is characterized by high mechanical strength, a wide temperature range of operation (from −270 °C to +250 °C), chemical resistance, and relative ease of manufacturability. These unique properties cause an interest in the manufacturing membranes based on PAI for ultra- and nanofiltration [9], gas separation [10], and pervaporation [11,12,13,14,15,16]. However, published works on pervaporation with PAI membranes had aimed alcohol dehydration and data on the performance that were not impressive. To enhance performance, the PAI membranes have been modified by blending with polyesterimide [8], polyimide [9], and gelatin [10]. In the present work, Torlon® PAI membrane and its modified form were studied in pervaporation of organic mixture for the first time. We developed a new composite membrane by PAI modification with the perovskite-type layered oxide K2La2Ti3O10 (Scheme 1).
Perovskite oxides are mixed metal compounds with a well-defined cubic structure and the general formula ABO3, where A and B are various cations, while the B cation is surrounded by an octahedron of O anions [17]. Perovskite-type layered oxide represents solid crystalline species consisting of two-dimensional nanosized plates of perovskite alternating with cations or cationic structural elements. Oxides of this type exhibit ion exchange and intercalation properties including reversible hydration of interlayer space [18,19]. These materials are attractive as components of composites due to the possibility of rearrangement of the metal sections, as well as their stability at elevated temperatures [20,21,22,23].
Currently, inorganic membranes based on perovskites are widely used for oxygen purification in catalytic membrane reactors and there are several attempts to use exfoliated perovskite nanosheets as membrane material [24]. However, there is no data on the use of perovskites as modifiers for polymer membranes for pervaporation. Perovskite-type layered oxides seem promising materials for incorporation into membranes due to combination of two factors: (i) the possibility of creating selective transport channels through the interlayer space of the perovskite structure—the presence of ionic compounds in the polymer matrix leads to an increase in the solubility of polar molecules and a decrease in the solubility of nonpolar media, and (ii) the possibility of increasing the distance between polymer chains, which should increase the permeability of the membranes due to the frame of inorganic materials that can prevent the tight packing of polymer chains [25,26,27].
This work aims to develop a method for manufacturing PAI/Per composites and dense film membranes based on them. The new composite membrane developed in this work is intended for the membrane process of pervaporation, which is used to separate liquid mixtures, including azeotropic, closely boiling, and heat-sensitive liquids [28]. The study on pervaporation separation of the ethanol–ethyl acetate mixture is of particular interest [29,30,31] because these liquids have close boiling points (EtOH–78.3 °C, EtOAc–77.0 °C) and form an azeotropic mixture of 26 wt.% ethanol and 74 wt.% ethyl acetate [32]. The impact of perovskites on morphology and physical properties are thoroughly discussed.

2. Materials and Methods

2.1. Materials

PAI Torlon® 4000TF, Mn = 30000 g/mole, Tc = 285 °C, (Solvay Specialty Polymers, Brussel, Belgium) and N-methyl pyrrolidone (NMP) were obtained from Aldrich; ethanol (EtOH) and ethyl acetate (EtOAc) were purchased from Vekton (Vekton, Saint Petersburg, Russia). The substances were of chemically pure (CP) grade and used without further purification

2.2. Synthesis of Perovskite

The perovskite K2La2Ti3O10 (Per) was synthesized by a hydrothermal method from commercially available precursors. The starting materials were La2O3, 99.9% (Vekton, Saint Petersburg, Russia), TiO2 P-25 (Evonik, Essen, Germany) and KOH, 99.9% (Vekton, Saint Petersburg, Russia) as both precursor and reaction media. The synthesis was performed in optimized conditions based on the method described earlier [33]. The stoichiometric amounts of TiO2 and La2O3 calculated for 250 mg of the product were weighed and suspended in 40 mL of 4 M KOH. The suspension was sonicated for 15 min and placed in a 50 mL Teflon-lined vessel. The vessel was then inserted into a stainless steel autoclave and heated at 230 °C. After the 72 h reaction period the vessel was naturally cooled in the air for ~3 h and the precipitate was separated by centrifugation. The obtained slurry material was washed once with 40 mL of distilled water and twice with 40 mL of ethanol and dried in air overnight.
As K2La2Ti3O10 is known to undergo partial substitution of interlayer potassium cations by protons in water solutions and humid air atmosphere [34], to obtain the stable phase product for membrane preparation the as-synthesized K2La2Ti3O10·yH2O powder was maintained in distilled water for 24 h forming a partly substituted stable form.

2.3. Perovskite Characterization

The phase compositions characterization was studied by powder XRD-analysis using a Rigaku MiniFlex II diffractometer with Cu Kα radiation source in the range of 3–60°, and at a scan speed of 10° min−1. The results of XRD-analysis show the formation of a well-crystallized, single-phase compound after the hydrothermal procedure (see Figure S1).
The morphologies of the obtained samples were carried out by scanning electron microscopy (SEM) (Zeiss Merlin, Oxford Instruments).

2.4. Membrane Preparation

The PAI/Per composites were prepared by mixing solutions of 3 wt.% PAI in NMP and Per in 10 mL of NMP in different amounts to obtain 1–3 wt.% Per in the resulting membranes. The produced solution was sonicated by ultrasound disperser Hielscher UP200St with a 7 mm probe (Hielscher Ultrasonics, Teltow, Germany) for 20 min at 40% amplitude. Then it was coated on a glass plate and left in an oven at 50 °C for a solvent evaporation. In 72 h, membranes were peeled off and placed in a vacuum oven at 60 °C to constant weight. The thicknesses of the resulted membranes were found to be ~33 μm.

2.5. Membrane Characterizations

Membrane morphology was observed using scanning electron microscope SEM Zeiss SUPRA 55VP (Carl Zeiss AG, Oberkochen, Germany). Before the test samples were coated with a platinum 20 nm layer by the Q150R S Plus coater (Quorum Technologies Ltd., Lewes, UK).
The platinum layer of 20 nm thick was coated on the sample surface by cathode sputtering using the Quorum 150 (Great Britain) installation.
Energy dispersive microanalysis (EDS) for study on the elemental composition of samples was performed using a system of microanalysis INCA Energy with an XMax 80 OXFORD detector. Spectra of film surface were examined for phase identification in a sample. Quantitative analysis was performed using the method of fundamental parameters. The standard spectrum accumulation time was 60 s. To determine small amounts of elements, the accumulation time was increased to 300 s.
The contact angle measurements were carried out by a DSA 10-drop shape analyzer (KRÜSS GmbH, Hamburg, Germany) at room temperature and atmospheric pressure using water and n-propanol. Based on measured contact angles, the critical surface tension (σs) including dispersion σ s d and polar σ s p components were calculated by the Owens–Wendt method [35]:
σ S = σ S d + σ S p
Membrane density (ρ) was measured by an immersing a sample (0.05–0.10 g) in chloroform–isopropanol solutions of various compounds in which the membrane remains suspended. Experiments were performed five times with at least two samples.
Fractional free volumes, FFV, of the PAI and composites PAI/Per were calculated by the following equations, respectively [36,37]:
F F V = ( V 0 1.3 V w ) / V 0
F F V c = 1 ρ c ( 1 w t P e r % 100 ) · ( 1 F F V ) ρ P A I 1 ρ c ρ P e r w t P e r % 100
where V0 = 1/ρ is the polymer specific volume and Vw is the van der Waals volume of the PAI macromolecule estimated via Askadskii’s group contribution method [38], ρPAI, ρPer, and ρC are densities of polymer, perovskite, and composites, respectively, and wtPer is the weight fraction of the Per modifier in the polymer matrix.

2.6. Pervaporation

The pervaporation experiment was carried out on a laboratory-scale unit and the details of the apparatus have been described in [39].
The pervaporation results were described in terms of total flux and separation factor. The total flux (J) was determined as the amount of liquid penetrated through membrane area per time interval:
J = Q S · t
where Q is the total weight of the permeate collected at time interval t and S is the effective surface area of the membrane.
The separation factor (α) was calculated with the following equation:
α = y / ( 1 y ) x / ( 1 x )
where x and y are the ethanol concentration in the feed and in the permeate, respectively.

3. Results

3.1. Membrane Structure

The morphology of the PAI dense films modified with different amounts of perovskite was studied by scanning electron microscopy (SEM). In Figure 1 micrographs of the cross-section of PAI/Per membranes containing 0, 1, 2, and 3 wt.% Per are presented. The inclusion of perovskite into the PAI changes the film cross-section images—the morphology of the membrane is complicated. Particular attention is drawn to the film structure inhomogeneous in thickness and a significant difference in morphology near the top and bottom surfaces of the membrane.
The top and bottom surfaces of the PAI/Per membranes (Figure 2b–d) differed markedly: a uniform top and a two-component bottom surfaces. The morphology of the top surface of all membranes remains identical while the morphology of the bottom surface changes after modification by perovskites. Figure 2e shows SEM images of as-prepared K2La2Ti3O10 oxide particles which have a plate-like morphology typical for layered compounds with lateral sizes of about 2–5 µm and ~4–5 nm thickness. According to Figure 2b–d), the sizes of the perovskite inclusions increase and plate profiles of perovskite inclusions (similar to initial oxide particle morphology) appear.
The elemental composition of the samples was investigated by energy dispersive analysis. It was observed that PAI/Per membranes made of composite materials have different compositions of the top and bottom surfaces. Table 1 shows the data of electron dispersion spectra (EDS) of the elemental analysis of the PAI/Per (2%) membrane along lines (1–5) of the cross-section near the top and bottom surfaces of the membrane. Figure 3 presents a micrograph of the cross-section of the PAI/Per (2%) membrane, where the position of these (1–5) spectral lines is indicated. For all PAI/Per membranes, the EDS of the top surface contains elements (C, N, O) that constitute the chemical formula of PAI. The composition of the bottom surface, in addition to reduced amounts (C, N, O), contains elements K, Ti, and La, which are components of the perovskite. Thus, the data of Table 1 and Figure 3 indicate the nonuniform distribution of perovskite in the PAI matrix along with the depth of the membrane. The spectrum along line 1 (top) does not contain perovskite, which appears in spectrums 3 and 4, and spectrum 5 (bottom) contains the maximum amount of perovskite elements. The filler is concentrated on one side of the membrane, in a layer ~3 µm thick. This may be due to a significant difference in the density of the membrane components: PAI ~1.38 g/cm3; Per ~4.0 g/cm3. When preparing the casting solution using an ultrasonic treatment, a uniform transparent PAI/Per (2%) solution is formed. However, after the solution coating to the glass plate, during the evaporation of NMP at 50 °C for 3–4 h, perovskite diffuses to the bottom under the action of gravity.
A similar EDS analysis of the cross-section was carried out for membranes containing also 1 and 3 wt.% perovskite. The same tendency for an increased concentration of perovskite elements was observed on the bottom of all membranes; and in the membrane containing 3 wt.% Per, the concentration of these elements was the highest.
Figure 4 shows the results of the study on surfaces enriched in perovskite for membranes with 0, 1, 2, and 3 wt.% Per. With an increase of the perovskite content in the membrane, the concentration of C and N elements, which are included only in the PAI formula, decreases (Figure 4a); and vice versa, the concentration of K, Ti, and La elements that are components of perovskite increases (Figure 4b). In this case, the concentration of oxygen (O), which is a common element for the both PAI and perovskite, remains unchanged, during decreasing (with the polymer) and increasing (with the filler).
It was found that PAI/Per membranes have an extraordinary nonsymmetrical structure, unexpected for a manufacturing process such as casting on the plate. The thin perovskite-enriched layer (3–5 μm) is probably bonded by hydrogen and coordination links the polymer component (PAI) of the membrane (~30 μm), as evidenced by the mixed composition of the intermediate layers.

3.2. Surface Tension Properties

The most common characteristic of solid surface tension is the contact angle. Table 2 presents the contact angles of water and n-propanol on the surfaces of the PAI/Per membranes containing 0, 1, 2, and 3 wt.% Per. PAI is a rather hydrophilic polymer since the water contact angle on its surface is less than 90°. It should be noted that all measurements were performed on the surface of PAI/Per membranes enriched with perovskite. The inclusion of perovskite reduces the contact angles of water and alcohol for the membranes.
The data on contact angles were used for calculation of the critical surface tension of PAI/Per membranes. The polar (σps) and dispersion (σds) contributions to the critical surface tension (σs) were estimated separately. Figure 5 illustrates the dependence of the critical surface tension and its polar and dispersion components on perovskite content in the membrane. As the loading of perovskites was increased, the polar contribution increased but the dispersion contribution decreases. The critical surface tension (σs) increases in all cases. Thus, the inclusion of perovskite resulted in a hydrophilization of the composite membrane.
Fraction free volume (FFV) was calculated based on data from the densities of PAI/Per membranes measured by flotation method. FFV characterizes the elements of the free volume distributed in the polymer matrix. Their presence facilitates the diffusion of molecules of separating mixtures through polymer films and influence on flux in membrane processes. Figure 6 shows FFV values calculated for the PAI/Per films by Equation (3). The FFVs of the composites are higher than that of PAI. This can be referred to a reorganization of polymer chains because of Per incorporation and formation having of not-so-dense macromolecule packaging in the composite membranes as against the PAI membrane. This fact will play a significant role in the study of transport properties of the membranes.

3.3. Transport Properties

For nonsymmetrical PAI/Per membranes, transport properties were studied in the pervaporation of a two-component ethanol‒ethyl acetate mixture. Separation of this mixture may be of independent interest since these liquids form an azeotropic mixture of 26 wt.% ethanol and 74 wt.% ethyl acetate [25].
In addition, this separation task arises during the manufacture of ethyl acetate by esterification (ethanol + acetic acid ↔ ethyl acetate + water) and the manufacture of butyl acetate by transesterification (ethyl acetate + n-butanol ↔ butyl acetate + ethanol). Considering the peculiarity of the ethyl acetate synthesis, to isolate the target component it is necessary to separate the ethyl acetate from ethanol impurities. In the butyl acetate production, ethanol and ethyl acetate act as by-products and subsequent regeneration of these important organic solvents is required to make them reusable.
Furthermore, the separation of the alcohol‒ester system is interesting in view of the assessment of the possibility to shift a chemical equilibrium in the four-component system that occurs during the chemical reaction. This can be done by isolation of the low molecular weight component (alcohol) and thereby increasing the yield of the target product.
Table 3 lists some physical properties of the studied liquids. Ethanol and ethyl acetate have close boiling points and significantly different density, volume, and solubility parameters. Such a characteristic as a solubility parameter, δ, is often used to predict the interaction of a polymer and an organic solvent. In accordance with the solubility theory [40,41], the solubility of a liquid in a polymer is greater, the smaller the difference in their solubility parameters |Δδ|. The solubility parameter δ for pure PAI Torlon® is equal to 23.0 (J/cm3)1/2. Thus, the solubility of ethanol in the studied membranes should be higher than the solubility of esters.
Figure 7 shows the dependences of ethanol concentration in the permeate on ethanol concentration in the feed for the pervaporation of ethanol‒ethyl acetate mixture by using PAI and PAI/Per (2%) membranes. The behavior of the permeate concentration versus feed for pervaporation is significantly different from the vapor-liquid equilibrium curve (VLE) for the ethanol-ethyl acetate mixture. In pervaporation, the permeate is enriched with ethanol, and ethanol content in the permeate increases with ethanol concentration in the feed. For each composition of the feed, the position of PAI/Per (2%) membrane is slightly lower than that of the PAI membrane (88 and 86 wt.% ethanol in permeate for PAI and PAI/Per (2%) correspondingly in pervaporation of the azeotropic mixture).
Figure 8 demonstrates the main transport properties of PAI and PAI/Per (2%) membranes: total flux and separation factor (αEtOH/EtOAc), as well as their dependence on the ethanol concentration in the feed and on the modifier content in the membrane.
From the data in Figure 8 it follows that the total flux of the membranes increases with the growth of ethanol concentration in the feed, while the flux is higher up to 1.5 times for PAI/Per (2%) membrane as compared with the unmodified PAI membrane. An opposite relationship was obtained for the separation factor (αEtOH/EtOAc) which decreases with the growth of ethanol concentration in the feed. According to αEtOH/EtOAc calculated using data on component concentrations in the feed and permeate, the selectivity of the membrane containing 2% Per is slightly lower compared with the unmodified PAI membrane.
Transport properties of membranes developed in this work were compared with literature data (not numerous) for the pervaporation of the ethanol‒ethyl acetate mixture. Table 4 lists data on total flux and separation factors that have been obtained by the use of different membranes [23,24,34]. It should be noted that No 3 is nonpolymeric membranes [24]. Membranes from polydimethylsiloxane (PDMS) exhibit high total flux and a small separation factor [30]. The separation of ethanol‒ethyl acetate mixture is a difficult task. According to Table 4, separation factors of known membranes are equal to (1.1–3.25) however the results of our work are superior to these data. The separation factor of the ethanol‒ethyl acetate mixture, even of the azeotropic composition, is very high for the PAI and PAI/Per (2%) membranes and equal to 21 and 17, respectively. The total flux is higher for PAI/Per (2%) as compared with the unmodified PAI membrane.

4. Conclusions

A method for producing novel composite membranes by inclusion of up to 3 wt.% perovskite-type layered oxide K2La2Ti3O10 into the Torlon® PAI matrix was developed. This method consists of mixing solutions of these components in NMP, sonication, casting on a glass plate, followed by evaporation of the solvent.
The membrane structure was studied by SEM in combination with energy dispersive analysis of the elemental composition. It was found that PAI/Per membranes have an extraordinary nonsymmetrical structure, unexpected for a manufacturing process such as casting on the plate; and a thin perovskite-enriched layer (3–5 μm) combined with the polymer component of the membrane (PAI) (~30 μm).
The creation of a layer containing perovskite in the composite membrane leads to hydrophilization of the membrane surface, as evidenced by a decrease in the contact angles for water and alcohol with an increase in the perovskite content in the membrane.
The transport properties of the nonsymmetrical PAI/Per membrane were studied during the pervaporation of ethanol‒ethyl acetate mixture and compared with literature data. It was found that PAI/Per 2% membrane has a high separation factor, mainly removing ethanol, even from a mixture of azeotropic composition. The inclusion of perovskite in the Torlon® PAI membrane leads to an increase of the total flux.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-8994/12/7/1142/s1, Figure S1: XRD patterns for as-prepared K2La2Ti3O10 (KLT3) and the product after 24 h water treatment HKLT3.

Author Contributions

Membrane preparation, physicochemical investigations, analysis of transport properties in pervaporation were carried out by V.R., A.P., and G.P. Synthesis and investigation of perovskites was conducted by I.M. and O.S. N.S. held SEM and EDS analysis of the membranes. Writing—review & editing, A.P., V.R., I.M., O.S., M.T., N.S. and G.P. All authors have read and agreed to the published version of the manuscript.

Funding

Membrane formation, investigation of structure and transport properties of the membranes were funded by the Russian Science Foundation (RSF) [grant 18-79-10116]. M.T., A.P., and V.R. also thank the Russian Foundation for Basic Research [grant 18-33-20138] for financial support of studies of the physical parameters and vapor-liquid equilibrium of separating liquids. Layered perovskite oxide was synthesized under grant [18-03-00915] of Russian Foundation for Basic Research. V.R. thanks the Russian Foundation for Basic Research [grant 19-33-90048] for financial support of layered perovskite oxide -polymers composites formation studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baker, R.W. Membrane Technology and Applications, 3rd ed.; Wiley: Chichester, UK, 2012; ISBN 9780470743720. [Google Scholar]
  2. Genduso, G.; Luis, P.; Van der Bruggen, B. Pervaporation membrane reactors (PVMRs) for esterification. In Membrane Reactors for Energy Applications and Basic Chemical Production; Elsevier Inc.: Amsterdam, The Netherlands, 2015; pp. 565–603. ISBN 9781782422273. [Google Scholar]
  3. Buonomenna, M.G. Design Next Generation Membranes or Rethink the “Old” Asymmetric Membranes? Symmetry (Basel) 2020, 12, 270. [Google Scholar] [CrossRef] [Green Version]
  4. Buonomenna, M.G. Energy Storage and Conversion Technologies: The Nanotechnology Role in their Advances. Recent Pat. Nanotechnol. 2017, 11. [Google Scholar] [CrossRef] [PubMed]
  5. Pulyalina, A.Y.; Polotskaya, G.A.; Toikka, A.M. Membrane materials based on polyheteroarylenes and their application for pervaporation. Russ. Chem. Rev. 2016, 85. [Google Scholar] [CrossRef]
  6. Kreiter, R.; Wolfs, D.; Engelen, C.; Vanveen, H.; Vente, J. High-temperature pervaporation performance of ceramic-supported polyimide membranes in the dehydration of alcohols. J. Memb. Sci. 2008, 319, 126–132. [Google Scholar] [CrossRef] [Green Version]
  7. Pulyalina, A.Y.; Toikka, A.M.; Polotskaya, G.A. Investigation of pervaporation membranes based on polycarbamide: Effect of residual solvent. Pet. Chem. 2014, 54, 573–579. [Google Scholar] [CrossRef]
  8. Pulyalina, A.Y.; Polotskaya, G.A.; Kalyuzhnaya, L.M.; Sushchenko, I.G.; Meleshko, T.K.; Yakimanskii, A.V.; Chislov, M.V.; Toikka, A.M. Sorption and transport of aqueous isopropanol solutions in polyimide-poly(aniline-co-anthranilic acid) composites. Russ. J. Appl. Chem. 2011, 84, 840–846. [Google Scholar] [CrossRef]
  9. Sun, S.P.; Wang, K.Y.; Peng, N.; Hatton, T.A.; Chung, T.S. Novel polyamide-imide/cellulose acetate dual-layer hollow fiber membranes for nanofiltration. J. Memb. Sci. 2010, 363, 232–242. [Google Scholar] [CrossRef]
  10. Peng, N.; Chung, T.S. The effects of spinneret dimension and hollow fiber dimension on gas separation performance of ultra-thin defect-free Torlon® hollow fiber membranes. J. Memb. Sci. 2008, 310, 455–465. [Google Scholar] [CrossRef]
  11. Wang, Y.; Jiang, L.; Matsuura, T.; Chung, T.S.; Goh, S.H. Investigation of the fundamental differences between polyamide-imide (PAI) and polyetherimide (PEI) membranes for isopropanol dehydration via pervaporation. J. Memb. Sci. 2008, 318, 217–226. [Google Scholar] [CrossRef]
  12. Wang, Y.; Goh, S.H.; Chung, T.S.; Na, P. Polyamide-imide/polyetherimide dual-layer hollow fiber membranes for pervaporation dehydration of C1-C4 alcohols. J. Memb. Sci. 2009, 326, 222–233. [Google Scholar] [CrossRef]
  13. Teoh, M.M.; Chung, T.S.; Wang, K.Y.; Guiver, M.D. Exploring Torlon/P84 co-polyamide-imide blended hollow fibers and their chemical cross-linking modifications for pervaporation dehydration of isopropanol. Sep. Purif. Technol. 2008, 61, 404–413. [Google Scholar] [CrossRef] [Green Version]
  14. Yoshikawa, M.; Higuchi, A.; Ishikawa, M.; Guiver, M.D.; Robertson, G.P. Vapor permeation of aqueous 2-propanol solutions through gelatin/Torlon® poly(amide imide) blended membranes. J. Memb. Sci. 2004, 243, 89–95. [Google Scholar] [CrossRef] [Green Version]
  15. Higuchi, A.; Yoshikawa, M.; Guiver, M.D.; Robertson, G.P. Vapor permeation and pervaporation of aqueous 2-propanol solutions through the Torlon® poly(amide imide) membrane. Sep. Sci. Technol. 2005, 40, 2697–2707. [Google Scholar] [CrossRef] [Green Version]
  16. Adoor, S.G.; Manjeshwar, L.S.; Bhat, S.D.; Aminabhavi, T.M. Aluminum-rich zeolite beta incorporated sodium alginate mixed matrix membranes for pervaporation dehydration and esterification of ethanol and acetic acid. J. Memb. Sci. 2008, 318, 233–246. [Google Scholar] [CrossRef]
  17. Nair, S.P.N.; Murugavel, P. Oxides: Their Properties and Uses. In Comprehensive Inorganic Chemistry II (Second Edition): From Elements to Applications; Elsevier Ltd.: Amsterdam, The Netherlands, 2013; Volume 4, pp. 47–72. ISBN 9780080965291. [Google Scholar]
  18. Uppuluri, R.; Sen Gupta, A.; Rosas, A.S.; Mallouk, T.E. Soft chemistry of ion-exchangeable layered metal oxides. Chem. Soc. Rev. 2018, 47, 2401–2430. [Google Scholar] [CrossRef]
  19. Schaak, R.E.; Mallouk, T.E. Perovskites by design: A toolbox of solid-state reactions. Chem. Mater. 2002, 14, 1455–1471. [Google Scholar] [CrossRef]
  20. Vente, J.F.; Haije, W.G.; Rak, Z.S. Performance of functional perovskite membranes for oxygen production. J. Memb. Sci. 2006, 276, 178–184. [Google Scholar] [CrossRef]
  21. Song, Z.; Zhang, Z.; Zhang, G.; Liu, Z.; Zhu, J.; Jin, W. Effects of polymer binders on separation performance of the perovskite-type 4-bore hollow fiber membranes. Sep. Purif. Technol. 2017, 187, 294–302. [Google Scholar] [CrossRef]
  22. Guironnet, L.; Geffroy, P.M.; Tessier-Doyen, N.; Boulle, A.; Richet, N.; Chartier, T. The surface roughness effect on electrochemical properties of La0.5Sr0.5Fe0.7Ga0.3O3-Δ perovskite for oxygen transport membranes. J. Memb. Sci. 2019, 588, 117199. [Google Scholar] [CrossRef]
  23. Prajongtat, P.; Sriprachuabwong, C.; Wongkanya, R.; Dechtrirat, D.; Sudchanham, J.; Srisamran, N.; Sangthong, W.; Chuysinuan, P.; Tuantranont, A.; Hannongbua, S.; et al. Moisture-Resistant Electrospun Polymer Membranes for Efficient and Stable Fully Printable Perovskite Solar Cells Prepared in Humid Air. ACS Appl. Mater. Interfaces 2019, 11, 27677–27685. [Google Scholar] [CrossRef]
  24. Nakagawa, K.; Yamashita, H.; Saeki, D.; Yoshioka, T.; Shintani, T.; Kamio, E.; Kreissl, H.T.; Tsang, S.C.E.; Sugiyama, S.; Matsuyama, H. Niobate nanosheet membranes with enhanced stability for nanofiltration. Chem. Commun. 2017, 53, 7929–7932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Avagimova, N.V.; Pulyalina, A.Y.; Toikka, A.M.; Suvorova, O.M.; Vilesov, A.D.; Polotskaya, G.A. Controlling the barrier properties of polymer composites containing montmorillonite. Pet. Chem. 2013, 53, 559–563. [Google Scholar] [CrossRef]
  26. Gao, C.; Zhang, M.; Jiang, Z.; Liao, J.; Xie, X.; Huang, T.; Zhao, J.; Bai, J.; Pan, F. Preparation of a highly water-selective membrane for dehydration of acetone by incorporating potassium montmorillonite to construct ionized water channel. Chem. Eng. Sci. 2015, 135, 461–471. [Google Scholar] [CrossRef]
  27. Choudhari, S.K.; Kariduraganavar, M.Y. Development of novel composite membranes using quaternized chitosan and Na+-MMT clay for the pervaporation dehydration of isopropanol. J. Colloid Interface Sci. 2009, 338, 111–120. [Google Scholar] [CrossRef]
  28. Roy, S.; Singha, N. Polymeric Nanocomposite Membranes for Next Generation Pervaporation Process: Strategies, Challenges and Future Prospects. Membranes (Basel) 2017, 7, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Sato, K.; Sugimoto, K.; Nakane, T. Separation of ethanol/ethyl acetate mixture by pervaporation at 100–130 °C through NaY zeolite membrane for industrial purpose. Microporous Mesoporous Mater. 2008, 115, 170–175. [Google Scholar] [CrossRef]
  30. Hasanoǧlu, A.; Salt, Y.; Keleşer, S.; Özkan, S.; Dinçer, S. Pervaporation separation of ethyl acetate-ethanol binary mixtures using polydimethylsiloxane membranes. Chem. Eng. Process. Process Intensif. 2005, 44, 375–381. [Google Scholar] [CrossRef]
  31. Ong, Y.T.; Tan, S.H. Pervaporation separation of a ternary azeotrope containing ethyl acetate, ethanol and water using a buckypaper supported ionic liquid membrane. Chem. Eng. Res. Des. 2016, 109, 116–126. [Google Scholar] [CrossRef]
  32. PD. CRC Handbook of Chemistry and Physics. J. Mol. Struct. 1992, 268, 320. [Google Scholar] [CrossRef]
  33. Huang, Y.; Wu, J.; Wei, Y.; Lin, J.; Huang, M. Hydrothermal synthesis of K2La2Ti3O10 and photocatalytic splitting of water. J. Alloys Compd. 2008, 456, 364–367. [Google Scholar] [CrossRef]
  34. Rodionov, I.A.; Mechtaeva, E.V.; Burovikhina, A.A.; Silyukov, O.I.; Toikka, M.A.; Zvereva, I.A. Effect of protonation on the photocatalytic activity of the K2La2Ti3O10 layered oxide in the reaction of hydrogen production. Monatshefte für Chemie-Chem. Mon. 2018, 149, 475–482. [Google Scholar] [CrossRef]
  35. Owens, D.K.; Wendt, R.C. Estimation of the surface free energy of polymers. J. Appl. Polym. Sci. 1969, 13, 1741–1747. [Google Scholar] [CrossRef]
  36. Boyer, R.F. Physical properties of molecular crystals, liquids and glasses. J. Polym. Sci. Part A-1 Polym. Chem. 1969, 7, 2466. [Google Scholar] [CrossRef]
  37. Zhang, H.; Wang, S.; Weber, S.G. Morphology and free volume of nanocomposite Teflon AF 2400 films and their relationship to transport behavior. J. Memb. Sci. 2013, 443, 115–123. [Google Scholar] [CrossRef]
  38. Askadskii, A.A. Computational Materials Science of Polymers; Cambridge International Science Publishing Conditions: Cambridge, UK, 2003; ISBN 1898326622. [Google Scholar]
  39. Pulyalina, A.; Tataurov, M.; Faykov, I.; Rostovtseva, V.; Polotskaya, G. Polyimide Asymmetric Membrane vs. Dense Film for Purification of MTBE Oxygenate by Pervaporation. Symmetry (Basel) 2020, 12, 436. [Google Scholar] [CrossRef] [Green Version]
  40. Belmares, M.; Blanco, M.; Goddard, W.A.; Ross, R.B.; Caldwell, G.; Chou, S.H.; Pham, J.; Olofson, P.M.; Thomas, C. Hildebrand and hansen solubility parameters from molecular dynamics with applications to electronic nose polymer sensors. J. Comput. Chem. 2004, 25, 1814–1826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Barton, A.F.M. CRC Handbook of Solubility Parameters and Other Cohesion Parameters; CRC Press: Boca Raton, FL, USA, 1991; ISBN 9780849301766. [Google Scholar]
  42. Knozowska, K.; Li, G.; Kujawski, W.; Kujawa, J. Novel heterogeneous membranes for enhanced separation in organic-organic pervaporation. J. Memb. Sci. 2020, 599, 117814. [Google Scholar] [CrossRef]
Scheme 1. (a) Crystal structure of perovskite-type layered oxide K2La2Ti3O10. (b) Formulas of Torlon® polyamide-imide (PAI).
Scheme 1. (a) Crystal structure of perovskite-type layered oxide K2La2Ti3O10. (b) Formulas of Torlon® polyamide-imide (PAI).
Symmetry 12 01142 sch001
Figure 1. SEM micrographs of membrane cross-section: (a) PAI, (b) PAI/Per (1%), (c) PAI/Per (2%), and (d) PAI/Per (3%). PAI: polyamide-imide; Per: perovskite-type layered oxide K2La2Ti3O10.
Figure 1. SEM micrographs of membrane cross-section: (a) PAI, (b) PAI/Per (1%), (c) PAI/Per (2%), and (d) PAI/Per (3%). PAI: polyamide-imide; Per: perovskite-type layered oxide K2La2Ti3O10.
Symmetry 12 01142 g001
Figure 2. SEM micrographs of the top and bottom surfaces of membranes: (a) PAI, (b) PAI/Per (1%), (c) PAI/Per (2%), (d) PAI/Per (3%), (e) as-prepared K2La2Ti3O10 oxide.
Figure 2. SEM micrographs of the top and bottom surfaces of membranes: (a) PAI, (b) PAI/Per (1%), (c) PAI/Per (2%), (d) PAI/Per (3%), (e) as-prepared K2La2Ti3O10 oxide.
Symmetry 12 01142 g002aSymmetry 12 01142 g002b
Figure 3. SEM micrographs of PAI/Per (2%) cross-section, where the position of 1–5 spectral lines is indicated.
Figure 3. SEM micrographs of PAI/Per (2%) cross-section, where the position of 1–5 spectral lines is indicated.
Symmetry 12 01142 g003
Figure 4. Change in the concentration of chemical elements (a) C, N, O, and (b) K, Ti, La on the bottom surface, depending on the perovskite content (0%, 1%, 2%, and 3%) in the PAI/Per membrane.
Figure 4. Change in the concentration of chemical elements (a) C, N, O, and (b) K, Ti, La on the bottom surface, depending on the perovskite content (0%, 1%, 2%, and 3%) in the PAI/Per membrane.
Symmetry 12 01142 g004
Figure 5. Dependence of the surface tension on perovskite content in the membrane.
Figure 5. Dependence of the surface tension on perovskite content in the membrane.
Symmetry 12 01142 g005
Figure 6. Dependence of fractional free volume (FFV) on perovskite content in membrane.
Figure 6. Dependence of fractional free volume (FFV) on perovskite content in membrane.
Symmetry 12 01142 g006
Figure 7. Dependence of ethanol concentration in the permeate on ethanol concentration in the feed for pervaporation using PAI and PAI/Per (2%) membranes; the vapor-liquid equilibrium (VLE) curve for ethanol‒ethyl acetate mixture, 40 °C.
Figure 7. Dependence of ethanol concentration in the permeate on ethanol concentration in the feed for pervaporation using PAI and PAI/Per (2%) membranes; the vapor-liquid equilibrium (VLE) curve for ethanol‒ethyl acetate mixture, 40 °C.
Symmetry 12 01142 g007
Figure 8. Dependence of the total flux and separation factor (αEtOH/EtOAc) on the ethanol concentration in the feed for pervaporation of the ethanol‒ethyl acetate mixture using PAI and PAI/Per (2%) membranes, 40 °C.
Figure 8. Dependence of the total flux and separation factor (αEtOH/EtOAc) on the ethanol concentration in the feed for pervaporation of the ethanol‒ethyl acetate mixture using PAI and PAI/Per (2%) membranes, 40 °C.
Symmetry 12 01142 g008
Table 1. EDS elemental analysis of the cross-section near the top and bottom surfaces of the PAI/Per (2%) membrane.
Table 1. EDS elemental analysis of the cross-section near the top and bottom surfaces of the PAI/Per (2%) membrane.
SpectrumCNOKTiLaTotal
wt.%wt.%wt.%wt.%wt.%wt.%wt.%
Spectrum 1 (Top)77.018.0814.910.000.000.00100.0
Spectrum 275.059.0615.790.100.000.00100.0
Spectrum 374.499.0715.660.170.000.60100.0
Spectrum 477.105.8514.930.300.731.09100.0
Spectrum 5 (Bottom)73.395.9316.630.421.212.42100.0
Table 2. Contact angles of PAI/Per membranes.
Table 2. Contact angles of PAI/Per membranes.
MembraneContact Angle, Degree
Watern-Propanol
PAI79.4 ± 0.614.7 ± 0.3
PAI/Per (1%)77.1 ± 0.713.2 ± 0.2
PAI/Per (2%)76.8 ± 0.512.7 ± 0.2
PAI/Per (3%)73.4 ± 0.411.5 ± 0.1
Table 3. Physical properties of liquids under the study.
Table 3. Physical properties of liquids under the study.
LiquidMol. Weight,
g/mol
Density,
g/cm3
Molar Volume,
cm3/mol
Tboiling,
°C
Hildebrand Solubility Parameter,
δ (J/cm3)1/2
Ethanol46.070.78957.578.425.9
Ethyl acetate88.110.90197.877.118.5
Table 4. Comparison of transport properties of the present membranes with literature data on pervaporation of the ethanol‒ethyl acetate azeotropic mixture. PDMS: Polydimethylsiloxane.
Table 4. Comparison of transport properties of the present membranes with literature data on pervaporation of the ethanol‒ethyl acetate azeotropic mixture. PDMS: Polydimethylsiloxane.
NoMembraneEthanol in the Feed,
wt.%
Total Flux, g/m2hSeparation
Factor
Ref.
1PDMS2048841.5[30]
2BP-SILM-7030703[31]
3PEBAX60 1.1[42]
4PDMS60 3.25[42]
5PERVAP 406060 2[42]
6PAI262021Present work
7PAI/Per(2%)262817Present work
PDMS – polydimethylsiloxane; BP-SILM-70 – buckypaper supported ionic liquid membrane; PEBAX – polyether block amide; PERVAP 4060 – commercial composite membrane based on polysiloxane.

Share and Cite

MDPI and ACS Style

Pulyalina, A.; Rostovtseva, V.; Minich, I.; Silyukov, O.; Toikka, M.; Saprykina, N.; Polotskaya, G. Specific Structure and Properties of Composite Membranes Based on the Torlon® (Polyamide-imide)/Layered Perovskite Oxide. Symmetry 2020, 12, 1142. https://doi.org/10.3390/sym12071142

AMA Style

Pulyalina A, Rostovtseva V, Minich I, Silyukov O, Toikka M, Saprykina N, Polotskaya G. Specific Structure and Properties of Composite Membranes Based on the Torlon® (Polyamide-imide)/Layered Perovskite Oxide. Symmetry. 2020; 12(7):1142. https://doi.org/10.3390/sym12071142

Chicago/Turabian Style

Pulyalina, Alexandra, Valeriia Rostovtseva, Iana Minich, Oleg Silyukov, Maria Toikka, Nataliia Saprykina, and Galina Polotskaya. 2020. "Specific Structure and Properties of Composite Membranes Based on the Torlon® (Polyamide-imide)/Layered Perovskite Oxide" Symmetry 12, no. 7: 1142. https://doi.org/10.3390/sym12071142

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