Next Article in Journal
Efficient Separation of a Novel Microbial Chassis, Vibrio natriegens, from High-Salt Culture Broth Using Ceramic Ultrafiltration Membranes
Next Article in Special Issue
Applications of Reverse Osmosis and Nanofiltration Membrane Process in Wine and Beer Industry
Previous Article in Journal
Changes in Tubular PVDF Membrane Performance During Initial Period of Pilot Plant Operation
Previous Article in Special Issue
Efficiency of an Ultrafiltration Process for the Depollution of Pretreated Olive Mill Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ultrasound on Dissolution of Polymeric Blends and Phase Inversion in Flat Sheet and Hollow Fiber Membranes for Ultrafiltration Applications

by
Gilberto Katmandú Méndez-Valdivia
1,
María De Lourdes Ballinas-Casarrubias
2,*,
Guillermo González-Sánchez
3,
Hugo Valdés
4,
Efigenia Montalvo-González
1,
Martina Alejandra Chacón-López
1,
Emmanuel Martínez-Montaño
5,
Beatriz Torrestiana-Sánchez
6,
Herenia Adilene Miramontes-Escobar
1 and
Rosa Isela Ortiz-Basurto
1,*
1
Laboratorio Integral de Investigación en Alimentos, Tecnológico Nacional de México-Instituto Tecnológico de Tepic, Tepic 63175, Mexico
2
Facultad de Ciencias Químicas, Universidad Autónoma de Chihuahua, Chihuahua 31125, Mexico
3
Departamento de Medio Ambiente y Energía, Centro de Investigación en Materiales Avanzados, Chihuahua 31136, Mexico
4
Centro de Innovación en Ingeniería Aplicada, Departamento de Computación e Industrias, Universidad Católica de Maule, Talca 3460000, Chile
5
Maestría en Ciencias Aplicadas, Unidad Académica de Ingeniería en Biotecnología, Universidad Politécnica de Sinaloa, Mazatlan 82199, Mexico
6
Unidad de Investigación y Desarrollo de Alimentos, Tecnológico Nacional de México-Instituto Tecnológico de Veracruz, Veracruz 91897, Mexico
*
Authors to whom correspondence should be addressed.
Membranes 2025, 15(4), 120; https://doi.org/10.3390/membranes15040120
Submission received: 1 March 2025 / Revised: 19 March 2025 / Accepted: 6 April 2025 / Published: 10 April 2025
(This article belongs to the Special Issue Membrane Processes for Water Recovery in Food Processing Industries)

Abstract

:
In seeking alternatives for reducing environmental damage, fabricating filtration membranes using biopolymers derived from agro-industrial residues, such as cellulose acetate (CA), partially dissolved with green solvents, represents an economical and sustainable option. However, dissolving CA in green solvents through mechanical agitation can take up to 48 h. An ultrasonic probe was proposed to accelerate mass transfer and polymer dissolution via pulsed interval cavitation. Additionally, ultrasound-assisted phase inversion (UAPI) on the external coagulation bath was assessed to determine its influence on the properties of flat sheet and hollow fiber membranes during phase inversion. Results indicated that the ultrasonic pulses reduced dissolution time by up to 98% without affecting viscosity (3.24 ± 0.06 Pa·s), thermal stability, or the rheological behavior of the polymeric blend. UAPI increased water permeability in flat sheet membranes by 26% while maintaining whey protein rejection above 90%. For hollow fiber membranes, UAPI (wavelength amplitude of 0 to 20%) improved permeability by 15.7% and reduced protein retention from 90% to 70%, with MWCO between 68 and 240 kDa. This report demonstrates the effectiveness of ultrasonic probes for decreasing the dissolution time of dope solution with green cosolvents and its potential to change the structure of polymeric membranes by ultrasound-assisted phase inversion.

1. Introduction

Filtration using polymeric membranes has a broad range of applications in science and industry due to its low energy consumption, continuous separation capability, and ease of scalability, among other advantages [1,2]. The base polymer of membranes defines key properties, such as hydrophobicity, surface charge, and selectivity [3]. Most polymeric membranes are made from polymers like polyethersulfone (PES) [4], polysulfone (PSF) [5], polyvinylidene fluoride (PVDF) [6], or polypropylene (PP) [7], as well as cellulose [8] and its derivatives, such as cellulose acetate (CA) [9]. CA membranes offer distinct advantages, including high hydrophilicity, which facilitates the separation of water-soluble compounds, maintains significant flux, and reduces fouling propensity. These benefits are complemented by good mechanical strength and thermal stability [10,11,12]. Our focus is on producing membranes made from cellulose acetate derived from natural sources. Several studies have investigated the synthesis of cellulose triacetate from various natural biomass sources, highlighting its potential as a valuable material. Meireles et al. [13] successfully produced cellulose acetate from mango seeds, achieving a high degree of substitution (DS) of 2.65. Similarly, Das et al. [14] synthesized cellulose acetate from rice husks using various methods, obtaining a notable DS. Another study demonstrated that solid waste from the olive industry could serve as an important source of cellulose powder and its derivatives [15]. Soto-Salcido et al. [16] focused on extracting cellulose acetate from agave bagasse for film production, showcasing the versatility of this biomass source. Additionally, Villanueva-Solís et al. [17] worked with oak sawdust to obtain cellulose acetates, further broadening the range of potential feedstocks for cellulose derivative synthesis.
The primary method for the elaboration of ultrafiltration polymeric membranes is non-solvent-induced phase separation (NIPS) [18], using a homogeneous solution (dope) composed of the base polymer and polymeric additives, such as methylcellulose, polyvinyl pyrrolidone (PVP), and polyethylene glycol (PEG); this could dissolve in water during the NIPS process (a pore-forming agent) or remain in the membrane matrix (hydrophilic and antifouling agent) or could have both roles [19]. The most commonly used solvents for preparing the dope solution with cellulose acetate by NIPS include N-methyl-2-pyrrolidone (NMP), dimethylacetamide (DMAc), dimethyl sulfoxide (DMSO), and N,N-dimethylformamide (DMF) [20]. However, these solvents pose significant health and environmental risks. Consequently, replacing them with non-toxic, biodegradable green solvents is essential [21]. Candidates for green solvents include γ-butyrolactone, γ-valerolactone, methyl acetate, triethyl phosphate, and glycerol derivatives like glycerol diacetate (diacetin) and glycerol triacetate (triacetin) have been suggested [22,23,24,25]. Unfortunately, green solvents often dissolve polymer blends and “dope” solutions less efficiently, as their Hildebrand and Hansen solubility parameters are significantly lower than those of conventional solvents. This means that the solubility parameters of green solvents and polymers are more different, resulting in lower cohesion energy (thermodynamic affinity) [24,25,26].
Various methods, including ultrasound, can increase the polymer’s dissolution rate in the solvent under standard temperature and pressure conditions. High-power ultrasound produces mechanical, acoustic, thermal, and chemical effects through cavitation, enhancing polymer-solvent contact and accelerating the dissolution rate [27,28]. The mechanical effects of cavitation generate high temperatures (up to 1000 °K), localized shear forces, microjets (up to 100 m/s), and shockwaves (up to 30 MPa) [29]. Cavitation involves the violent collapse of bubbles, resulting in molecular fragmentation, ultrasonic diffusion, and enhanced mass transfer [30], which accelerates the dissolution of homogeneous polymer blends for membrane production [31]. This is particularly important given cavitation’s reported effects on viscosity, polymerization, and potentially irreversible polymer chain degradation [32,33,34]. The ability of ultrasonic baths to dissolve polymers, disperse additives in dope solutions, and degas them to avoid imperfections during membrane manufacturing has been reported. Table 1 presents the effects of the US on the dissolution of polymeric mixtures and the production of flat membranes by US-assisted phase inversion (UAPI) in an ultrasonic bath. However, it has also been reported that ultrasonic probes can depolymerize cellulose, causing a reduction in viscosity and alteration of solution properties [35]. It is, therefore, of interest to verify that applying a high-power ultrasonic probe to reduce the dissolution time of polymeric mixtures does not affect their properties that could change the properties of the membrane.
The dope solution is spread or extruded by NIPS, inducing phase inversion and membrane formation into a non-solvent coagulation bath, typically water [42]. The active layer structure depends on system thermodynamics [43]. Studies using ultrasound in the dope solution and membrane phase inversion with other materials are limited and presented in Table 1. Therefore, studying the impact of ultrasound on a polymer blend based on cellulose acetate, where DMF is partially replaced by diacetin, is essential for evaluating rheological and thermogravimetric properties. This is particularly important given cavitation’s ultrasound probe-reported effects on viscosity, polymerization, and potentially irreversible polymer chain degradation [32,34].
This study aims to (1) evaluate the use of a high-power ultrasonic probe to dissolve a polymer blend containing CA, PEG, DMF, and diacetin to reduce dissolution time and verify the stability of the properties of the polymer mixture. And (2) assess the ultrasound-assisted phase inversion on the membrane structure, permeability, and changes in static (structure, MWCO) or dynamic (permeability and selectivity) properties of flat sheet and hollow fiber membranes for ultrafiltration applications.

2. Materials and Methods

2.1. Materials

The polymeric mixture was prepared using commercially available CA (acetyl content 39.8 wt%, Mw = 30,000 g/mol, Sigma Aldrich, St. Louis, MO, USA) and PEG (PEG 6000, Mw = 6000 g/mol, Sigma Aldrich, St. Louis, MO, USA) as a polymeric additive for membrane pore formation. Two solvents were used for polymer dissolution: diacetin (CH3COOCH2)2CHOH (50% technical-grade diacetin, Mw = 176.17 g/mol, Sigma Aldrich, St. Louis, MO, USA) as a green solvent and DMF (Mw = 73.09 g/mol, J.T. Baker, Phillipsburg, NJ, USA) as a cosolvent.

2.2. Dissolution of the Polymeric Mixture by Stirring and Ultrasound

Five treatments were conducted to compare dissolution times. For stirring-based dissolution, a solution containing 12.5% CA (previously dried at 100 °C for 24 h), 42.25% DMF, 42.25% diacetin, and 5% PEG was heated to 55 °C until complete dissolution was observed. A factorial design was applied for ultrasound-assisted dissolutions, evaluating two methods using identical proportions and two ultrasonic amplitudes. The first method involved dissolving PEG with the solvents before adding CA, while the second dissolved all components simultaneously. A stainless-steel ultrasonic probe with a diameter of 0.5 and a length of 2.5 inches was used, connected to a high-power (700 W) ultrasonic generator QSonica-700 (Newtown, CT, USA) operated at 20 kHz for both methods, with amplitudes of 50% and 100%, applied in pulsed intervals of 30 s/min until total dissolution of polymer agglomerates was achieved (Figure A1).

2.3. Characterization of the Polymeric Mixture

2.3.1. Viscosity and Rheological Behavior

As Miramontes-Escobar [1] reported, apparent viscosity was determined using a Discovery DHR 1 hybrid rheometer (TA Instruments, New Castle, DE, USA) with a 40 mm serrated cone-plate geometry at 25 °C. Dynamic rheological studies were performed on mixtures dissolved via ultrasound, involving a strain sweep from 1% to 100% deformation at a fixed frequency within the linear viscoelastic region. Angular frequency sweeps from 1 to 100 rad/s were conducted to evaluate viscoelastic behavior by monitoring the storage modulus (G′) and loss modulus (G″). Flow behavior was assessed by measuring viscosity as a function of shear rate from 1 to 100 s−1.
The flow behavior was analyzed using the Cross Equation (1) [44].
η ( γ   ˙ ) = η 0 η η 0 1 λ γ ˙ n
where η ( γ   ˙ ) represents apparent viscosity (Pa·s), η is viscosity at infinite shear rate (Pa·s), η 0 is viscosity at zero shear rate (Pa·s), λ is the relaxation time (s), n is the flow index (dimensionless), and γ   ˙ is the shear rate (s−1).

2.3.2. Calorimetric Analysis of Polymeric Blends

Thermogravimetric analysis (TGA) was performed using a TGA-550 instrument (TA Instruments, New Castle, DE, USA) with sample weights ranging from 15 to 20 mg. Temperature was ramped from 25 °C to 800 °C at 10 °C/min under a nitrogen atmosphere [45].
Differential scanning calorimetry (DSC) was carried out using a DSC-250 instrument (TA Instruments, New Castle, DE, USA) with 10 mg samples. Temperature ranged from 25 °C to 400 °C at a heating rate of 10 °C/min under nitrogen.

2.4. Membrane Fabrication

2.4.1. Ultrasound-Assisted Flat Sheet Membranes

Membranes were produced using the non-solvent-induced phase separation (NIPS) method (Figure 1), following the procedure by Ballinas-Casarrubias et al. [46], with modifications for ultrasound-assisted phase inversion. Approximately 20 mL of dope polymeric mixture solution was deposited on the upper end of a glass plate (14 × 19 cm); the solution was immediately spread using a film extensor model 3580 (Elcometer, Manchester, UK) calibrated to 150 µm. The glass plate with the extended film of 150 µm was quickly placed on the surface of the sonicated coagulant bath (with water as a nonsolvent) for induced phase separation and membrane structure formation. The glass plate was immersed by gravity (in 1–2 s) in a stainless steel tank (40 cm diameter, 55 cm height, with a double jacket for temperature control at 25 °C) filled with 40 L of filtered water (non-solvent). The coagulation tank (Figure A2) contained 25 cm from the bottom a stainless-steel ultrasonic probe with a diameter of 0.5 and a length of 2.5 inches, connected to a high power (700 W) ultrasonic generator QSonica-700 (Newtown, CT, USA). A factorial design was applied, evaluating three ultrasonic amplitudes (25, 50, and 75%) at different times (1, 3, and 5 min). In all cases, the ultrasound was turned on just before spreading the soap solution film on the glass; the time was counted from the moment the glass was deposited on the surface of the coagulant bath. The US probe allows the glass plate to reach the bottom of the external coagulant tank. The membranes were subjected to two washes (24 h each), the first with water and the second with 30% glycerol, to prevent structural collapse. They were dried at 40 °C before being stored in plastic bags until characterization.

2.4.2. Ultrasound-Assisted Hollow Fiber Membranes

Hollow fiber membranes were prepared following Torrestiana-Sánchez et al. [47], with modifications mainly in ultrasound-assisted phase inversion (Figure 2). Twenty-five milliliters of the dope polymeric mixture liquid were placed in a sono-extruder (designed in the laboratory, see Figure A3) with internal and external coagulation temperatures set to 25 °C. The air gap was fixed at 15 cm, internal coagulant flow at 15 mL/min, and extrusion pressure at 0.5 bar. A one-way statistical design was applied, evaluating three ultrasonic amplitudes (5, 10, and 20%) with the ultrasonic probe (on during all the extrusion) placed 50 cm away. After extrusion, fibers were submerged in water for 24 h, in glycerol for 24 h to prevent structural collapse, and dried for 24 h at 40 °C. Membranes were stored in bags for subsequent characterization.

2.5. Characterization of the Manufactured Filtration Membranes

2.5.1. Internal Structure of the Membranes

The methodology described by Terrazas-Bandala et al. [48] was followed to determine the internal structure of the produced membranes. The membranes were exposed to liquid nitrogen, fractured, and coated with a thin gold layer using a Denton Desk-II Gatan (Moorestown, NJ, USA) coating system. Micrographs were captured at 10 kV using a scanning electron microscope (SEM SU3500, Hitachi High-Tech, Tokyo, Japan) in a cross-sectional view at the Advanced Materials Research Center.

2.5.2. Filtration Performance (Flux and Retention)

The filtration performance analysis was conducted following the methodology of Rahimpour and Madaeni [49] with minor modifications. A 2% whey protein solution was filtered at room temperature for 60 min through all obtained membranes. Permeate samples were collected every 5 min, and turbidity was measured using a Turbidimeter HACH (Model 2100Qis, HACH COMPANY, Guiping Road, Shanghai, China).
The flux (J) was calculated using Equation (2).
J = Q p A Δ P
where Q p represents the permeate flow (L), A is the membrane surface area (m2), and ΔP denotes the transmembrane pressure (bar). The results were expressed in L·h−1·m−2.
Protein retention capacity (R, %) was determined using Equation (3).
R = 1 N T U p N T U f × 100
where NTUp is the particle count in the permeate, and NTUf is the particle count in the feed solution.
The Volume Reduction Factor (VRF) for different membranes was evaluated using Equation (4).
V R F = V 0 V f V 0
where V0 is the initial sample volume (mL), and Vf is the filtrate volume (mL) [50].

2.5.3. Protein Retention by Electrophoretic Analysis

The electrophoretic analysis followed Laemmli’s method [51]. A stacking and separating gel (4% and 12% polyacrylamide, respectively) with sodium dodecyl sulfate (SDS) (Sigma Aldrich, St. Louis, MO, USA) was prepared. T0-A0 and T5-A75 permeate flat membrane samples were diluted (1:1) in the sample buffer and boiled for 3 min. Aliquots of 10–15 μL at a 2 mg/mL concentration were loaded onto the gel. Electrophoresis was performed using Tris-HCl buffer (Sigma Aldrich, St. Louis, MO, USA) (0.025 M, pH 8.8), glycine (Sigma Aldrich, St. Louis, MO, USA) (0.192 M), and SDS (0.1%) at 150 V for 30 min and 200 V for 50 min. The gel was stained with Coomassie Brilliant Blue R-250 (0.1% w/v) (Sigma Aldrich, St. Louis, MO, USA) in methanol-acetic acid (40% and 10% v/v) (Sigma Aldrich, St. Louis, MO, USA) for 24 h and destained in the same solution. Molecular weights were determined by applying five μL of a standard compared to a molecular weight marker (10–250 kDa) (SDS-PAGE Standards, Bio-Rad, Hercules, CA, USA).

2.5.4. MWCO Determination by HPLC

The molecular weight cut-off (MWCO) in hollow fiber membranes was determined by HPLC. The proteins in defatted and lyophilized samples (So and permeated solutions) were determined following the methodology reported by González-Felix et al. [52] with some modifications. A Varian™ Pro Star (Varian Inc., Walnut Creek, CA, USA) high-pressure liquid chromatography equipped with a diode array detector (DAD) and Galaxy™ software version 1.9.302.952 was used for the analysis. Each sample was dissolved (20 mg/mL) in a mobile phase of 150 mM sodium phosphate buffer, pH 7, at 25 °C. A peptide (catalog 151-1901; BIORAD, Hercules, CA, USA) containing five known MW compounds (thyroglobulin, 670 kDa; gamma-globulin, 158 kDa; ovalbumin, 44 kDa; myoglobin, 17 kDa; and vitamin B12, 1.35 kDa) was used. A 20 µL aliquot was injected at an isocratic flow of 0.4 mL/min in a size exclusion column (Bio Sec-5TM, 4.6 x 300 mm; Agilent, Santa Clara, CA, USA), and the absorbance was monitored at a wavelength of 254 nm. The membrane’s molecular weight cut-off (MWCO) determination was performed by comparing the area under the curve determined in the chromatogram of the fed solution (So) and the 10% in the permeate obtained from each of the membranes.

2.6. Statistical Analysis

The results were analyzed using ANOVA with a significance value of p < 0.05 with Statistica v12 software. All analyses were performed in triplicate.

3. Results and Discussion

3.1. Dissolution of Polymer Mixture by Agitation and Ultrasound

3.1.1. Dissolution Time and Viscosity of Polymer Blends

The dissolution time (min) and apparent viscosity (Pa·s) for different ultrasound intensities (W/cm2) are presented in Table 2. Ultrasound intensity, which measures energy applied per unit area, increases with amplitude, affecting dissolution efficiency. Various ultrasound intensities were used for comparative analysis.
The use of ultrasound reduced the effective dissolution time of the polymer blend from two days (48 h) under stirring to 35 min using ultrasonic pulses with an intensity between 55 and 60 W/cm2, regardless of the dissolution order, and to 15 min at an intensity between 118 and 123 W/cm2. This reduction is attributed to increased temperature and the effects of cavitation, such as acoustic streaming, microstreamers, microjets, and microstreaming, which enhance physical agitation and collisions within the container, generating micro vortices that promote improved mass transfer and diffusion. This process enables better solvent penetration into the polymers, significantly reducing dissolution time [53]. Other researchers reported similar findings. Lan et al. [31] used ultrasound at 20 to 75 W and 110 °C for 5 to 20 min to dissolve 2% cellulose in an ionic liquid (C4mim) Cl, reducing the dissolution time from 190 to 60 min. The improvement was attributed to ultrasound effects, such as cavitation and mechanical impact, which enhanced mass transfer between the ionic liquid and cellulose. Despite Ávila-Orta et al. [54] noting that increased ultrasonic intensity affects polymer viscosity, the apparent viscosity of the different samples in this study showed no significant differences (p < 0.05) compared to non-sonicated samples, indicating the polymer blend’s stability under applied ultrasonic intensities.

3.1.2. Rheological Behavior of Polymer Blends

Frequent sweep tests with 25% deformation were conducted on all mixtures to confirm that ultrasound did not alter the polymer blend. Figure 3 shows storage modulus (G′) and loss modulus (G″) as functions of angular frequency.
The results showed that G″ predominated over G′, a characteristic of viscoelastic liquids [55]. G′ represented the energy stored within the blend during deformation, which increased with more intense interactions, while G″ represented dissipated energy as heat during deformation, allowing the most fluidity [56]. The predominance of G″ over G′ was primarily due to the compound’s functionality, where cellulose acetate’s -COCH3 groups prevent interaction with PEG’s OH-groups, reducing intermolecular interactions; consequently, the blend stored less energy, making G′ lower than G″ [57]. Additionally, G″ increased with rising angular frequency, reflecting increased friction between polymers, proportional to viscosity [58]. This behavior was consistent in all sonicated samples, confirming that ultrasound enhanced mass transfer and accelerated polymer dissolution without breaking functional groups or altering the blend’s behavior [28].
To determine the viscosity behavior under applied shear stress (Newtonian or Non-Newtonian), a flow curve (see Figure 4) was generated for the different treatments. Flow curves were used to examine the viscosity behavior under shear stress. Comparing sonicated and stirred blends revealed Newtonian behavior at shear rates below 100 s−1, while high rates showed a viscosity decrease, indicating shear-thinning behavior.
To confirm this without sample degradation, data were fitted to the Cross model, which accounted for Newtonian and shear-thinning behaviors. This polymer viscosity model allows the description of a Newtonian fluid, a shear-thinning fluid, or a shear-thickening fluid based on the parameter “n” [59].
Table 3 presents the fitted model constants using the TRIOS version 3 software (TA Instruments, New Castle, DE, USA). The viscosity trend η 0   >   η matched Figure 4, where initial viscosity values decreased with shear rate. Relaxation time (λ) values indicated low relaxation times for all blends, suggesting uniform recovery after stress removal [60]. All blends exhibited shear-thinning behavior (n < 1), confirming that viscosity decreased with increasing shear rate; this confirmed the observation in Figure 4, where an increase in shear rate tended to result in a decrease in apparent viscosity [61]. The Cross model confirmed shear-thinning behavior with low relaxation times, which is favorable for hollow fiber membrane extrusion processes due to enhanced spinning properties [62]. Overall, ultrasound-treated blends exhibited rheological behavior comparable to stirred blends, confirming no adverse effects on the polymer mixture.

3.1.3. Calorimetric Analysis: TGA and DSC of Polymer Blends

Calorimetric analyses were performed on polymer blends treated with maximum ultrasonic intensity (120 W/cm2) to assess molecular stability (see Figure 5). DSC curves (Figure 5A) revealed two endothermic transitions for all samples (Table 4). The first transition corresponded to the glass transition temperature (Tg), with similar values across blends: PEG-CA (165.94 °C), PEG-CA-100% (166.79 °C), and PEG-100%-CA (165.66 °C). Statistical analysis showed no significant effect (α < 0.05) on Tg, confirming that ultrasound did not alter intermolecular interactions affecting Tg. Vinodhini et al. [63] reported a Tg of 165 °C for chitosan-CA-PEG-DMF blends, consistent with this study’s findings, confirming minimal ultrasound impact on polymer chain arrangement. The second DSC transition (Figure 5A) represents the melting temperature (Tm). Tm values varied slightly among blends: PEG-CA (216.61 °C), PEG-CA-100% (226.58 °C), and PEG-100%-CA (221.82 °C), suggesting ultrasound-induced polymer chain reorganization, increasing density and thermal stability [64]. Rodrigues et al. [65] reported a Tm close to 225 °C for commercial cellulose acetate, aligning with these results and indicating ultrasound promoted structural reorganization without degradation. On the other hand, TGA (Figure 5B) further evaluated polymer chain integrity and transition temperatures (Table 4), confirming ultrasound effects without compromising polymer stability. In addition to the DSC analysis, thermogravimetric analysis (TGA) was performed (Figure 5B) to study the polymeric chains and evaluate potential degradation at the transition temperature (Table 4), which may have been influenced by ultrasound. The first weight loss (80%) of the PEG-CA mixture occurred at 178 °C, whereas the mixtures subjected to ultrasound exhibited lower temperatures: 171 °C for the PEG-CA-100% sample and 163 °C for the PEG-100%-CA sample. This phenomenon is associated with the solvents, which have evaporation temperatures of 99 °C and 65 °C, respectively, values lower than the polymer degradation temperatures. The second weight loss (90%) was linked to the thermal resistance of PEG and CA, occurring at 320 °C for PEG-CA, 300 °C for PEG-CA-100%, and 290 °C for PEG-100%-CA. During the third loss (95%), carbonization was observed at 404, 396, and 395 °C, respectively. These differences, particularly in the second weight loss, suggested that ultrasound might have caused chain breakage, primarily due to homolytic scissions in polymeric chains and solvent oncolysis. These reactions promoted radical formation, which impacted the thermal stability of the polymeric mixture, causing the observed decrease [34]. However, previous studies [66,67] indicated that cellulose acetate remained stable between 260 °C and 450 °C, suggesting that ultrasound did not significantly affect the thermal stability of this polymer.

3.2. Effect of Ultrasound on the Elaboration of Flat Membranes

3.2.1. Flat Membranes Internal Structure Micrographs

Cross-sectional micrographs of flat membranes elaborated by ultrasound-assisted phase inversion under different US amplitudes (Figure 6) demonstrated that this method facilitated the formation of enlarged pores or anisotropic structures. This structure consisted of a dense layer providing membrane selectivity, followed by an extension of pores known as “finger-like” structures, which improve permeability. Membranes subjected to 25% and 50% US amplitudes for 5 min exhibited this structure. Huan et al. [38] showed that when flat PVDF membranes were produced using an ultrasonic bath with a power less than 60 W during phase inversion, macropore formation was reduced, and the tendency for finger-like structures with greater water flux, high porosity, and sizeable mean pore size of membranes was enhanced. This highlights that the applied power defines finger-like pore structures or teardrop-like voids.
Internal membrane structures exhibited more uniform cavity extensions at longer times and lower amplitudes, as seen in membrane T5-A25. However, increasing the amplitude caused more significant tortuosity in the support structure, observed in membranes treated at 75% amplitude. This effect was attributed to ultrasonic intensity generating higher vibrations and microjets that modified the structure, promoting irregular cavity formation. Furthermore, ultrasound enhanced mass transfer between solvents and water, accelerating phase inversion and solidification. This rapid solidification prevented pore-forming additives from creating finger-like structures in the support layer, altering the internal architecture [36].

3.2.2. Whey Protein Filtration Capabilities of Flat Membranes

The permeability behavior of membranes in a stirred tank with frontal filtration of a 2% whey protein solution at 1.5 bar pressure is illustrated in Figure 7. Initial permeability decline, characteristic of this technology, resulted from fouling layer formation on the membrane surface due to protein and ion (calcium, sodium, and zinc) adsorption. The flux reduction and consequently the permeability decrease correlated with an increasing volumetric reduction factor (VRF). Higher sodium concentrations increased protein size, exacerbating fouling [68]. However, membranes treated with ultrasound exhibited higher flux than untreated membranes (T0-A0).
Ultrasonically treated membranes exhibited permeability increases from 18.43% to 96.59%, while VRF increased from 4.46% to 15.14%, with significant differences (p < 0.05). This suggested that ultrasound-induced internal structural changes, such as increased extension pores and reduced macropores in the support layer. Increases in filtration performance in protein separation have been reported by reducing the tortuosity of the membrane due to the reduction in protein retention in the internal structure [69].
To assess whether increased permeability compromised molecular retention, turbidity reduction was measured using nephelometric turbidity units (NTU) (Table 5). The initial solution had a turbidity value of 903 NTU.
All flat membranes achieved over 97% turbidity reduction. This protein retention capacity was confirmed by gel electrophoresis (Figure A4), demonstrating effectiveness in retaining whey proteins with molecular weights between 60 and 250 kDa [70], which indicates that they all have a retention capacity of at least 60 kDa. That is, in the range for ultrafiltration separation. Huan et al. [38] reported that ultrasonically treated PVDF membranes exhibited improved filtration performance and molecular retention due to morphological changes. Under the evaluated conditions, no effect of ultrasound was observed on flat membranes’ filtration performance and retention capacity.

3.3. Effect of Ultrasound on the External Coagulant Bath in Hollow Fiber Membrane Fabrication

Due to technical challenges, it was not possible to replicate the experimental design to apply the same ultrasonic amplitude to both flat and hollow fiber membranes. Additionally, achieving homogeneous and well-defined membranes at amplitudes >20% proved difficult. As a result, the study of the impact of applied amplitude on the ultrasound-assisted phase inversion process for hollow fiber membranes was limited.

3.3.1. Hollow Fiber Membranes Internal Structure Micrographs

Scanning electron microscopy (SEM) micrographs (Figure 8) revealed the internal structure of hollow fiber membranes produced at different amplitudes. The internal structure comprised a thin active layer for separation, followed by macropores, and a second layer with external pores limited by macropores. This double-layer, finger-like structure results from rapid phase inversion using water as a non-solvent [71]. Ultrasound at 10% and 20% amplitudes expanded macropores, indicating morphological changes in hollow fiber membranes.
Cross-sections revealed higher ultrasound intensities correlated with larger macropores and reduced structural uniformity. The formation of larger macropores was directly proportional to the ultrasonic intensity applied. These alterations in the membranes exposed to ultrasound appear to be caused by the impact of sound waves on the internal structure, primarily due to the position at which they were received. The membrane subjected to 20% amplitude did not achieve concentricity, possibly due to the transition of the membrane from a liquid to a solid state during phase inversion, making it susceptible to external forces [72]. Porous materials absorb acoustic energy, causing wave collisions that enlarge macropores and displace coagulants [73]. This study is the first to apply ultrasound during hollow fiber membrane phase inversion, recommending ultrasonic application with a maximum intensity of 3.55 W/cm2 (20% amplitude).

3.3.2. Whey Protein Filtration Capabilities of Hollow Fiber Membranes

The membranes’ permeability varied according to the application of ultrasound; Figure 9 depicts filtration performance at 1 bar. Untreated membranes had low permeability (0.3 ± 0.01 L·h−1·m−2, VRF = 1.04). Ultrasonically treated membranes increased permeability as the amplitude rose, reaching 0.7 ± 0.02 L·h−1·m−2 and VRF = 1.11.
Increased filtration capacity was attributed to ultrasound-induced porosity and macrospace changes. Turbidity reduction post-filtration was measured (Table 6). Initial turbidity So (2% solution proteins) was 920 NTU. Ultrasound reduced protein retention from 90% (0% or control, 5%, 10% amplitude), while there was only 70% protein retention when applying an amplitude of 20%, where larger pores increased permeability but decreased retention. Control membranes exhibited higher flux stability and whey protein retention (90.57 ± 0.914), indicating smaller pore sizes relative to proteins and that the tangential filtration of hollow fiber membranes, unlike the cross-flow of flat membranes, could prevent the formation of any fouling [73,74,75]. In the A-20% membrane, a permeability loss of 42.42 ± 1.24% was observed, with no fouling formation, primarily due to the increase in pore size caused by ultrasound. This was confirmed by the low protein retention (70.08 ± 2.42%), allowing proteins to pass through easily without blocking the pores. Initially, the membrane showed higher flux, but as filtration progressed, some proteins and ions in milk protein solutions, such as sodium, zinc, and calcium, adhered, reducing permeability as reported by Wang et al. [68,76]. Given the excellent dispersion in the protein retention capacity of these hollow fiber membranes, HPLC analysis was performed to determine the MWCO of each membrane obtained.

3.3.3. Hollow Fiber MWCO

The membrane’s molecular weight cut-off (MWCO) determination was performed by comparing the area under the curve determined in the chromatogram (Figure 10) of the fed solution (So) and the 10% in the permeate obtained from each of the membranes in which ultrasonic amplitudes of 0, 5, 10, and 20% were applied during phase inversion. The retention time (RT) of 10% of the initial area of the protein fraction in the permeate was identified (corresponding to 90% retention). The MWCO of the membrane was estimated using the equation obtained from a calibration curve (LogMW = −0.6174 RT + 8.8441; r2 = 0.97) with the molecular weight of the markers (RT670 kDa = 5.05 min; RT158 kDa = 5.62 min; RT 44 kDa = 6.71 min; RT 17 kDa = 7.47 min; RT 1.35 = 110.69 min).
The chromatographic profiles (Figure 10) clearly show that the membrane without US treatment has higher retention, and as the ultrasonic amplitude in the external coagulant bath increases, lower retention of low MW proteins is observed, i.e., the MWCO of the membranes increases. This paper reports the effect of a range of ultrasonic amplitudes during ultrasound-assisted phase inversion using an ultrasonic probe, thereby expanding the technological bases to carry out studies aimed at the preparation of membranes with ultrasonic pulses during phase inversion. Further studies are being performed in our research group on this effect to expand this knowledge.

4. Conclusions

These results confirm the hypothesis about the efficient use of pulse using ultrasonic probes connected to high-power generators to accelerate the dissolution of polymer blends and improve the performance of phase inversion membranes. It was established in this system how to significantly reduce polymer dissolution time without compromising viscosity, thermal stability, or rheological behavior, which was beneficial for polymer dissolution using diacetin as a green co-solvent. Ultrasound-assisted phase inversion improved the ultrafiltration permeability of flat membranes and scattered the retention capacity of hollow fiber membranes, observing a direct relationship between the applied ultrasonic amplitude and the increase in MWCO. This technique effectively produced flat and hollow fiber membranes with MWCO for ultrafiltration applications, which allows the separation, purification, and/or concentration of compounds of interest in fluid foods and agro-industrial effluents such as proteins, peptides, and dyes, among others, which, if not removed, become environmental pollutants. However, further studies are still needed on this system, for example, on the effects of changing the MWCO of the additives on the structure and properties of the membranes, as well as the possible incorporation of composites or new biopolymers to produce functionalized membranes prepared by ultrasound-assisted phase inversion using ultrasonic probes with high-power generators. These perspectives are in order to achieve membranes with better properties and greater targeted applications.

Author Contributions

Conceptualization, M.D.L.B.-C. and R.I.O.-B.; formal analysis, G.K.M.-V., G.G.-S., E.M.-M. and H.A.M.-E.; funding acquisition, M.D.L.B.-C. and R.I.O.-B.; project administration, M.D.L.B.-C. and R.I.O.-B.; supervision, G.G.-S., E.M.-G., M.A.C.-L. and B.T.-S.; validation, H.V.; writing—original draft, G.K.M.-V.; writing—review and editing, G.G.-S., H.V., E.M.-G., M.A.C.-L., E.M.-M., B.T.-S., H.A.M.-E. and R.I.O.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Fondo Institucional para el Desarrollo Científico, Tecnológico y de Innovación” F/3578, FORDECYT-PRONACES, grant number 2558579 and the APC was funded by the Universidad Autónoma de Chihuahua “Más ciencia, más investigación”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contribution presented in the study is included in the article; further inquiries can be directed to the corresponding authors.

Acknowledgments

All authors thank CONAHCyT for the financial project 2558579 and the master scholarship to Gilberto Katmandú Méndez-Valdivia. The authors thank the Universidad Autónoma de Chihuahua, Centro de Investigación en Materiales Avanzados y TecNM-ITTepic for their support in the administration of project resources. To MMinoxidables y Servicios SA de CV (Guadalajara, Jalisco) for their support in the assembly of the sono-extruder. Special thanks to Erika Ivonne López-Martínez and Karla Campos-Venegas for their experimental contributions.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

CAcellulose acetate
PESpolyethersulfone
PSFpolysulfone
PVDFpolyvinylidene
PPpolypropylene
NIPSnon-solvent-induced phase separation
PVPpolyvinyl pyrrolidone
PEGpolyethylene glycol
MWCOmolecular weight cut-off
NMP N-methyl-2-pyrrolidone
DMAcdimethylacetamide
DMFN,N-dimethylformamide
TGAthermogravimetric analysis
DSCdifferential scanning calorimetry
SEMscanning electron microscope
J permeate flux (L·h−1·m−2)
Rprotein retention capacity
NTUnephelometric turbidity unit
VRFvolume reduction factor
SDSsodium dodecyl sulfate
G′storage modulus
G″loss modulus
Tgglass transition temperature
Tmmelting temperature
MWmolecular weight

Appendix A

Figure A1. Dope solution: (A) with agglomerates; (B) total dissolution of polymers.
Figure A1. Dope solution: (A) with agglomerates; (B) total dissolution of polymers.
Membranes 15 00120 g0a1
Figure A2. Phase inversion tank coupled with an ultrasonic probe.
Figure A2. Phase inversion tank coupled with an ultrasonic probe.
Membranes 15 00120 g0a2
Figure A3. Graphic representation of the sono-extrusion pilot used for the elaboration of hollow fiber membranes.
Figure A3. Graphic representation of the sono-extrusion pilot used for the elaboration of hollow fiber membranes.
Membranes 15 00120 g0a3
Figure A4. Electrophoresis conducted on the initial solutions and permeates of whey proteins, compared with a protein standard. Lane 1: Protein standard with a molecular weight (MW) of 10 to 250 kDa. Lane 2: Whey proteins in the initial solution. Lane 3: Permeate whey proteins from flat membranes with a 97% retention. Lane 4: Permeate whey proteins without ultrasound-assisted phase inversion and with 5 min of ultrasonic amplitude of 75%.
Figure A4. Electrophoresis conducted on the initial solutions and permeates of whey proteins, compared with a protein standard. Lane 1: Protein standard with a molecular weight (MW) of 10 to 250 kDa. Lane 2: Whey proteins in the initial solution. Lane 3: Permeate whey proteins from flat membranes with a 97% retention. Lane 4: Permeate whey proteins without ultrasound-assisted phase inversion and with 5 min of ultrasonic amplitude of 75%.
Membranes 15 00120 g0a4

References

  1. Miramontes-Escobar, H.A.; Hengl, N.; Dornier, M.; Montalvo-González, E.; Chacón-López, M.A.; Achir, N.; Vaillant, F.; Ortiz-Basurto, R.I. Coupling Low-Frequency Ultrasound to a Crossflow Microfiltration Pilot: Effect of Ultrasonic Pulse Application on Sono-Microfiltration of Jackfruit Juice. Membranes 2024, 14, 192. [Google Scholar] [CrossRef] [PubMed]
  2. Bera, S.P.; Godhaniya, M.; Kothari, C. Emerging and advanced membrane technology for wastewater treatment: A review. J. Basic Microbiol. 2022, 62, 245–259. [Google Scholar] [CrossRef] [PubMed]
  3. Othman, N.H.; Alias, N.H.; Fuzil, N.S.; Marpani, F.; Shahruddin, M.Z.; Chew, C.M.; Ismail, A.F. A review on the use of membrane technology systems in developing countries. Membranes 2021, 12, 30. [Google Scholar] [CrossRef] [PubMed]
  4. Dmitrieva, E.S.; Anokhina, T.S.; Novitsky, E.G.; Volkov, V.V.; Borisov, I.L.; Volkov, A.V. Polymeric membranes for oil-water separation: A review. Polymers 2022, 14, 980. [Google Scholar] [CrossRef]
  5. Mamah, S.C.; Goh, P.S.; Ismail, A.F.; Suzaimi, N.D.; Yogarathinam, L.T.; Raji, Y.O.; El-badawy, T.H. Recent development in modification of polysulfone membrane for water treatment application. J. Water Process. Eng. 2021, 40, 101835. [Google Scholar] [CrossRef]
  6. Saxena, P.; Shukla, P. A comprehensive review on fundamental properties and applications of poly (vinylidene fluoride) (PVDF). Adv. Compos. Hybrid Mater. 2021, 4, 8–26. [Google Scholar] [CrossRef]
  7. Guo, M.; Kanezashi, M. Recent progress in a membrane-based technique for propylene/propane separation. Membranes 2021, 11, 310. [Google Scholar] [CrossRef]
  8. Vatanpour, V.; Pasaoglu, M.E.; Barzegar, H.; Teber, O.O.; Kaya, R.; Bastug, M.; Khataee, A.; Koyuncu, I. Cellulose acetate in fabrication of polymeric membranes: A review. Chemosphere 2022, 295, 133914. [Google Scholar] [CrossRef]
  9. Goswami, K.P.; Pugazhenthi, G. Credibility of polymeric and ceramic membrane filtration in the removal of bacteria and virus from water: A review. J. Environ. Manag. 2020, 268, 110583. [Google Scholar] [CrossRef]
  10. Asiri, A.M.; Petrosino, F.; Pugliese, V.; Khan, S.B.; Alamry, K.A.; Alfifi, S.Y.; Marwani, H.M.; Alotaibi, M.M.; Algieri, C.; Chakraborty, S. Synthesis and Characterization of Blended Cellulose Acetate Membranes. Polymers 2022, 14, 4. [Google Scholar] [CrossRef]
  11. Wang, J.; Song, H.; Ren, L.; Talukder, M.E.; Chen, S.; Shao, J. Study on the preparation of cellulose acetate separation membrane and new adjusting method of pore size. Membranes 2021, 12, 9. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, S.; Tang, R.; Dai, Y.; Wang, T.; Zeng, Z.; Zhang, L. Fabrication of cellulose acetate membrane with advanced ultrafiltration performances and antibacterial properties by blending with HKUST-1@ LCNFs. Sep. Purif. Technol. 2021, 279, 119524. [Google Scholar] [CrossRef]
  13. Meireles, C.d.S.; Filho, G.R.; Fernandes Ferreira, M.; Cerqueira, D.A.; Assunção, R.M.N.; Ribeiro, E.A.M.; Poletto, P.; Zeni, M. 2010. Characterization of asymmetric membranes of cellulose acetate from biomass: Newspaper and mango seed. Carbohyd. Polym. 2010, 80, 954–961. [Google Scholar] [CrossRef]
  14. Das, A.M.; Ali, A.A.; Hazarika, M.P. Synthesis and characterization of cellulose acetate from rice husk: Eco-friendly condition. Carbohyd. Polym. 2014, 112, 342–349. [Google Scholar] [CrossRef]
  15. Hamed, O.A.; Jodeh, S.; Al-Hajj, N.; Hamed, E.M.; Abo-Obeid, A.; Fouad, Y. Cellulose acetate from biomass waste of olive industry. J. Wood Sci. 2015, 61, 45–52. [Google Scholar] [CrossRef]
  16. Soto-Salcido, L.A.; González-Sánchez, G.; Rocha-Gutierrez, B.; Peralta-Perez, R.; Zavala-Díaz, F.J.; Ballinas-Casarrubias, L. Preparation, characterization and performance of acetylated cellulignin membranes obtained by green methods from biomass. Desalination 2018, 430, 186–196. [Google Scholar] [CrossRef]
  17. Villanueva-Solís, L.A.; Ruíz-Cuilty, K.; Camacho-Dávila, A.; Espinoza-Hicks, J.C.; González-Sánchez, G.; Ballinas-Casarrubias, L. Lignocellulosic waste pretreatment and esterification using green solvents. Sep. Purif. Technol. 2020, 250, 117102. [Google Scholar] [CrossRef]
  18. Tibi, F.; Charfi, A.; Cho, J.; Kim, J. Fabrication of polymeric membranes for membrane distillation process and application for wastewater treatment: Critical review. Process Saf. Environ. Prot. 2020, 141, 190–201. [Google Scholar] [CrossRef]
  19. Nasrollahi, N.; Ghalamchi, L.; Vatanpour, V.; Khataee, A.; Yousefpoor, M. Novel polymeric additives in the preparation and modification of polymeric membranes: A comprehensive review. J. Ind. Eng. Chem. 2022, 109, 100–124. [Google Scholar] [CrossRef]
  20. Khan, B.; Zhan, W.; Lina, C. Cellulose acetate (CA) hybrid membrane prepared by phase inversion method combined with chemical reaction with enhanced permeability and good anti-fouling property. J. Appl. Polym. Sci. 2020, 137, 49556. [Google Scholar] [CrossRef]
  21. Jiang, S.; Ladewig, B.P. Green synthesis of polymeric membranes: Recent advances and future prospects. Curr. Opin. Green Sustain. Chem. 2020, 21, 1–8. [Google Scholar] [CrossRef]
  22. Hong, S.U.; Wang, Y.; Soh, L.S.; Yong, W.F. Are green solvents truly green? Integrating life cycle assessment and techno-economic analysis for sustainable membrane fabrication. Green Chem. 2023, 25, 4501–4512. [Google Scholar] [CrossRef]
  23. Mehrabani, S.A.N.; Vatanpour, V.; Koyuncu, I. Green solvents in polymeric membrane fabrication: A review. Sep. Purif. Technol. 2022, 298, 121691. [Google Scholar] [CrossRef]
  24. Rasool, M.A.; Vankelecom, I.F. Preparation of full-bio-based nanofiltration membranes. J. Membr. Sci. 2021, 618, 118674. [Google Scholar] [CrossRef]
  25. Zou, D.; Nunes, S.P.; Vankelecom, I.F.; Figoli, A.; Lee, Y.M. Recent advances in polymer membranes employing non-toxic solvents and materials. Green Chem. 2021, 23, 9815–9843. [Google Scholar] [CrossRef]
  26. Díaz de los Ríos, M.; Hernández Ramos, E. Determination of the Hansen solubility parameters and the Hansen sphere radius with the aid of the solver add-in of Microsoft Excel. SN Appl. Sci. 2020, 2, 1–7. [Google Scholar] [CrossRef]
  27. Prasad, R.; Dalvi, S.V. Sonocrystallization: Monitoring and controlling crystallization using ultrasound. Chem. Eng. Sci. 2020, 226, 115911. [Google Scholar] [CrossRef]
  28. Cai, B.; Mazahreh, J.; Ma, Q.; Wang, F.; Hu, X. Ultrasound-assisted fabrication of biopolymer materials: A review. Int. J. Biol. Macromol. 2022, 209, 1613–1628. [Google Scholar] [CrossRef]
  29. Petkovšek, M.; Kržan, A.; Šmid, A.; Žagar, E.; Zupanc, M. Degradation of water-soluble poly (vinyl alcohol) with acoustic and hydrodynamic cavitation: Laying foundations for microplastics. NPJ Clean Water 2023, 6, 35. [Google Scholar] [CrossRef]
  30. Dehane, A.; Merouani, S.; Hamdaoui, O.; Alghyamah, A. A comprehensive numerical analysis of heat and mass transfer phenomenons during cavitation sono-process. Ultrason. Sonochem. 2021, 73, 105498. [Google Scholar] [CrossRef]
  31. Lan, W.; Liu, C.F.; Yue, F.X.; Sun, R.C.; Kennedy, J.F. Ultrasound-assisted dissolution of cellulose in ionic liquid. Carbohydr. Polym. 2011, 86, 672–677. [Google Scholar] [CrossRef]
  32. Wang, X.; Majzoobi, M.; Farahnaky, A. Ultrasound-assisted modification of functional properties and biological activity of biopolymers: A review. Ultrason. Sonochem. 2020, 65, 105057. [Google Scholar] [CrossRef]
  33. Lupacchini, M.; Mascitti, A.; Giachi, G.; Tonucci, L.; d’Alessandro, N.; Martinez, J.; Colacino, E. Sonochemistry in non-conventional, green solvents or solvent-free reactions. Tetrahedron 2017, 73, 609–653. [Google Scholar] [CrossRef]
  34. McKenzie, T.G.; Karimi, F.; Ashokkumar, M.; Qiao, G.G. Ultrasound and sonochemistry for radical polymerization: Sound synthesis. Chem. Eur. J. 2019, 25, 5372–5388. [Google Scholar] [CrossRef] [PubMed]
  35. Haouache, S.; Karam, A.; Chave, T.; Clarhaut, J.; Amaniampong, P.N.; Fernandez, J.M.G.; Vigier, K.D.O.; Capron, I.; Jérôme, F. Selective radical depolymerization of cellulose to glucose induced by high frequency ultrasound. Chem. Sci. 2020, 11, 2664–2669. [Google Scholar] [CrossRef]
  36. Tao, M.M.; Liu, F.; Xue, L.X. Poly (vinylidene fluoride) membranes by an ultrasound assisted phase inversion method. Ultrason. Sonochem. 2013, 20, 232–238. [Google Scholar] [CrossRef]
  37. Qu, T.; Pan, K.; Li, L.; Liang, B.; Wang, L.; Cao, B. Influence of Ultrasonication Conditions on the Structure and Performance of Poly(vinylidene fluoride) Membranes Prepared by the Phase Inversion Method. Ind. Eng. Chem. Res. 2014, 53, 8228–8234. [Google Scholar] [CrossRef]
  38. Huan, Y.; Li, Z.; Li, C.; Li, G. Adsorption performances of Methylene blue by poly(vinylidene fluoride)/MWCNT membranes via ultrasound-assisted phase inversion method. Desal. Water Treat. 2019, 163, 83–95. [Google Scholar] [CrossRef]
  39. Ionita, M.; Crica, L.E.; Voicu, S.I.; Pandele, A.M.; Iovu, H. Fabrication of cellulose triacetate/graphene oxide porous membrane. Polym. Adv. Technol. 2016, 27, 350–357. [Google Scholar] [CrossRef]
  40. Kuzminova, A.; Dmitrenko, M.; Dubovenko, R.; Puzikova, M.; Mikulan, A.; Korovina, A.; Koroleva, A.; Selyutin, A.; Semenov, K.; Su, R.; et al. Development and Study of Novel Ultrafiltration Membranes Based on Cellulose Acetate. Polymers 2024, 16, 1236. [Google Scholar] [CrossRef]
  41. Ren, S.; Huang, G.; Yao, Y.; Zhang, P.; Zhang, Z.; Wang, Y. Integrating radical polymerization and non-solvent induced phase inversion strategy for functionalized ultrafiltration membrane fabrication. Desalination 2024, 573, 117220. [Google Scholar] [CrossRef]
  42. Ahmad, A.L.; Otitoju, T.A.; Ooi, B.S. Hollow fiber (HF) membrane fabrication: A review on the effects of solution spinning conditions on morphology and performance. J. Ind. Eng. Chem. 2019, 70, 35–50. [Google Scholar] [CrossRef]
  43. Karimi, A.; Khataee, A.; Vatanpour, V.; Safarpour, M. The effect of different solvents on the morphology and performance of the ZIF-8 modified PVDF ultrafiltration membranes. Sep. Purif. Technol. 2020, 253, 117548. [Google Scholar] [CrossRef]
  44. Xie, J.; Jin, Y.C. Parameter determination for the Cross rheology equation and its application to modeling non-Newtonian flows using the WC-MPS method. Eng. Appl. Comput. Fluid Mech. 2016, 10, 111–129. [Google Scholar] [CrossRef]
  45. Calderón-Chiu, C.; Calderón-Santoyo, M.; Barros-Castillo, J.C.; Díaz, J.A.; Ragazzo-Sánchez, J.A. Structural Modification of Jackfruit Leaf Protein Concentrate by Enzymatic Hydrolysis and Their Effect on the Emulsifier Properties. Colloids Interfaces 2022, 6, 52. [Google Scholar] [CrossRef]
  46. Ballinas, L.; Torras, C.; Fierro, V.; Garcia-Valls, R. Factors influencing activated carbon-polymeric composite membrane structure and performance. J. Phys. Chem. Solids 2004, 65, 633–637. [Google Scholar] [CrossRef]
  47. Torrestiana-Sánchez, B.; Ortiz-Basurto, R.I.; Brito-De La Fuente, E. Effect of nonsolvents on properties of spinning solutions and polyethersulfone hollow fiber ultrafiltration membranes. J. Membr. Sci. 1999, 152, 19–28. [Google Scholar] [CrossRef]
  48. Terrazas-Bandala, L.P.; Gonzalez-Sanchez, G.; Garcia-Valls, R.; Gumi, T.; Beurroies, I.; Denoyel, R.; Torras, C.; Ballinas-Casarrubias, L. Influence of humidity, temperature, and the addition of activated carbon on the preparation of cellulose acetate membranes and their ability to remove arsenic from water. J. Appl. Polym. Sci. 2014, 131. [Google Scholar] [CrossRef]
  49. Rahimpour, A.; Madaeni, S.S. Polyethersulfone (PES)/cellulose acetate phthalate (CAP) blend ultrafiltration membranes: Preparation, morphology, performance and antifouling properties. J. Membr. Sci. 2007, 305, 299–312. [Google Scholar] [CrossRef]
  50. Jang, H.; Kang, S.; Kim, J. Identification of Membrane Fouling with Greywater Filtration by Porous Membranes: Combined Effect of Membrane Pore Size and Applied Pressure. Membranes 2024, 14, 46. [Google Scholar] [CrossRef]
  51. Laemmli, U.K. Cleavage of structure proteins during the assembly of the head of bacteriophage T4. Nature 1970, 227, 680–685. [Google Scholar] [CrossRef] [PubMed]
  52. González-Félix, G.K.; Luna-Suárez, S.; García-Ulloa, M.; Martínez-Montaño, E.; Barreto-Curiel, F.; Rodríguez-González, H. Extraction methods and nutritional characterization of protein concentrates obtained from bean, chickpea, and corn discard grains. Curr. Res. Food Sci. 2023, 7, 100612. [Google Scholar] [CrossRef] [PubMed]
  53. Aliyu, M.; Hepher, M.J. Effects of ultrasound energy on degradation of cellulose material. Ultrason. Sonochem. 2000, 7, 265–268. [Google Scholar] [CrossRef] [PubMed]
  54. Ávila-Orta, C.; Espinoza-González, C.; Martínez-Colunga, G.; Bueno-Baqués, D.; Maffezzoli, A.; Lionetto, F. An overview of progress and current challenges in ultrasonic treatment of polymer melts. Adv. Polym. Technol. 2013, 32, E582–E602. [Google Scholar] [CrossRef]
  55. Ramli, H.; Zainal, N.F.A.; Hess, M.; Chan, C.H. Basic principle and good practices of rheology for polymers for teachers and beginners. Chem. Teach. Int. 2022, 4, 307–326. [Google Scholar] [CrossRef]
  56. Zhang, L.; Jiang, Z.; Yang, S.; Zeng, Z.; Zhang, W.; Zhang, L. Different rheological behaviours of cellulose/tetrabutylammonium acetate/dimethyl sulfoxide/water mixtures. Cellulose 2020, 27, 7967–7978. [Google Scholar] [CrossRef]
  57. Digaitis, R.; Wadsö, L.; Fredriksson, M.; Thybring, E.E. Effect of acetylation on wood-water interactions studied by sorption calorimetry. Cellulose 2024, 31, 7325–7334. [Google Scholar] [CrossRef]
  58. Utracki, L.A. Rheology of polymer blends. Encycl. Polym. Blends 2011, 27–108. [Google Scholar] [CrossRef]
  59. Osswald, T.A.; Rudolph, N. Polymer rheology: Fundamentals and Applications; Hanser Publications Springer: Munich, Germany, 2014. [Google Scholar] [CrossRef]
  60. Ebagninin, K.W.; Benchabane, A.; Bekkour, K. Rheological characterization of poly (ethylene oxide) solutions of different molecular weights. J. Colloid Interface Sci. 2009, 336, 360–367. [Google Scholar] [CrossRef]
  61. Turan, O.; Sachdeva, A.; Chakraborty, N.; Poole, R.J. Laminar natural convection of power-law fluids in a square enclosure with differentially heated side walls subjected to constant temperatures. J. Non-Newtonian Fluid Mech. 2011, 166, 1049–1063. [Google Scholar] [CrossRef]
  62. Mousavi, S.M.; Raveshiyan, S.; Amini, Y.; Zadhoush, A. A critical review with emphasis on the rheological behavior and properties of polymer solutions and their role in membrane formation, morphology, and performance. Adv. Colloid Interface Sci. 2023, 319, 102986. [Google Scholar] [CrossRef] [PubMed]
  63. Vinodhini, P.A.; Sangeetha, K.; Thandapani, G.; Sudha, P.N.; Jayachandran, V.; Sukumaran, A. FTIR, XRD and DSC studies of nanochitosan, cellulose acetate and polyethylene glycol blend ultrafiltration membranes. Int. J. Biol. Macromol. 2017, 104, 1721–1729. [Google Scholar] [CrossRef]
  64. Hu, W. The melting point of chain polymers. J. Chem. Phys. 2000, 113, 3901–3908. [Google Scholar] [CrossRef]
  65. Rodrigues Filho, G.; Monteiro, D.S.; da Silva Meireles, C.; de Assunção, R.M.N.; Cerqueira, D.A.; Barud, H.S.; Messadeq, Y. Synthesis and characterization of cellulose acetate produced from recycled newspaper. Carb. Polym. 2008, 73, 74–82. [Google Scholar] [CrossRef]
  66. Arthanareeswaran, G.; Thanikaivelan, P.; Srinivasn, K.; Mohan, D.; Rajendran, M. Synthesis, characterization and thermal studies on cellulose acetate membranes with additive. Eur. Polym. J. 2004, 40, 2153–2159. [Google Scholar] [CrossRef]
  67. Cavalcante, L.; de Alencar, A.E.V.; Mazzeto, S.E.; de A Soares, S. The effect of additives on the thermal degradation of cellulose acetate. Polym. Degrad. Stab. 2003, 80, 149–155. [Google Scholar] [CrossRef]
  68. Wang, W.Q.; Zhou, J.Y.; Li, J.J.; Cong-Cong, T. The Mechanism of Whey Protein on Membrane Surface Fouling During Ultrafiltration Process. Food Biophys. 2024, 19, 143–159. [Google Scholar] [CrossRef]
  69. Abdelrasoul, A.; Doan, H.; Lohi, A.; Cheng, C.H. Morphology control of polysulfone membranes in filtration processes: A critical review. ChemBioEng Rev. 2015, 2, 22–43. [Google Scholar] [CrossRef]
  70. Ali, A.; Ain, Q.; Saeed, A.; Khalid, W.; Ahmed, M.; Bostani, A. Bio-Molecular Characteristics of Whey Proteins with Relation to Inflammation; In New advances in the dairy industry. IntechOpen 2022. [Google Scholar] [CrossRef]
  71. Alsalhy, Q.F.; Rashid, K.T.; Noori, W.A.; Simone, S.; Figoli, A.; Drioli, E. Poly (vinyl chloride) hollow-fiber membranes for ultrafiltration applications: Effects of the internal coagulant composition. J. Appl. Polym. Sci. 2012, 124, 2087–2099. [Google Scholar] [CrossRef]
  72. Ruckdashel, R.R. The Role of Spinneret Configurations on Hollow Fiber Formation and Impacts of Hollow Fiber Geometry on Thermal and Acoustic Properties. Ph.D. Thesis, North Carolina State University, Raleigh, NC, USA, 2019; p. 27700918. [Google Scholar]
  73. Kuczmarski, M.A.; Johnston, J.C. Acoustic Absorption in Porous Materials; NASA Glenn Research Center: Cleveland, OH, USA, 2011. [Google Scholar]
  74. Pereira, G.L.D.; Cardozo-Filho, L.; Jegatheesan, V.; Guirardello, R. Generalization and expansion of the Hermia model for a better understanding of membrane fouling. Membranes 2023, 13, 290. [Google Scholar] [CrossRef] [PubMed]
  75. Coronel, M. Microfiltración Tangencial. Enfoque UTE [online] 2012, 3, 1–7. Available online: https://www.redalyc.org/articulo.oa?id=572260835001 (accessed on 25 January 2025). [CrossRef]
  76. Wang, Q.; Dai, F.; Zhang, S.; Chen, C.; Yu, Y. Fabrication of ultrafiltration membranes by poly (aryl ether nitrile) with poly (ethylene glycol) as additives. Water Sci. Technol. 2020, 82, 2847–2856. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagram of flat membrane fabrication by ultrasound-assisted phase inversion.
Figure 1. Diagram of flat membrane fabrication by ultrasound-assisted phase inversion.
Membranes 15 00120 g001
Figure 2. Graphic representation of ultrasound-assisted phase inversion in the elaboration of hollow fiber membranes.
Figure 2. Graphic representation of ultrasound-assisted phase inversion in the elaboration of hollow fiber membranes.
Membranes 15 00120 g002
Figure 3. Effect of ultrasound on the polymer blend. Storage modulus G′ (□) and loss modulus G″ (■).
Figure 3. Effect of ultrasound on the polymer blend. Storage modulus G′ (□) and loss modulus G″ (■).
Membranes 15 00120 g003
Figure 4. Dope solutions’ viscosity vs. shear rate curve for the treatments.
Figure 4. Dope solutions’ viscosity vs. shear rate curve for the treatments.
Membranes 15 00120 g004
Figure 5. (A) Differential scanning calorimetry (DSC) curves and (B) thermogravimetric analysis (TGA) curves for stirred and sonicated polymer blends to be used for the membrane’s preparation.
Figure 5. (A) Differential scanning calorimetry (DSC) curves and (B) thermogravimetric analysis (TGA) curves for stirred and sonicated polymer blends to be used for the membrane’s preparation.
Membranes 15 00120 g005aMembranes 15 00120 g005b
Figure 6. Micrographs evidencing the ultrasonic effect on the internal structure of flat membranes. T: time (ultrasound application in minutes); A: ultrasound amplitude (%).
Figure 6. Micrographs evidencing the ultrasonic effect on the internal structure of flat membranes. T: time (ultrasound application in minutes); A: ultrasound amplitude (%).
Membranes 15 00120 g006
Figure 7. Performance and volumetric reduction factor (VRF) during whey protein filtration with flat membranes subjected to different ultrasonic amplitudes. T: time (ultrasound application in minutes), percentage of applied ultrasonic amplitude (A) 25 (B) 50 (C) 75 %. VRF (□). Flux (●).
Figure 7. Performance and volumetric reduction factor (VRF) during whey protein filtration with flat membranes subjected to different ultrasonic amplitudes. T: time (ultrasound application in minutes), percentage of applied ultrasonic amplitude (A) 25 (B) 50 (C) 75 %. VRF (□). Flux (●).
Membranes 15 00120 g007
Figure 8. Micrographs of hollow fiber membranes with (A) 0% (control membrane); (B) 5%; (C) 10%; (D) 20% of amplitude ultrasound effect.
Figure 8. Micrographs of hollow fiber membranes with (A) 0% (control membrane); (B) 5%; (C) 10%; (D) 20% of amplitude ultrasound effect.
Membranes 15 00120 g008
Figure 9. Flux and volumetric reduction factor (VRF) comparisons for whey protein filtration with hollow fiber membranes. A: ultrasound amplitude (%).
Figure 9. Flux and volumetric reduction factor (VRF) comparisons for whey protein filtration with hollow fiber membranes. A: ultrasound amplitude (%).
Membranes 15 00120 g009
Figure 10. Chromatographic profiles of UV protein signal at 254 nm. (a) Protein solution fed to determine the molecular weight cut-off (MWCO); (b) 0%, (c) 5%, and (d) 20% of ultrasonic amplitude applied during phase inversion.
Figure 10. Chromatographic profiles of UV protein signal at 254 nm. (a) Protein solution fed to determine the molecular weight cut-off (MWCO); (b) 0%, (c) 5%, and (d) 20% of ultrasonic amplitude applied during phase inversion.
Membranes 15 00120 g010
Table 1. Effects of ultrasound on polymeric mixtures’ dissolution and flat membranes’ production by ultrasound-assisted phase inversion in ultrasonic baths.
Table 1. Effects of ultrasound on polymeric mixtures’ dissolution and flat membranes’ production by ultrasound-assisted phase inversion in ultrasonic baths.
Composition (Polymer/Additive/Solvent)Process USConditions (kHz/Hzmax/W/Wmax)Application TimeObservationsReference
PVDF/LiCl/DMFPhase inversion 40 kHz/40 kHz/180–300 W/300 W60 sImproves interdiffusion between solvent and nonsolvent, affecting membrane morphology and performance.[36]
PVDF/NMPPhase inversion 45, 80, 100 kHz/100 kHz/120–300 W/300 W20 minPromotes the formation of cellular pores and improves membrane porosity and permeability.[37]
PVDF/MWCNT/DMFPhase inversion Not reported/100 kHz/60 W, 80 W, and 100 W/300 W20 s, 40 s, and 60 sModified membrane structure by promoting the formation of finger-like pores instead of macropores. Improved permeability.[38]
CA/ZnO/DMFPhase inversionNot reported8 hIncreases pore size, improves hydrophilicity, and enhances mechanical stability.[10]
CTA/Graphene oxide/DMFPolymer
disolution
Not reported90 minFacilitates graphene dispersion in the polymer matrix, improving structural orientation.[39]
CA/PEG-400, PS, or PL/DMAcPolymer
disolution
Not reportedNot reportedAids in the dissolution of additives in the polymer solution promote homogeneity.[40]
PVDF/PVP/DMFDegassing polymer solutionNot reported10 minAllowed the removal of bubbles from the polymer mixture.[41]
(CA) Cellulose acetate; (CTA) Cellulose triacetate; (PVDF) polyvinylidene fluoride; (DMF) dimethylformamide; (MWCNT) multiwalled carbon nanotubes; (PEG) polyethylene glycol; (PS) polysulfone; (PL) Pluronic F127; (DMAc) N,N-dimethylacetamide; (PVP) polyvinylpyrrolidone.
Table 2. Effect of ultrasonic intensity on polymer blend dissolution.
Table 2. Effect of ultrasonic intensity on polymer blend dissolution.
Polymer
Blends
Ultrasonic Intensity (W/cm2)Effective Dissolution Time (min)Apparent Viscosity (Pa·s)
PEG-CA0 ± 0.00 a2880 ± 30.01 a3.24 ± 0.06 ab
PEG-CA-A50%59.61 ± 3.89 b35.0 ± 2.50 b3.17 ± 0.04 a
PEG-CA-A100%123.23 ± 4.19 c17.5 ± 1.01 b3.43 ± 0.02 b
PEG-A50% + CA55.81 ± 4.6 b32.5 ± 2.50 b3.23 ± 0.03 ab
PEG-A100% + CA118.6 ± 4.00 c15.0 ± 1.01 b3.25 ± 0.07 ab
(PEG) polyethylene glycol; (CA) cellulose acetate; (A) US amplitude. The results are expressed as the mean ± standard deviation (n = 3). Different letters by columns represent statistically significant differences (p < 0.05).
Table 3. Rheological parameters of the Cross model for stirred and sonicated polymer blends.
Table 3. Rheological parameters of the Cross model for stirred and sonicated polymer blends.
BlendsR2 η 0 (Pa·s) η (Pa·s)λ (s)n
PEG-CA0.974 ± 0.0014.25 ± 0.250.73 ± 0.260.002 ± 0.00020.381 ± 0.169
PEG-CA-A50%0.978 ± 0.0143.55 ± 0.130.18 ± 0.040.002 ± 0.00010.257 ± 0.200
PEG-CA-A100%0.987 ± 0.0103.88 ± 0.190.21 ± 0.120.002 ± 0.00010.371 ± 0.141
PEG-A50%-CA0.996 ± 0.0014.32 ± 0.060.17 ± 0.010.002 ± 0.00010.444 ± 0.323
PEG-A100%-CA0.988 ± 0.0023.69 ± 0.080.21 ± 0.330.002 ± 0.00020.371 ± 0.202
The results are expressed as the mean ± standard deviation (n = 3).
Table 4. Weight loss temperatures and residue for polymer blends under maximum ultrasonic intensity, including endothermic transitions.
Table 4. Weight loss temperatures and residue for polymer blends under maximum ultrasonic intensity, including endothermic transitions.
Polymer
Blends
T 80% (°C)T 90% (°C)T 95% (°C)Glass Transition Temperature Tg (°C)Melting Temperature Tm (°C)
PEG-CA176.50 ± 2.12 a321.5 ± 2.12 a399.0 ± 2.41 a165.98 ± 0.06 a217.56 ± 1.35 a
PEG-CA-A100%172 ± 1.41 a302.5 ± 3.53 b397.0 ± 1.70 a166.07 ± 0.46 a225.52 ± 1.49 b
PEG-A100%-CA164 ± 1.41 b292.5 ± 1.53 c395.5 ± 1.94 a165.63 ± 0.27 a220.78 ± 1.47 a
The results are expressed as the mean ± standard deviation (n = 3). Different letters indicate statistically significant differences.
Table 5. Reduction in turbidity and retention of flat membranes.
Table 5. Reduction in turbidity and retention of flat membranes.
Membrane NTU in PermeateRetention (%)
T0-A019.5 ± 0.7 a97.22 ± 0.1 a
T1-A2520.5 ± 0.7 a97.08 ± 0.1 a
T3-A2519.5 ± 0.7 a98.22 ± 0.1 a
T5-A2520.5 ± 0.7 a97.08 ± 0.1 a
T1-A5020.5 ± 0.7 a97.08 ± 0.1 a
T3-A5020.5 ± 0.7 a97.08 ± 0.1 a
T5-A5019.5 ± 0.7 a97.22 ± 0.1 a
T1-A7520.5 ± 0.7 a97.08 ± 0.1 a
T3-A7519.5 ± 0.7 a97.22 ± 0.1 a
T5-A7519.0 ± 1.4 a97.29 ± 0.2 a
The results are expressed as the mean ± standard deviation (n = 3). Different letters indicate statistically significant differences. NTU: nephelometric turbidity units; T: time (ultrasound application in minutes); A: ultrasound amplitude (%).
Table 6. Turbidity reduction in NTU by the different membranes evaluated.
Table 6. Turbidity reduction in NTU by the different membranes evaluated.
Membrane NTU in PermeateReduction (%)
A-0% (Control) 86.67 ± 8.40 a90.57 ± 0.914 a
A-5%50.33 ± 6.65 b94.52 ± 0.72 b
A-10%103.67 ± 3.05 a88.72 ± 0.33 a
A-20%275.00 ± 22.27 c70.08 ± 2.42 c
The results are expressed as the mean ± standard deviation (n = 3). Different letters indicate statistically significant differences. A: Ultrasound amplitude. Initial turbidity of 920 NTU (nephelometric turbidity units) of 2% whey protein solution.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Méndez-Valdivia, G.K.; Ballinas-Casarrubias, M.D.L.; González-Sánchez, G.; Valdés, H.; Montalvo-González, E.; Chacón-López, M.A.; Martínez-Montaño, E.; Torrestiana-Sánchez, B.; Miramontes-Escobar, H.A.; Ortiz-Basurto, R.I. Effect of Ultrasound on Dissolution of Polymeric Blends and Phase Inversion in Flat Sheet and Hollow Fiber Membranes for Ultrafiltration Applications. Membranes 2025, 15, 120. https://doi.org/10.3390/membranes15040120

AMA Style

Méndez-Valdivia GK, Ballinas-Casarrubias MDL, González-Sánchez G, Valdés H, Montalvo-González E, Chacón-López MA, Martínez-Montaño E, Torrestiana-Sánchez B, Miramontes-Escobar HA, Ortiz-Basurto RI. Effect of Ultrasound on Dissolution of Polymeric Blends and Phase Inversion in Flat Sheet and Hollow Fiber Membranes for Ultrafiltration Applications. Membranes. 2025; 15(4):120. https://doi.org/10.3390/membranes15040120

Chicago/Turabian Style

Méndez-Valdivia, Gilberto Katmandú, María De Lourdes Ballinas-Casarrubias, Guillermo González-Sánchez, Hugo Valdés, Efigenia Montalvo-González, Martina Alejandra Chacón-López, Emmanuel Martínez-Montaño, Beatriz Torrestiana-Sánchez, Herenia Adilene Miramontes-Escobar, and Rosa Isela Ortiz-Basurto. 2025. "Effect of Ultrasound on Dissolution of Polymeric Blends and Phase Inversion in Flat Sheet and Hollow Fiber Membranes for Ultrafiltration Applications" Membranes 15, no. 4: 120. https://doi.org/10.3390/membranes15040120

APA Style

Méndez-Valdivia, G. K., Ballinas-Casarrubias, M. D. L., González-Sánchez, G., Valdés, H., Montalvo-González, E., Chacón-López, M. A., Martínez-Montaño, E., Torrestiana-Sánchez, B., Miramontes-Escobar, H. A., & Ortiz-Basurto, R. I. (2025). Effect of Ultrasound on Dissolution of Polymeric Blends and Phase Inversion in Flat Sheet and Hollow Fiber Membranes for Ultrafiltration Applications. Membranes, 15(4), 120. https://doi.org/10.3390/membranes15040120

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