Next Article in Journal
Effects of Apple Vinegar, Mouthwashes, and Bleaching on Color Stability and Surface Properties of Fiber-Reinforced and Non-Reinforced Restorative Materials
Previous Article in Journal
Scratch Resistance and Tribological Enhancement of Epoxy Composites Reinforced with Chopped Glass Fiber and Nano Silica Through Taguchi Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interphase-Resolved Performance in PA6/TiO2 Nanocomposite Fibers: Four-Phase Geometry Linking Structure to Mechanical and UV Protection

1
State Key Laboratory of Advanced Fiber Materials, College of Materials Science and Engineering, Donghua University, Shanghai 201620, China
2
Shaoxing Huiqun New Material Technology Co., Ltd., Shaoxing 312000, China
*
Authors to whom correspondence should be addressed.
Polymers 2025, 17(18), 2551; https://doi.org/10.3390/polym17182551
Submission received: 28 August 2025 / Revised: 16 September 2025 / Accepted: 17 September 2025 / Published: 21 September 2025
(This article belongs to the Section Polymer Fibers)

Abstract

Melt-spun PA6/TiO2 fibers with TiO2 modified by silane coupling agents KH550 and KH570 at 0, 1.6, and 4 wt% provide a practical testbed to address three fiber-centric gaps: transferable interphase quantification, interphase-resolved indications of compatibility, and a reproducible kinetics–structure–property link. This work proposes, for the first time at fiber scale, a four-phase partition into crystal (c), crystal-adjacent rigid amorphous fraction (RAF-c), interfacial rigid amorphous fraction (RAF-i), and mobile amorphous fraction (MAF), and extracts an interfacial triad consisting of the specific interfacial area (Sv), polymer-only RAF-i fraction expressed per composite volume (Γi), and interphase thickness (ti) from SAXS invariants to establish a quantitative interphase-structure–property framework. A documented SAXS/DSC/WAXS workflow partitions the polymer into the above four components on a polymer-only basis. Upon filling, Γi increases while RAF-c decreases, leaving the total RAF approximately conserved. Under identical cooling, DSC shows the crystallization peak temperature is higher by 1.6–4.3 °C and has longer half-times, indicating enhanced heterogeneous nucleation together with growth are increasingly limited by interphase confinement. At 4 wt% loading, KH570-modified fibers versus KH550-modified fibers exhibit higher α-phase orientation (Hermans factor f(α): 0.697 vs. 0.414) but an ~89.4% lower α/γ ratio. At the macroscale, compared to pure (neat) PA6, 4 wt% KH550- and KH570-modified fibers show tenacity enhancements of ~9.5% and ~33.3%, with elongation decreased by ~31–68%. These trends reflect orientation-driven stiffening accompanied by a reduction in the mobile amorphous fraction and stronger interphase constraints on chain mobility. Knitted fabrics achieve a UV protection factor (UPF) of at least 50, whereas pure PA6 fabrics show only ~5.0, corresponding to ≥16-fold improvement. Taken together, the SAXS-derived descriptors (Sv, Γi, ti) provide transferable interphase quantification and, together with WAXS and DSC, yield a reproducible link from interfacial geometry to kinetics, structure, and properties, revealing two limiting regimes—orientation-dominated and phase-fraction-dominated.

1. Introduction

Polyamide-6 (PA6)/TiO2 nanocomposite fibers are compelling for advanced textiles, filtration, and protective garments because they can combine mechanical robustness, thermal stability, and ultraviolet (UV) shielding within a single platform. Beyond overall loading, macroscopic behavior reflects the combined effects of interfacial chemistry and interphase geometry, via constrained chain dynamics that influence non-isothermal crystallization, orientation during drawing, and long-term durability. Within semicrystalline polymers, the rigid amorphous fraction (RAF) provides a physically grounded descriptor for such constrained domains and has long been linked to crystallization pathways and morphology development [1,2,3,4,5,6]. This conceptualization aligns with long-standing SAXS/WAXS and crystallization practice in polyamides, including precedents on non-isothermal kinetics and X-ray-based crystallinity quantification [7,8,9]; and is also consistent with metastable-state interpretations of semicrystalline transitions [10,11,12]. Organic–inorganic hybrid fiber systems therefore provide an effective route to integrate functional synergy with interfacial regulation, and recent reviews have summarized hybridization and interface-engineering strategies for multifunctional fibers [13,14]. These studies underline the centrality of interfacial design, and complementary analyses of TiO2 nanocomposites emphasize that an enlarged interfacial area produces a significant polymer fraction with properties distinct from the bulk, yet reproducible quantification remains under development [15]. A systematic framework that resolves interphase geometry and connects it to mechanical and UV-protection performance is therefore needed to advance PA6/TiO2 fibers.
Recent works have expanded RAF analysis to diverse polymer systems, covering its temperature-dependent quantification [16], development under cold-crystallization conditions [17], and even its use as an indicator for polymer–polymer interactions [18]. These developments, however, remain rooted in the classical three-phase picture crystal(c), RAF, and MAF—which continues to provide powerful framework for semicrystalline polymers [1,2,5,8,19]. In nanocomposites, RAF can be apportioned by origin into RAF-c and RAF-i, enabling interface-resolved analysis of constrained domains without abandoning the three-phase foundation [20]. Four-phase RAF partitioning has been implemented in bulk/thin-film nanocomposites and linked to kinetic/dynamical readouts [18], yet its application to melt-spun fibers is technically challenging: invariant-based SAXS workflows require stringent procedures, and fiber anisotropy couples Porod terms with α/γ orientation. Precedents exist in thin films and electrospun mats [21,22], but fiber studies rarely adopt geometry-normalized and interphase-resolved descriptors that are transferable across processing conditions [23,24]. In parallel, the confined mechanical response of RAF has been directly quantified [25], and its thermal stability recently linked to lamellar thickening in nascent iPPs [26].
In melt-spun PA6/TiO2 fibers, high TiO2 loadings induce aggregation and weak interfaces, undermining strength and UV gains. Silane surface functionalization mitigates these drawbacks [27,28,29], and polymeric compatibilizers reported for PA6 can further reinforce the interface [30]. Recent PA6-fiber studies have mapped process–property windows, grafting-based thermal/UV stabilization, and interphase tailoring via macromolecular anchoring [31,32,33]. Yet fiber-relevant and reproducible descriptors that capture the geometry and continuity of the interfacial constrained layer under melt-spinning remain limited [15,34]. Against this backdrop, three gaps remain: (i) transferable fiber-scale quantification of the interphase-constrained amorphous fraction is scarce (downsizing amplifies interfacial area and yields fractions distinct from bulk [20,24]); (ii) routine XRD/WAXS, SAXS, DSC, and microscopy assess dispersion and compatibility but rarely provide interphase-resolved, geometry-normalized metrics for the geometry/redistribution of constrained amorphous material [35,36,37,38,39]; and (iii) a reproducible, in-one-study bridge linking non-isothermal crystallization kinetics [40,41,42,43], WAXS-derived orientation and polymorphism, and properties remains uncommon [3,23,44]. Recent studies further highlight these issues: interphase-modulated RAF effects were observed in PA66/MMT nanocomposites [45], strain-dependent RAF evolution was revealed under deformation [46]. Together with recent insights into RAF stability and confinement, these findings reinforce the relevance of addressing the three gaps identified here at the fiber scale.
Here, we address these gaps by extending the classical three-phase picture to a polymer-only four-phase partition (c, RAF-c, RAF-i, MAF). On this basis, SAXS Porod–Debye invariants after crystalline-peak subtraction provide the specific interfacial area (Sv) [47,48,49], which, combined with DSC crystallinity, yields an interfacial-allocation descriptor (Γi); together, they are summarized by an effective thickness (ti) as geometry-normalized readouts at the fiber scale. These descriptors are then linked to non-isothermal crystallization kinetics, WAXS-derived α/γ phase fractions and Hermans orientation [50], and non-isothermal kinetics are evaluated via Nakamura global fitting with complementary Kissinger/KAS closures consistent with ICTAC recommendations [40,41,42,43,51]. Complementary characterizations (mechanical, rheological, SEM, UPF [52], and qualitative coarse-grained MD) and sensitivity checks are summarized in Methods/SI.
To rationalize the material selection, PA6 was chosen as a representative semi-crystalline fiber matrix for its spinnability and established structure–property framework; TiO2 nanoparticles were incorporated for their dual role as UV-shielding fillers and nucleating agents; KH550 (amine-functional) and KH570 (methacrylate-functional) silanes were selected to contrast interfacial chemistries. The novelty of this work lies in introducing interphase-resolved four-phase geometry at the fiber scale and establishing transferable SAXS-derived descriptors (Sv, Γi, ti) that reproducibly link interphase geometry to crystallization kinetics, microstructural orientation, and macroscopic performance. The overarching objective is to demonstrate that such geometry provides a unifying and transferable basis for understanding and tailoring mechanical and UV-protection properties in PA6/TiO2 nanocomposite fibers.

2. Materials and Methods

2.1. Materials and Surface Modification

Polyamide 6 (PA6, grade M2400; Xinhui Meida Chemical Co., Ltd., Jiangmen, China) from a single production lot was used for all experiments. According to the manufacturer’s specification for grade M2400, this is a fiber-grade, unfilled PA6 suitable for melt processing. Pellets were pre-crystallized in a vacuum rotary drum oven (JM-500ZDGX, Shanghai Jinma Electric Co., Ltd., Shanghai, China) at 90 °C for 18 h, followed by vacuum drying at 135 °C for 18 h immediately prior to processing to minimize residual moisture. Titanium dioxide (TiO2, TA-300, Fuji Titanium, Osaka, Japan; mean particle size—0.613 μm) was dried at 80 °C for 12 h and stored under nitrogen. The silane coupling agents γ-aminopropyltriethoxysilane (KH550, ≥98%) and γ-methacryloxypropyltrimethoxysilane (KH570, ≥98%) were purchased from Shanghai Chemical Reagent Co., Ltd., Shanghai, China and tetrabutyl titanate (TBOT, ≥99%) was supplied by Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. The mass ratio of silane to TiO2 was fixed at 5 wt%. Antioxidant Irganox 1010 (BASF SE, Ludwigshafen, Germany, 0.1 wt%), release agent HLL-1 (Covestro, Leverkusen, Germany, 0.05 wt%), and flow modifier Lotader AX8900 (Arkema S.A., 0.2 wt%, Colombes, France) were used as processing aids. Solvents included anhydrous ethanol and n-hexane (both ≥99%, Sinopharm Chemical Reagent, Shanghai, China), as well as deionized water prepared in-house.
For the surface modification of TiO2, 8.00 kg of pre-dried TiO2 powder was dispersed in 25.0 L anhydrous ethanol in a 50 L glass reactor, and 1.60 L TBOT was added. The mixture was stirred at room temperature for 60 min and left standing for 24 h. After filtration and washing with ethanol and deionized water, the product was vacuum dried at 80 °C for 12 h to yield TBOT-coated TiO2 (p-TiO2). For silane modification, 2.32 kg KH550 or KH570 was dissolved in 25.0 L ethanol, and 5.6 L deionized water was slowly added (silane: water: ethanol ≈ 1:2:8). The solution was hydrolyzed at pH 5.5 for 1 h under stirring, then p-TiO2 was gradually added. The suspension was heated to 70 °C and stirred for 4 h, then filtered, washed, and vacuum dried at 80 °C for 12 h to obtain silane-modified TiO2 nanoparticles. Although the silane-to-TiO2 ratio during modification was fixed at 5 wt%, only DLS was performed on the modified powders. The effective surface coupling was then assessed after compounding—FTIR-ATR on PA6/TiO2 masterbatches and Raman/DMA/melt rheology on composite specimens—and is reported in the Results.
The overall modification mechanism, interfacial interactions, and subsequent processing strategy are summarized schematically in Scheme 1.
In panel (a), the hydrolysis and condensation of silane coupling agents on TiO2 are illustrated. The alkoxy groups (–OR) of KH550 or KH570 undergo hydrolysis to form silanol groups (–Si–OH), which then condense with surface hydroxyl groups on TiO2 to yield covalent Si–O–Ti bonds. This reaction ensures stable anchoring of the silane molecules on the oxide surface, while the organic functional groups (–NH2 for KH550, –C(=O)O–C=C– for KH570) remain exposed for subsequent interaction with the polymer matrix.
In panel (b), the interfacial hydrogen bonding between KH550 and PA6 is depicted. The amino end group (–NH2) of KH550 acts as a hydrogen-bond donor, forming bifurcated hydrogen bonds with the carbonyl oxygen atoms of PA6 chains. This dual hydrogen-bonding interaction enhances interfacial compatibility and immobilizes the amorphous PA6 fraction adjacent to TiO2, strengthening stress transfer across the filler–matrix interface.
In panel (c), the polar interaction between KH570 and PA6 is shown. Unlike KH550, KH570 does not contain an amino group, but instead a methacrylate ester moiety. The ester carbonyl oxygen can serve as a weak hydrogen-bond acceptor with PA6 amide N–H groups. This results in less pronounced interfacial binding compared to KH550, which is reflected in weaker confinement of the amorphous PA6 phase.
Finally, panel (d) presents the overall processing route from surface modification of TiO2 powders, through masterbatch compounding at high filler loadings (up to 60 wt%), to dilution with neat PA6 and subsequent melt spinning at ≤6 wt% TiO2, followed by weaving into fabrics. This schematic highlights how surface chemistry at the nanoparticle level is integrated into fiber-scale processing, ultimately linking interfacial interactions to macroscopic textile structures.

2.2. Fiber Preparation and Processing

Dried PA6 and silane-modified TiO2 powders were premixed in specific mass ratios, together with Irganox 1010 (0.1 wt%) and HLL-1 (0.05 wt%) as processing aids. The mixtures were dried at 120 °C under vacuum for 12 h prior to compounding.
Masterbatches containing 50 wt% and 60 wt% TiO2 were prepared using a CTE20 twin-screw extruder (Nanjing KEYA, Nanjing, China; screw diameter 20 mm, L/D = 25). The barrel temperature profile was set at 235–248 °C, and the screw speed was 350 rpm. The melt was filtered through 100-mesh and 200-mesh screens, extruded through a 2 mm die, water-cooled, pelletized (2–3 mm), and vacuum dried at 110 °C for 6 h. The masterbatches were designated as “PA6–50%TiO2-KH550/570” and “PA6–60%TiO2-KH550/570”.
For fiber spinning, the masterbatches were diluted with pure (neat) PA6 to obtain different TiO2 contents. The dried blends were processed on an ABE-25 single-screw extruder (operated at ~60% channel fill, 100 rpm). The temperature zones were set at 280–300 °C, with a 36-hole spinneret (0.15 mm diameter), dual-layer filters (50- and 200-mesh), and side air cooling. The as-spun fibers were collected at 800–1000 m/min and immediately drawn and heat-set using a UDY+DT stretching machine (hot plate 60 °C, hot roll 120 °C to a fixed draw ratio (DR) 3.0. The resulting yarns were named according to the coupling agent and their TiO2 content. The samples are named and referenced as shown in Table 1 for direct comparison of TiO2 content, coupling agent, and draw ratio.
The KH550 series (0, 1.6, and 4.0 wt%) was selected to resolve concentration-dependent interfacial effects, given its stronger hydrogen-bonding affinity with PA6. In contrast, KH570 was included at 4.0 wt% as a representative high-loading case to highlight differences between coupling agents; this loading was chosen to maximize contrast in interfacial behavior, rather than to form a full concentration series.

2.3. Characterization Techniques

2.3.1. Dynamic Light Scattering (DLS)

Particle size distributions of silane-modified TiO2 powders (KH550 or KH570) were measured using a Nano ZS90 instrument (Malvern Panalytical Ltd., Malvern, UK) equipped with a 633 nm laser at a scattering angle of 90°. Powders were dispersed in absolute ethanol and ultrasonicated for 10 min prior to measurement at 25 ± 0.1 °C. Z-average diameter and polydispersity index (PDI) were obtained from intensity-weighted distributions (n = 3 independent runs) using the instrument software and in-house scripts, and results are reported as mean ± standard error of the mean (SE). A PDI value below 0.2 was considered indicative of monodispersity, following Bhattacharjee [53].

2.3.2. Filter Pressure Value (FPV) Testing

The filter pressure value was determined according to BS EN 13900-5:2005 [39] using standard melt-filtration equipment (Zibo Fangchen Masterbatch Factory, Zibo, China), without deviations from the prescribed test and calculation procedures. PA6/TiO2 masterbatches were diluted to the target TiO2 contents prior to testing [35]. FPV testing was conducted once per condition, consistent with the TBOOT (test based on one trial) approach previously validated for comparable masterbatch systems [54,55], where repeatability is ensured by the standardized procedure and narrow process tolerances. Reported values are therefore represent single standardized measurements (see Figure 1b in the Results section).

2.3.3. Fourier-Transform Infrared (FTIR) Spectroscopy

FTIR spectra were recorded using a Nicolet iS10 spectrometer (Thermo Fisher Scientific Inc., Waltham, MA, USA) in attenuated total reflectance (ATR) mode over 400–4000 cm−1, with 32 scans averaged at 4 cm−1 resolution. Samples were PA6 masterbatches containing unmodified or silane-modified TiO2 (60 wt% TiO2 in a PA6 carrier), dried at 80 °C for 12 h prior to measurement. Spectra were normalized to the 0–1 range and analyzed between 4000 and 650 cm−1. Difference spectra were obtained by least-squares scaling outside 930–1100 cm−1 to isolate silane–TiO2-related absorptions. Each spectrum represents the averaged response of multiple scans, and replicate acquisitions under identical conditions yielded consistent peak positions and relative intensities, confirming reproducibility (see representative spectra in the Results section).

2.3.4. Raman Spectroscopy

Raman spectra were collected on a Renishaw inVia Reflex confocal Raman microscope (Renishaw plc., Wotton-under-Edge, UK) using 532 nm laser excitation (50× objective, spectral range 100–3200 cm−1). Injection-molded specimens (pure PA6 and composites containing 1.6 and 4.0 wt% TiO2 with KH550 modification) were polished for reproducible optical contact. Spectra were acquired with an integration time of 10 s, with three accumulations, at a laser power of ~5 mW measured at the sample surface. Data were normalized to the 95th percentile intensity over 100–700 cm−1. The amide I region (1600–1700 cm−1) was additionally inspected by magnified plots to compare peak profiles among samples. eCharacteristic bands agree with literature assignments for PA6 and TiO2, supporting the reliability of the measurements, while the representative spectra are provided in the Results section (Figure 2).

2.3.5. Scanning Electron Microscopy (SEM) and Energy-Dispersive X-Ray Spectroscopy (EDS)

Fiber microstructure and TiO2 dispersion were examined by SEM (JSM-5600LV, JEOL Ltd., Tokyo, Japan) operated at 1–10 kV accelerating voltage. Cross-sections of melt-spun PA6/TiO2 fibers were prepared by cryofracture in liquid nitrogen and sputter-coated with a thin Au layer. Elemental mapping and semi-quantitative analysis of Ti and O were performed using an attached EDS detector under identical imaging conditions. For each condition, at least multiple regions were examined, and characteristic morphologies and Ti/O distributions were found to be consistent across replicates, ensuring reproducibility. Representative SEM/EDS micrographs from related melt-spun fibers are provided in the Supporting Information (Figure S1).

2.3.6. Differential Scanning Calorimetry (DSC)

DSC is performed on a Pyris-1 (PerkinElmer Inc., Waltham, MA, USA) using ~5–10 mg specimens sealed in standard aluminum pans. Cooling scans are conducted at β = 5, 10, 20, 30, and 40 °C·min−1 under nitrogen. Heat-flow signals are blank-corrected for TiO2 and normalized to PA6 mass (W g−1). Representative cooling curves are shown in the Results section (Figure 3a) for pure PA6 and PA6/TiO2 fibers (0, 1.6, and 4 wt% KH550; 4 wt% KH570) at 20 °C·min−1.
Relative crystallinity Xt(t) is obtained by integrating the exotherm between onset and endset determined by the tangent-intercept method, followed by 0–1 normalization. Representative plots are presented in the Results section (Figure 3b),where symbols indicate experimental data and dashed lines for Nakamura [42] fits based on the isothermal-equivalent time θ *, performed over X = 0.20–0.60, with activation energies fixed from independent isoconversional analysis.
Apparent activation energies are evaluated by Kissinger [40] using peak temperatures Tp (linear regression of ln [β/Tp2] versus 1/Tp) and by the isoconversional KAS [41,51,56] method at fixed conversions X = 0.30, 0.40, 0.50, and 0.55. Representative regressions are presented in the Results section (Figure 3c), where ln[β/Tx2] versus 1/Tx plots at different conversion levels and cooling rates, where the slopes were used to obtain Ea(X). Temperatures in regressions are expressed in Kelvin. Parameters are reported as mean ± standard deviation; uncertainties for Kissinger/KAS follow linear-regression statistics, and uncertainties for the Nakamura fit are derived from the covariance of the global model [43].
Supplementary Information provides the full DSC cooling curves (Figure S3), per-conversion KAS regressions (Figure S4), single-point Kissinger regressions (Figure S5), and a parameter/uncertainty summary (Table S1). Each DSC cooling experiment was performed once per sample per rate. This follows common practice in polymer crystallization studies, where instrument stability and standardized protocols provide reproducible results. Reproducibility is further supported by the consistency between Kissinger and KAS kinetic analyses, and uncertainties are quantified from regression statistics rather than replicate runs.

2.3.7. Synchrotron Small- and Wide-Angle X-Ray Scattering (SAXS/WAXS)

Unless otherwise noted, all phase fractions φ—φc, φRAF-c, φRAF-i, φMAF, φA—are reported on a polymer-only basis (normalized to the polymer internal volume). The interfacial volume Γi is reported per composite volume as Γi = (1 − φif)·φRAF-i, where φif is the filler volume fraction. This convention is used consistently throughout all figures and tables.
SAXS experiments were conducted at the BL16B1 beamline of the Shanghai Synchrotron Radiation Facility (SSRF), Shanghai, China (λ = 0.124 nm) [57]. Fiber bundles were mounted on aluminum frames and equilibrated at 25 °C and 45% RH prior to measurement. Ag behenate was used for calibration. Raw images were reduced in pyFAI [58] with solid-angle and polarization corrections and exposure-scaled air subtraction, followed by azimuthal averaging to obtain I(q) (0.02–0.50 Å−1).
Representative SAXS profiles were obtained for pure PA6, KH550-modified fibers with 1.6 and 4 wt% TiO2, and KH570-modified fibers with 4 wt% TiO2. Shaded regions (Q1: 0.08–0.14 Å−1, Q2: 0.14–0.23 Å−1, Q3: 0.23–0.50 Å−1) were integrated to obtain Porod invariants. These data are presented in the Results section (Figure 4a,b). Quantitative analysis used the Porod constant B and invariant Q: the specific interfacial area was obtained from Sv = 2π2 B/Q [47,48,49]. Polymer-internal phase fractions were resolved into four components—crystal (φc), rigid amorphous fraction (φRAF–c + φRAF–i), and mobile amorphous fraction (φMAF)—normalized to the polymer internal volume [3]. The interface fraction was defined as Γi = φRAF–i·(1–φf), and the effective interfacial-layer thickness as ti = Γi/Sv, enabling interface-resolved metrics [51,59].
Bragg peaks above q ≈ 0.14 Å−1 were masked/fitted and removed; the Porod invariant was integrated from the residuals over 0.02–0.50 Å−1, and the high-q Porod plateau was estimated on 0.32–0.40 Å−1. Each sample was measured once per condition, consistent with beamline practice where instrument stability and standardized reduction protocols ensure reproducibility; repeatability was supported by internal consistency between Porod-based invariants and DSC-derived crystallinity.
Details of masking, baseline handling, and residual-based Q integration for SAXS are provided in the SI (Figure S6; Table S2).
WAXS CeO2 was used for q/geometry calibration. After masking, solid-angle and polarization corrections, and exposure-scaled air subtraction, diffraction images were binned onto a q–χ grid [58]. Hermans orientation factors were evaluated from narrow q-bands centered at the PA6 α(200) (~1.717 Å−1) and γ(200/020) (~2.776 Å−1) reflections. For each band, intensities across Δq were summed (main analysis) to preserve peak-height information, and the normalized I(χ) was converted to f following the Hermans formalism and modern implementations [50,60]. Azimuthal sectors affected by instrumentation (−180° to −160°, 0° to 20°) were excluded. For γ, notch masks were applied near TiO2 lines and a fixed narrow Δq = 0.010–0.012 Å−1 was used. The azimuthal zero was fixed at 90° based on an 85–95° scan.
The α/γ crystalline phase ratio and the representative γ-phase crystallite size Dγ(rep) were obtained from the integrated peak areas and Scherrer-type analysis, respectively, for pure PA6 and TiO2-filled fibers (550-1.6%, 550-4%, 570-4%). Each sample was measured once per condition, consistent with synchrotron WAXS practice, where high beam stability and standardized reduction pipelines ensure reproducibility; repeatability was supported by consistent trends across α/γ ratios, crystallite sizes, and Hermans orientation factors.
Error bars represent standard deviations estimated from Monte-Carlo resampling (N = 300) over band center/width and noise injection, as presented in the Results section (Figure 5). Robustness checks using band-wise means (instead of sums), alternative sector exclusions, and baseline options, as well as representative patterns and azimuthal profiles, are provided in Figures S8–S10.
A targeted robustness analysis of the α(200) Hermans factor under single-ring, excluded-sector settings is summarized in Table S3.

2.3.8. Dynamic Mechanical Analysis (DMA)

DMA was conducted on a Q800 analyzer (TA Instruments, New Castle, DE, USA) in single-cantilever mode using injection-molded bars (35 × 10 × 2 mm). Temperature sweeps were run from −150 °C to 150 °C at 3 °C·min−1 and 1 Hz frequency after conditioning at 25 °C. Representative storage modulus curves were obtained; the results are presented in the Results section (Figure 6a). Each condition was tested once, consistent with DMA practice where instrument stability and standardized thermal protocols ensure reproducibility.

2.3.9. Rotational Rheology

Rotational rheology was performed on a Haake MARS rheometer (Thermo Fisher Scientific GmbH, Karlsruhe, Germany) with 20 mm in diameter, 1 mm thick disk specimens at 230 °C under nitrogen. Complex viscosity η*(ω) curves were obtained from frequency sweeps, as presented in the Results section (Figure 7a). Each sample was measured once per condition, consistent with rheology practice where instrument stability and standardized protocols ensure reproducibility.

2.3.10. Mechanical Testing

Tensile properties of multifilament yarns were measured using an XL-2 instrument (Shanghai Xinfang Instrument Co., Ltd., Shanghai, China) according to the standard test method. At least 5 specimens per condition were tested, and results are reported as mean ± SD. Numerical labels correspond to average tensile strength values, while percentages denote mean elongation at break, as presented in the Results section (Figure 6b). Linear densities (dtex) are reported in Table S4.

2.3.11. Ultraviolet Protection Factor (UPF)

Plain-woven fabric swatches (10 × 10 cm) were conditioned at 25 °C and measured on a fabric ultraviolet transmittance analyzer equipped with an integrating sphere (UV1000F, Labsphere Inc., North Sutton, NH, USA). Spectra were recorded over 250–450 nm; UPF was computed from the 290–400 nm region in accordance with EN 13758-1 [4] and AATCC TM183 [52], using the standard solar irradiance profile in EN 13758-1:2002. Each specimen was measured at multiple positions using a 3 × 3 grid across the swatch (n = 4–5 per specimen), and results are reported as the mean UPF together with the 95% one-sided lower confidence bound (UPFLCB). UPF measurements reflect fabric-level UV protection of plain-woven textiles prepared from the fibers. For visualization, log10(UPF) (dimensionless) is used, and samples with UPF ≥ 50 are denoted as 50+. UVA (315–400 nm) and UVB (290–315 nm) transmittance statistics are provided in the Supplementary Information.

2.3.12. Molecular Dynamics (MD) Simulations

A two-stage, two-resolution coarse-grained (CG) workflow is employed to compare how silane coupling affects PA6–TiO2 crystallization. First, atomistic PA6 oligomers are mapped to Martini beads [61,62] (four beads per repeat unit) and equilibrated in GROMACS 2023 (Royal Institute of Technology, Stockholm, Sweden) [63] to yield equilibrium CG melt trajectories via polymerization simulation [64], from which chain statistics were computed. Equilibrated CG melt trajectories were remapped to a one-bead-per-repeat PA6 (CGPA) model and transferred to LAMMPS version 3 Mar 2020 (Sandia National Laboratories, Albuquerque, NM, USA) [65] for production runs. Two TiO2 surface models representing KH550-anchored and KH570-passivated interfaces were implemented [28,29,66]. At the Martini level, standard Martini parameters are used. In the CGPA model, bonded terms (bonds/angles) are obtained by Boltzmann inversion of distributions from the Martini trajectories [67,68,69,70], whereas non-bonded interactions are fitted to interaction-energy–distance data generated by ORCA 5.0.3 (Max Planck Institute for Coal Research, Mülheim an der Ruhr, Germany) [71] DLPNO-CCSD(T) [72]/aug-cc-pVTZ [73] scans of two PA6 repeat-unit fragments (inputs prepared by Multiwfn 3.8 (Beijing Kein Research Center for Natural Sciences, Beijing, China) [74]), and are implemented in LAMMPS as lj96/cut. Production runs follow an anneal–cool schedule consistent with processing; simulated heat-flow–temperature curves derived from energy–temperature traces are used to extract crystallization peak temperatures for comparison between the two surfaces. Representative snapshots and the simulated heat-flow traces for both surfaces are provided in the Supplementary Information (Figure S11). The simulations are used qualitatively to contextualize experimental trends rather than to assign quantitative causality. Simulations with independent initial configurations yielded consistent equilibrium descriptors, confirming reproducibility.

2.3.13. Data Integration and Correlation Analysis

Eight parameters were selected to represent different levels: interphase (RAF), kinetics (Tp), structure (Xc, α/γ, f(α)), and performance (tenacity σ, elongation ε, and log10(UPF)). Units and definitions are listed in Table S2c. Spearman correlation coefficients were computed (n = 4 samples) using Python’s SciPy 1.12 (Python Software Foundation, Wilmington, DE, USA), with significance treated as exploratory due to sample size. Heatmaps visualizing correlation strength and sign are presented in the Results section (Figure 8a). The plots of α/γ ratios and Hermans orientation factors were plotted against TiO2 loading in the KH550 series are presented in the Results section (Figure 8b), with tensile strength and elongation annotated near each point for context. Structural descriptors (α/γ, f(α)) against performance indicators are provided in the Results section (Figure 8c), where marker size scales with σ and color with ε. Dashed arrows denote the shift from neat PA6 to nanocomposites. The geometric drivers Sv, ti, and Γi are reported separately in Table S2c and discussed in Section 3.4.

2.3.14. Generative AI Disclosure

ChatGPT (OpenAI, San Francisco, CA, USA; accessed June and July 2025) was used for minor language polishing and to assist in drafting and refining code snippets for data processing and visualization. No data or figures were generated by AI. All AI-assisted outputs were reviewed by the authors.

3. Results

3.1. Nanoparticle Dispersion and Interfacial Chemistry

Anatase TiO2 (TA-300) was surface-modified with KH550 or KH570 silane coupling agents to enhance nanoparticle dispersion and interfacial compatibility in the PA6 matrix. KH550 (aminosilane) introduces –NH2 groups capable of forming hydrogen bonds with the amide groups of PA6, thereby promoting strong interfacial adhesion, whereas KH570 (methacryloxy silane) offers weaker polar interactions, serving as a comparative modifier. This dual choice enables systematic evaluation of how interfacial chemistry governs dispersion, crystallization, and performance. The resulting powders and PA6 masterbatches were examined by DLS, FPV testing, FTIR, Raman spectroscopy, and SEM/EDS (Figure S1, Supporting Information). These complementary techniques yielded both quantitative and qualitative insights into particle size distribution, melt processability, chemical bonding at the polymer–filler interface, and spatial distribution of TiO2 in the matrix. Representative SEM/EDS micrographs from related melt-spun PA6/TiO2 fibers corroborated the dispersion and elemental distribution trends inferred from DLS, FPV, and spectroscopic analyses.

3.1.1. TiO2 Dispersion and Interfacial Modification

Dynamic light scattering (Figure 1a) showed monomodal particle size distributions with narrow peaks for both KH550- and KH570-modified TiO2 powders, indicating uniform dispersion. The Z-average diameters were 359.6 nm for KH550-modified and 455.0 nm for KH570-modified TiO2, with corresponding PDIs of 0.120 and 0.115 (n = 3). Both PDI values were below the 0.2 threshold for monodispersity [53]. The slightly larger size of KH570 powders arises from looser silane coverage and minor re-agglomeration in ethanol, which did not affect dispersion in the final fibers, as confirmed by FPV and SEM/EDS.
Figure 1. (a) DLS particle size distributions of KH550- and KH570-modified TiO2 powders in ethanol. (b) Filtration pressure values (FPV) of PA6/TiO2 masterbatches at 4 wt% and 8 wt% TiO2.
Figure 1. (a) DLS particle size distributions of KH550- and KH570-modified TiO2 powders in ethanol. (b) Filtration pressure values (FPV) of PA6/TiO2 masterbatches at 4 wt% and 8 wt% TiO2.
Polymers 17 02551 g001
FPV testing of PA6/TiO2 masterbatches at 4 wt% and 8 wt% filler loadings yielded values within the acceptable limits defined by BS EN 13900-5:2005 [39]. For KH550-modified TiO2, FPVs were 0.60 ± 0.02 bar g−1 (4 wt%) and 0.85 ± 0.02 bar g−1 (8 wt%); for KH570-modified TiO2, values were 0.80 ± 0.02 bar g−1 (4 wt%) and 0.89 ± 0.02 bar g−1 (8 wt%). All tested samples met the melt-filtration criteria for both low and high filler contents.
SEM images of representative melt-spun PA6/TiO2 fibers (Figure S1, Supporting Information) showed that, at 1.8 wt% TiO2, particles were uniformly distributed within the matrix, with core–shell-like features of approximately 200–300 nm. At 4.0 wt%, the particle density increased and occasional local adjacency was observed, but without formation of continuous agglomerated domains, as confirmed by higher-magnification SEM (Figure S1b). EDS mapping confirmed Ti and O signals in particle-rich zones, consistent with the nominal TiO2 contents. Additional low-magnification SEM and EDS mapping (Figure S12) further confirm fibrous continuity and TiO2 distribution consistent with nominal loadings, supporting effective but not perfect dispersion at higher filler contents.

3.1.2. Chemical Interactions

The FTIR spectra of KH550-modified and KH570-modified TiO2 masterbatches (60 wt% TiO2 in PA6) exhibited additional absorption bands in the 930–1100 cm−1 region, which were not present in pure PA6 (Figure 2a, shaded area). The normalized absorbance in this region was higher for MB-570 than for MB-550. The inset difference spectra show positive deviations for both modified samples relative to pure PA6 in this range, with a slightly larger deviation for MB-570. A small shift in the amide I band (~1650 cm−1) towards lower wavenumbers was observed for both modified samples.
Figure 2. (a) FTIR spectra of pure PA6 and silane-treated TiO2 masterbatches, highlighting O–H/N–H, amide I, and Si–O–Ti/Si–O–Si bands. (b) Raman spectra of pure PA6 and selected composites; insets show difference and magnified amide I region.
Figure 2. (a) FTIR spectra of pure PA6 and silane-treated TiO2 masterbatches, highlighting O–H/N–H, amide I, and Si–O–Ti/Si–O–Si bands. (b) Raman spectra of pure PA6 and selected composites; insets show difference and magnified amide I region.
Polymers 17 02551 g002
The Raman spectra (Figure 2b) were measured from injection-molded specimens prepared from the same masterbatch formulations but with different TiO2 loadings: KH550 series (0, 1.6, and 4 wt%) and KH570 series (4 wt%). Enhanced features in the low-frequency region (100–700 cm−1) were observed for all TiO2-containing samples compared with pure PA6. The inset focusing on the amide I region (1600–1700 cm−1) shows increased normalized peak intensities for the modified samples, with the 550-4 wt% specimen exhibiting the highest intensity and a slight shift in the peak position to lower wavenumbers relative to pure PA6. The full Raman spectra are provided in Figure S2 (Supplementary Information).
These spectral differences will be further considered in relation to rigid amorphous fraction (RAF) formation and crystallization behavior in Section 3.2.

3.2. Multi-Scale Structure and Crystallization Behavior

3.2.1. Non-Isothermal Crystallization Kinetics: Avrami–Nakamura and Isoconversional Analyses

The crystallization kinetics of pure PA6 and PA6/TiO2 fiber composites were quantified using a θ*-based Nakamura framework (with activation energies fixed from independent isoconversional analysis) together with an isoconversional KAS approach.
Figure 3a shows the non-isothermal DSC cooling curves at 20 °C·min−1. All samples exhibit a single, well-defined exothermic peak. The peak crystallization temperature (Tp) increases from 181.68 °C for pure PA6 to 183.23 °C (1.6 wt% TiO2–KH550), 183.74 °C (4 wt% TiO2–KH550), and 185.97 °C (4 wt% TiO2–KH570). The full width at half maximum (FWHM) narrows from 6.31 °C (pure PA6) to 5.43–5.83 °C for the composites. The crystallization enthalpy (ΔHc), normalized per gram of PA6 after TiO2 blank correction, is 55.10 J·g−1 for pure PA6 and 60.67–61.08 J·g−1 for the composites.
Figure 3. (a) Non-isothermal DSC cooling curves of pure PA6 and PA6/TiO2 fibers. (b) Relative crystallinity as a function of time with Nakamura fits. (c) Isoconversional KAS analysis at multiple conversions and cooling rates.
Figure 3. (a) Non-isothermal DSC cooling curves of pure PA6 and PA6/TiO2 fibers. (b) Relative crystallinity as a function of time with Nakamura fits. (c) Isoconversional KAS analysis at multiple conversions and cooling rates.
Polymers 17 02551 g003
Figure 3b presents the evolution of relative crystallinity (Xt) as a function of time at 20 °C·min−1. Using the Nakamura model with isothermal-equivalent time (θ*), global fits were performed over 0.20 ≤ Xt ≤ 0.60 with E fixed from the independent isoconversional analysis and rate-specific K0*. The fitted Avrami-type exponents were n = 3.147 (pure PA6), 2.871 (1.6 wt% TiO2–KH550), 2.762 (4 wt% TiO2–KH550), and 2.360 (4 wt% TiO2–KH570), with R2 = 0.864–0.925 across formulations. Despite the stricter cross-rate normalization, the rate scaling remained internally consistent: regression of ln K0* versus ln β gave αˆ ≈ 2.10–2.12 with R2 ≥ 0.989, and the predicted curves closely tracked the experimental Xt(t) within the fit window. Uncertainty quantified by nonparametric bootstrap (B = 200) yielded 95% confidence intervals of 3.019–3.273, 2.691–3.107, 2.633–2.900, and 2.277–2.446 for the four samples, respectively; leave-one-rate validation gave nˉ = 3.144, 2.893, 2.765, and 2.356 (Table S6).
Figure 3c summarizes the isoconversional KAS (multi-point) analysis at X = 0.30, 0.40, 0.50, and 0.55 using cooling rates of 5–40 °C·min−1. Within this window, the apparent activation energy Ea(X) shows weak conversion dependence; formulation-level medians are 208.2 kJ·mol−1 (pure PA6), 202.9 kJ·mol−1 (1.6 wt% TiO2–KH550), 229.6 kJ·mol−1 (4 wt% TiO2–KH550), and 237.4 kJ·mol−1 (4 wt% TiO2–KH570). Complementary single-point Kissinger regressions are provided in Figure S5, with full-rate DSC curves and per-conversion KAS fits in Figures S3 and S4, and parameter uncertainties (95% confidence intervals from bootstrap/leave-one-out) summarized in Table S6.

3.2.2. Rigid Amorphous Fraction (RAF) Quantification

SAXS intensity profiles for four samples—Pure PA6, 550 1.6% (KH-550, 1.6 wt% TiO2), 550 4% (KH-550, 4 wt% TiO2), and 570 4% (KH-570, 4 wt% TiO2)—are shown in Figure 4a. All TiO2-filled samples exhibit higher scattering intensity in the low-q range (0.08–0.14 Å−1) than Pure PA6. In the mid-q range (0.14–0.23 Å−1), the four curves are generally comparable, whereas in the high-q range (0.23–0.50 Å−1) the TiO2-filled samples show slightly lower intensities than pure PA6.
Figure 4. (a) SAXS intensity profiles of pure PA6 and selected nanocomposites. (b) Porod invariant integrals over defined q-regions. (c) Phase fractions resolved into φc, φRAF−c, φRAF−i, and φMAF.
Figure 4. (a) SAXS intensity profiles of pure PA6 and selected nanocomposites. (b) Porod invariant integrals over defined q-regions. (c) Phase fractions resolved into φc, φRAF−c, φRAF−i, and φMAF.
Polymers 17 02551 g004
As presented in Figure 4b, Porod invariant integration shows that Q1 (low-q) is consistently higher for all TiO2-filled samples than for Pure PA6. Q2 (mid-q) remains similar across the four samples, while Q3 (high-q) is lower for the TiO2-filled samples. The total Porod invariant (Qtotal) is comparable among all samples, indicating that the redistribution of scattering intensity occurs mainly between the low- and high-q regions.
Using DSC-derived crystalline fractions together with SAXS Porod-invariant analysis, we quantified the phase fractions on a polymer-only basis with a fixed pipeline (Figure 4c): first, the interfacial rigid amorphous fraction (φRAF-i) was obtained by the low-q difference method, ΔQ1 = Q1, Sample − Q1, (pure PA6 constrained to φRAF-i = 0); second, φRAF-i was held fixed while the Q23 band was used to partition the crystal-adjacent RAF (φRAF-c) and the mobile amorphous fraction (φMAF). Point estimates and 95% bootstrap CIs are compiled in SI Table S2 (panels b and e); the Porod-invariant q-bands used are listed in Table S2a; the low-q baselines and increments that determine φRAF-i are in Table S2d; and interfacial metrics (Sv, Γi, ti) are given in Table S2c. Relative to pure PA6, all TiO2-filled samples show a non-zero, formulation-dependent φRAF-i and a compensating decrease in φMAF, whereas φc (DSC) and φRAF-c remain within a similar range. Among the three filled formulations, the total constrained fraction normalized by amorphous content, RAF-tot/φA = (φRAF-c + φRAF-i)/φA, lies within a narrow band and is statistically equivalent under a ±0.01 TOST margin (Figure S7b), while pure PA6 exhibits the expected lower baseline; formulation-level differences in φRAF-i are visualized with 95% CIs in Figure S7a. This partition aligns with literature that distinguishes crystal-induced and interfacial rigid fractions [20].

3.2.3. WAXS Analysis of Polymorphs and Orientation

The α/γ crystalline phase ratio (Figure 5a) was 0.205 for pure PA6 and increased to 0.291 for the 550-1.6% TiO2 fiber and 0.360 for the 550-4% TiO2 fiber. In contrast, the 570-4% TiO2 fiber exhibited a sharply reduced value of 0.038. The γ-phase crystallite size, Dγ(rep), varied from 2.28 to 3.42 nm across the samples, with the largest value observed for the 550-4% TiO2 fiber and the smallest for pure PA6 (peak deconvolution shown in Figure S9).
Figure 5. (a) α/γ crystalline phase ratio and γ-phase crystallite size. (b) Hermans orientation factors for α- and γ-phases.
Figure 5. (a) α/γ crystalline phase ratio and γ-phase crystallite size. (b) Hermans orientation factors for α- and γ-phases.
Polymers 17 02551 g005
For the Hermans orientation factors (Figure 5b), f(α) spans a wider range (0.41–0.70) compared with f(γ) (0.53–0.59). The highest α-phase orientation was observed in the 570-4 wt% TiO2 fiber (f(α) = 0.697), whereas the lowest occurred in the 550-4 wt% TiO2 fiber (f(α) = 0.414). By contrast, the γ-phase orientation exhibits only minor differences among the four samples (0.532–0.587). For 570-4 wt%, the α/γ ratio reaches the lowest value, while the Hermans factor of the α(200) reflection remains high and numerically stable (methodological validation in Figure S10, and extraction robustness is summarized in Figure S8). Sensitivity analyses (band width/center and noise injection; Table S3) confirm that the high f(α) observed for 570-4 wt% is robust against WAXS processing choices.

3.3. Macroscopic Functional and Mechanical Performance

3.3.1. Stiffness and Strength

Figure 6a shows the temperature dependence of the storage modulus (E′) for pure PA6 and PA6/TiO2 fibers with different silane modifications. Across the measured range from –150 to 120 °C, all TiO2-filled samples maintained higher E′ values than pure PA6, with the 550-4% sample exhibiting the highest modulus over most of the range. At low temperatures (<–50 °C), E′ values were in the range of 4000–5000 MPa, while in the intermediate region (–50 °C to 50 °C) the modulus decreased progressively with increasing temperature. The 570-4% sample consistently showed lower E′ values compared to other filled samples. The glass transition region was indicated by a rapid modulus drop between approximately 40 °C and 60 °C, consistent with Tg values determined from tan δ measurements.
Figure 6. (a) Storage modulus as a function of temperature. (b) Tensile strength and elongation at break of pure PA6 and nanocomposite fibers.
Figure 6. (a) Storage modulus as a function of temperature. (b) Tensile strength and elongation at break of pure PA6 and nanocomposite fibers.
Polymers 17 02551 g006
Figure 6b summarizes the tensile properties measured at room temperature. Tensile strength ranged from 2.9 cN/dtex for the 550-1.6% sample to 5.6 cN/dtex for the 570-4% sample. Pure PA6 exhibited a strength of 4.2 cN/dtex, while the 550-4% sample reached 4.6 cN/dtex. Elongation at break varied inversely with strength, from 47.9% for 550-1.6% to 13.9% for 570-4%. Pure PA6 showed 42.9% elongation, higher than all samples except 550-1.6%. Overall, high-modulus samples tended to exhibit reduced elongation.

3.3.2. Flowability and UV Protection

The complex viscosity (η*) curves of pure PA6 and TiO2-filled fibers (550-1.6% and 550-4%) are presented in Figure 7a. All samples exhibit typical shear-thinning behavior, where η* remains nearly constant at low angular frequencies and decreases progressively with increasing ω, consistent with typical shear-thinning behavior reported for polymer melts [75,76]. At ω = 100 rad·s−1, the η* values are approximately 4.10 × 102 Pa·s for pure PA6, 4.14 × 102 Pa·s for the 550-1.6% fiber, and 3.24 × 102 Pa·s for the 550-4% fiber. The differences between samples become more pronounced at higher frequencies, with the 550-4% fiber showing the steepest reduction in viscosity, indicating weakened interfacial restriction and enhanced chain mobility at short time scales.
Figure 7. (a) Complex viscosity as a function of angular frequency at 230 °C. (b) Ultraviolet protection reported as log10(UPF) of plain-woven fabrics from pure PA6 and nanocomposite fibers.
Figure 7. (a) Complex viscosity as a function of angular frequency at 230 °C. (b) Ultraviolet protection reported as log10(UPF) of plain-woven fabrics from pure PA6 and nanocomposite fibers.
Polymers 17 02551 g007
Figure 7b shows that all TiO2-filled fabrics reach UPF 50+, whereas pure PA6 provides only minimal protection. Pure PA6 (UPF = 5.02; UPF_LCB95% = 4.78; n = 4) is far below the threshold. Among filled samples, 550-1.6% (666.33; 636.86; n = 5) shows the highest protection, followed by 550-4% (246.61; 224.28; n = 4). 570-4% (81.30; 77.41; n = 5) is the lowest among the filled samples yet still exceeds the threshold. The ranking is consistent with UVB transmittance—0.05% (550-1.6%), 0.27% (550-4%), and 1.12% (570-4%)—whereas UVA varies less; full UVA/UVB statistics are listed in Table S5.

3.4. Interface–Structure–Performance Coupling

The Spearman correlation heatmap summarizes monotonic associations among interface (RAF), kinetics (Tp), structure (Xc, α/γ, f(α)), and performance (tenacity σ, elongation ε, log10(UPF)). On a polymer-only basis, Xc and Tp are strongly positively associated (rs = 1.00). Phase composition and orientation are orthogonal in this set: α/γ and f(α) vary in opposite directions across the four specimens. Tenacity is positively associated with Tp and Xc, whereas elongation shows the opposite tendencies. The log10(UPF) metric shows positive associations with RAF and α/γ and weaker associations with Tp/Xc and f(α).
Figure 8. (a) Correlation map of interface, kinetics, structure, and performance descriptors. (b) Evolution of α/γ phase fraction and Hermans orientation factor f(α) in the KH550 series. (c) Structure–performance map linking f(α), α/γ, and mechanical/UV outcomes. Asterisks indicate statistically significant correlations based on Spearman tests (p < 0.05).
Figure 8. (a) Correlation map of interface, kinetics, structure, and performance descriptors. (b) Evolution of α/γ phase fraction and Hermans orientation factor f(α) in the KH550 series. (c) Structure–performance map linking f(α), α/γ, and mechanical/UV outcomes. Asterisks indicate statistically significant correlations based on Spearman tests (p < 0.05).
Polymers 17 02551 g008
In the KH550 series (0, 1.6, and 4 wt% TiO2), α/γ increases from 0.21 to 0.29 and 0.36, whereas the Hermans factor of the α phase decreases from 0.511 to 0.416 and 0.414. The mechanical responses are non-monotonic: σ is 4.2, 2.9, and 4.6 cN/dtex, and ε is 42.9%, 47.9%, and 14.4%, respectively. The sequence denotes the series evolution; no optimality is implied.
At the same loading (4 wt% TiO2), KH570 shows a higher f(α) (0.697) and a lower α/γ (0.038) than KH550 (f(α) 0.414, α/γ 0.360). Among the three points displayed, KH570 exhibits the highest σ. Marker size encodes σ and color encodes ε; dashed arrows indicate the shift from the reference (Pure PA6).

4. Discussion

Building on the integrated dataset, we rationalize how interfacial chemistry restructures the constrained amorphous fraction (RAF) spatially, how such constraints couple to non-isothermal crystallization and phase orientation, and how these structural states relate to mechanical, rheological, and UV protection metrics, thereby providing transferable interphase quantification and interphase-resolved readouts of dispersion and compatibility. We emphasize mechanism-centric contrasts across silane types and loadings rather than optimality claims. Within this processing window and on a polymer-only basis, two equivalent routes emerge on the common outcome space: a geometry path (KH550 concentration series) and a chemistry-mediated friction path (4 wt% KH570 and KH550), whose endpoints are compactly visualized in Figure 8b,c, closing a reproducible kinetics–structure–property loop.

4.1. Role of Interfacial Chemistry in RAF Redistribution

Spectroscopic and dispersion evidence supports enhanced polymer–filler interactions with silane coupling agents. FTIR and Raman spectra show characteristic Si–O–C/Si–O–Si (and Ti–O–Si shoulder) features consistent with covalent grafting on TiO2 surfaces (Figure 2), in line with prior reports on silane-functionalized titania [27,36]. Complementary DLS, FPV, and SEM analyses demonstrate improved particle dispersion (Figure 1; Figure S1). On a polymer-only basis, SAXS-invariant analysis reveals that the total rigid amorphous fraction (RAF) remains approximately conserved, while its spatial allocation shifts from crystal-adjacent (φRAF-c) to filler-proximal (φRAF-i) zones (Figure 4c; Table S2b). This redistribution highlights that interfacial chemistry primarily relocates constrained amorphous domains rather than enlarging their overall amount, consistent with the established three-phase view of semicrystalline polymers [1,2,5,19,20,59,77].
To rationalize how a minor φRAF-i exerts measurable influence, we transform it into system-level descriptors. The interfacial RAF volume is defined as Γi = φRAF-i(poly-only)·(1 − φf), i.e., the portion of the entire composite volume occupied by polymer in the interfacial RAF. Dividing Γi by the specific interfacial area Sv yields an effective interfacial thickness ti = Γi/Sv. Here, Sv is obtained from SAXS via the Porod constant and invariant (Sv = 2π2B/Q) (Methods; Figure 4a,b and Figure S6; Table S2a) [47,48,49]. The resulting ti values are sub-nanometric (see Table S2c) and are treated as an effective descriptor rather than an absolute physical thickness; thin but spatially extensive interfacial skins can disproportionately affect chain mobility and non-isothermal crystallization. Such interfacial skins are known to impose disproportionate constraints on chain mobility and to modulate non-isothermal crystallization in semicrystalline polymers [9,51,66].
Along the KH550 concentration series (0, 1.6, and 4 wt%), the system evolves from an orientation-dominated regime—characterized by higher Sv and thinner ti—towards a phase-fraction-dominated regime at higher loading, where Sv decreases and ti thickens (Table S2c). At comparable interfacial descriptors (4 wt% KH550 vs. 4 wt% KH570), the downstream kinetic–orientation responses diverge, suggesting that distinct coupling pathways are activated depending on the silane chemistry rather than interfacial area/skin thickness alone [28,29].
We therefore interpret the observed small-in-volume yet surface-abundant RAF-i through the (Sv, ti, Γi) triad, consistent with the interphase perspective in polymer nanocomposites [34,77,78].
At similar specific interfacial area Sv but distinct effective thickness ti at 4 wt% loading, the outcomes diverge by chemistry: KH570 (passivated/low-friction) follows an orientation-dominated route (higher f(α) with lower α/γ), whereas KH550 (anchored/high-friction) follows a phase-fraction-dominated route (higher α/γ with lower f(α)) [30,79]. These contrasts do not require changing the total constrained fraction; they reflect RAF redistribution under comparable Sv and chemistry-dependent ti and Γi.

4.2. Influence of Interfacial RAF on Non-Isothermal Kinetics and Orientation

Increasing the interfacial RAF volume (Γi) is associated with systematic shifts in non-isothermal crystallization metrics [42]: the supercooling ΔTp decreases (higher Tp) while −log10(t1/2) decreases (i.e., t1/2 lengthens) across formulations (Figure 3a,b). Partial Spearman correlations controlling TiO2 wt% and silane type support these associations (Table S5a: ρ(ΔTp, Γi) < 0; ρ(−log10t1/2, Γi) < 0). Taken together, these trends are consistent with an interfacial-constraint scenario [59] in which an ultrathin, high-coverage interfacial constrained layer provides earlier effective nucleation sites (reduced ΔTp) while constraining growth/transport (longer t1/2) in the Xt = 0.20–0.60 window used for Nakamura fitting [9,40,41,43] (Figure 3b; Table S1). Nakamura global fits (fixed Ea from isoconversional closure) show rate-dependent K* shifts with n varying modestly within uncertainty across formulations (Table S1), indicating changes in the overall time scale rather than a wholesale change in growth dimensionality.
WAXS reveals that phase composition (α/γ) and orientation (f(α) at α(200)) are orthogonal descriptors: a lower α/γ can coexist with a higher f(α)(Figure 5a,b). The Hermans factor is robust to q-band width, band-center, and noise injection (Table S3; Figures S8–S10), ensuring that observed differences reflect genuine orientation contrasts rather than processing artifacts [60]. Partial correlations indicate a negative association between f(α) and Γi (filled + pure set), but a positive association between f(α) and the interfacial thickness ti (filled subset) (Table S5a). This pattern is consistent with two coupling pathways: orientation-dominated skins (large Sv, small ti) correlate with higher f(α) and lower α/γ, whereas phase-fraction-dominated skins (smaller Sv, larger ti) correlate with higher α/γ and lower f(α).
Consistent with 4.1, partial Spearman correlations (controlling TiO2 wt% and silane type) show that larger Γi associates with smaller ΔTp (higher Tp) and longer t1/2 (Table S5a; Figure 3a,b). Along the KH550 series, Sv decreases while ti increases, with Γi varying modestly within the small-n uncertainty (Table S2c); α/γ rises whereas f(α) falls, consistent with a shift toward the phase-fraction-dominated side. We refrain from causal claims beyond the reported statistics and use these outcome labels to summarize directionality.

4.3. Structural States and Their Relation to Macroscopic Performance

The interfacial descriptors identified above—α-phase orientation (f(α), interfacial RAF volume (Γi), interfacial coverage (Sv), and the mobile amorphous fraction (MAF)—provide a common coordinate system to interpret stiffness, strength/ductility, viscoelasticity, and UV protection.
Across filled formulations, the storage modulus E′ increases relative to pure PA6 (Figure 6a), consistent with additional interfacial constraints on segmental mobility. Within the same structural coordinate, tensile strength tracks with f(α) and Γi, whereas elongation at break decreases in line with reduced MAF (Figure 6b) [80,81,82]. Partial Spearman correlations (controlling TiO2 wt% and silane type) and descriptive ΔR2 confirm that f(α) and Γi retain independent explanatory power beyond composition (Table S5a,b). This joint trend reflects a common trade-off: enhanced orientation/constraint bolsters strength but narrows large-strain compliance.
The low-frequency viscosity plateau remains comparable among formulations, while high-frequency shear-thinning diverges (Figure 7a), consistent with constraint-sensitive fast modes [81,82] in the viscoelastic spectrum; high-frequency response is governed by local segmental constraints and entanglement dynamics rather than fully relaxed network motions. Directional statistics on η* support these associations within the available subset (KH550 series + pure; Table S5a). Note that Figure 7a does not include the 570-4 wt% sample; hence, rheological comparisons involving KH570 rely on structural/mechanical/UV triangulation rather than direct η* statistics, and all controls are stated at the subset level.
log10(UPF) increases with improved dispersion/coverage (Sv) and higher Γi (Figure 7b; Table S5a,b), indicating that thin yet pervasive interfacial skins extend scattering/absorption pathways at the fiber scale. Several formulations meet or surpass the practical benchmark log10(UPF) ≈ 1.70 (UPF = 50) [4]; replicate-level lower confidence bounds (UPFLCB) and UVA/UVB transmittance are reported in Table S4 under EN 13758-1 [4]/AATCC TM183 protocols [52,83]. These enhancements are consistent with classical photostabilization mechanisms of polymer matrices [83].
At 4 wt% loading, KH570 attains the highest tenacity despite a lower α/γ, consistent with an orientation-dominated reinforcement scenario where higher f(α), better dispersion, and fewer stress concentrators outweigh the benefit of higher phase fraction under anchored interfaces. Conversely, the KH550 series increases α/γ but suppresses f(α) as loading rises, aligning with a phase-fraction-dominated constraint that limits draw-induced orientation. In this compact view, KH570-4% represents the orientation-dominated endpoint (higher f(α), lower α/γ, high σ with better dispersion), whereas the KH550 loading series progressively shifts toward the phase-fraction-dominated endpoint (higher α/γ, lower f(α), increased σ but reduced ε), consistent with the two equivalent routes summarized in Figure 8b,c.

4.4. Interfacial Regimes: Geometric Descriptors and Structural Outcomes

We distinguish between geometric descriptors and structural outcomes. The interfacial triad (Sv, ti, Γi) quantifies geometry (area, effective thickness, and interfacial volume), whereas the outcome regimes describe how the structure responds: orientation-dominated versus phase-fraction-dominated. In practice, high Sv with smaller ti tends to favor orientation-dominated responses, while lower Sv with larger ti tends to favor phase-fraction-dominated responses. Across the filled formulations, RAF-tot normalized by the amorphous phase (φRAF-totA) remains statistically equivalent, indicating that interfacial chemistry redistributes RAF between RAF-i and RAF-c instead of increasing the total constrained fraction (Table S2c). These outcome regimes summarize how interfacial geometry biases nucleation versus alignment pathways in semicrystalline polymer–particle systems [82,84].
Along the KH550 loading series, Sv decreases and ti increases, and the outcome shifts toward a phase-fraction-dominated response (α/γ increases while f(α) decreases). At 4 wt%, KH570 exhibits smaller ti at similar Sv (Table S2c) and follows an orientation-dominated response (high f(α) with low α/γ), whereas KH550 shows the complementary phase-fraction-dominated pattern (high α/γ with low f(α)). These outcome regimes rationalize the distinct strength–ductility balances seen in Figure 8b,c without implying recipe optimality.
At comparable specific interfacial area (Sv within 95% CIs), the functional-group chemistry of the silane coupling agent modulates adhesion versus interfacial slip, biasing the structural route. KH550 (3-aminopropyl; –NH2) can form stronger H-bond/Lewis acid–base interactions with PA6 amide carbonyls (and may condense with –COOH chain ends), consistent with higher interfacial anchoring/friction [30,79]; KH570 (methacryloxypropyl; –COO–) interacts more weakly and permits easier slip. Accordingly, at 4 wt% TiO2 and comparable Sv, KH570 exhibits smaller ti and Γi and follows an orientation-dominated response (higher f(α), lower α/γ), whereas KH550 shows larger ti and Γi and a phase-fraction-dominated response (higher α/γ, lower f(α)) (Figure 5a,b; Table S2c) [44]. Cross-stage, qualitative side-evidence is consistent with this picture: the Raman amide-I band in 550-4% is stronger and slightly red-shifted; DMA storage modulus E′ peaks at 550-4%; while 570-4% attains the highest tenacity despite lower α/γ, consistent with a low-friction/thin-skin, orientation-led route. These trends are mutually coherent with the SAXS interfacial metrics (ti, Γi), WAXS polymorphs and orientation, and DSC kinetics (see operational definitions and statistical notes of Table S5).
Spearman rank correlations (Table S5a) summarize directional associations among Γi, kinetic/structural responses (ΔTp, −log10 t1/2, f(α)), and interfacial geometry (Sv/ti), while descriptive ΔR2 (Table S5b) provides effect-size context under small n. Together with Porod-based surface quantification and RAF partitioning (Table S2c), the orientation- vs. phase-fraction-dominated regimes reconcile the trends observed in Figure 8a–c without invoking ranking or optimal recipes [66,69].

4.5. Geometry-Based Summary of RAF–Kinetics–Structure–Performance Relations

The interfacial triad (Sv, ti, Γi) is purely geometric, providing a transferable, reproducible way to parameterize constrained skins at filler surfaces. In practical terms, concentration mainly “turns” the geometry axis (Sv with induced changes in Γi), while silane chemistry mainly “turns” the friction axis (effective ti, hence Γi) at comparable Sv. Both routes are equivalent in that they relocate samples on the same low-dimensional manifold and project onto the orientation- vs. phase-fraction-dominated outcomes captured by Figure 8b,c, while RAFtot/ϕA remains statistically equivalent within this window.
Within the three-phase picture of semicrystalline polymers (crystal/RAF/MAF), this triad explains how a small φRAF-i can yield measurable effects when widely distributed as a sub-nanometric, high-coverage interfacial layer [1,2,59,85]. The approach builds on established SAXS invariants and Porod asymptotics to quantify specific interface area [47,48,86], with phase fractions expressed on a polymer-only basis for internal consistency.
Rather than relying on potentially misleading scatter fits at small n, we summarize directional relationships using Spearman rank correlations (exploratory, n = 4) and silane type, and we report descriptive ΔR2 for nested models (Figure 8a; Table S5a,b). In this compact view, Γi aligns with non-isothermal kinetics (ΔTp and −log10t1/2), while Xc co-varies strongly with Tp on a polymer-only basis; RAF shows weak positive associations with Xc and α/γ. Sv and ti show tendencies consistent with orientation and phase ratio (f(α), α/γ). Tenacity increases with Tp and Xc, whereas elongation shows the opposite tendency; log10(UPF) trends with RAF and α/γ and exhibits near-zero association with f(α). For completeness, model-based analyses—Nakamura global fits and isoconversional methods (Kissinger/KAS)—are reported in the SI (see Figure S3 for multi-rate DSC, Figure S4 for KAS isoconversional results, Figure S5 for Kissinger plots, and Table S1 for Nakamura parameters); these reproduce the rate-dependent time-scale shifts consistent with the Tp trends used here, without requiring rate-wise changes in growth dimensionality within the analyzed conversion window [6,9,40,41,42,43,51].
The regime view—coverage-dominated (high Sv, thin ti) vs. thickness-dominated (lower Sv, thicker ti)—rationalizes concentration trends and silane-dependent pathways. At fixed loading, KH550 and KH570 can share similar Sv and ti yet exhibit distinct kinetic–orientation signatures, indicating that interfacial chemistry modulates the activation of nucleation and alignment routes beyond geometry alone (Figure 8b,c). This is consistent with literature on polymer–nanoparticle interphases where interfacial layer thickness and interaction strength co-govern dynamics and properties [34].
Coarse-grained MD (Figure S11) illustrates stronger interfacial adhesion and more upright chain alignment near KH550-modified surfaces compared with the passivated KH570 surfaces, qualitatively consistent with the kinetic shifts inferred from DSC. We emphasize that these simulations visualize plausible local motifs; absolute temperatures depend on mapping and potential simplifications.
Overall, the (Sv, ti, Γi) triad, coupled with small-sample-robust statistics, offers a reproducible and transferable protocol to connect interface-resolved RAF with kinetics, structure, and performance across polymer–nanoparticle fiber systems [5]. Building on the interphase perspective emphasized by Huang et al. [34], our study quantifies and reproducibly partitions RAF into interfacial and crystal-adjacent components. By integrating SAXS invariants and DSC kinetics within fiber systems, we operationalize the interphase concept into transferable descriptors (Sv, ti, Γi), thereby providing a practical, cross-scale basis for design.
For operational clarity, we use a minimal two-readout notion of “compatibility”: (i) nucleation-compatibility, summarized by ΔTp (and apparent Ea from KAS/Kissinger) evaluated at comparable Sv (or expressed per unit Sv when appropriate); and (ii) constraint-compatibility, summarized by Γi and ti (=Γi/Sv). We employ these strictly as operational readouts, not as theory claims; full operational definitions and statistical notes are provided in Table S5 notes.

5. Conclusions

We established an interface-to-performance link for PA6/TiO2 fibers that is consistent across kinetics, structure, and properties. On a polymer-only basis, the total rigid amorphous fraction remains approximately conserved, while its allocation redistributes from crystal-adjacent RAF to filler-proximal interfacial RAF, thereby addressing transferable interphase quantification at the fiber scale. To summarize interfacial participation without over-emphasizing thickness magnitudes, we use a geometry-aware pair (Sv, Γi) derived from SAXS invariants: specific interfacial area and the interfacial RAF volume within the composite—with ti treated as an effective descriptor.
  • Γi correlates with crystallization metrics (higher Tp, longer t1/2), while α/γ and f(α) provide orthogonal structural coordinates.
  • “Compatibility” is described operationally as nucleation-compatibility and constraint-compatibility, consistent with RAF views.
  • KH550 and KH570 yield distinct coupling routes (phase-fraction–dominated vs. orientation-dominated).
  • Performance changes arise from redistribution and pathway choice, not from a net increase in RAF.
The (Sv, Γi) descriptors and the associated workflow based on SAXS invariant together with WAXS and DSC provide a transferable, reproducible basis for interphase-aware design of polymer–nanoparticle fibers, thereby closing the kinetics–structure–property loop articulated in the Introduction.
This work examined three compositions including pure PA6 (0, 1.6, and 4.0 wt% TiO2) in a single PA6 matrix under a fixed melt-spinning window (temperature, take-up speed, DR ≈ 3.0) and two coupling agents (KH550 vs. KH570 at a representative loading); thus composition–processing interactions are bounded by this design. Several selected datasets were acquired as single-run; robustness checks and cross-method closure were applied. The SAXS-derived interphase metrics (Sv, Γi, ti) rely on Porod invariants after Bragg subtraction and a polymer-only basis, whose sensitivity to model choices and particle polydispersity warrants further tests (e.g., contrast variation, tomography). Application-side validation (operando/in situ crystallization, long-term mechanical/UV durability, and laundering stability for UPF) was outside the present scope. Future work will broaden filler fractions and surface chemistries, map the spinning parameter space (T, ω, DR) using designed experiments, include independent replicates (n ≥ 3) and inter-lab checks, and integrate MD/DFT-guided interface design with process optimization to generalize the interphase-geometry–performance framework across polymer–nanoparticle fiber systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym17182551/s1, The following supporting information can be provided upon acceptance. Figure S1: SEM micrographs of melt-spun PA6/TiO2 fibers at representative loadings; Figure S2: Full-range Raman spectra (100–1800 cm−1) of injection-molded PA6/TiO2 composites (KH550/KH570 series); Figure S3: Full non-isothermal DSC cooling curves at β = 5–40 °C·min−1; Figure S4: Per-conversion KAS regressions (ln[β/Tx2] vs. 1/Tx) at X = 0.30–0.55; Figure S5: Kissinger regressions (ln[β/Tp2] vs. 1/Tp) and fitted Ek; Figure S6: Raw SAXS profiles and Bragg-peak subtraction used for Porod invariants; Figure S7: Interfacial rigid amorphous fraction (φRAF-i) and normalized total RAF (RAFtot/φA, polymer-only basis) with 95% CIs and TOST equivalence assessment; Figure S8: Robustness of the Hermans factor extraction at the PA6 α(200) reflection; Figure S9: Representative peak deconvolution of the WAXS pattern for the 550-1.6% sample; Figure S10: Azimuthal profile at PA6 α(200) for Hermans factor evaluation; Figure S11: Coarse-grained MD snapshots and DSC-like heat-flow curves (−dU/dT); Figure S12: SEM–EDS analyses confirming fibrous morphology and TiO2 dispersion at 1.8% and 6% loadings. Figure S13. Robustness check of pure PA6 DSC curve. The 165 °C jitter is due to overlap, not data loss; raw and processed curves give consistent ΔHc and Tp. Table S1: Nakamura (θ*) parameters and uncertainties under non-isothermal cooling; Table S2: SAXS results panel—(a) Porod invariants by q-range; (b) phase fractions (φc, φRAF-c, φRAF-i, φMAF) on a polymer-only basis; (c) interfacial-layer metrics (Sv, Γi, ti); (d) low-q baselines and increments (ΔQ1) used to fix φRAF-i; (e) master summary with 95% CIs and absolute scale factor α; Table S3: Robustness of the α(200) Hermans factor; Table S4: Ultraviolet protection metrics (UPF, T(UVA), T(UVB)) for woven fabrics; Table S5a,b: Partial correlations and descriptive ΔR2 linking interfacial metrics (Γi, Sv, ti) to proximal responses (ΔTp, t1/2, f(α), η*).

Author Contributions

Conceptualization, H.Y. and B.S.; Resources, B.S.; Methodology, Investigation, Data Curation, Visualization, and Formal Analysis, H.Y.; Writing—Original Draft Preparation, H.Y.; Writing—Review and Editing, H.Y., X.J. (Xiaoze Jiang), P.L. and X.J. (Xiaohuan Ji); Supervision, B.S.; Project Administration, X.J. (Xiaoze Jiang) and B.S.; Funding Acquisition, B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Special Excellent PhD International Visit Program (DHU Grant No. 2021).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The datasets generated and analyzed during this study are available in the Zenodo repository at DOI: 10.5281/zenodo.16983099 (all versions), with Version v1.0 archived at DOI: 10.5281/zenodo.16983100.

Acknowledgments

The authors acknowledge the Shanghai Synchrotron Radiation Facility (SSRF) for beamtime at BL16B1 under the proposal “Structure Regulation of Condensed Matter Interfaces for the Fabrication of High-Performance Thermoplastic Polymer Nanocomposite Fibers” (Project No. 2020-SSRF-PT-012686). We gratefully acknowledge Xiaran Miao and the BL16B1 beamline staff for their assistance during the SAXS/WAXS measurements, as well as Jing Shen and Jia Chen from our research group for their participation and support throughout the project. The authors acknowledge the use of local computing resources for the molecular dynamics simulations.

Conflicts of Interest

Author Xiaohuan Ji was employed by Shaoxing Huiqun New Material Technology Co., Ltd., Shaoxing 312000, China. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Wunderlich, B. Reversible crystallization and the rigid–amorphous phase in semicrystalline macromolecules. Prog. Polym. Sci. 2003, 28, 383–450. [Google Scholar] [CrossRef]
  2. Di Lorenzo, M.L.; Righetti, M.C. Crystallization-induced formation of rigid amorphous fraction. Polym. Cryst. 2018, 1, e10023. [Google Scholar] [CrossRef]
  3. Strobl, G. From the melt via mesomorphic and granular crystalline layers to lamellar crystallites: A major route followed in polymer crystallization? Eur. Phys. J. E 2000, 3, 165–183. [Google Scholar] [CrossRef]
  4. Gambichler, T.; Laperre, J.; Hoffmann, K. The European standard for sun-protective clothing: EN 13758. J. Eur. Acad. Dermatol. Venereol. 2006, 20, 125–130. [Google Scholar] [CrossRef]
  5. Strobl, G. The Physics of Polymers: Concepts for Understanding Their Structures and Behavior; Springer: Berlin/Heidelberg, Germany, 2007; pp. 1–518. [Google Scholar]
  6. Strobl, G. Colloquium: Laws controlling crystallization and melting in bulk polymers. Rev. Mod. Phys. 2009, 81, 1287–1300. [Google Scholar] [CrossRef]
  7. Mo, Z. A Method for the Non-Isothermal Crystallization Kinetics of Polymers. Acta Polym. Sin. 2008, 656–661. [Google Scholar] [CrossRef]
  8. Cheng, S.Z.D.; Wunderlich, B. Thermal analysis of thermoplastic polymers. Thermochim. Acta 1988, 134, 161–166. [Google Scholar] [CrossRef]
  9. Jeziorny, A. Parameters characterizing the kinetics of the non-isothermal crystallization of poly(ethylene terephthalate) determined by d.s.c. Polymer 1978, 19, 1142–1144. [Google Scholar] [CrossRef]
  10. Keller, A. A note on single crystals in polymers: Evidence for a folded chain configuration. Philos. Mag. 1957, 2, 1171–1175. [Google Scholar] [CrossRef]
  11. Porter, R.S. Macromolecular physics, volume 3—crystal melting, Bernhard Wunderlich, Academic Press, New York, 1980, 363 pp. Price: $42.50. J. Polym. Sci. Polym. Lett. Ed. 1980, 18, 824. [Google Scholar] [CrossRef]
  12. Cheng, S.Z.D. Phase Transitions in Polymers: The Role of Metastable States; Elsevier: Oxford, UK, 2008. [Google Scholar]
  13. Zhang, F.; Sherrell, P.C.; Luo, W.; Chen, J.; Li, W.; Yang, J.; Zhu, M. Organic/Inorganic Hybrid Fibers: Controllable Architectures for Electrochemical Energy Applications. Adv. Sci. 2021, 8, e2102859. [Google Scholar] [CrossRef]
  14. Cheng, Y.; Wang, X.; Sun, B.; Chen, L.; Xiang, H.; Zhu, M.; Chen, Z.; Wang, S.; Zhou, Z. Organic-inorganic hybrid material and multifunctional fibers. Sci. Sin. Technol. 2014, 44, 1137–1144. [Google Scholar] [CrossRef]
  15. Cazan, C.; Enesca, A.; Andronic, L. Synergic Effect of TiO2 Filler on the Mechanical Properties of Polymer Nanocomposites. Polymers 2021, 13, 2017. [Google Scholar] [CrossRef]
  16. Righetti, M.C.; Di Lorenzo, M.L.; Cinelli, P.; Gazzano, M. Temperature dependence of the rigid amorphous fraction of poly(butylene succinate). RSC Adv. 2021, 11, 25731–25737. [Google Scholar] [CrossRef]
  17. Chen, B.; Torkelson, J.M. Rigid amorphous fraction and crystallinity in cold-crystallized syndiotactic polystyrene: Characterization by differential scanning calorimetry. Polymer 2021, 230, 124044. [Google Scholar] [CrossRef]
  18. Benz, J.; Bonten, C. Rigid amorphous fraction as an indicator for polymer-polymer interactions in highly filled plastics. Polymers 2021, 13, 3349. [Google Scholar] [CrossRef] [PubMed]
  19. Wu, Q. Polymer Condensation Process and Phase Transition; Higher Education Press: Beijing, China, 2016. [Google Scholar]
  20. Klonos, P.; Pissis, P. Effects of interfacial interactions and of crystallization on rigid amorphous fraction and molecular dynamics in polylactide/silica nanocomposites: A methodological approach. Polymer 2017, 112, 228–243. [Google Scholar] [CrossRef]
  21. Ma, Q.; Pyda, M.; Mao, B.; Cebe, P. Relationship between the rigid amorphous phase and mesophase in electrospun fibers. Polymer 2013, 54, 2544–2554. [Google Scholar] [CrossRef]
  22. Klonos, P.; Terzopoulou, Z.; Koutsoumpis, S.; Zidropoulos, S.; Kripotou, S.; Papageorgiou, G.Z.; Bikiaris, D.N.; Kyritsis, A.; Pissis, P. Rigid amorphous fraction and segmental dynamics in nanocomposites based on poly(l–lactic acid) and nano-inclusions of 1–3D geometry studied by thermal and dielectric techniques. Eur. Polym. J. 2016, 82, 16–34. [Google Scholar] [CrossRef]
  23. Shayestehfar, S.; Yazdanshenas, M.E.; Khajavi, R.; Rashidi, A.-S. Physical and mechanical properties of nylon 6/titanium dioxide micro and nano-composite multifilament yarns. J. Eng. Fibers Fabr. 2014, 9, 155892501400900319. [Google Scholar] [CrossRef]
  24. Cot, M.; Mijas, G.; Prieto-Fuentes, R.; Riba-Moliner, M.; Cayuela, D. The Influence of Titanium Dioxide (TiO2) Particle Size and Crystalline Form on the Microstructure and UV Protection Factor of Polyester Substrates. Polymers 2024, 16, 475. [Google Scholar] [CrossRef] [PubMed]
  25. Polinska, M.; Rozanski, A.; Kozanecki, M.; Galeski, A. Elastic modulus of the amorphous phase confined between lamellae: The role of crystalline component. Polymer 2023, 269, 125753. [Google Scholar] [CrossRef]
  26. Zhao, J.; Alsubhi, A.; Rastogi, S. Differences in thermal stability of Rigid Amorphous Fraction in nascent UHMWiPPs and its implication on lamellae thickening. Polymer 2025, 331, 128489. [Google Scholar] [CrossRef]
  27. Dalod, A.R.; Henriksen, L.; Grande, T.; Einarsrud, M.A. Functionalized TiO2 nanoparticles by single-step hydrothermal synthesis: The role of the silane coupling agents. Beilstein J. Nanotechnol. 2017, 8, 304–312. [Google Scholar] [CrossRef]
  28. Nguyen, T.-C.; Nguyen, T.-D.; Vu, D.-T.; Dinh, D.-P.; Nguyen, A.-H.; Ly, T.-N.-L.; Dao, P.-H.; Nguyen, T.-L.; Bach, L.-G.; Thai, H.; et al. Modification of Titanium Dioxide Nanoparticles with 3-(Trimethoxysilyl)propyl Methacrylate Silane Coupling Agent. J. Chem. 2020, 2020, 1381407. [Google Scholar] [CrossRef]
  29. Meroni, D.; Lo Presti, L.; Di Liberto, G.; Ceotto, M.; Acres, R.G.; Prince, K.C.; Bellani, R.; Soliveri, G.; Ardizzone, S. A Close Look at the Structure of the TiO2-APTES Interface in Hybrid Nanomaterials and Its Degradation Pathway: An Experimental and Theoretical Study. J. Phys. Chem. C 2017, 121, 430–440. [Google Scholar] [CrossRef]
  30. Chen, Q.; Li, F.; Zhai, Z.; Li, S.; Cai, Y.; Li, Q. Effect of Interfacial Compatibility on Mechanical Property of Polyamide 6 Modified by Polyborosiloxane. Polymers 2025, 17, 392. [Google Scholar] [CrossRef]
  31. Ludas Dujmic, A.; Radicic, R.; Ercegovic Razic, S.; Cingesar, I.K.; Glogar, M.; Jurov, A.; Krstulovic, N. Characterization of Melt-Spun Recycled PA 6 Polymer by Adding ZnO Nanoparticles during the Extrusion Process. Polymers 2024, 16, 1883. [Google Scholar] [CrossRef] [PubMed]
  32. Du, S.; Wei, Y.; Ahmed, S.; Zhou, F.; Tan, Y.; Li, Y.; Wang, M.; Chen, X.; Zhou, W. Enhanced thermal stability and UV resistance of polyamide 6 filament fabric via in-situ grafting with methyl methacrylate. Colloids Surf. A Physicochem. Eng. Asp. 2022, 651, 129371. [Google Scholar] [CrossRef]
  33. Yu, X.; Zhang, Y.; Fan, X.; Lv, Y.; Wang, Z.; Zhang, X.; Liu, T. Macromolecular PEI-modified carbon nanotubes as multifunctional additives in controlling crystallization and enhancing comprehensive performance of polyamide 6 nanocomposites. Compos. Commun. 2024, 51, 102057. [Google Scholar] [CrossRef]
  34. Huang, J.; Zhou, J.; Liu, M. Interphase in Polymer Nanocomposites. JACS Au 2022, 2, 280–291. [Google Scholar] [CrossRef]
  35. Buccella, M.; Dorigato, A.; Rizzola, F.; Caldara, M.; Fambri, L. Influence of the processing parameters on the dispersion and coloration behavior of a halogenated copper phthalocyanine-based masterbatch. Adv. Polym. Technol. 2018, 37, 778–785. [Google Scholar] [CrossRef]
  36. Wang, L.; Xie, G.; Mi, X.; Zhang, B.; Du, Y.; Zhu, Q.; Yu, Z. Surface-Modified TiO2@SiO2 Nanocomposites for Enhanced Dispersibility and Optical Performance to Apply in the Printing Process as a Pigment. ACS Omega 2023, 8, 20116–20124. [Google Scholar] [CrossRef]
  37. Zhu, R.; Yadama, V.; Liu, H.; Lin, R.J.T.; Harper, D.P. Fabrication and characterization of Nylon 6/cellulose nanofibrils melt-spun nanocomposite filaments. Compos. Part A Appl. Sci. Manuf. 2017, 97, 111–119. [Google Scholar] [CrossRef]
  38. Fu, S.; Sun, Z.; Huang, P.; Li, Y.; Hu, N. Some basic aspects of polymer nanocomposites: A critical review. Nano Mater. Sci. 2019, 1, 2–30. [Google Scholar] [CrossRef]
  39. EN 13900-5:2005; Pigments and Extenders—Methods of Dispersion and Assessment of Dispersibility in Plastics—Part 5: Determination by Filter Pressure Value Test. CEN: Brussels, Belgium, 2005.
  40. Kissinger, H.E. Reaction Kinetics in Differential Thermal Analysis. Anal. Chem. 1957, 29, 1702–1706. [Google Scholar] [CrossRef]
  41. Akahira, T.; Sunose, T. Method of determining activation deterioration constant of electrical insulating materials. Res. Rep. Chiba. Inst. Technol. (Sci. Technol.) 1971, 16, 22–31. [Google Scholar]
  42. Nakamura, K.; Watanabe, T.; Katayama, K.; Amano, T. Some aspects of nonisothermal crystallization of polymers. I. Relationship between crystallization temperature, crystallinity, and cooling conditions. J. Appl. Polym. Sci. 1972, 16, 1077–1091. [Google Scholar] [CrossRef]
  43. Vyazovkin, S.; Burnham, A.K.; Criado, J.M.; Pérez-Maqueda, L.A.; Popescu, C.; Sbirrazzuoli, N. ICTAC Kinetics Committee recommendations for performing kinetic computations on thermal analysis data. Thermochim. Acta 2011, 520, 1–19. [Google Scholar] [CrossRef]
  44. Huang, Y.-P.; Chen, T.-K.; Tang, J.-W.; Yeh, C.; Tien, C.-H. Effect of PET melt spinning on TiO2 nanoparticle aggregation and friction behavior of fiber surface. Ind. Eng. Chem. Res. 2007, 46, 5548–5554. [Google Scholar] [CrossRef]
  45. Ismail, M.A.; Yousef, E.A.; Nasr, G.M. Surface modification of montmorillonite—MMT nanofiller: How it affects both the rigid amorphous fraction (RAF) and the physical properties of polyamide 66. Polym. Bull. 2025, 82, 3553–3611. [Google Scholar] [CrossRef]
  46. Li, C.; Xia, Z.; Wu, L.; Xiong, Y.; Chen, W. Strain-Dependent Evolution of the Rigid Amorphous Fraction of Low-Density Polyethylene under Deformation. Macromolecules 2025, 58, 2320–2335. [Google Scholar] [CrossRef]
  47. Feigin, L.; Svergun, D.I. Structure Analysis by Small-Angle X-Ray and Neutron Scattering; Springer: Boston, MA, USA, 1987; Volume 1. [Google Scholar]
  48. Rambo, R.P.; Tainer, J.A. Characterizing flexible and intrinsically unstructured biological macromolecules by SAS using the Porod-Debye law. Biopolymers 2011, 95, 559–571. [Google Scholar] [CrossRef]
  49. Glatter, O. Small Angle X-Ray Scattering; Academic Press: London, UK, 1982. [Google Scholar]
  50. Hermans, P.; Weidinger, A. Quantitative x-ray investigations on the crystallinity of cellulose fibers. A background analysis. J. Appl. Phys. 1948, 19, 491–506. [Google Scholar] [CrossRef]
  51. Di Lorenzo, M.L.; Silvestre, C. Non-isothermal crystallization of polymers. Prog. Polym. Sci. 1999, 24, 917–950. [Google Scholar] [CrossRef]
  52. AATCC TM183; Transmittance or Blocking of Erythemally Weighted Ultraviolet Radiation through Fabrics. American Association of Textile Chemists and Colorists (AATCC): Research Triangle Park, NC, USA, 2020.
  53. Bhattacharjee, S. DLS and zeta potential—What they are and what they are not? J. Control. Release 2016, 235, 337–351. [Google Scholar] [CrossRef] [PubMed]
  54. Scientific, T.F. Filter Pressure Value Test (FPV test)—A New Standard to Examine Master Batches; Application Note LR49-e; Thermo Fisher Scientific: Waltham, MA, USA, 2007. [Google Scholar]
  55. Müller, A.R.K.; Rupp, H. Quality of Dispersion—Standardised: The Filter Pressure Value as an Evaluation Tool. Kunststoffe Int. 2004, 94, 70–73. [Google Scholar]
  56. Zhang, X. Applications of Kinetic Methods in Thermal Analysis: A Review. Eng. Sci. 2020, 14, 1–13. [Google Scholar] [CrossRef]
  57. Feng, T.; Xiu-Hong, L.; Yu-Zhu, W.; Chun-Ming, Y.; Ping, Z.; Jin-You, L.; Zeng, J.; Chun-Xia, H.; Wen-Qiang, H.; Xiao-Yun, L.; et al. Small angle X-ray scattering beamline at SSRF. Nucl. Sci. Tech. 2015, 26. [Google Scholar] [CrossRef]
  58. Kieffer, J.; Karkoulis, D. PyFAI, a versatile library for azimuthal regrouping. J. Phys. Conf. Ser. 2013, 425, 202012. [Google Scholar] [CrossRef]
  59. Xu, H.; Ince, B.S.; Cebe, P. Development of the crystallinity and rigid amorphous fraction in cold-crystallized isotactic polystyrene. J. Polym. Sci. Part B Polym. Phys. 2003, 41, 3026–3036. [Google Scholar] [CrossRef]
  60. Ogle, J.; Powell, D.; Amerling, E.; Smilgies, D.-M.; Whittaker-Brooks, L. Quantifying multiple crystallite orientations and crystal heterogeneities in complex thin film materials. CrystEngComm 2019, 21, 5707–5720. [Google Scholar] [CrossRef]
  61. Marrink, S.J.; Risselada, H.J.; Yefimov, S.; Tieleman, D.P.; de Vries, A.H. The MARTINI force field: Coarse grained model for biomolecular simulations. J. Phys. Chem. B 2007, 111, 7812–7824. [Google Scholar] [CrossRef]
  62. Alessandri, R.; Barnoud, J.; Gertsen, A.S.; Patmanidis, I.; De Vries, A.H.; Souza, P.C.T.; Marrink, S.J. Martini 3 coarse-grained force field: Small molecules. Adv. Theory Simul. 2022, 5, 2100391. [Google Scholar] [CrossRef]
  63. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1-2, 19–25. [Google Scholar] [CrossRef]
  64. Mousavifard, S.M.; Ghermezcheshme, H.; Mirzaalipour, A.; Mohseni, M.; de With, G.; Makki, H. PolySMart: A general coarse-grained molecular dynamics polymerization scheme. Mater. Horiz. 2023, 10, 2281–2296. [Google Scholar] [CrossRef] [PubMed]
  65. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; in ’t Veld, P.J.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; et al. LAMMPS—A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comput. Phys. Commun. 2022, 271, 108171. [Google Scholar] [CrossRef]
  66. Luo, C.; Kröger, M.; Sommer, J.-U. Molecular dynamics simulations of polymer crystallization under confinement: Entanglement effect. Polymer 2017, 109, 71–84. [Google Scholar] [CrossRef]
  67. Mortezazadeh, S.; Jamali, Y.; Naderi-Manesh, H.; Lyubartsev, A.P. Implicit solvent systematic coarse-graining of dioleoylphosphatidylethanolamine lipids: From the inverted hexagonal to the bilayer structure. PLoS ONE 2019, 14, e0214673. [Google Scholar] [CrossRef] [PubMed]
  68. Pereira, G.P.; Alessandri, R.; Dominguez, M.; Araya-Osorio, R.; Grunewald, L.; Borges-Araujo, L.; Wu, S.; Marrink, S.J.; Souza, P.C.T.; Mera-Adasme, R. Bartender: Martini 3 Bonded Terms via Quantum Mechanics-Based Molecular Dynamics. J. Chem. Theory Comput. 2024, 20, 5763–5773. [Google Scholar] [CrossRef]
  69. Reith, D.; Putz, M.; Muller-Plathe, F. Deriving effective mesoscale potentials from atomistic simulations. J. Comput. Chem. 2003, 24, 1624–1636. [Google Scholar] [CrossRef]
  70. Moore, T.C.; Iacovella, C.R.; McCabe, C. Derivation of coarse-grained potentials via multistate iterative Boltzmann inversion. J. Chem. Phys. 2014, 140, 224104. [Google Scholar] [CrossRef]
  71. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2011, 2, 73–78. [Google Scholar] [CrossRef]
  72. Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural triple excitations in local coupled cluster calculations with pair natural orbitals. J. Chem. Phys. 2013, 139, 134101. [Google Scholar] [CrossRef]
  73. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef]
  74. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  75. Wilkes, G.L. An overview of the basic rheological behavior of polymer fluids with an emphasis on polymer melts. J. Chem. Educ. 1981, 58, 880. [Google Scholar] [CrossRef]
  76. Sangroniz, L.; Fernández, M.; Santamaria, A. Polymers and rheology: A tale of give and take. Polymer 2023, 271, 125811. [Google Scholar] [CrossRef]
  77. Pissis, P.; Klonos, P.; Kyritsis, A. Interfacial effects in polymer nanocomposites studied by dielectric and thermal techniques. In Proceedings of the 2011—14th International Symposium on Electrets, Piscataway, NJ, USA, 28–31 August 2011; pp. 67–68. [Google Scholar]
  78. Keddie, J.L.; Jones, R.A.L.; Cory, R.A. Interface and surface effects on the glass-transition temperature in thin polymer films. Faraday Discuss. 1994, 98, 219–230. [Google Scholar] [CrossRef]
  79. Yu, D.; Wang, J.; Wang, G. Glass fiber treated with a glycine bridged silane coupling agent reinforcing polyamide 6(PA6): Effect of hydrogen bonding. RSC Adv. 2025, 15, 3331–3338. [Google Scholar] [CrossRef]
  80. Doi, M.; Edwards, S.F.; Edwards, S.F. The Theory of Polymer Dynamics; Oxford University Press: Oxford, UK, 1988; Volume 73. [Google Scholar]
  81. Ferry, J.D. Viscoelastic Properties of Polymers; John Wiley & Sons: New York, NY, USA, 1980. [Google Scholar]
  82. Landel, R.F.; Nielsen, L.E. Mechanical Properties of Polymers and Composites; CRC Press: Boca Raton, FL, USA, 1993. [Google Scholar]
  83. Allen, N.S.; Chirinis-Padron, A.; Henman, T.J. The photo-stabilisation of polypropylene: A review. Polym. Degrad. Stab. 1985, 13, 31–76. [Google Scholar] [CrossRef]
  84. Ward, I.M.; Sweeney, J. Mechanical Properties of Solid Polymers; Wiley Online Library: Chichester, UK, 1971; Volume 313. [Google Scholar]
  85. Forrest, J.A.; Dalnoki-Veress, K. The glass transition in thin polymer films. Adv. Colloid Interface Sci. 2001, 94, 167–195. [Google Scholar] [CrossRef]
  86. Blachnio, M.; Zienkiewicz-Strzalka, M.; Derylo-Marczewska, A.; Nosach, L.V.; Voronin, E.F. Chitosan-Silica Composites for Adsorption Application in the Treatment of Water and Wastewater from Anionic Dyes. Int. J. Mol. Sci. 2023, 24, 11818. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic illustration of (a) hydrolysis and condensation of silane coupling agents on TiO2 surface, (b) interfacial hydrogen bonding between KH550 and PA6, (c) polar interaction between KH570 and PA6, and (d) processing route from surface modification to masterbatch compounding, fiber spinning, and fabric weaving.
Scheme 1. Schematic illustration of (a) hydrolysis and condensation of silane coupling agents on TiO2 surface, (b) interfacial hydrogen bonding between KH550 and PA6, (c) polar interaction between KH570 and PA6, and (d) processing route from surface modification to masterbatch compounding, fiber spinning, and fabric weaving.
Polymers 17 02551 sch001
Table 1. Fiber samples used for quantitative analysis and their preparation parameters.
Table 1. Fiber samples used for quantitative analysis and their preparation parameters.
Sample IDSilane Coupling AgentTiO2 Content (wt%)Draw Ratio
pure PA6None03.0
PA6-1.6 wt% TiO2-KH550KH5501.63.0
PA6-4.0 wt% TiO2-KH550KH5504.03.0
PA6-4.0 wt% TiO2-KH570KH5704.03.0
Sample IDs (e.g., 550-1.6%, 570-4%) are shorthand notations used consistently in figures and tables to denote PA6/TiO2 fibers with specified TiO2 content (wt%) and silane coupling agents (KH550 or KH570).
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

Yu, H.; Liu, P.; Ji, X.; Jiang, X.; Sun, B. Interphase-Resolved Performance in PA6/TiO2 Nanocomposite Fibers: Four-Phase Geometry Linking Structure to Mechanical and UV Protection. Polymers 2025, 17, 2551. https://doi.org/10.3390/polym17182551

AMA Style

Yu H, Liu P, Ji X, Jiang X, Sun B. Interphase-Resolved Performance in PA6/TiO2 Nanocomposite Fibers: Four-Phase Geometry Linking Structure to Mechanical and UV Protection. Polymers. 2025; 17(18):2551. https://doi.org/10.3390/polym17182551

Chicago/Turabian Style

Yu, Hailong, Ping Liu, Xiaohuan Ji, Xiaoze Jiang, and Bin Sun. 2025. "Interphase-Resolved Performance in PA6/TiO2 Nanocomposite Fibers: Four-Phase Geometry Linking Structure to Mechanical and UV Protection" Polymers 17, no. 18: 2551. https://doi.org/10.3390/polym17182551

APA Style

Yu, H., Liu, P., Ji, X., Jiang, X., & Sun, B. (2025). Interphase-Resolved Performance in PA6/TiO2 Nanocomposite Fibers: Four-Phase Geometry Linking Structure to Mechanical and UV Protection. Polymers, 17(18), 2551. https://doi.org/10.3390/polym17182551

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