<<

polymers

Article Fabrication and Characterization of Electrospun Poly(acrylonitrile-co-Methyl Acrylate)/Lignin Nanofibers: Effects of Lignin Type and Total Polymer Concentration

Suchitha Devadas 1 , Saja M. Nabat Al-Ajrash 1 , Donald A. Klosterman 1 , Kenya M. Crosson 2,3, Garry S. Crosson 4 and Erick S. Vasquez 1,3,*

1 Department of Chemical and Materials Engineering, University of Dayton, Dayton, OH 45469, USA; [email protected] (S.D.); [email protected] (S.M.N.A.-A.); [email protected] (D.A.K.) 2 Department of Civil and Environmental Engineering and Engineering Mechanics, University of Dayton, Dayton, OH 45469, USA; [email protected] 3 Integrative Science and Engineering Center, University of Dayton, Dayton, OH 45469, USA 4 Department of Chemistry, University of Dayton, Dayton, OH 45469, USA; [email protected] * Correspondence: [email protected]; Tel.: +1-(937)-229-2627

Abstract: Lignin macromolecules are potential precursor materials for producing electrospun nanofibers for composite applications. However, little is known about the effect of lignin type

 and blend ratios with synthetic polymers. This study analyzed blends of poly(acrylonitrile-co-methyl  acrylate) (PAN-MA) with two types of commercially available lignin, low sulfonate (LSL) and alkali,

Citation: Devadas, S.; kraft lignin (AL), in DMF solvent. The electrospinning and polymer blend solution conditions were Al-Ajrash, S.M.N.; Klosterman, D.A.; optimized to produce thermally stable, smooth lignin-based nanofibers with total polymer content Crosson, K.M.; Crosson, G.S.; of up to 20 wt % in solution and a 50/50 blend weight ratio. Microscopy studies revealed that AL Vasquez, E.S. Fabrication and blends possess good solubility, miscibility, and dispersibility compared to LSL blends. Despite the Characterization of Electrospun lignin content or type, rheological studies demonstrated that PAN-MA concentration in solution Poly(acrylonitrile-co-Methyl dictated the blend’s viscosity. Smooth electrospun nanofibers were fabricated using AL depending Acrylate)/Lignin Nanofibers: Effects upon the total polymer content and blend ratio. AL’s addition to PAN-MA did not affect the glass of Lignin Type and Total Polymer transition or degradation temperatures of the nanofibers compared to neat PAN-MA. We confirmed Concentration. Polymers 2021, 13, 992. the presence of each lignin type within PAN-MA nanofibers through infrared spectroscopy. PAN- https://doi.org/10.3390/ MA/AL nanofibers possessed similar morphological and thermal properties as PAN-MA; thus, these polym13070992 lignin-based nanofibers can replace PAN in future applications, including production of carbon fibers

Academic Editor: Zeynep Aytac and supercapacitors.

Received: 28 February 2021 Keywords: electrospinning; alkali; kraft lignin; low sulfonate lignin; poly(acrylonitrile-co-methyl Accepted: 22 March 2021 acrylate); nanofibers; modulated DSC; lignin-based nanofibers Published: 24 March 2021

Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in 1. Introduction published maps and institutional affil- Many of today’s leading industries—such as aerospace, automobile, military, and iations. energy—are increasingly pursuing sustainable materials with enhanced properties, includ- ing low density, good thermal stability, corrosion resistance, and high stiffness and strength. These performance enhancements traditionally have been achieved by using composite materials, most notably carbon fiber composites, offering an alternative to denser and Copyright: © 2021 by the authors. more expensive metallic and ceramic material [1,2]. Commercially available precursor Licensee MDPI, Basel, Switzerland. materials for carbon fiber production are synthesized from petroleum feedstocks, such This article is an open access article as polyacrylonitrile (PAN) [3–5]. Recently, natural polymers such as cellulose, curdlan, distributed under the terms and and lignin emerged as alternatives for preparing composite materials due to their wide conditions of the Creative Commons availability, cost-effectiveness, and recyclability [6–9]. For instance, lignin is a traditional Attribution (CC BY) license (https:// by-product of the paper and pulp industry with high carbon content and multiple aromatic creativecommons.org/licenses/by/ structures. In addition to these attributes, lignin macromolecules can exhibit hydrophilic 4.0/).

Polymers 2021, 13, 992. https://doi.org/10.3390/polym13070992 https://www.mdpi.com/journal/polymers Polymers 2021, 13, 992 2 of 16

or hydrophobic properties and are used as UV-protecting materials [6,10]. Hence, lignin has emerged as a potential sustainable precursor for fabricating carbon-based fibers, com- posites, or nanomaterials products. However, due to its low molecular weight, lignin usage is limited and is currently used only as an additive material or filler in synthetic polymers. Despite this limitation, partially bio-based carbon nanofibers made from blends of lignin macromolecules and other polymers such as PAN are cost-effective and could be used as value-added products [11,12]. For example, alkali lignin/PAN blends exhibited higher elastic modulus and tensile strength than pure PAN [13]. These enhanced properties could allow the fabrication and design of improved membranes, adsorbents, or electrical materials [14]. As recently demonstrated through scale-up processes, electrospinning has emerged as a viable tool in the manufacturing of natural nanocomposite fibers [15–18]. In the electro- spinning process, a high electrical field is applied to obtain a jet flow of polymer solution by rapid solvent evaporation, resulting in the formation of microfibers or nanofibers [19,20]. Conditions such as polymer concentration, molecular weight of polymers, viscosity, elec- trical conductivity, and surface tension affect nanofiber formation [21,22]. Additionally, electrospinning has enabled the fabrication of nanofibers comprised of polymer blends, including natural polymers such as lignin [23,24]. These exhibit high porosity, higher electrical conductivity, improved mechanical properties, and high surface area compared to the neat polymer. For example, biorefinery lignin/PAN based carbon fibers enhanced via graphene modification used in electrodes improved electrical conductivity and quan- tification of acetaminophen [25]. Electrospun organosolv lignin/PAN nanofibers flexibility and thermochemical properties were improved at higher lignin contents [26]. Adsorptive removal of Pb(II) contaminants from wastewater was successful using sago-lignin/PAN electrospun carbon nanofibers [27]. In addition to PAN, lignin blends with other polymer matrices such as polyvinyl alcohol (PVA) or polyethylene oxide (PEO) exhibited enhanced properties [28–30]. For example, phosphoric acid modified kraft lignin has been shown to increase the mechanical properties of electrospun lignin/PEO fibers due to an increase in lignin’s molecular weight [31]. The rheological, electrical conductivity and surface tension of polymer solutions heav- ily influence the morphological properties of electrospun nanofibers [32]. For instance, the rheological properties of low molecular weight organosolv lignin/PAN polymer solutions were analyzed over a range of polymer concentrations and blend ratios to optimize the fabrication of defect-free, smooth nanofibers [33]. Electrospun nanofibers comprised of alkali lignin/PVA and PAN had larger pore sizes compared to traditional activated carbon, and the alkali lignin/PVA nanofibers demonstrated enhanced and faster adsorption of a large molecule adsorbate [34]. Dallmeyer et al. [35] examined the physical properties and chemical structure differences of electrospun kraft lignin that influence shape change and recovery, as well as moisture responses of fractions of lignin. Cho et al. [36] analyzed functional groups on solvent cast film and nanofibers using Fourier transform infrared spectroscopy (FTIR) and polarized light optical microscopy analysis in order to understand alignment of lignin within the structured films. They also analyzed the birefringence of lignin nanofibers under changing angles of polarized light. PAN homopolymers, blended with lignin, are the traditional precursor materials for the preparation of carbon nanofibers through electrospinning as discussed recently [23]; however, PAN lacks drawability and possess poor heat treatment tolerance due to the interactions among nitrile groups [37]. As a result, PAN copolymerization with methyl acrylate (MA), vinyl acetate (VAc) or itaconic acid (IA) has been used to enhance nanofiber spinnability, as well as thermal performance [38]. Although lignin cannot be electrospun directly due to its molecular weight branched structure and lack of chain entanglement, PAN-co-MA copolymer blended with lignin fillers could produce stable nanofibers. Fur- thermore, the quality and spinnability of nanofibers are also governed by lignin type. Lignins are classified by their extraction methods and also on the plant resource such as organosolv lignin, alkali, Kraft lignin, and lignosulfonates. Polymers 2021, 13, 992 3 of 16

The current study intends to shed light on the potential for producing thermally stable, uniform nanofibers from blends of the most common types of commercially available lignin, i.e., low sulfonate lignin (LSL) and alkali, kraft lignin (AL), with the copolymer PAN-co-MA rather than with pure PAN. We prepared PAN-co-MA/lignin blends at different weight ratios (50/50 and 80/20) and total polymer content (10%, 12%, and 20%) in DMF solvents for each lignin type. Using these polymers and blends, for the first time, we investigated the correlations between the polymer solution behavior (i.e., lignin distribution within PAN-co- MA/lignin solvent cast films and polymer solution rheological studies) with the fabricated electrospun nanofibers morphological, thermal, and chemical properties. Furthermore, we discussed the effects of lignin sulfonation by using LSL and AL in the preparation and characterization of lignin-based electrospun nanofibers through microscopy, thermal, and spectroscopy analyses.

2. Materials and Methods 2.1. Materials Alkali Lignin, low sulfonate content (LSL), alkali, kraft lignin (AL), and N,N-Dimethy- lformamide (DMF) anhydrous, 99.8% were purchased from Sigma-Aldrich (Saint Louis, MO, USA). Poly (acrylonitrile-co-methyl acrylate) with acrylonitrile ~94 wt %, MW 100,000 was procured from Scientific Polymer Products, Inc. (Ontario, NY, USA). Hereafter, the copolymer is referred to as PAN-MA. Chemicals were used as received, without further purification.

2.2. Preparation of Polymer Blends and Electrospinning Parameters PAN-MA was initially dissolved in DMF at 60 ◦C and mixed in 20 mL scintillation vials with magnetic stirring at 600 rpm for 30 min. Next, specific masses of LSL or AL powder were added to the PAN-MA solution and magnetically stirred. PAN-MA/LSL mixtures were stirred for 12 h, whereas PAN-MA/AL solutions were mixed for 2 h due to greater solubility. Table1 shows the specific lignin types and concentration used for each blend. Three types of electrospun nanofibers were prepared using these polymer solutions: PAN- MA (control), PAN-MA/LSL, and PAN-MA/AL. Based on the PAN-MA/AL solubility in DMF, the total polymer content, including both PAN-MA and lignin, varied from concentrations of 10, 12, and 20 wt % (Table1).

Table 1. Weight ratios of PAN-MA:Lignin polymer blends prepared based on total polymer concen- tration in wt %.

Total Polymer PAN Content Lignin Content Lignin Type Sample ID Content in wt % 100 0 10 PAN-MA 80 20 10 Low sulfonate LSL-1 50 50 10 Low sulfonate LSL-2 80 20 10 Alkali, Kraft AL-1 50 50 10 Alkali, Kraft AL-2 80 20 12 Alkali, Kraft AL-3 50 50 20 Alkali, Kraft AL-4

Each prepared polymer blend was filled transferred a 10 mL plastic syringe with a blunt 18G needle (~1 mm inner diameter) and mounted on a syringe pump (New Era Pump Systems, Inc., NE-300, Farmingdale, NY, USA). Electrospinning was conducted at a 0.003 mL/min flow rate. A voltage generator (Stanford Research Systems, Inc. model P.S375, Sunnyvale, CA, USA) was operated at 15 kV using a positive voltage supplied to the needle. A negative voltage (6–8 volts) was used to control a rotatory cylindrical drum collector (Dayton DC Motor, OH, USA 4Z145). Electrospun nanofibers were col- lected on aluminum foil. The working distance between the collector and the needle was fixed at 21 cm. Fabricated nanofibers were heat-treated in a vacuum oven (Fisher Polymers 2021, 13, 992 4 of 16

Isotemp® standard capacity) at 60 ◦C for 6 h to evaporate the residual DMF solvent after the electrospinning process.

2.3. Polymer Solutions and Films: Rheological and Microscopy Analysis The polymer blends’ rheological behavior was determined at room temperature (25 ◦C) using flow and frequency sweep tests. These measurements were conducted using an Anton Paar MCR 302 modular compact rheometer equipped with a cone (25 mm, 1◦) and plate fixture. Using flow sweep tests, the shear rate varied between 1–1000 s−1. Frequency sweep tests in oscillation mode were conducted using the same cone & plate geometry in the range of 100–0.1 rad/s (strain amplitude 1%) at 25 ◦C. All measurements were collected within the linear viscoelastic regime (LVE) of each sample. The distribution of LSL and AL nanofibers within the PAN-MA polymer solution were macroscopically analyzed with polarized optical microscopy using a polarized light microscope (Zeiss AX10 with Axio Cam HRc). To analyze lignin distribution within dried PAN-MA/lignin films, a film solvent casting approach was used. The films were prepared by placing each polymer solution on a glass microscope slide followed by solvent evaporation at 60 ◦C for 12 h in an oven. The films were examined with the same microscope and also with a scanning electron microscope (SEM; Phenom ProX, Nanoscience Instruments, Eindhoven, The Netherlands) with an operating voltage of 5 kV or 15 kV.

2.4. Morphological, Chemical, and Thermal Characterization of Electrospun Nanofibers The electrospun nanofibers’ morphology and diameter were examined using scanning electron microscopy (SEM; Phenom ProX) with an operating voltage of 15 kV and a high- resolution SEM (Hitachi S-4800) at an operating voltage of 30 kV. The mean diameter and standard deviation of the nanofiber thickness were estimated from the average of at least 50 measurements using ImageJ software. Attenuated Total Reflectance Fourier Transform Infrared (ATR-FTIR) spectra were collected using a Nicolet iS50 instrument (Thermo Scien- tific). A minimum of 64 scans were collected for all spectra recorded in the 4000–500 cm−1 range using a Smart Golden Gate ZnSe ATR accessory. The thermal stability of each sample was evaluated using Thermogravimetric analysis (TGA; Q500-0845-TGA model, TA Instru- ments). Thermograms were obtained using 4 mg samples heated on a platinum pan at a heating rate of 10 ◦C/min from 30 to 800 ◦C under a nitrogen atmosphere. Each thermo- gram was normalized with 100% weight percentage at ~130 ◦C to remove solvent effects (i.e., DMF residues or adsorbed water onto the nanofibers). Thermal transitions—glass transition temperature (Tg), melting temperature (Tm) and crystallization temperature (Tc)—were identified for each electrospun nanofiber via Differential Scanning Calorimetry (DSC) and Modulated Differential Scanning Calorimetry (mDSC) experiments using a Q2000 instrument (TA Instruments). Samples were placed in hermetically sealed T-zero low mass aluminum pans for DSC and mDSC measurements with sample weights between 1–3.5 mg. DSC measurements were made while heating each sample from 40 to 400 ◦C with a heating rate of 10 ◦C/min while mDSC measurements were performed at 2◦/min, 1 ◦C amplitude, and 60 s periods up to 200 ◦C. All tests were performed under an inert nitrogen atmosphere.

3. Results and Discussion PAN-MA/lignin polymer weight ratios were selected for electrospinning after a thorough microscopy inspection of solvent cast films (Table1). In a previous literature study, solvent cast films provided detailed information about the interactions of polymer blends, or composites, prepared with lignin in a polymer matrix [39]. In the current study, the PAN-MA/lignin films were prepared in an effort to assess the effects of the lignin type and concentration on the distribution, phase separation, orientation, or aggregation of lignin within PAN-MA. As expected, the neat PAN-MA solvent cast film had a smooth surface without significant defects or roughness (Figure S1). Through the addition of lignin to PAN-MA, a birefringence response was expected for the films under polarized light, as Polymers 2021, 13, x FOR PEER REVIEW 5 of 17

and concentration on the distribution, phase separation, orientation, or aggregation of lig- Polymers 2021, 13, 992 5 of 16 nin within PAN-MA. As expected, the neat PAN-MA solvent cast film had a smooth sur- face without significant defects or roughness (Figure S1). Through the addition of lignin to PAN-MA, a birefringence response was expected for the films under polarized light, as reported elsewhereelsewhere [ 36[36].]. AL AL containing containing sample sample micrographs micrographs showed showed well-dispersed, well-dispersed, round roundlignin aggregateslignin aggregates (~5 micrometer (~5 micrometer diameter) diameter) within within PAN-MA PAN-MA at a PAN-MA/AL at a PAN-MA/AL ratio ra- of tio80/20 of 80/20 as shown as shown in Figure in Figure1A for 1A 10 for wt %10 totalwt % polymer total polymer (AL-1 sample)(AL-1 sample) and in Figureand in1 FigureC for 1Ca 12 for wt a %12 sample wt % sample total polymer total polymer (AL-3 sample).(AL-3 sample). At a 50/50 At a PAN-MA/AL50/50 PAN-MA/AL ratio (AL-2ratio (AL- and 2AL-4), and AL-4), lignin birefringencelignin birefringence was more was evident more withinevident PAN-MA within PAN-MA films and films the formation and the offor- a mationporous of structure a porous was structure noted (Figure was noted1B,D). (Fig Whenure 1B,D). the total When polymer the total wt %polymer increased wt from% in- creased10 wt % from (AL-2) 10 towt 20 % wt(AL-2) % (AL-4), to 20 wt the % lignin(AL-4), aggregate the lignin size aggregate increased. size Despiteincreased. the Despite rough thesurface rough observed surface atobserved 50/50 blendat 50/50 ratios, blend AL ratios, seems AL to seems have to distributed have distributed homogeneously homoge- neouslywithin PAN-MA within PAN-MA independent independent of the total of the polymer total polymer concentration concentration of 10 wt %of or10 20wt wt% or %. 20 It wtis important %. It is important to note that to note the PAN-MA/AL that the PAN-MA/AL polymer solutionspolymer weresolutions magnetically were magnetically stirred for stirredonly 2 h,for as only discussed 2 h, as indiscussed the Methods in the section. Methods section.

Figure 1. PolarizedPolarized optical optical light light micrographs micrographs of of solvent solvent casted casted polymer polymer films: films: (A ()A AL-1;) AL-1; (B) ( BAL-2;) AL-2; (C) (AL-3;C) AL-3; (D) (AL-4,D) AL-4, (E) LSL-1; and (F) LSL-2, revealed phase separation and aggregation of lignin within PAN-co-MA. [Scale bar = 50 microns]. (E) LSL-1; and (F) LSL-2, revealed phase separation and aggregation of lignin within PAN-co-MA. [Scale bar = 50 microns].

Preliminary PAN-MA/LSL solvent solvent evaporated evaporated films films were were obtained; however, phase separation was observed for short mixing times.times. Upon further inspection, an overnight mixing step (12 h) was implemented to improveimprove the low-sulfonatelow-sulfonate lignin’s dispersion within the the PAN-MA PAN-MA films. films. The The additional additional mixing mixing time time significantly significantly improved improved lignin lignin dis- persiondispersion for foran an80/20 80/20 PAN-MA/LSL PAN-MA/LSL content, content, as asshown shown in in Figure Figure 1E1E (LSL-1). As the LSLLSL content increased from 20% to 50%, large aggregatesaggregates formation continued to be observed within the PAN-MA filmsfilms shown in Figure1 1EE (LSL-2).(LSL-2). ForFor thisthis reason,reason, bothboth LSLLSL samplessamples were only producedproduced atat aa totaltotal polymer polymer content content of of 10 10 wt wt % % in in this this study. study. Further Further increases increases in inLSL LSL ratio ratio or totalor total polymer polymer content content (e.g., (e.g., 20 wt 20 %) wt was %) unfeasiblewas unfeasible as more as more aggregates aggregates were wereprone prone to form. to form. SEM micrographsSEM micrographs were collectedwere collected to further to further confirm confirm the presence the presence of lignin of ligninagglomerated agglomerated distribution distribution within within the PAN-MA the PAN-MA films films and filmand film morphology morphology (Figure (Figure S2). S2).SEM SEM microscopy microscopy confirmed confirmed that that LSL LSL samples samples showed showed larger larger cluster cluster formation,formation, flaky flaky structures, and large pores in contrastcontrast to AL containing samples. The viscosity results for the PAN-MA/ligninPAN-MA/lignin solutions are shown in Figure2 2.. TheThe curves illustrateillustrate thethe shear-thinning shear-thinning behavior behavior of theof polymerthe polymer solutions solutions which which was attributed was at- tributedto the macromolecular to the macromolecular nature of nature the polymers of the polymers [40]. At a [40]. 10 wt At % a polymer10 wt % concentration,polymer con- centration,the control samplethe control (pure sample PAN-MA/DMF (pure PAN-MA solution,/DMF no lignin) solution, exhibited no lignin) the highest exhibited viscosity the due to a ten-fold high molecular weight of PAN-MA relative to lignin. The viscosity steadily decreased as the lignin content increased in the polymer blend solution (compare LSL-1 to LSL-2 and AL-1 to AL-2). Note that, for the same PAN-MA/lignin ratios, the lignin type

did not affect the viscosity results. For example, LSL-1 and AL-1 showed similar viscosity Polymers 2021, 13, x FOR PEER REVIEW 6 of 17

highest viscosity due to a ten-fold high molecular weight of PAN-MA relative to lignin. Polymers 2021, 13, 992 The viscosity steadily decreased as the lignin content increased in the polymer blend6 of so- 16 lution (compare LSL-1 to LSL-2 and AL-1 to AL-2). Note that, for the same PAN-MA/lig- nin ratios, the lignin type did not affect the viscosity results. For example, LSL-1 and AL- values.1 showed However, similar variationsviscosity values. in viscosity However, were observedvariations as in the viscosity total polymer were observed concentration as the intotal solution polymer increased concentration to 12 wt in % solution and 20 wt increased % (see AL-3 to 12 and wt AL-4).% and AL-320 wt contained % (see AL-3 slightly and lessAL-4). PAN-MA AL-3 contained in solution slightly as the less control PAN-MA sample in solution (9.6 wt %as vs.the control 10 wt %) sample in addition (9.6 wt to% 2.4vs. 10 wt wt % AL,%) in but addition the viscosity to 2.4 wt was % AL, lower but than the viscosity the control was sample. lower than AL-4 the contained control sam- the sameple. AL-4 amount contained of PAN-MA the same in amount solution of as PAN-MA the control in solution (10 wt %) as but the alsocontrol contained (10 wt %) 10 but wt %also AL. contained Its viscosity 10 wt curve % AL. was Its essentially viscosity the cu samerve was as theessentially control despitethe same containing as the control alkali lignin.despite These containing results alkali imply lignin. that the These viscosity results of theimply solution that the is dominated viscosity of by the the solution PAN-MA is concentration,dominated by andthe PAN-MA the amount concentration, of lignin is inconsequential. and the amount This of lignin behavior is inconsequential. implies that the highThis behavior molecular implies weight that PAN-MA the high serves molecular as a type weight of scaffold PAN-MA in solution,serves as whicha type carriesof scaffold the lowin solution, molecular which weight carries lignin. the Thislow molecular study confirms weight that lignin. lignin This biopolymers study confirms can be that added lignin at variousbiopolymers concentrations can be added of interest at various without concentrations impacting of the interest solution without viscosity impacting as long the as theso- PAN-MAlution viscosity concentration as long as is fixed.the PAN-MA concentration is fixed.

FigureFigure 2.2. Steady-flowSteady-flow rheologyrheology resultsresults ofof allall polymerpolymer solutionssolutions revealedrevealed thatthat viscosityviscosity cancan bebe tuned basedtuned onbased the on lignin the contentlignin content and the and total the polymer total polymer concentration. concentration. [PAN-MA [PAN-MA in DMF in (squares)DMF was prepared(squares) atwas a 10 prepared wt % in at DMF a 10 and wt % used in DMF as acontrol and used sample]. as a control sample].

PreviousPrevious studiesstudies suggestedsuggested that thethe electrospinnabilityelectrospinnability of polymer solutions, com- prisedprised ofof lignin,lignin, dependsdepends onon thethe viscosity behaviorbehavior ofof thethe polymer solution,solution, whichwhich dictatesdictates thethe nanofibers’nanofibers’ morphologicalmorphological propertiesproperties [[41].41]. Furthermore,Furthermore, PAN-MA andand 80/2080/20 PAN- MA/ligninMA/lignin blend blend viscosity viscosity curves curves exhibited exhibited a aNewtonian Newtonian plateau plateau from from low low shear shear rate rate up −1 upto at to least at least 10 s 10−1 (LSL1, s (LSL1, AL-1, AL-1, AL-3; AL-3;Figure Figure 2). Note2). that Note larger that largerlignin ligninratios (50/50) ratios (50/50)showed showeda slight increase a slight increase in viscosities in viscosities at low shear at low rates; shear however, rates; however, this was this probably was probably due to due the torheometer the rheometer reaching reaching the lower the limit lower of limittorque of and torque was and not wasconsidered not considered significant. significant. Thus, all −1 Thus,samples all were samples safely were within safely their within zero-shear their zero-shear viscosity viscosityrange under range 1 s under−1 with 1 the s exceptionwith the exceptionof PAN-MA. of PAN-MA.The shear Therate shearin the rateelectrospi in thenning electrospinning needle, estimated needle, estimatedaccording accordingto laminar −1 toflow laminar in a capillary flow in a (shear capillary rate (shear = 4 Q/p rate R3), = 4 Q/pwas R3),approaching was approaching 0.1 s−1; therefore, 0.1 s ; therefore, the flow thebehavior flow behavior during electr duringospinning electrospinning was considered was considered to be Newtonian to be Newtonian for all for samples. all samples. Fre- Frequencyquency sweep sweep tests tests in oscillation in oscillation mode mode confirmed confirmed the thepolymer polymer solutions’ solutions’ liquid-like liquid-like be- behaviorhavior at atdifferent different lignin lignin concentrations concentrations and and total total polymer polymer concentrations. concentrations. At angular At angular fre- frequenciesquencies below below 100 100 rad/s, rad/s, G’’ G”> G’ > for G’ forall the all thesamples samples (Figure (Figure S3). S3). TheThe controlcontrol samplesample PAN-MAPAN-MA (PAN-MA(PAN-MA inin DMFDMF solutionsolution atat 1010 wtwt %)%) exhibitedexhibited thethe highesthighest viscosityviscosity of of all all samples samples and and resulted resulted in a in stable a stable jet during jet during electrospinning electrospinning and defect- and free continuous nanofiber mats as shown in Figure3A. Additionally, these fibers had the defect-free continuous nanofiber mats as shown in Figure 3A. Additionally, these largest average diameter for the samples prepared at a 10 wt % polymer content (Figure4A ). had the largest average diameter for the samples prepared at a 10 wt % polymer content AL-1 and LSL-1 solutions also showed a continuous fiber morphology (Figure3B,C ), and (Figure 4A). AL-1 and LSL-1 solutions also showed a continuous morphology (Fig- their average diameters were slightly less than PAN-MA. A high-resolution SEM image of ure 3B,C), and their average diameters were slightly less than PAN-MA. A high-resolution AL-1 demonstrated a smooth electrospun nanofiber surface (Figure S4). When the ratio of LSL or AL to PAN-MA was increased to 50/50 (LSL-2 and AL-2), the fiber morphology resulted in nodules, or beads, along its length (Figure3D,E) and their average diameters

were the lowest of all samples (Figure4A). Moreover, LSL-2 and AL-2 coincided with the Polymers 2021, 13,Polymers x FOR PEER 2021 ,REVIEW 13, x FOR PEER REVIEW 7 of 17 7 of 17

SEM image ofSEM AL-1 image demonstrated of AL-1 demonstrated a smooth electrospun a smooth nanofiber electrospun surface nanofiber (Figure surface S4). (Figure S4). Polymers 2021, 13, 992 7 of 16 When the ratioWhen of LSL the or ratio AL ofto LSLPAN-MA or AL wasto PAN-MA increased was to 50/50 increased (LSL-2 to and50/50 AL-2), (LSL-2 the and AL-2), the fiber morphologyfiber resultedmorphology in nodules, resulted or in beads, nodules, along or itsbeads, length along (Figure its length 3D,E) (Figureand their 3D,E) and their average diametersaverage were diameters the lowest were of allthe samples lowest of (Figure all samples 4A). Moreover, (Figure 4A). LSL-2 Moreover, and AL-2 LSL-2 and AL-2 coincided withcoincidedlowest the lowest solution with solution the viscosities lowest viscosities solution of all samplesof allviscosities samples tested. of tested. all During samples During electrospinning, tested. electrospin- During these electrospin- samples ning, these samplesning,were observedthese were samples observed to form were dropletsto form observed atdroplets the to needle form at the tip,dr needleoplets and the attip, processthe and needle the gradually process tip, and electrosprayed the process gradually electrosprayedgraduallywith improper electrosprayed with formation improper with of formation fibers improper (Figure of fibersformation3D,E (Figure insets). of fibers3D,E A few insets).(Figure droplets A3D,E few and insets). aggregated A few droplets and aggregateddropletsstructures and werestructures aggregated observed were structures withinobserved the were within electrospun observed the electrospun nanofibers,within the nanofibers, electrospun in particular in nanofibers, for the LSL in particular for particularthesamples, LSL samples, confirming for the LSLconfirming the samples, aggregation the confirming aggregation phenomena the phenomena aggregation observed observed phenomena in films in produced films observed via in solvent films produced via producedsolventcasting (Figurecasting via solvent S5A).(Figure casting S5A). (Figure S5A).

Figure 3. SEM images of electrospun PAN-MA/lignin nanofibers at a total polymer concentration of 10 wt % for (A) neat Figure 3. SEM images of electrospun PAN-MA/ligninPAN-MA/lignin nanofibers nanofibers at at a a total total polymer concentration of 10 wt % for ( A) neat PAN-MA (control),PAN-MA (B) AL-1, (control), (C) LSL-1, (B) AL-1, (D) AL-2, (C) LSL-1, and ( (ED) )LSL-2; AL-2, andand at(E ()F LSL-2;) 12 wt and % AL-3, at (F) and 12 wt (G %) 20 AL-3, wt % and AL-4. (G) 20 wt % AL-4. PAN-MA (control), (B) AL-1, (C) LSL-1, (D) AL-2, and (E) LSL-2; and at (F) 12 wt % AL-3, and (G) 20 wt % AL-4.

Figure 4. CorrelationsFigure 4. Correlations obtainedFigure between 4.obtained Correlations PAN-MA/lignin between obtained PAN-MA/ligni electrospun between nPAN-MA/ligni nanofibers’electrospun diameter nanofiben electrospun withrs’ diameter polymer nanofibe with solutionrs’ diameter viscosity with at 10 s−1 forpolymer (A) fibers solution of differentpolymer viscosity lignin solution at types 10 s viscosity− and1 for blend (A) atfibers ratios 10 s −of1 atfor different a total(A) fibers 10 lignin wt of % differenttypes and (B and) fibers lignin blend at types higherratios and at polymer ablend total ratios content at ofa total 12 wt % (AL-3)10 wt and % 20 and wt ( %B)10 (AL-4).fibers wt % at and [Points higher (B) with fiberspolymer error at higher content bars representpolymer of 12 wt content nanofiber % (AL-3) of diameters].12and wt 20 % wt (AL-3) % (AL-4). and 20 [Points wt % (AL-4). [Points with error bars withrepresent error nanofiberbars represent diameters]. nanofiber diameters]. As the polymer content increased, AL-3 and AL-4 fibers were observed to have a As the polymercontinuous,As contentthe polymer smooth increased, fibercontent surfaceAL-3 increased, and (Figure AL-4 AL-33 F,G),fibers and referred were AL-4 observed tofibers in some were to literaturehave observed a studies to have as a continuous, smooth“beads-free”continuous, fiber surfacesmooth morphology (Figurefiber [surface36 3F,G),]. For (Figure AL-4referred fibers, 3F,G), to in inter-fiber referredsome literature to bonding in some studies was literature present as studies with fiber as “beads-free” morphology“beads-free”diameters in [36]. themorphology micron-sized For AL-4 [36]. fibers scale For, inter-fiber AL-4 (Figure fibers4B bonding and, inter-fiber Figure was S5B); presentbonding thus, with was the fiber-likefiber present with structure fiber diameters in thediametersis alsomicron-sized affected in the bymicron-sizedscale an (Figure increase 4Bscale in and AL (Figure Figure lignin 4B S5B); content. and thus, Figure In the this S5B); fiber-like study, thus, astruc-the polymer fiber-like solution struc- ture is also affectedtureviscosity is byalso an range affected increase of approximately by in an AL increase lignin content. 0.45–3.5 in AL lignin In Pa-s this wascontent. study, observed Ina polymer this to study, be solution a threshold a polymer of solution achiev- viscosity rangeviscositying of beads-freeapproximately range morphologyof approximately 0.45–3.5 Pa-s in the was 0.45–3.5 current observed Pa-s study. wasto be Currently, observed a threshold to whether be of aachieving threshold the viscosity of achieving results beads-free morphologybeads-freeare coincidental inmorphology the current or directly in study. the related current Currently, is study. unknown. whether Currently, The the phenomenonviscosity whether resultsthe ofviscosity howare the results nodular are (bead) morphology is obtained is not well understood in the literature, but the results of this work suggest that bead formation is related to lignin phase separation within PAN-MA. For instance, PAN-MA nanofibers exposed to polarized light revealed a sin- gle polymer phase (Figure S6A). Polarized micrographs of electrospun PAN-MA/lignin Polymers 2021, 13, 992 8 of 16

fibers revealed discontinuities within AL samples suggesting phase separation of the two polymers (Figure S6B,C). Although the polymer solution rheology was expected to im- pact electrospun nanofiber formation, the difference in solution viscosities between the produced beaded fibers and beads-free fibers was not significant (Figure4, red squares). Moreover, electrospun nanofiber average diameters were significantly different for almost all the electrospun fibers (alpha = 0.05, Figure S7, One-way ANOVA, Tukey’s test). We only found non-significant differences in average diameter for three samples: PAN-MA and AL-3, LSL-1 and AL-1, and LSL-2 and AL-2. Additional factors experimentally studied in the literature include electrical conductivity, applied voltage, and surface tension [42]. Fu- ture studies could also correlate additional polymer solution parameters, such as electrical conductivity and surface tension [32]. In the industrial processes of polymer extrusion and fiber spinning, complex fiber morphologies such as sharkskin and melt fracture have been observed for pure polymers extruded at high shear rates, in some cases as low as 69 s−1 for the appearance of shark- skin [43]. The phenomenon has been closely linked to normal forces in the melt [44] and high shear stress at the wall, with critical value being on the order of 0.1 to 0.4 MPa [45]. However, the fiber morphology in the current study is not likely related to this due to the low shear rates, solution phase extrudate, and the presence of other factors related to electrical conductivity and rapid evaporation. What can be said for the current student is that, at the same total polymer content (10%), the average electrospun nanofibers’ diameters decreased as the lignin content increased, which corresponded to lower viscosity values (Figure4A). As the total polymer content increased, at a fixed PAN-MA concentration, the nanofiber thickness and viscosity of the solution also increased (Figure4B). The purpose of TGA testing was to evaluate the decomposition behavior of the various powders and electrospun nanofiber materials, with a focus on how the material converts to a char which is important for the production of carbon fibers. First, the individual responses of the three raw materials in powder form are given in Figure5 along with the wt % remaining curves and derivative curves. PAN-MA exhibited three primary stages (peaks at 303, 316, 427 ◦C) and a char yield of 47.4 wt %. This is reasonably consistent with other published studies that revealed PAN-MA undergoes degradation of the C-N bond at 297.5 ◦C[39]. In contrast, AL and LSL decomposition began a broad decomposition event around 150 ◦C. This is also consistent with previous studies, in which this effect was attributed to aromatic groups present in lignin and various molecular interactions between the bonds that are responsible for an extensive range of degradation [46]. The decomposition of AL powder was essentially single stage with a peak rate loss at 376 ◦C and a char yield of 37.0 wt %. LSL powder exhibited two stages (270, 350 ◦C) and a fairly high wt % plateau until after 750 ◦C when the weight remaining dropped to 51.5 wt %. The TGA results for AL-containing electrospun nanofibers are given in Figure6. The results for the control fiber (PAN-MA) were comparable to the PAN-MA powder, with derivative peaks (points 1, 2, 4) located at the same temperature as in Figure5. With the addition of 20% AL, the first two derivative peaks for PAN-MA (points 1,2) condensed into one (point 3), and the higher temperature peak (point 4) shifted to slightly lower temperature (point 5). No new peaks appeared (for example at 376 ◦C), perhaps because MA was present in such a low quantity. However, with the addition of 50% AL, a new derivative peak appeared around 350 ◦C (point 7), which may be attributed to the main AL decomposition (point 8) shift to a lower temperature. The peak corresponding to point 6 was attributed to PAN-MA decomposition, as was the case in the AL-1 fiber (i.e., point 3). The main conclusion drawn from analyzing the derivative peaks is that PAN-MA and AL do not decompose independently when blended together. This is not usual for polymer blends, in which case three possibilities exist that is difficult or impossible to predict: one of the components increases the stability of the other, one of the components reduces the stability of the components, or the blend has equal stability as the individual components [47]. Perhaps the most striking result of the AL containing fibers is that they produced a higher Polymers 2021, 13, 992 9 of 16

Polymers 2021, 13, x FOR PEER REVIEW 9 of 17

char yield than the AL powder and approximately the same char yield as the PAN-MA fibers. This positive synergy was considered to be a good result.

Polymers 2021, 13, x FOR PEER REVIEW 10 of 17

◦ FigureFigure 5.5. TGA resultsresults forfor powderpowder rawraw materialsmaterials (10(10 °C/min,C/min, nitrogen). nitrogen).

The TGA results for AL-containing electrospun nanofibers are given in Figure 6. The results for the control fiber (PAN-MA) were comparable to the PAN-MA powder, with derivative peaks (points 1, 2, 4) located at the same temperature as in Figure 5. With the addition of 20% AL, the first two derivative peaks for PAN-MA (points 1,2) condensed into one (point 3), and the higher temperature peak (point 4) shifted to slightly lower tem- perature (point 5). No new peaks appeared (for example at 376 °C), perhaps because MA was present in such a low quantity. However, with the addition of 50% AL, a new deriv- ative peak appeared around 350 °C (point 7), which may be attributed to the main AL decomposition (point 8) shift to a lower temperature. The peak corresponding to point 6 was attributed to PAN-MA decomposition, as was the case in the AL-1 fiber (i.e., point 3). The main conclusion drawn from analyzing the derivative peaks is that PAN-MA and AL do not decompose independently when blended together. This is not usual for polymer blends, in which case three possibilities exist that is difficult or impossible to predict: one of the components increases the stability of the other, one of the components reduces the stability of the components, or the blend has equal stability as the individual components [47]. Perhaps the most striking result of the AL containing fibers is that they produced a higher char yield than the AL powder and approximately the same char yield as the PAN- MA fibers. This positive synergy was considered to be a good result.

FigureFigure 6.6. TGATGA resultsresults forfor AL-containingAL-containing nanofibersnanofibers as as compared compared to to pure pure PAN-MA PAN-MA fibers. fibers. Numbers Num- denotebers denote points points of discussion of discussion in the in text. the text.

TheThe TGATGA resultsresults forfor LSL-containingLSL-containing fibersfibers areare shownshown inin FigureFigure7 .7. With With the the addition addition ofof 20%20% LSL,LSL, thethe firstfirst twotwo derivativederivative peakspeaks forfor PAN-MAPAN-MA (points(points 1,2)1,2) condensedcondensed intointo oneone (point(point 3)3) atat slightlyslightly lowerlower temperature,temperature, andand thethe higherhigher temperaturetemperature peakpeak (point(point 5)5) shiftedshifted toto slightlyslightly lowerlower temperaturetemperature (point (point 6). 6). WithWith thethe additionaddition ofof 50%50% LSL,LSL, thethe lowlow temperaturetemperature peakpeak fromfrom LSL-2LSL-2 shiftedshifted furtherfurther toto thethe leftleft (lower(lower temperature,temperature, comparecompare pointpoint 44 toto 3).3). InIn addition, the high temperature PAN-MA peak (point 5) does not appear in the LSL-2 addition, the high temperature◦ PAN-MA peak (point 5) does not appear in the LSL-2 re- response,sponse, but but the the 350 350 °CC peak peak observed observed for for LSL LSL powder powder (point (point 8) does appear asas aa shouldershoulder inin LSL-2.LSL-2. TheThe derivativederivative peak peak results results are are similar similar to to AL-containing AL-containing fibers fibers as as seen seen in inFigure Figure6 . 6. However, the effect on wt % remaining was opposite, where the addition of LSL reduced the level of the high temperature plateau observed in LSL powder between 500– 750 °C. On the other hand, the late drop in mass recorded for LSL powder (750–780 °C) placed all four materials in a similar range of final char yield. A comparison of the weight loss curves for all five fibers is offered in Figure S8. De-

spite the fact that the low temperature weight loss behavior (for example, onset of weight loss) was significantly different, the final char yields at 780 °C were all in a similar range. The average char yield was about 49%. However, it is interesting that AL-2 fiber had the highest value (50.5%), while AL-1 fiber had the lowest (47.3%). Additional research would be needed to determine the statistical significance of these differences. Additionally, two of the materials (PAN-MA, LSL-2) lost weight at a higher rate than others at the cut-off temperature of 780 °C, so additional insight could be gained by investigating a higher final temperature, perhaps in the 800–900 °C range.

Polymers 2021, 13, 992 10 of 16

However, the net effect on wt % remaining was opposite, where the addition of LSL reduced ◦ Polymers 2021, 13, x FOR PEER REVIEWthe level of the high temperature plateau observed in LSL powder between 500–750 11C. of On 17 the other hand, the late drop in mass recorded for LSL powder (750–780 ◦C) placed all four materials in a similar range of final char yield.

FigureFigure 7.7. TGA resultsresults forfor LSL-containingLSL-containingnanofibers nanofibers as as compared compared to to pure pure PAN-MA PAN-MA fibers. fibers. Numbers Num- denotebers denote points points of discussion of discussion in the in text. the text.

ADifferential comparison Scanning of the weightCalorimetry loss curves(DSC) was for allused five to fibersfurther is study offered the in decomposi- Figure S8. Despitetion behavior the fact of the that blends the low at temperaturethe relatively weightlow end loss of the behavior spectrum (for (T example, < 400 °C), onset as well of ◦ weightas to measure loss) was the significantly glass transition different, temperatur the finale of char the yields blends at (Table 780 C 2). were Generally, all in a similar the Tg range.value(s) The for average a polymer char blend yield can was often about be 49%. used However, to infer information it is interesting about that the AL-2 phase fiber be- hadhavior, the highesti.e., whether value the (50.5%), blend while is homogeneous AL-1 fiber had (one the T lowestg value) (47.3%). or heterogeneous Additional research (two Tg wouldvalues). be The needed DSC toresults determine for the the powders statistical are significance given in Figure of these 8A. differences.PAN-MA powder Additionally, shows twoan exothermic of the materials peak (PAN-MA,at 309 °C, in LSL-2) agreement lost weight with the at a onset higher decomposition rate than others peaks at the from cut-off the temperature of 780 ◦C, so additional insight could be gained by investigating a higher final TGA data observed in Figure 5. This sharp◦ peak is consistent with nitrile group cyclization temperature,[48]. The peak perhaps at 353.9 in °C the for 800–900 LSL powderC range. is also consistent with TGA results from Figure Differential Scanning Calorimetry (DSC) was used to further study the decomposition 5. Interestingly, AL powder did not exhibit a DSC peak at 376 °C, although the◦ curve ap- behaviorpears to be of trending the blends upward at the as relatively it approached low end 400 of°C. the At spectrumlower temperatures, (T < 400 C),the asAL well and as to measure the glass transition temperature of the blends (Table2). Generally, the LSL powder curves both show a broad endotherm between 40–120 °C. With further in- T value(s) for a polymer blend can often be used to infer information about the phase vestigation,g this was attributed to the evaporation of water absorbed in the polymer (ap- behavior, i.e., whether the blend is homogeneous (one T value) or heterogeneous (two proximately 5 wt %). g T values). The DSC results for the powders are given in Figure8A. PAN-MA powder g DSC results for the electrospun fibers are given in Figure 8C. The main decomposi- shows an exothermic peak at 309 ◦C, in agreement with the onset decomposition peaks tion peak of PAN-MA fiber is the same as the powder (309 °C). When PAN-MA was from the TGA data observed in Figure5. This sharp peak is consistent with nitrile group blended with increasing amounts of AL, this peak temperature increased, while PAN-MA cyclization [48]. The peak at 353.9 ◦C for LSL powder is also consistent with TGA results blended with LSL resulted in a decreasing peak temperature. This trend is consistent with from Figure5. Interestingly, AL powder did not exhibit a DSC peak at 376 ◦C, although the the TGA results given in Figures 6 and 7. To more closely evaluate Tg results, the curves curve appears to be trending upward as it approached 400 ◦C. At lower temperatures, the were enlarged and offset vertically to give Figure 8C. The irregularity of these curves AL and LSL powder curves both show a broad endotherm between 40–120 ◦C. With further investigation,made it impossible this was to attributeddetermine tothe the Tg evaporation of each material. of water Therefore, absorbed modulated in the polymer DSC (approximately(mDSC) was conducted 5 wt %). to help resolve these values, as given in Figure 8D. These results show that all the fibers exhibited a clear Tg around 97 °C. There was no apparent shift with the AL-containing fibers relative to pure PAN-MA, while the LSL containing fibers exhib- ited a 1 or 2 °C shift to higher temperatures (although this may be within the uncertainty of the measurement). Surprisingly, no indication of a glass transition was observed for the MA or LSL component of the blend, which was expected to occur around 140 °C based on the results of the pure powders (Figure 8E). The possible exception was for the AL-2 fiber (50% AL) in which it may be possible to interpret a broad step down between 110–160 °C. The Tg event for the PAN-MA powder and AL powder were of similar strength, while the

Polymers 2021, 13, 992 11 of 16

Table 2. Glass transition temperature (Tg) and decomposition temperature (Td) [endothermic peaks] of 10 wt % electrospun PAN-MA/lignin nanofibers.

◦ ◦ Sample ID Tg, CTd (Exo), C

Polymers 2021, 13, x FOR PEER REVIEW PAN-MA 97.4 309.1 12 of 17 AL-1 98.0 311.6 AL-2 97.4 314.4 LSL-1 98.4 299.7 LSL powder Tg was weaker and somewhat broader. The Tg results, while interesting, were LSL-2 99.9 281.5 not sufficient to allow us to make conclusions on the heterogeneity of the fibers.

FigureFigure 8. (A )8. DSC (A) DSC results results for for powder powder samples, samples, ( B(B)) DSC DSC resultsresults for for fiber fiber samples, samples, (C ()C close) close−up− ofup DSC of DSCresults results for fiber for fiber samplessamples in Tg in range, Tg range, illustrating illustrating difficulty difficulty of of assigning assigning TgTg values, (D (D) )MDSC MDSC results results for forfiber fiber samples, samples, with withclear glass clear glass transitions,transitions, and (andE) MDSC (E) MDSC results results for for powder powder samples. samples.

Polymers 2021, 13, 992 12 of 16

DSC results for the electrospun fibers are given in Figure8C. The main decomposition peak of PAN-MA fiber is the same as the powder (309 ◦C). When PAN-MA was blended with increasing amounts of AL, this peak temperature increased, while PAN-MA blended with LSL resulted in a decreasing peak temperature. This trend is consistent with the TGA results given in Figures6 and7. To more closely evaluate T g results, the curves were enlarged and offset vertically to give Figure8C. The irregularity of these curves made Polymers 2021, 13, x FOR PEER REVIEW 13 of 17 it impossible to determine the Tg of each material. Therefore, modulated DSC (mDSC) was conducted to help resolve these values, as given in Figure8D. These results show ◦ that all the fibers exhibited a clear Tg around 97 C. There was no apparent shift with the AL-containingTable 2. Glass transition fibers relative temperature to pure (Tg PAN-MA,) and decomposition while the temperature LSL containing (Td) [endothermic fibers exhibited a 1peaks] or 2 ◦ofC 10 shift wt to% electros higherpun temperatures PAN-MA/lignin (although nanofibers. this may be within the uncertainty of the

measurement).Sample Surprisingly, ID no indication Tg of, °C a glass transition was observed Td (Exo), °C for the MA ◦ or LSL componentPAN-MA of the blend, which was97.4 expected to occur around 140 309.1C based on the results of the pure powders (Figure8E). The possible exception was for the AL-2 fiber (50% AL-1 98.0 311.6 AL) in which it may be possible to interpret a broad step down between 110–160 ◦C. The AL-2 97.4 314.4 Tg event for the PAN-MA powder and AL powder were of similar strength, while the LSL LSL-1 98.4 299.7 powder Tg was weaker and somewhat broader. The Tg results, while interesting, were not sufficient toLSL-2 allow us to make conclusions on99.9 the heterogeneity of the fibers. 281.5 The chemical characterization of the polymerpolymer powderspowders andand electrospunelectrospun nanofibersnanofibers were performed using ATR-FTIR. The FTIR spectrumspectrum for aa PAN-MAPAN-MA powderpowder samplesample − − shows characteristic stretching vibration for -C≡NN at at 2243 2243 cm cm−1, 1C=O, C=O at at1732 1732 cm cm−1, and1, and C- − C-OO at at1170 1170 cm cm−1 (Figure1 (Figure 9A)9 [49].A) [ 49These]. These vibrations vibrations werewere not observed not observed in either in either lignin lignin pow- powdersders (Figure (Figure 9A; LSL9A; LSLand andAL). AL).Within Within the fingerprint the fingerprint region, region, the LSL the powder LSL powder sample sample spec- −1 spectrumtrum shows shows peaks peaks at 1583, at 1583,1514, 1514,and 1420 and cm 1420−1, which cm , were which assigned were assigned to aromatic to aromatic skeletal skeletalvibrations vibrations [33,50]. [Similar33,50]. Similarpeaks were peaks also were found also foundfor AL for powders AL powders at 1583, at 1583,1514, 1514,and 1420 and −1 1420cm−1. cmTwo .additional Two additional bands bands characteristic characteristic of lignin of lignin were wereobserved observed for LSL for powders LSL powders at 1260 at −1 −1 1260cm−1 cmfor C-Ofor stretching C-O stretching and andat 1218 at 1218 cm− cm1 for C-C/C=O/C-Ofor C-C/C=O/C-O [50]. Slight [50]. Slight shifts shifts were were also −1 −1 alsofound found for AL for powder AL powder samples samples to 1267 to 1267 cm cm−1 andand 1218 1218 cm− cm1. These. These slight slight differences differences be- betweentween LSL LSL and and AL AL powders powders could could result result from from the thedesulfonation desulfonation process, process, required required to ob- to obtaintain low low sulfonate sulfonate lignin lignin powders powders [51]. [ 51A ].comp A completelete IR spectra IR spectra for all for powders all powders can be can found be foundin Supplemental in Supplemental Figure FigureS9. S9.

FigureFigure 9.9. FTIR spectra of (A)) LSL,LSL, AL,AL, andand PANPAN−−MA (control) powders and (B) electrospun nanofibersnanofibers ofof PANPAN−−MA fibersfibers (control),(control), LSLLSL−−1 and AL −1.1. Both Both lignin lignin samples samples had had the the same same PAN PAN−MA/ligninMA/lignin blend blend ratio ratio (80/20) (80/20) and and were were prepared prepared at at a 10 wt % concentration in DMF. a 10 wt % concentration in DMF.

FTIR analyses of PAN-MA/ligninPAN-MA/lignin electros electrospunpun nanofibers nanofibers were were undertaken to con- firmfirm the the presence presence of of lignin lignin and and to togain gain insights insights into into lignin lignin and andPAN-MA PAN-MA interactions. interactions. Fig- ure 9B displays the IR spectra for the 80/20 PAN-MA lignin fibers prepared at a 10% total polymer content (AL-1 and LSL-1). Additionally, an IR spectrum for a PAN-MA nano- fibers control sample is shown. As expected, all electrospun PAN-MA/lignin fibers pos- sess the characteristic PAN-MA peaks (2243 cm−1, 1732 cm−1, and 1170 cm−1). Broad -OH (~3300 cm−1) and CH2 stretching (2930 cm−1) peaks are typically found in lignin samples [52], and were observed in LSL-1 and AL-1 but not in PAN-MA (Figure 9B). These results confirm the presence of lignin within PAN-MA/lignin blends’ electrospun nanofibers. Note that the PAN-MA/lignin blends show peak shifting as compared to neat powder samples. For instance, the carbonyl stretching vibration associated with aromatic rings

Polymers 2021, 13, 992 13 of 16

Figure9B displays the IR spectra for the 80/20 PAN-MA lignin fibers prepared at a 10% total polymer content (AL-1 and LSL-1). Additionally, an IR spectrum for a PAN-MA nanofibers control sample is shown. As expected, all electrospun PAN-MA/lignin fibers possess the characteristic PAN-MA peaks (2243 cm−1, 1732 cm−1, and 1170 cm−1). Broad −1 −1 -OH (~3300 cm ) and CH2 stretching (2930 cm ) peaks are typically found in lignin samples [52], and were observed in LSL-1 and AL-1 but not in PAN-MA (Figure9B). These results confirm the presence of lignin within PAN-MA/lignin blends’ electrospun nanofibers. Note that the PAN-MA/lignin blends show peak shifting as compared to neat powder samples. For instance, the carbonyl stretching vibration associated with aromatic rings was shifted to a higher wavenumber of ~1600 cm−1 for both electrospun lignin fibers (AL-1 and LSL-1). This shift is attributed to potential interactions between lignin and PAN-MA [53]. The C-O stretching peak at 1260 cm−1, characteristic for lignin, was also found in both PAN-MA/lignin electrospun nanofibers. Compared to the samples’ powder IR spectra, a red shift from 1267 cm−1 to 1260 cm−1 was observed only for the AL-1 blend; however, this shift was not observed for the LSL-1 sample. This finding supports that AL compatibility with PAN-MA is superior as compared to PAN-MA/LSL blends, confirming microscopy observations.

4. Conclusions In this study, we demonstrated the successful production of electrospun nanofibers with the addition of alkali, Kraft lignin (AL) to PAN-MA at blend ratios of up to 50 wt % with a total polymer content up to 20 wt % in DMF solvent. These nanofibers exhibited PAN-MA/AL compatibility, a smooth morphology, and nanofiber diameters in the sub- micron scale region, as confirmed via microscopy and infrared spectroscopy measurements. As compared to neat PAN-MA, AL presence did not affect the fibers’ glass transition temperature (~97 ◦C) and did not affect the char yield (~47%) when heated in nitrogen to 800 ◦C. For PAN-MA/lignin blends, we found a minimum viscosity threshold of ~0.3 Pa-s to produce smooth electrospun nanofibers without a nodular morphology, independently of the lignin type. For LSL, however, we found poor compatibility with PAN-MA, which resulted in polymer phase separation, electrospun nanofibers with a nodular morphology, and polymer solutions prone to electrospraying. Ultimately, the degree of lignin sulfonation dictates the properties of high-quality lignin-based electrospun nanofibers. We conclude that PAN-MA/AL nanocomposite fibers could replace PAN or PAN-MA as precursors to producing carbon fibers or directly in other applications such as membranes, absorbents, supercapacitors, or electrical materials.

Supplementary Materials: The following are available online at https://www.mdpi.com/2073-4 360/13/7/992/s1, Figure S1: Polarized light optical microscopy image of PAN-co-MA dried films. Figure S2: SEM image of polymer solution blend film; (a) AL-1, (b) LSL-1, (c) AL-2 and (d) LSL-2. Inset in b and d shows enlarged LSL clusters at 10 µm. Insets in b and d show clusters formed at 10 µm. Figure S3: Frequency sweep data of storage modulus (G’) (black dots) and loss modulus (G”) (red dots) for PAN-MA/lignin blends, collected at a 1% strain amplitude. Figure S4: HR-SEM image of AL-1 nanofibers showing smooth structure. Figure S5: Low-magnification SEM images of LSL-2 and AL-4 samples show (A) large aggregates were present during the electrospinning of LSL-2 samples and (B) inter-fiber bonding was observed for sample AL-4. Figure S6: Polarized optical light microscopy images of electrospun nanofibers at 50 magnification for (a) PAN-MA; (b) AL-1; and (c) AL-3 samples. Figure S7: One-way ANOVA (Tukey’s) test to estimate significant differences for electrospun nanofiber average diameters based on SEM measurements. Figure S8: Compiled TGA weight loss results for electrospun fibers. Figure S9: Full FTIR spectra for AL-1, LSL-1, and PAN-MA control sample (referred to as PAN in the plot). Polymers 2021, 13, 992 14 of 16

Author Contributions: Conceptualization, methodology, experimental execution, and data collection, writing the initial draft and preparation, visualization, and figures preparation, E.S.V. and S.D.; methodology and experimental planning, initial draft contribution, S.M.N.A.-A.; review, editing, DSC data collection, thermal analysis interpretation, and graph preparation, D.A.K. and IR and microscopy analysis discussion, review and editing, project administration, E.S.V., G.S.C., and K.M.C. All authors have read and agreed to the published version of the manuscript. Funding: This research received no external funding. Data Availability Statement: The data presented in this study are available on request from the corresponding author. Acknowledgments: We acknowledge the Chemical and Materials Engineering Department at the University of Dayton for providing access to research instrumentation. Funding for Open Access provided by the University of Dayton Open Access Fund. Conflicts of Interest: The authors declare no conflict of interest.

References 1. Shanmugam, K.; Gadhamshetty, V.; Yadav, P.; Athanassiadis, D.; Tysklind, M.; Upadhyayula, V.K.K. Advanced high-strength steel and carbon fiber reinforced polymer composite body in white for passenger cars: Environmental performance and sustainable return on investment under different propulsion modes. ACS Sustain. Chem. Eng. 2019, 7, 4951–4963. [CrossRef] 2. Zhang, X.; Fan, X.; Yan, C.; Li, H.; Zhu, Y.; Li, X.; Yu, L. Interfacial microstructure and properties of carbon fiber composites modified with graphene oxide. ACS Appl. Mater. Interfaces 2012, 4, 1543–1552. [CrossRef] 3. Al-Ajrash, S.M.N.; Lafdi, K.; Vasquez, E.S.; Chinesta, F.; Le Coustumer, P. Experimental and Numerical Investigation of the Silicon Particle Distribution in Electrospun Nanofibers. Langmuir 2018, 34.[CrossRef] 4. Sabantina, L.; Rodríguez-Cano, M.Á.; Klöcker, M.; García-Mateos, F.J.; Ternero-Hidalgo, J.J.; Mamun, A.; Beermann, F.; Schwakenberg, M.; Voigt, A.-L.; Rodríguez-Mirasol, J. Fixing PAN nanofiber mats during stabilization for carbonization and creating novel metal/carbon composites. Polymers 2018, 10, 735. [CrossRef] 5. Zheng, Y.; Chen, L.; Wang, X.; Wu, G. Modification of renewable cardanol onto carbon fiber for the improved interfacial properties of advanced polymer composites. Polymers 2020, 12, 45. [CrossRef] 6. Ago, M.; Okajima, K.; Jakes, J.E.; Park, S.; Rojas, O.J. Lignin-based electrospun nanofibers reinforced with cellulose nanocrystals. Biomacromolecules 2012, 13, 918–926. [CrossRef][PubMed] 7. Liu, H.; Dai, Z.; Cao, Q.; Shi, X.; Wang, X.; Li, H.; Han, Y.; Li, Y.; Zhou, J. Lignin/Polyacrylonitrile : The E ff ect of Fractionation and Purification on Properties of Derived Carbon Fibers. ACS Sustain. Chem. Eng. 2018.[CrossRef] 8. El-Naggar, M.E.; Abdelgawad, A.M.; Salas, C.; Rojas, O.J. Curdlan in fibers as carriers of tetracycline hydrochloride: Controlled release and antibacterial activity. Carbohydr. Polym. 2016, 154, 194–203. [CrossRef][PubMed] 9. El-Newehy, M.H.; El-Naggar, M.E.; Alotaiby, S.; El-Hamshary, H.; Moydeen, M.; Al-Deyab, S. Preparation of biocompatible system based on electrospun CMC/PVA nanofibers as controlled release carrier of diclofenac sodium. J. Macromol. Sci. Part A Pure Appl. Chem. 2016, 53, 566–573. [CrossRef] 10. Figueiredo, P.; Lintinen, K.; Hirvonen, J.T.; Kostiainen, M.A.; Santos, H.A. Properties and chemical modifications of lignin: Towards lignin-based nanomaterials for biomedical applications. Prog. Mater. Sci. 2018, 93, 233–269. [CrossRef] 11. Baker, D.A.; Rials, T.G. Recent advances in low-cost carbon fiber manufacture from lignin. J. Appl. Polym. Sci. 2013, 130, 713–728. [CrossRef] 12. Liengprayoon, S.; Suphamitmongkol, W.; Jantarasunthorn, S.; Rungjang, W.; Sunthornvarabhas, J.; Tanthana, J. Investigation of the potential for utilization of sugarcane bagasse lignin for carbon fiber production: Thailand case study. SN Appl. Sci. 2019, 1, 115. [CrossRef] 13. Kumar, A.; Sinha-Ray, S. A review on biopolymer-based fibers via electrospinning and solution blowing and their applications. Fibers 2018, 6, 45. [CrossRef] 14. Kumar, M.; Hietala, M.; Oksman, K. Lignin-based electrospun carbon nanofibers. Front. Mater. 2019, 6, 1–6. [CrossRef] 15. El-Newehy, M.H.; El-Naggar, M.E.; Alotaiby, S.; El-Hamshary, H.; Moydeen, M.; Al-Deyab, S. Green Electrospining of Hydroxypropyl Cellulose Nanofibres for Drug Delivery Applications. J. Nanosci. Nanotechnol. 2017, 18, 805–814. [CrossRef] [PubMed] 16. Sharaf, S.; El-Naggar, M.E. Eco-friendly technology for preparation, characterization and promotion of honey bee propolis extract loaded nanofibers in medical domains. Cellulose 2018, 25, 5195–5204. [CrossRef] 17. Abdelgawad, A.M.; El-Naggar, M.E.; Hudson, S.M.; Rojas, O.J. Fabrication and characterization of bactericidal thiol-chitosan and chitosan iodoacetamide nanofibres. Int. J. Biol. Macromol. 2017, 94, 96–105. [CrossRef] 18. Aldalbahi, A.; El-Naggar, M.E.; Ahmed, M.K.; Periyasami, G.; Rahaman, M.; Menazea, A.A. Coreeshell Au@Se nanoparticles embedded in cellulose acetate/polyvinylidene fluoride scaffold for wound healing. J. Mater. Res. Technol. 2020, 9, 15045–15056. [CrossRef] 19. Jin, J.; Yu, B.J.; Shi, Z.Q.; Wang, C.Y.; Chong, C. Bin Lignin-based electrospun carbon nanofibrous webs as free-standing and binder-free electrodes for sodium ion batteries. J. Power Sources 2014, 272, 800–807. [CrossRef] Polymers 2021, 13, 992 15 of 16

20. Lu, Y.; Liu, Z.; Woon, S.; Mcloughlin, L.; Bridgers, B.; Hayes, S.; Wang, X.; Wang, R.; Guo, Z.; Wujcik, E.K. Electrospun carbon/iron nano fi bers: The catalytic effects of iron and application in Cr (VI) removal. Carbon N. Y. 2020, 166, 227–244. [CrossRef] 21. SantosMiranda, M.E.; Marcolla, C.; Rodriguez, C.A.; Wilhelm, H.M.; Sierakowski, M.R.; BelleBresolin, T.M.; Alves de Freitas, R.I. The role of N-carboxymethylation of chitosan in the thermal stability and dynamic. Polym. Int. 2006, 55, 961–969. [CrossRef] 22. Matulevicius, J.; Kliucininkas, L.; Martuzevicius, D.; Krugly, E.; Tichonovas, M.; Baltrusaitis, J. Design and characterization of electrospun polyamide nanofiber media for air filtration applications. J. Nanomater. 2014, 2014.[CrossRef] 23. Svinterikos, E.; Zuburtikudis, I.; Al-Marzouqi, M. Electrospun Lignin-Derived Carbon Micro- And Nanofibers: A Review on Precursors, Properties, and Applications. ACS Sustain. Chem. Eng. 2020, 8, 13868–13893. [CrossRef] 24. Adam, A.A.; Dennis, J.O.; Al-Hadeethi, Y.; Mkawi, E.M.; Abdulkadir, B.A.; Usman, F.; Hassan, Y.M.; Wadi, I.A.; Sani, M. State of the art and new directions on electrospun lignin/cellulose nanofibers for supercapacitor application: A systematic literature review. Polymers 2020, 12, 2884. [CrossRef][PubMed] 25. Demiro˘gluMustafov, S.; Mohanty, A.K.; Misra, M.; Seydibeyo˘glu,M.Ö. Fabrication of conductive Lignin/PAN carbon nanofiber with enhanced graphene for the modified electrode. Carbon N. Y. 2019, 147, 262–275. [CrossRef] 26. Dalton, N.; Lynch, R.P.; Collins, M.N.; Culebras, M. Thermoelectric properties of electrospun carbon nanofibres derived from lignin. Int. J. Biol. Macromol. 2019, 121, 472–479. [CrossRef] 27. Nordin, N.A.; Rahman, N.A.; Abdullah, A.H. Effective Removal of Pb(II) Ions by Electrospun PAN/Sago Lignin-Based Activated Carbon Nanofibers. Molecules 2020, 25, 3081. [CrossRef] 28. Domínguez, J.C.; Oliet, M.; Alonso, M.V.; Gilarranz, M.A.; Rodríguez, F. Thermal stability and pyrolysis kinetics of organosolv lignins obtained from Eucalyptus globulus. Ind. Crops Prod. 2008, 27, 150–156. [CrossRef] 29. Aadil, K.R.; Mussatto, S.I.; Jha, H. Synthesis and characterization of silver nanoparticles loaded poly(vinyl alcohol)-lignin electrospun nanofibers and their antimicrobial activity. Int. J. Biol. Macromol. 2018, 120, 763–767. [CrossRef] 30. Poursorkhabi, V.; Mohanty, A.K.; Misra, M. Electrospinning of aqueous lignin/poly(ethylene oxide) complexes. J. Appl. Polym. Sci. 2015, 132, 1–9. [CrossRef] 31. Vivo-Vilches, J.F.; Celzard, A.; Fierro, V.; Devin-Ziegler, I.; Brosse, N.; Dufour, A.; Etienne, M. Lignin-based carbon nanofibers as electrodes for vanadium redox couple electrochemistry. Nanomaterials 2019, 9, 106. [CrossRef] 32. Du, L.; Xu, H.; Zhang, Y.; Zou, F. Electrospinning of polycaprolatone nanofibers with DMF additive: The effect of solution proprieties on jet perturbation and fiber morphologies. Fibers Polym. 2016, 17, 751–759. [CrossRef] 33. Oroumei, A.; Fox, B.; Naebe, M. Thermal and rheological characteristics of biobased carbon fiber precursor derived from low molecular weight organosolv lignin. ACS Sustain. Chem. Eng. 2015, 3, 758–769. [CrossRef] 34. Beck, R.J.; Zhao, Y.; Fong, H.; Menkhaus, T.J. Journal of Water Process Engineering Electrospun lignin carbon nanofiber membranes with large pores for highly efficient adsorptive water treatment applications. J. Water Process Eng. 2017, 16, 240–248. [CrossRef] 35. Dallmeyer, I.; Chowdhury, S.; Kadla, J.F. Preparation and characterization of kraft lignin-based moisture-responsive films with reversible shape-change capability. Biomacromolecules 2013, 14, 2354–2363. [CrossRef][PubMed] 36. Cho, M.; Ko, F.K.; Renneckar, S. Molecular Orientation and Organization of Technical Lignin-Based Composite Nanofibers and Films. Biomacromolecules 2019.[CrossRef] 37. Zhang, H.; Quan, L.; Gao, A.; Tong, Y.; Shi, F. Thermal Analysis and Crystal Structure of Poly(Acrylonitrile-Co-Itaconic Acid) Copolymers Synthesized in Water. Polymers. 2020, 12, 221. [CrossRef] 38. Park, D.U.; Ryu, J.H.; Han, N.K.; Park, W.H.; Jeong, Y.G. Thermal Analysis on the Stabilization Behavior of Ternary Copolymers Based on Acrylonitrile, Methyl Acrylate and Itaconic Acid. Fibers Polym. 2018, 19, 2439–2448. [CrossRef] 39. Seydibeyolu, M.Ö. A novel partially biobased pan-lignin blend as a potential carbon fiber precursor. J. Biomed. Biotechnol. 2012, 2012.[CrossRef] 40. Zhang, D.; Karki, A.B.; Rutman, D.; Young, D.P.; Wang, A.; Cocke, D.; Ho, T.H.; Guo, Z. Electrospun polyacrylonitrile nanocom- posite fibers reinforced with Fe3O4 nanoparticles: Fabrication and property analysis. Polymer 2009, 50, 4189–4198. [CrossRef] 41. Rošic, R.; Pelipenko, J.; Kocbek, P.; Baumgartner, S.; Bešter-Rogaˇc,M.; Kristl, J. The role of rheology of polymer solutions in predicting nanofiber formation by electrospinning. Eur. Polym. J. 2012, 48, 1374–1384. [CrossRef] 42. Suresh, S.; Becker, A.; Glasmacher, B. Impact of apparatus orientation and gravity in electrospinning—A review of empirical evidence. Polymers 2020, 12, 2448. [CrossRef][PubMed] 43. Painter, P.C.; Coleman, M.M. Fundamentals of Polymer Science: An Introductory Text; Technomic: Chicago, IL, USA, 1997; ISBN 1566765595. 44. Rauwendaal, C. Polymer Extrusion 5E; Hanser Publishers: Munich, Germany, 2014. 45. Whittingstall, P. Pitfalls in Measurement of non-Newtonian Materials—“It’s a Line, Not a Point!”. In Limitations of Test Methods for Plastics; ASTM International: West Conshohocken, PSA, USA, 2000. 46. Yang, H.; Yan, R.; Chen, H.; Lee, D.H.; Zheng, C. Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel 2007, 86, 1781–1788. [CrossRef] 47. Day, M.; Cooney, J.D.; Klein, C.; Fox, J. Thermal degradation of automotive plastics: A possible recycling opportunity. Adv. Chem. Ser. 1996, 249, 56–57. [CrossRef] 48. Maradur, S.P.; Kim, C.H.; Kim, S.Y.; Kim, B.H.; Kim, W.C.; Yang, K.S. Preparation of carbon fibers from a lignin copolymer with polyacrylonitrile. Synth. Met. 2012, 162, 453–459. [CrossRef] Polymers 2021, 13, 992 16 of 16

49. Jiang, X.; Ouyang, Q.; Liu, D.; Huang, J.; Ma, H.; Chen, Y.; Wang, X. Preparation of low-cost carbon fiber precursors from blends of wheat straw lignin and commercial -grade polyacrylonitrile (PAN). Holzforschung 2018, 72, 727–734. [CrossRef] 50. Horikawa, Y.; Hirano, S.; Mihashi, A.; Kobayashi, Y.; Zhai, S.; Sugiyama, J. Prediction of Lignin Contents from Infrared Spectroscopy: Chemical Digestion and Lignin/Biomass Ratios of Cryptomeria japonica. Appl. Biochem. Biotechnol. 2019, 188, 1066–1076. [CrossRef] 51. Aro, T.; Fatehi, P. Production and Application of Lignosulfonates and Sulfonated Lignin. ChemSusChem 2017, 10, 1861–1877. [CrossRef] 52. Lu, Y.; Lu, Y.C.; Hu, H.Q.; Xie, F.J.; Wei, X.Y.; Fan, X. Structural characterization of lignin and its degradation products with spectroscopic methods. J. Spectrosc. 2017.[CrossRef] 53. Sargent, M.; Koenig, J.L. Fourier Transform IR Spectroscopic Analysis of the Molecular Structure of Compatible Polymer Blends. Adv. Chem. 1993, 236, 191–219.