Fact-checked by Grok 2 weeks ago

Polymer characterization

Polymer characterization encompasses the suite of analytical techniques employed to elucidate the molecular , , physical properties, and attributes of polymeric materials, both synthetic and natural. These methods are indispensable for assessing factors such as molecular weight distribution, crystallinity, transitions, and behavior, which directly influence a polymer's processability, , and functionality in applications ranging from packaging to biomedical devices. The importance of polymer characterization lies in its role in , , , and efforts, such as and , by providing insights into elemental makeup, phase composition, and structural integrity. Due to the inherent complexity of polymers—which can exhibit polydispersity, branching, and varying degrees of crystallinity—multiple complementary techniques are typically required to obtain a comprehensive profile, categorized broadly into chemical, molecular, and bulk property analyses. Key chemical characterization techniques include (NMR) spectroscopy, which identifies chemical bonds, functional groups, and sequence distributions in polymer chains, though it is best suited for shorter chains with fewer than 100 repeat units; Fourier-transform infrared (FTIR) and , which detect functional groups based on dipole moments and vibrational modes, respectively; and variants like (MALDI-MS) for precise molecular weight determination. For molecular size and distribution, (SEC) or (GPC), often coupled with (SEC-MALS), measures number-average (M_n) and weight-average (M_w) molecular weights, as well as polydispersity index (M_w/M_n), which typically approaches 2 for condensation polymers near reaction completion. (SLS) and (DLS) further quantify and , respectively, aiding in conformational analysis such as models with end-to-end distances proportional to the of chain length. Thermal and bulk properties are evaluated using (DSC), which determines temperature (T_g), (T_m), and crystallization behavior to assess purity and processing conditions; (TGA) for decomposition profiles and stability; and (XRD) to quantify crystalline versus amorphous content, influencing tensile strength and . Morphological insights come from techniques like (SAXS) for nanoscale structures and (SEM/TEM) for surface and internal features, while (DMA) measures storage modulus and viscoelastic responses. These methods, often applied in combination, enable precise control over properties, from the rotational isomeric states affecting short-range interactions to long-range effects characterized by theta temperatures where behavior occurs.

Introduction

Definition and Scope

Polymer characterization is the analytical process of determining the molecular and supramolecular properties of polymers, encompassing measurements of molecular weight, , , thermal transitions, and mechanical behavior to correlate synthesis conditions, processing parameters, and material performance. This discipline addresses the inherent heterogeneity of polymers, which consist of non-identical macromolecules varying in chain length, branching, and , unlike small-molecule compounds. The field originated in the early 20th century through the foundational work of Hermann Staudinger, who in 1920 proposed that polymers such as natural rubber are composed of high-molecular-weight macromolecules linked by covalent bonds, challenging prevailing views of colloidal aggregates. Staudinger's macromolecular hypothesis, substantiated through experiments like viscometry and hydrogenation, established the basis for modern polymer science and earned him the Nobel Prize in Chemistry in 1953 for discoveries in macromolecular chemistry. His contributions enabled the systematic study of both natural polymers, like proteins and polysaccharides, and synthetic ones, such as polyesters and polyolefins. The scope of polymer characterization spans diverse techniques, including , , , and , to provide insights into polymer architecture and behavior across scales from individual chains to bulk materials. A central metric within this scope is the polydispersity index (PDI), now officially termed (Đ) by IUPAC, defined as the ratio of the weight-average molecular weight to the number-average molecular weight (M_w / M_n), which quantifies the breadth of the molecular weight distribution and influences properties like and strength. Techniques are categorized broadly into molecular characterization (e.g., assessing average molecular weights and distributions), (e.g., evaluating and ), morphological examination (e.g., probing crystallinity), and performance evaluation (e.g., measuring and responses). This classification facilitates targeted investigations into how polymer features dictate functionality in applications ranging from packaging to biomedical devices.

Importance and Applications

Polymer characterization plays a pivotal role in by enabling the establishment of structure-property relationships, which allow researchers to link molecular-level features, such as chain length and branching, to macroscopic behaviors like mechanical strength and thermal stability. This correlation is essential for troubleshooting issues during synthesis, such as unintended side reactions that alter molecular weight distributions, and for predicting end-use performance in diverse applications from to . Without thorough , it becomes challenging to optimize synthesis conditions or scale up production while maintaining desired properties. In industrial settings, polymer characterization is indispensable for during , exemplified by the monitoring of molecular weight consistency in production to ensure uniform tensile strength and processability for pipes and films. It also supports , particularly in , where the U.S. (FDA) mandates detailed material characterization for medical devices to verify , stability, and absence of leachables that could cause adverse reactions. For instance, characterization of polymers like polyurethanes in implants helps confirm long-term performance and safety under physiological conditions. In research, characterization drives innovation in sustainable materials, such as biodegradable polymers where techniques assess kinetics and mechanical integrity to tailor environmental breakdown without compromising functionality. It is equally vital for nanocomposites, revealing filler and interfacial interactions that enhance properties like or barrier performance in eco-friendly composites. Post-2020 trends emphasize characterizing recycled polymers to support the , evaluating purity, molecular weight , and compatibility for reprocessing into high-value products, thereby reducing reliance on virgin feedstocks. As of 2025, advancements include AI-assisted characterization for faster of recycled polymer compositions and strategies for waste plastics into high-performance materials. However, incomplete characterization poses significant challenges, often leading to material failures such as those observed in polyurethane-based implants due to undetected resulting in reduced molecular weight and compromised mechanical performance. These issues underscore the need for comprehensive to prevent costly recalls and environmental harm in applications ranging from consumer goods to structural components.

Molecular Weight Characterization

Average Molecular Weights and Distributions

In , synthetic polymers are inherently heterogeneous in chain length due to the statistical nature of polymerization reactions, necessitating the use of multiple average molecular weights to fully characterize their size distribution. This polydispersity arises from variations in , , and termination steps, leading to a population of chains with differing degrees of rather than uniform lengths. Unlike small molecules, where a single molecular weight suffices, polymers require averages that weight chains differently based on number, mass, or higher moments to capture their bulk accurately. The number-average molecular weight, M_n, is defined as the total mass of all chains divided by the total number of chains, emphasizing shorter chains in the distribution. In contrast, the weight-average molecular weight, M_w, weights each chain by its mass, giving prominence to longer chains and typically yielding a higher value than M_n. The viscosity-average molecular weight, M_v, arises in viscometric measurements and lies between M_n and M_w, reflecting hydrodynamic interactions in . Higher-order averages, such as the z-average molecular weight, M_z, further weight by the square of chain mass, providing insight into the tail of the distribution. The breadth of this is quantified by the polydispersity index (PDI), defined as \text{PDI} = M_w / M_n, where a value of 1 indicates a monodisperse with all chains of equal , and PDI > 1 signifies polydispersity, common in most synthetic polymers. Monodisperse polymers, often achieved through living techniques, exhibit uniform properties, while polydisperse ones show broader variations. , such as polycondensation, typically yields a Flory-Schulz , characterized by an in chain probabilities, leading to PDI ≈ 2 at high conversion. Chain-growth mechanisms, like free-radical , often produce log-normal distributions, skewed toward higher weights due to termination events. These averages profoundly influence polymer properties and performance. The weight-average M_w dominates rheological behavior, particularly melt viscosity \eta, which scales as \eta \propto M_w^{3.4} in the entangled regime above the critical entanglement molecular weight, due to dynamics constraining chain motion. Conversely, the number-average M_n controls thermodynamic properties like , where lower M_n enhances chain mobility and reduces entanglement, facilitating dissolution in solvents. For crystallinity, higher M_n promotes ordered packing in semicrystalline polymers by increasing chain overlap, though excessive polydispersity can disrupt perfection. Thus, understanding these distributions is essential for tailoring polymers to applications ranging from adhesives to structural materials.

Absolute Determination Methods

Absolute determination methods enable the direct measurement of molecular weights without the need for against standards, providing intrinsic values such as the number-average molecular weight M_n or weight-average molecular weight M_w, which are essential for understanding properties independent of relative techniques. These approaches rely on thermodynamic, scattering, or mass-based principles and are particularly valuable for validating results from faster, -dependent methods like . Key techniques include , light scattering, ultracentrifugation, and , each suited to specific molecular weight ranges and types. Osmotic pressure measurement determines M_n by exploiting the colligative property of solutions, where the pressure \pi across a separating polymer solution from pure is related to concentration c via the : \pi = RT \frac{c}{M_n} + A_2 RT c^2 + \cdots Here, R is the , T the absolute , and A_2 the second virial coefficient reflecting polymer-solvent interactions. Plotting \pi / c versus c yields a straight line at low concentrations, with the intercept equal to RT / M_n, allowing direct calculation of M_n. This method is ideal for low molecular weight polymers (typically M_n < 50,000 Da) because high-MW species contribute negligibly to \pi, ensuring sensitivity to the entire distribution, though it requires careful membrane selection to avoid leaks and is time-intensive for equilibrium attainment. Static light scattering (SLS) provides absolute M_w and the radius of gyration R_g, characterizing polymer chain dimensions in dilute solutions by measuring the angular dependence of scattered laser light intensity. The scattered intensity \Delta R_\theta relates to concentration and scattering vector via the Zimm plot, derived from the Debye approximation for the Rayleigh ratio: \frac{K c}{\Delta R_\theta} = \frac{1}{M_w} + 2 A_2 c + \cdots where K incorporates optical constants like the refractive index increment dn/dc. Data from multiple angles and concentrations are plotted as Kc / \Delta R_\theta versus \sin^2(\theta/2) + k c (with k a scaling factor), extrapolating to zero angle and concentration to obtain $1/M_w from the intercept and R_g from the initial slope; this graphical method, introduced by Zimm, corrects for intramolecular interference and multiple scattering. SLS excels for high-MW polymers (up to millions of Da) but demands dust-free samples and knowledge of dn/dc, with A_2 providing insights into chain conformation in good or theta solvents. Analytical ultracentrifugation via sedimentation equilibrium measures M_w by balancing centrifugal force with diffusion in a sector-shaped cell, achieving an exponential concentration profile at equilibrium described by: c(r) = c(r_0) \exp\left[ M (1 - \bar{v} \rho) \omega^2 (r^2 - r_0^2) / (2 RT) \right] where c(r) is concentration at radial position r, \bar{v} the partial specific volume, \rho the solvent density, and \omega the angular velocity. Fitting absorbance or interference data to this equation yields M_w directly, with multiple speeds used to resolve heterogeneity or associations; pioneered by Svedberg, this thermodynamic method is calibration-free and applicable to polymers up to $10^6 Da in solution, though it requires high rotor speeds (up to 60,000 rpm) and precise buoyancy corrections. It is especially useful for studying polymer self-association or polydispersity without fractionation. Matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry offers absolute molecular weight determination for synthetic polymers by ionizing intact chains in a matrix that facilitates soft desorption, producing singly charged ions whose flight time to the detector reveals mass-to-charge ratio m/z \approx M + cation mass. For polymers like polyethylene glycol or polystyrene, spectra show oligomer peaks separated by the repeat unit mass, enabling calculation of M_n and M_w from peak intensities, with polydispersity often near 1.0 for narrow distributions up to $10^4 to $10^5 Da. Matrix selection (e.g., 2,5-dihydroxybenzoic acid) and cationization (e.g., with Na^+) minimize fragmentation, but challenges include mass discrimination favoring lower-MW species in polydisperse samples and limited resolution for ultra-high MW due to ion transmission inefficiencies. This technique provides end-group analysis alongside MW, making it complementary for low-to-medium MW polymers where sequence information aids structural elucidation.

Relative Determination Methods

Relative determination methods for polymer molecular weight estimation rely on empirical calibrations using standards of known molecular weight, offering efficient and accessible alternatives to absolute techniques for routine analysis. These approaches, including , viscometry, and end-group analysis, provide estimates of average molecular weights such as the viscosity-average (M_v) or number-average (M_n), but require assumptions about polymer architecture and solvent interactions. They are particularly valuable in industrial settings for quality control, where high throughput and low cost outweigh the need for absolute precision. Gel permeation chromatography (GPC), also known as size-exclusion chromatography (SEC), separates polymer molecules based on their hydrodynamic volume as they pass through a porous stationary phase, with larger molecules eluting first due to limited access to the pores. Developed by J. C. Moore in 1964, GPC enables the determination of molecular weight distributions by calibrating elution volumes against narrow molecular weight standards, typically polystyrene in organic solvents. The relationship between elution volume and molecular weight is often established using the universal calibration principle, which plots the logarithm of the product of intrinsic viscosity and molecular weight (log([η]M)) versus elution volume, allowing application across different polymers provided their hydrodynamic behavior is comparable. In practice, calibration with polystyrene standards yields relative molecular weights, with detectors such as refractive index or light scattering providing distribution profiles; for example, polystyrene standards with molecular weights from 10^3 to 10^6 g/mol are commonly used to cover a broad range. The intrinsic viscosity plays a central role in GPC calibration through the Mark-Houwink equation, expressed as [\eta] = K M^a where [η] is the intrinsic viscosity (in dL/g), M is the molecular weight, and K and a are empirical constants dependent on the polymer, solvent, and temperature (typically a ranges from 0.5 for random coils to 0.8 for rigid rods). This equation, originally formulated by H. Mark and R. Houwink in the 1930s and refined by I. Sakurada, links solution viscosity to chain dimensions, enabling conversion of hydrodynamic volume to molecular weight estimates. For polystyrene in tetrahydrofuran at 25°C, representative values are K = 1.10 × 10^{-4} dL/g and a = 0.725, illustrating how calibration curves are constructed for specific systems. Viscometry measures the intrinsic viscosity [η] of dilute polymer solutions to estimate the viscosity-average molecular weight M_v via the , making it a simple, low-cost method suitable for routine laboratory use. The intrinsic viscosity is defined as [\eta] = \lim_{c \to 0} \frac{\eta_{sp}}{c} where η_sp = (η - η_0)/η_0 is the specific viscosity, η is the solution viscosity, η_0 is the solvent viscosity, and c is the polymer concentration (typically in g/dL); this limit is extrapolated from measurements at several low concentrations using capillary viscometers like the Ubbelohde type. Pioneered by H. Staudinger in the 1930s, viscometry correlates solution flow resistance to polymer chain length, with M_v providing a weight-biased average sensitive to long chains. For example, in poly(ethylene oxide) solutions, viscometry yields M_v values accurate to within 5-10% when Mark-Houwink parameters are known, though it requires prior calibration with absolute methods for K and a. End-group analysis determines the number-average molecular weight M_n by quantifying the concentration of functional end groups in telechelic or condensation polymers, assuming each chain has one or two reactive termini. For polymers like polyamides or polyesters, M_n is calculated as M_n = (concentration of repeating units × molecular weight of repeat unit) / (end-group concentration), often via titration or spectroscopy. Acid-base titration, for instance, measures carboxylic acid end groups in poly(ethylene terephthalate) by reacting with a base like potassium hydroxide, with equivalence points detected colorimetrically; this method, refined in the 1960s, achieves accuracy of ±2% for M_n up to 20,000 g/mol. Spectroscopy, such as NMR, can identify specific end groups like hydroxyls in polyesters, providing structural confirmation alongside M_n. This technique is most effective for low-molecular-weight polymers (M_n < 10,000 g/mol) where end groups are abundant relative to chain segments. These relative methods assume linear, monodisperse chains and similar conformational behavior to calibration standards, leading to errors of 10-20% or more in branched or rigid polymers due to deviations in hydrodynamic volume. For validation, they are often cross-checked with absolute methods like light scattering.

Molecular Structure Characterization

Chemical Composition Analysis

Chemical composition analysis in polymer characterization involves techniques that quantify the elemental makeup, functional groups, and monomer distributions within polymer chains, essential for understanding copolymer structures and properties. These methods provide insights into monomer ratios and sequence irregularities, which influence material performance in applications such as elastomers and plastics. Elemental analysis, particularly CHN combustion, determines the carbon, hydrogen, and nitrogen content of polymers by high-temperature combustion followed by gas detection, offering bulk composition data for copolymers. In this process, samples are combusted at temperatures around 900–1000°C, converting elements into detectable gases like CO₂, H₂O, and N₂, with results calibrated against standards for accuracy within 0.3%. For example, in ethylene-propylene copolymers, the C/H atomic ratio derived from CHN data allows calculation of the ethylene-to-propylene mole fraction, as ethylene (C₂H₄) has a higher carbon-to-hydrogen ratio than propylene (C₃H₆), enabling composition estimates without spectroscopic resolution of sequences. This technique is particularly valuable for verifying synthesis outcomes in polyolefins, though it requires corrections for oxygen or additives if present. Fourier-transform infrared (FTIR) spectroscopy identifies and quantifies functional groups through characteristic vibrational peaks, providing a rapid assessment of copolymer content via absorbance ratios. Common assignments include the C=O stretch at approximately 1700 cm⁻¹ for carbonyl groups in polyesters or acrylics, with peak intensity proportional to concentration after baseline correction and Beer-Lambert law application. For quantitative analysis, FTIR measures copolymer composition by comparing band areas, such as the 1730 cm⁻¹ ester peak versus a reference like 2920 cm⁻¹ C-H stretch in ethylene-methyl acrylate copolymers, achieving accuracies of 1–2 mol% with proper calibration curves. This method excels in detecting functional group heterogeneity but may require chemometric tools like partial least squares for complex spectra. Nuclear magnetic resonance (NMR) spectroscopy, using ¹H and ¹³C nuclei, offers detailed monomer sequencing by resolving chemical shifts corresponding to local environments in copolymer chains. In ¹H NMR, proton signals from distinct monomers are integrated to yield dyad (e.g., AA, AB, BB) fractions, while ¹³C NMR provides higher resolution for triad (e.g., AAA, AAB, ABB) distributions due to larger chemical shift dispersions, often exceeding 10 ppm for carbon atoms in polyolefins. For instance, in styrene-butadiene copolymers, ¹³C NMR integration of methine carbons at 40–45 ppm quantifies 1,2- versus 1,4-addition modes, with triad fractions calculated from peak areas after decoupling and relaxation corrections to ensure quantitative accuracy. These integrations follow the formula for sequence probability, where dyad content AB = 2 × (integral of AB peak) / total monomer integrals, enabling reactivity ratio estimations. Recent advances in high-resolution NMR, particularly post-2020, have enhanced sensitivity for sequence distribution in complex copolymers through techniques like dynamic nuclear polarization (DNP) and automated fast acquisition protocols. DNP boosts signal-to-noise ratios by 50–100 times via microwave-induced hyperpolarization, allowing detection of low-abundance triads in biobased copolymers at natural abundance levels. Additionally, ultra-fast ¹³C NMR instruments now resolve microstructures in polyolefins within minutes, improving throughput for industrial quality control of sequence irregularities in materials like polyethylene-co-1-hexene. These developments, including multidimensional HSQC for precise peak assignments, have expanded applicability to heterogeneous systems without isotopic labeling.

Chain Conformation and Tacticity

Chain conformation and tacticity are key aspects of polymer microstructure that influence local chain arrangement, flexibility, and overall material properties such as crystallinity and mechanical behavior. Tacticity describes the stereochemical configuration of repeat units along the polymer backbone, arising from the relative orientations of substituent groups around chiral centers. In vinyl polymers, three primary types exist: isotactic, where all substituents are on the same side of the zigzag chain; syndiotactic, where they alternate sides; and atactic, which lacks regularity and features random configurations. These configurations are determined during polymerization, often by catalysts like Ziegler-Natta systems that favor specific stereoregularity, and they directly affect chain packing and secondary interactions. Nuclear magnetic resonance (NMR) spectroscopy, particularly ^{13}C NMR, is the standard method for quantifying tacticity by resolving chemical shifts sensitive to stereosequences. In polypropylene, for instance, the methyl carbon resonances distinguish triad sequences: mm (isotactic), mr (heterotactic), and rr (syndiotactic), with peaks at approximately 21-22 ppm for mm, 20.5 ppm for mr, and 19-20 ppm for rr, allowing calculation of triad fractions to assess overall stereoregularity. High-resolution techniques, including one-dimensional and two-dimensional NMR, enable precise assignment of higher-order pentads and hexads, providing detailed microstructure analysis even in complex copolymers. This approach has been pivotal since early applications in the 1960s, revealing how defects in isotactic sequences impact properties like melting point. Branching introduces deviations from linear chain topology, particularly long-chain branching (LCB), which connects side chains longer than a critical length to the backbone, altering hydrodynamic volume and entanglement. LCB is detected using gel permeation chromatography (GPC) coupled with inline viscometry and multi-angle light scattering (MALS), which measure intrinsic viscosity [\eta] and radius of gyration R_g across molecular weight fractions. The viscosity branching index g' is defined as g' = [\eta]{br} / [\eta]{lin}, where [\eta]{br} is the intrinsic viscosity of the branched polymer and [\eta]{lin} that of a linear reference at equivalent molecular weight; values below 1 indicate contraction due to branching. Similarly, the size branching index g = \langle R_g^2 \rangle_{br} / \langle R_g^2 \rangle_{lin} follows Zimm-Stockmayer theory, relating branching density to reduced dimensions via combinatorial statistics for random branching. In polyethylene, for example, g' values as low as 0.7 signal significant LCB, correlating with enhanced melt strength. Polymer chain conformation is modeled using the rotational isomeric state (RIS) approximation, which discretizes torsional rotations around backbone bonds into discrete states (e.g., trans, gauche^+, gauche^-) with Boltzmann-weighted probabilities based on energy minima. Developed by , this model computes statistical averages like the characteristic ratio C_\infty = \langle R^2 \rangle / (n l^2), where n is the number of bonds and l the bond length, to quantify deviation from ideal Gaussian chains due to local stiffness and excluded volume. The persistence length \xi_p, defined as the projection of the chain along its initial tangent decaying to 1/e, serves as a measure of overall stiffness; for flexible polymers like polyethylene, \xi_p \approx 5-10 Å, while rigid rods exceed 100 Å. RIS parameters are derived from quantum calculations or experiments, enabling predictions of end-to-end distances and dipole moments. Small-angle neutron scattering (SANS) provides direct experimental validation of chain dimensions and conformation in dilute solutions, leveraging neutron contrast between deuterated solvents and protonated polymers. In the Guinier regime (q R_g < 1, where q is the scattering vector), the scattered intensity I(q) \approx I(0) (1 - q^2 R_g^2 / 3), yielding R_g from low-q slopes; for higher q, the Debye function describes Gaussian coils, I(q) = (2 / (q^2 R_g^2)^2) (\exp(-q^2 R_g^2) + q^2 R_g^2 - 1). SANS distinguishes swollen coils in good solvents (R_g \propto M^{0.588}) from theta conditions (R_g \propto M^{0.5}), as demonstrated in polystyrene solutions where chain expansion aligns with Flory exponents. This technique is especially valuable for inaccessible regimes like bulk melts via contrast matching.

Morphology Characterization

Crystallinity and Semicrystallinity

Semi-crystalline polymers exhibit a coexistence of ordered crystalline regions and disordered amorphous phases, where the degree of crystallinity, typically ranging from 20% to 80% depending on the polymer and processing conditions, profoundly affects properties such as stiffness, permeability, and thermal stability. Quantifying this crystallinity and associated lamellar structures is crucial for understanding morphology and optimizing material performance. Common methods include thermal analysis, X-ray diffraction, and density measurements, each providing complementary insights into crystalline fraction and perfection. Differential scanning calorimetry (DSC) is a widely used thermal technique to determine crystallinity by measuring the enthalpy of fusion during melting. In DSC, the sample is heated at a controlled rate, and the heat flow associated with the endothermic melting peak is integrated to yield the enthalpy of fusion, \Delta H_f = \int \Delta C_p \, dT, where \Delta C_p is the heat capacity difference. The degree of crystallinity X_c is then calculated as X_c = \frac{\Delta H_f}{\Delta H_f^\circ} \times 100\%, with \Delta H_f^\circ being the enthalpy for a 100% crystalline reference; for , this value is 293 J/g. This method assumes complete melting of crystals and negligible contributions from other transitions, though baseline construction and sample history can influence accuracy. X-ray diffraction (XRD) provides structural information on crystallite size and orientation through analysis of diffraction patterns from crystalline domains. In wide-angle XRD, the degree of crystallinity is estimated by peak deconvolution, separating crystalline reflections from amorphous scattering, often using methods like the Hermans-Weidinger approach. Crystallite size D is derived from peak broadening via the Scherrer equation, D = \frac{K \lambda}{\beta \cos \theta}, where K is a shape factor (typically 0.9), \lambda is the X-ray wavelength, \beta is the full width at half maximum, and \theta is the Bragg angle. This technique is particularly valuable for non-destructive assessment of lamellar thickness and perfection in polymers like polyamides. Density measurements offer a bulk method to infer crystallinity based on the additive volume contributions of crystalline and amorphous phases. Using a density gradient column, samples are floated in a liquid column with a controlled density gradient (e.g., via miscible solvents like carbon tetrachloride and n-heptane), and their equilibrium position yields the density \rho. The crystallinity is computed as X_c = \frac{\rho - \rho_a}{\rho_c - \rho_a} \times 100\%, where \rho_a and \rho_c are the densities of fully amorphous and crystalline phases, respectively; for PE, typical values are \rho_a = 0.855 g/cm³ and \rho_c = 1.000 g/cm³. This approach is precise for homogeneous samples but requires accurate reference densities and assumes ideal mixing. Several factors influence crystallinity and lamellar organization in semi-crystalline polymers. Annealing, a thermal treatment below the melting point, promotes crystal perfection and reorganization, increasing X_c by allowing chain diffusion and secondary crystallization. Spherulite growth, the radial assembly of lamellae from a central nucleus during cooling, dictates supermolecular morphology; growth rates depend on undercooling and follow kinetic models, with branching at lamellar tips leading to spherical domains up to hundreds of micrometers in size, impacting mechanical isotropy. The melting temperature observed in thermal analysis is linked to crystallinity levels, as detailed in the section on transition temperatures.

Microphase Separation and Domains

Microphase separation in block copolymers and polymer blends arises from the incompatibility between distinct segments, leading to the formation of nanoscale domains such as lamellae, cylinders, or spheres that dictate material properties like mechanical strength and transport behavior. These domains typically span 10-100 nm and exhibit long-range order influenced by factors including segment volume fraction, molecular weight, and interaction parameter χ. Characterization focuses on quantifying domain morphology, spacing, and interfacial characteristics to understand self-assembly mechanisms and enable applications in nanotechnology, such as templates for nanopatterning. Small-angle X-ray scattering (SAXS) is a primary technique for probing microphase-separated structures in bulk samples, providing statistical information on domain size, spacing, and order. In the low-q regime, Guinier analysis approximates the scattering intensity as I(q) ≈ I(0) exp(-q²Rg²/3), where Rg is the radius of gyration reflecting domain size, applicable for weakly segregated or disordered systems. At higher q, Porod's law describes the decay as I(q) ∝ q⁻⁴, enabling estimation of interfacial area per unit volume through the Porod constant, which correlates with domain sharpness and segregation strength in block copolymers like polystyrene-block-polybutadiene. Primary scattering peaks at q* yield the characteristic domain spacing d = 2π / q*, with higher-order peaks indicating lamellar or cylindrical morphologies; for instance, in symmetric diblock copolymers, well-ordered lamellae show peaks at integer multiples of q*. Atomic force microscopy (AFM) complements SAXS by offering surface-specific visualization of domain morphology with nanometer resolution, particularly through phase imaging in tapping mode, which exploits contrasts in viscoelastic properties between domains. In block copolymers such as polystyrene-block-poly(ethylene oxide), phase images reveal alternating lamellae or cylindrical features, where softer domains appear brighter due to higher energy dissipation, allowing mapping of morphologies like gyroids or hexagons without sample preparation beyond spin-coating. Height and modulus contrasts further distinguish domains, enabling quantitative analysis of domain width and orientation, though care is needed to account for tip convolution effects in sub-20 nm features. Transmission electron microscopy (TEM) provides high-resolution cross-sectional views of internal domain structures, essential for confirming three-dimensional arrangements in block copolymer thin films or blends. Selective staining with agents like osmium tetroxide or ruthenium tetroxide enhances electron density contrast between domains, such as staining polybutadiene segments in polystyrene-block-polybutadiene to visualize sharp interfaces. Domain spacing is directly measured from micrographs or Fourier transforms, aligning with SAXS-derived d = 2π / q* values; for example, in poly(styrene-block-isoprene), staining reveals cylindrical domains with spacings of 20-50 nm, revealing defects like dislocations that SAXS averages over. The order-disorder transition (ODT) marks the boundary between microphase-separated and disordered states, characterized by a critical value of the product χN, where χ is the Flory-Huggins interaction parameter and N is the total degree of polymerization. Leibler's mean-field theory predicts (χN)_ODT ≈ 10.5 for symmetric , above which microphase separation drives domain formation, as verified experimentally in polystyrene-block-polyisoprene melts where ODT temperatures correspond to this threshold. Techniques like SAXS track the ODT through the emergence of structure factor peaks and reduced scattering at q=0, while rheological measurements detect viscosity drops, providing insights into segregation strength and phase stability in blends. In semicrystalline block copolymers, microphase domains can template crystallinity, influencing overall morphology.

Thermal Properties Characterization

Transition Temperatures

Transition temperatures in polymers, particularly the glass transition temperature (Tg) and melting temperature (Tm), are critical thermal properties that define the material's mechanical behavior and phase changes. The glass transition temperature marks the reversible transition from a hard, glassy state to a soft, rubbery state in amorphous regions, associated with a change in heat capacity (ΔCp), while the melting temperature indicates the endothermic disruption of crystalline order in semicrystalline polymers. These temperatures are measured using techniques like differential scanning calorimetry (DSC) and dynamic mechanical analysis (DMA), providing insights into polymer structure and enabling predictions of performance under processing and service conditions. Differential scanning calorimetry (DSC) is a primary method for determining both Tg and Tm. In DSC thermograms, Tg appears as a step change in the baseline due to the ΔCp increase, with the midpoint defined as the temperature at the inflection point or half-height of this step, offering a reproducible measure of the transition. For semicrystalline polymers, Tm is identified as the peak temperature of the endothermic melting transition, reflecting the energy required to disrupt crystalline domains; the peak onset and shape can indicate crystal perfection. Heating rate significantly influences these measurements, as faster rates (e.g., 20°C/min versus 10°C/min) shift Tg to higher values due to reduced relaxation time, typically by 1-2°C per °C/min increase, while Tm shows lesser sensitivity but broader peaks at higher rates. Crystallinity plays a key role in Tm, with higher degrees elevating the temperature as more stable crystals require greater thermal input for melting. Dynamic mechanical analysis (DMA) complements DSC by probing Tg through mechanical response under oscillatory stress. Tg is determined from the peak in the loss tangent (tan δ), which corresponds to maximum energy dissipation during segmental motion, often occurring 5-10°C above the DSC midpoint due to the dynamic nature of the test. Unlike DSC, DMA reveals frequency dependence of Tg, where increasing frequency (e.g., from 0.1 to 10 Hz) raises Tg by 2-5°C per decade, as described briefly by the , which models viscoelastic relaxation via free volume changes near Tg. Several factors influence , including composition and molecular weight. For copolymers, the Gordon-Taylor modification of the Fox equation predicts as T_g = \frac{w_1 T_{g1} + k w_2 T_{g2}}{w_1 + k w_2}, where w_i are weight fractions and k is a fitting constant reflecting component interactions, assuming miscibility; for polystyrene-poly(methyl methacrylate) copolymers, k ≈ 0.18 fits experimental data well. Molecular weight dependence follows an approximate logarithmic relationship, ∞ - A / Mn, but often linearized as increasing with log Mw for Mw > 10^4 g/mol, with chain ends reducing by diluting segmental packing in lower Mw polymers. These transition temperatures guide polymer applications by predicting processability and end-use properties. For instance, polymers with Tg > room temperature, such as (Tg ≈ 100°C), maintain rigidity for structural uses, while those below, like polyisobutylene (Tg ≈ -70°C), enable flexible elastomers; exceeding Tg during processing risks deformation, whereas Tm sets limits for melt in semicrystalline resins like . Accurate Tg and Tm characterization thus ensures optimal formulation for injection molding or , balancing flow and stability.

Thermal Stability and Degradation

Thermal stability and degradation characterization in polymers focuses on evaluating the temperature at which irreversible occurs, providing critical insights into material longevity under heat exposure. Techniques in this domain quantify mass loss, identify decomposition products, and elucidate underlying reaction kinetics, essential for applications in processing, recycling, and high-temperature environments. Thermogravimetric analysis (TGA) is a primary method for assessing polymer thermal stability, measuring continuous mass changes as a function of under controlled heating rates and atmospheres. curves from reveal decomposition stages, with the decomposition (Td) commonly defined as the point of 5% mass loss, serving as a for onset of degradation. For instance, Td values help compare stability across types, such as polyolefins exhibiting Td above 400°C in inert conditions. To determine activation energy (Ea) for degradation kinetics, the Ozawa-Flynn-Wall (OFW) method analyzes TGA data from multiple heating rates (β), plotting log β versus 1/T at fixed conversion levels to yield Ea from the slope. The method employs the isoconversional equation: \log \beta = \text{const} - 0.456 \frac{E_a}{RT} where R is the gas constant and T is the absolute temperature at a given conversion α. This model-free approach avoids assuming reaction mechanisms, making it widely adopted for polymers like polyimides, where Ea values typically range from 150-250 kJ/mol. Evolved gas analysis (EGA), often coupled with TGA to mass spectrometry (MS) or Fourier-transform infrared spectroscopy (FTIR), identifies volatile products released during decomposition, enabling mechanistic insights. For example, TGA-FTIR detects CO₂ evolution from polyesters via ester bond cleavage, while TGA-MS provides sensitive quantification of fragments like monomers or oligomers in polyamides. These hyphenated techniques distinguish decomposition pathways by correlating gas profiles with mass loss events. Polymer degradation mechanisms under heat include random scission, where bonds break statistically along the , reducing molecular weight without specific end-group involvement, as seen in . Unzipping , conversely, proceeds sequentially from chain ends, predominant in polyesters like , yielding cyclic monomers. Oxidative degradation in air accelerates these processes via formation, leading to carbonyl byproducts and lower Td compared to pyrolytic conditions in inert atmospheres, where scission dominates without oxygen. Recent studies emphasize thermal stability of bio-based polymers, such as (), which exhibits Td around 300°C at 5% mass loss in , attributed to random scission and forming oligomers. Enhancements via nanofillers like can raise Td,max to 396°C, improving suitability for sustainable applications while highlighting the need for tailored degradation control.

Mechanical Properties Characterization

Static Mechanical Testing

Static mechanical testing evaluates the equilibrium mechanical properties of polymers under quasi-static loading conditions, providing essential data on load-bearing capacity, , and deformation behavior for applications in structural components and . These tests are conducted at low strain rates to minimize viscoelastic effects, focusing on metrics such as strength, , and that inform and processing optimization. Common standards ensure , with specimens typically molded or machined to precise geometries to isolate intrinsic material responses from artifacts like defects or . Tensile testing, the cornerstone of static mechanical characterization, measures the response of polymers to uniaxial pulling forces using dogbone-shaped specimens as specified in ASTM D638. The applied force F generates stress \sigma = \frac{F}{A}, where A is the initial cross-sectional area, while the resulting \Delta L yields engineering \varepsilon = \frac{\Delta L}{L_0}, with L_0 as the gauge length. E, representing initial stiffness, is calculated as E = \frac{\sigma}{\varepsilon} in the linear elastic region, typically up to 0.1-1% for thermoplastics. Key outcomes include yield strength (onset of plastic deformation for ductile polymers like ), ultimate (peak stress before fracture), and at break (total to failure, indicating ; e.g., >200% for elastomers versus <10% for brittle polystyrene). These properties distinguish ductile necking behavior in semicrystalline polymers from brittle failure in glassy ones, guiding design for impact-resistant parts. Compression and flexural tests complement tensile data by assessing behavior under opposing or bending loads, revealing differences in ductile versus brittle responses across loading modes. In compression testing per , cylindrical or cubic specimens are compressed between platens to determine compressive strength (stress at failure) and modulus, often higher than tensile values due to suppressed buckling in confined geometries; for example, rigid plastics like exhibit compressive yields around 80-100 MPa. Flexural testing via employs three-point or four-point bending on rectangular bars to evaluate flexural modulus and strength, where the maximum stress \sigma_f = \frac{3FL}{2bd^2} (for three-point loading, with L, b, and d as span, width, and thickness) highlights shear effects in thicker samples. These tests are critical for polymers in load-bearing roles, such as automotive panels, where flexural strength might reach 50-150 MPa for reinforced composites. Hardness testing quantifies surface resistance to indentation, correlating with overall rigidity for quality control in molded parts. For elastomers and soft thermoplastics, the Shore durometer (ASTM D2240) uses a spring-loaded indenter; Shore A scale applies to flexible materials like rubber (values 20-90), while Shore D suits semirigid plastics like polyurethane (50-100), with higher numbers indicating greater hardness via minimal penetration depth. Rigid engineering plastics, such as nylons or acetals, are assessed by Rockwell hardness (ASTM D785), employing ball or conical indenters under major and minor loads; scales like R (Rockwell R, 0-150) or M (for softer plastics) yield values where, for instance, high-density polyethylene rates around 70, reflecting wear resistance. These non-destructive metrics enable rapid screening without full specimen failure. Polymer microstructure profoundly influences static mechanical properties, linking molecular architecture to macroscopic performance. Increased crystallinity in semicrystalline polymers, such as , elevates Young's modulus by stiffening the material through ordered chain packing, with E rising from ~0.2 GPa at low crystallinity levels to ~1 GPa at high crystallinity (e.g., 80%), as crystalline lamellae restrict amorphous chain mobility. Conversely, molecular weight (MW) enhances toughness by promoting chain entanglements that dissipate energy during deformation; for , toughness (measured as area under stress-strain curve) increases with MW above the entanglement threshold (~3.5 \times 10^4 g/mol), though elongation at break remains low (typically 3-5%), the increased entanglements enhance overall toughness without significantly altering modulus. These relationships underscore the need for tailored synthesis and processing to balance stiffness and ductility in applications like fibers or films.

Dynamic and Rheological Properties

Dynamic mechanical analysis (DMA) is a technique used to characterize the viscoelastic properties of polymers by applying a sinusoidal stress or strain and measuring the resulting response as a function of temperature, frequency, or time. In DMA, the storage modulus E' represents the elastic component, quantifying the energy stored and recovered during deformation, while the loss modulus E'' indicates the viscous component, reflecting energy dissipation as heat. The ratio \tan \delta = E'' / E' provides insight into the balance between elastic and viscous behavior, with peaks in \tan \delta often marking the glass transition temperature T_g, where the polymer shifts from a glassy to a rubbery state. This method is particularly valuable for identifying secondary transitions and assessing molecular mobility in both amorphous and semicrystalline polymers. Rheometry complements DMA by evaluating the flow and deformation behavior of polymer melts and solutions under shear, especially through oscillatory shear measurements that probe linear viscoelasticity without causing structural breakdown. In oscillatory rheometry, the storage modulus G' measures the elastic recovery, akin to a spring, while the loss modulus G'' captures viscous flow, similar to a dashpot; crossover points where G' = G'' indicate transitions between viscous-dominated and elastic-dominated regimes. The complex |\eta^*(\omega)| from these measurements relates to steady-state shear \eta(\dot{\gamma}) via the Cox-Merz rule, which empirically states that |\eta^*(\omega)| = \eta(\dot{\gamma}) when \omega = \dot{\gamma}, enabling broader profiling from oscillatory data. This superposition holds well for many entangled polymer melts, facilitating predictions of processing behavior. The melt flow index (MFI), determined per , assesses polymer processability by measuring the mass of molten material extruded through a die under specified temperature and load conditions over 10 minutes, providing a simple indicator of melt viscosity. Higher MFI values correspond to lower viscosities, with viscosity inversely proportional to MFI (\eta \propto 1/MFI), allowing quick comparisons for injection molding or extrusion suitability, though it approximates low-shear behavior and should be supplemented with full rheological data. In entangled polymer systems, the reptation model describes chain dynamics where molecules are confined to tube-like regions formed by neighboring chains, leading to curvilinear motion for relaxation. According to this model, the longest relaxation time \tau scales with molecular weight M_w as \tau \propto M_w^{3.4}, reflecting the time for a chain to reptate out of its tube, which governs viscoelastic response in melts above the entanglement molecular weight. This scaling, derived from the Doi-Edwards theory, explains the dramatic increase in viscosity and relaxation times with chain length in processing-relevant regimes.

Spectroscopic Techniques

Nuclear Magnetic Resonance Spectroscopy

Nuclear Magnetic Resonance (NMR) spectroscopy provides atomic-level insights into the structure, composition, and dynamics of polymers by exploiting the magnetic properties of nuclei such as ¹H and ¹³C. In polymer characterization, it reveals microstructural details like monomer sequencing, tacticity, branching, and end-group functionalities, which influence macroscopic properties such as mechanical strength and thermal behavior. Unlike scattering techniques, NMR offers chemical specificity through resonance frequencies that depend on local electronic environments, making it indispensable for both academic research and industrial quality control. Solution-state NMR requires polymers to be dissolved in deuterated solvents to minimize solvent signal interference and enable high-resolution spectra. ¹H NMR spectra display peaks whose chemical shifts and integrals quantify end-group concentrations, such as hydroxyl or vinyl termini in polyesters, by comparing their intensities to backbone signals. For tacticity, ¹³C NMR is preferred due to its greater spectral dispersion; triad probabilities (e.g., mm, mr, rr in polypropylene) are calculated from the relative areas of the relevant carbon peaks, such as methyl carbons, providing statistical models of stereoregularity. Diffusion-ordered spectroscopy (DOSY), a pulsed-field gradient variant, separates signals by diffusion coefficients, allowing non-standard-based estimation of molecular weight (M_w) from hydrodynamic radii via the Stokes-Einstein relation, with applications in polydisperse systems like polyethylene glycols up to 10 kDa. Advanced multidimensional NMR enhances resolution for complex polymers. COSY correlates protons through J-couplings to map intra-monomer connectivities, aiding sequence assignment in copolymers like . HSQC extends this to heteronuclear correlations, overlaying ¹H and ¹³C dimensions for unambiguous carbon-proton pairing, which is critical for quantifying branch content in low-density polyethylene via methyl side-chain signals. Quantitative ¹³C NMR, with inverse-gated decoupling and long relaxation delays, achieves accuracy within 1-2% for branch frequencies, as demonstrated in where short branches (e.g., 5-10 mol%) are resolved. These techniques have been pivotal in characterizing sequence-defined polymers, revealing dyad and triad distributions that dictate phase behavior. Solid-state NMR circumvents solubility issues for intractable polymers by employing magic-angle spinning (MAS) at 54.74° to the magnetic field, averaging dipolar and chemical shift anisotropies for spectra approaching solution-like resolution. In semicrystalline materials like , ¹³C CP-MAS spectra differentiate crystalline (sharp peaks at ~25 ppm for CH₂) and amorphous domains (broader at ~24 ppm) via linewidth analysis, quantifying crystallinity up to 60%. Cross-polarization (CP) transfers magnetization from abundant ¹H to rare ¹³C, boosting sensitivity by a factor of up to ~4 while relaxation measurements (e.g., T_{1ρ}) probe dynamics, such as segmental motions in rubbery phases with correlation times of 10^{-5} to 10^{-8} s. This approach has elucidated microphase separation in block copolymers, where domain-specific shifts indicate confinement effects. Limitations of NMR include the need for soluble, often deuterated, samples in solution mode, restricting analysis of crosslinked or ultrahigh-molecular-weight polymers (>10^6 Da) due to viscosity broadening and long acquisition times (hours to days). Solid-state spectra, while versatile, exhibit reduced sensitivity for dilute nuclei, necessitating larger samples (50-100 mg). Recent high-field instruments (>600 MHz, e.g., 800-1000 MHz) in the have mitigated these by enhancing signal-to-noise ratios by 2-4 times through increased Boltzmann polarization, enabling faster experiments and finer resolution of overlapping peaks in heterogeneous plastics like recycled polyolefins.

Vibrational Spectroscopy

Vibrational spectroscopy, encompassing infrared (IR) and Raman techniques, provides non-destructive methods to probe molecular vibrations in polymers, enabling identification of functional groups, chemical bonding, and chain orientation. These methods rely on the interaction of light with molecular vibrations: IR absorption occurs when dipole moments change, while Raman scattering arises from polarizability variations. In polymer characterization, Fourier-transform IR (FTIR) spectroscopy is widely used for its sensitivity to functional groups, with absorption bands such as the C-H stretching vibrations appearing between 2800 and 3000 cm⁻¹ in aliphatic polymers like polyethylene. Similarly, Raman spectroscopy complements IR by highlighting symmetric vibrations, such as the C=C stretching mode around 1600 cm⁻¹ in conjugated polymers like polystyrene. Both techniques are essential for analyzing bulk and surface properties without sample preparation, offering rapid insights into polymer composition and structure. FTIR excels in detecting polar bonds and is particularly valuable for solid polymers through (ATR) mode, which allows analysis of thick or irregular samples by evanescent wave penetration of 0.5–5 μm into the material. In ATR-FTIR, characteristic bands like the CF₂ stretching at 1284 cm⁻¹ help identify crystalline phases in fluoropolymers such as PVDF. For studies, FTIR dichroism measures differences in absorption between and polarized , quantifying via dichroic ratios (e.g., R = A∥/A⊥) in stretched polymers; this is especially useful in ATR configuration for surface-sensitive analysis of rubbers and composites. modulation enhances precision, detecting subtle effects in copolymers and blends. Raman spectroscopy provides complementary information on non-polar bonds and is advantageous for aqueous or fluorescent samples, with scattering shifts revealing molecular . For instance, the G-band at ~1580 cm⁻¹ indicates sp² carbon bonding in polymer nanocomposites. Confocal Raman achieves spatial resolution down to 200 nm for 2D/3D mapping of domain structures in phase-separated polymers, using single-acquisition techniques for efficient profiling. Polarization-resolved Raman exploits vibrational (e.g., Ag vs. B modes) to determine , with intensity variations as a function of angle enabling <1° accuracy in crystalline polymers like fibers. Key applications include monitoring , such as oxidation, where FTIR tracks the growth of O-H stretching bands (3700–3000 cm⁻¹) and H-O-H bending (~1640 cm⁻¹) in exposed elastomers like rubbers. In copolymers, band intensity ratios—e.g., C=O at 1720 cm⁻¹ relative to CH₂ at 1453 cm⁻¹—quantify compositions, as demonstrated in PVDF-based systems with varying side-chain ratios. Hyphenated FTIR-microscopy extends this to , combining imaging with for micrometer-scale analysis of inhomogeneous blends and composites, revealing local distributions without sectioning.
TechniqueExample BandAssignmentPolymer Context
FTIR2800–3000 cm⁻¹C-H stretchAliphatic chains in polyethylene
FTIR1720 cm⁻¹C=O stretchCarbonyl groups in oxidized or copolymerized chains
Raman~1600 cm⁻¹C=C stretchAromatic or unsaturated bonds in polystyrene
Raman~1580 cm⁻¹G-band (sp² C)Graphitic domains in nanocomposites

Scattering and Microscopic Techniques

X-ray and Neutron Scattering

X-ray and neutron scattering techniques are essential for probing the nanoscale structure of polymers, providing bulk-averaged information on crystallinity, domain sizes, and chain conformations that complement other characterization methods. These non-destructive methods exploit the elastic scattering of s or neutrons by or scattering length density variations, respectively, to reveal structural features over length scales from angstroms to hundreds of nanometers. In polymer characterization, wide-angle scattering captures atomic-scale order, while elucidates mesoscale organization such as lamellar stacks or phase-separated domains. Wide-angle X-ray scattering (WAXS) is primarily used to determine the in semi-crystalline polymers, identifying phases, degree of crystallinity, and parameters. By analyzing peaks according to , n\lambda = 2d \sin\theta, where n is an , \lambda is the , d is the interplanar spacing, and \theta is the scattering angle, WAXS enables the calculation of lattice dimensions from peak positions. For instance, in polyolefins like isotactic , WAXS distinguishes α- and β-crystalline forms and quantifies their relative contents based on peak intensities. This technique has been instrumental in studying shear-induced , where preferred orientations alter parameters during processing. Small-angle X-ray scattering (SAXS) and small-angle neutron scattering (SANS) extend analysis to larger scales, measuring domain spacings and morphological features through the scattering vector q = \frac{4\pi \sin\theta}{\lambda}, which relates to real-space distances via d = \frac{2\pi}{q}. SAXS is particularly effective for electron density contrasts in polymer blends or block copolymers, revealing periodic structures like lamellar domains with spacings of 10–100 nm. SANS complements this by offering tunable contrast through deuteration, where replacing hydrogen with deuterium alters the scattering length density, enabling contrast matching to isolate specific components such as polymer chains in a matrix. For example, in deuterated polyethylene blends, SANS suppresses scattering from the hydrogenated phase, highlighting chain conformations. Applications include determining lamellar thickness L via one-dimensional correlation function analysis of SAXS profiles, which deconvolves the electron density autocorrelation to yield long periods and crystal thicknesses in semi-crystalline polymers like polyethylene. In polymer melts, both techniques measure the radius of gyration R_g, a key indicator of chain dimensions, using the Guinier approximation I(q) \approx I(0) \exp(-q^2 R_g^2 / 3) at low q, with values around 5–10 nm for typical molecular weights. Advances in sources have enabled studies of dynamic processes, such as during processing, with high post-2015. For instance, simultaneous SAXS/WAXS at facilities tracks the evolution of oriented β-crystals in isotactic under thermal treatments from 155–170°C, revealing temperature-dependent transitions to α-phases and enhanced heat resistance in oriented forms. More recent developments as of 2025 include resonant soft (RSoXS), which provides chemically and orientationally resolved nano-to-mesoscale of morphologies, such as in organic photovoltaics and self-assembled block copolymers, and accelerated SANS algorithms that optimize data collection for probing dynamics and hierarchical structures. These capabilities provide quantitative insights into structure-property relationships, such as interdiffusion at weld zones in 3D-printed parts, informing mechanical performance.

Electron and Atomic Force Microscopy

Scanning electron microscopy (SEM) is a widely used imaging technique in polymer characterization that employs a focused beam of electrons to scan the surface of a sample, generating images based on interactions such as for topographic details and backscattered electrons for compositional contrasts. This method excels in visualizing surface features like fractures and in polymers, providing insights into failure mechanisms and material heterogeneity. Typical resolution in SEM for polymers achieves approximately 1 , enabling nanoscale surface analysis. Sample preparation often involves fracturing the polymer in to expose internal structures, followed by coating with a conductive layer such as via to prevent charging under . Transmission electron microscopy (TEM) complements SEM by transmitting electrons through ultrathin polymer sections to reveal internal structures at higher resolution, typically below 1 nm and up to atomic scales in advanced setups. In TEM, electrons interact with the sample to form contrast based on density and thickness differences, allowing detailed examination of nanoscale features such as crystalline lamellae and domain boundaries. Sample preparation requires sectioning polymers into ultrathin slices (e.g., 25-100 nm) using ultramicrotomy, often with staining agents like osmium tetroxide or ruthenium tetroxide to enhance contrast. For soft matter polymers, cryo-TEM preserves hydrated structures by rapid freezing in vitreous ice, minimizing artifacts and beam damage while enabling observation of dynamic morphologies in their native state. Recent advances as of 2025 include cryo-EM applications for visualizing cyclic macromolecular chains, revealing ring expansion mechanisms in synthetic polymers. Atomic force microscopy (AFM) provides nanoscale surface probing without vacuum requirements, using a sharp tip on a to raster-scan the polymer surface and measure forces for and property mapping. In tapping mode, the oscillating intermittently contacts the sample, allowing phase imaging that differentiates material phases based on viscoelastic responses, ideal for identifying nanoscale domains in blends. Force-distance curves from AFM enable modulus mapping by analyzing tip-sample interactions, quantifying local mechanical properties like and elasticity at the nanoscale. This technique is particularly valuable for observing textures in semicrystalline polymers, revealing growth patterns and defects with sub-nanometer lateral resolution. Recent developments as of 2025, such as photothermal AFM-IR, combine AFM with for sub-5 nm chemical mapping in blends and nanoparticles, enabling quantification of composition and supporting studies in sustainable materials and . These methods are often integrated in polymer studies; for instance, and TEM images can validate particle sizes obtained from techniques, ensuring correlation between surface and bulk features. Applications include visualizing spherulites in to understand and domain structures in polystyrene-polybutadiene blends for analysis.

References

  1. [1]
    Polymer characterization techniques – an introduction
    May 28, 2024 · Polymer characterization includes many chemical analysis techniques that examine the chemical properties of the polymer, such as its elemental, ...
  2. [2]
    Polymer Characterization Technique Overview - Measurlabs
    Aug 15, 2024 · Different polymer characterization techniques (NMR, SEC, FTIR, DSC, etc.) provide different types of information on chemical, molecular, ...
  3. [3]
    None
    Summary of each segment:
  4. [4]
  5. [5]
    Hermann Staudinger Foundation of Polymer Science - Landmark
    For his work in the field of polymers, Staudinger was awarded the Nobel Prize for chemistry in 1953. Acknowledgments. Adapted for the internet from “The ...
  6. [6]
    Hermann Staudinger – Facts - NobelPrize.org
    He showed how small molecules can join to form long chains and so become very large molecules—polymers. The result was the basis for the development of ...
  7. [7]
    Dispersity in polymer science (IUPAC Recommendations 2009)
    This recommendation defines just three terms, viz., (1) molar-mass dispersity, relative-molecular-mass dispersity, or molecular-weight dispersity.
  8. [8]
    (PDF) Polymer characterization (II) - ResearchGate
    Jul 20, 2016 · The polymer characterization technique categories are: chemical, electrical, mechanical, molecular, physical, rheological, spectroscopic, ...<|separator|>
  9. [9]
    Understanding and Modeling Polymers: The Challenge of Multiple ...
    Nov 14, 2022 · The study of inhomogeneous polymer solutions is particular challenging due to the vastly different mobilities of polymer and solvent molecules.
  10. [10]
    Polymer Science and Engineering: The Shifting Research Frontiers
    This chapter discusses some of the direct societal benefits that derive from the field of polymer science and engineering and illustrates how this ...
  11. [11]
    Structure-Property Relations in Polymers - ACS Publications
    ... structure-property relationship for the characterization of the ... Polymer Characterization Using Singlet Oxygen Phosphorescence as a Spectroscopic Probe.
  12. [12]
    3. Manufacturing: Materials and Processing | Polymer Science and ...
    A process known as "gel spinning" has been commercialized, which produces fibers of ultrahigh-molecular-weight polyethylene. The less crystalline LDPE and ...
  13. [13]
    Materials and Chemical Characterization and Medical Devices - FDA
    Jan 30, 2023 · The Materials and Chemical Characterization Program develops regulatory science tools for evaluating medical device material biostability ...
  14. [14]
    Longevity Expectations for Polymers in Medical Devices Demand ...
    Nov 23, 2020 · According to a recent Food and Drug Administration (FDA) report, materials were identified as the primary root cause of device failure for ∼40% ...
  15. [15]
    Sustainable Polymer Composites and Nanocomposites - SpringerLink
    The book presents the manufacturing methods, properties and characterization techniques of these eco-friendly composites. The respective chapters address ...
  16. [16]
    Biorenewable and circular polyolefin thermoplastic elastomers
    Oct 1, 2024 · Chemical recycling to monomer for an ideal, circular polymer economy ... W.S. contributed the methodology for monomer and polymer characterization ...Missing: post- | Show results with:post-
  17. [17]
    Molecular Size Distribution in Linear Condensation Polymers 1
    Polymer molecular weight distributions: a tutorial on transformations between number density, mass density, and linear/logarithmic axes.
  18. [18]
    Principles of Polymer Chemistry by Paul J. Flory Jr. | Hardcover
    by Paul J. Flory Jr. Guest Lecturer in: Format. Hardcover$113.95. Add to cart. Open Access. Open Access. This work can be downloaded for non-commercial ...Missing: PDF | Show results with:PDF
  19. [19]
    Confirmation of the Doi-Edwards model | Macromolecules
    Determining the dilution exponent for entangled 1,4-polybutadienes using blends of near-monodisperse star with unentangled, low molecular weight linear ...
  20. [20]
    [PDF] Polymers: Molecular Weight and its Distribution
    Jan 15, 2001 · The theory of step-growth (condensation) polymerization for flexible chain polymers led to the Flory-Schulz distribution. In the theory, p ...
  21. [21]
    Osmotic Pressure
    Osmometry provides absolute measurements of molecular weight. It relies on Flory-Huggins lattice theory for a connection between the chemical potential of a ...
  22. [22]
    Overview of Methods for the Direct Molar Mass Determination ... - NIH
    First, the osmotic pressure is inversely proportional to molecular weight, so molecules with a high molecular weight contribute very little. Therefore, the ...Missing: seminal | Show results with:seminal
  23. [23]
    Characterization of polymers by static light scattering - ScienceDirect
    The static light scattering (SLS) method is introduced with key equations used to extract quantitative information such as weight-average molecular weight.
  24. [24]
    SEDFIT-MSTAR: Molecular weight and molecular weight distribution ...
    Sedimentation equilibrium (SE) in the analytical ultracentrifuge is a well established method for obtaining the molecular weights of polymers. It has an ...
  25. [25]
    Determination of Absolute Mass Values in MALDI-TOF of Polymeric ...
    Determination of Absolute Mass Values in MALDI-TOF of Polymeric Materials by a Method of Self-Calibration of the Spectra. Click to copy article linkArticle ...
  26. [26]
    Gel Permeation Chromatography - an overview | ScienceDirect Topics
    Gel permeation chromatography (GPC) is defined as a technique used for the characterization of molecular weight and molecular weight distribution of ...
  27. [27]
    Gel permeation chromatography. I. A new method for molecular ...
    ... Gel permeation chromatography. I. A new method for molecular weight distribution of high polymers. J. C. Moore,. J. C. Moore. Texas Basic Research Department ...
  28. [28]
    The Mark–Houwink–Sakurada Equation for the Viscosity of Atactic ...
    Oct 1, 1985 · The viscosity–molecular weight (Mark–Houwink–Sakurada) relationships have been critically evaluated for atactic polystyrene for a variety of solvents.Missing: seminal | Show results with:seminal
  29. [29]
    [PDF] The Mark–Houwink–Sakurada Equation for the Viscosity of Linear ...
    Oct 15, 2009 · The Mark-Houwink-Sakurada equation relates viscosity to molecular weight, using constants K and a, specific to polymer, solvent, and ...Missing: seminal | Show results with:seminal
  30. [30]
    Intrinsic Viscosity - an overview | ScienceDirect Topics
    Molecular weight is related to the viscosity of polymer solution. The molecular weights of polymers have been correlated to the viscosity at specific ...
  31. [31]
    [PDF] Hermann Staudinger - Nobel Lecture
    The viscosity number of these compounds is proportional to the degree of polymerization as shown in Table 5. The end-group molecular weight of higher polymer ...
  32. [32]
    End-Group Analysis and Number-Average Molecular Weight ...
    End-Group Analysis and Number-Average Molecular Weight Determination of Some Polyalkylene Glycols and Glycol Polyesters Using Nuclear Magnetic Resonance ...
  33. [33]
    Gel Permeation Chromatography - an overview | ScienceDirect Topics
    GPC, or gel permeation chromatography is unique to polymers. It is the best method for characterizing the complete molecular weight distribution of a polymer.
  34. [34]
    [PDF] Characterization of Polymer Properties and Identification of ... - NREL
    Here we characterized 59 polymers from common commercial vendors across 20 different polymer classes, representing >95% of global plastic production by mass.
  35. [35]
    [PDF] Elemental Analysis: CHNS/O Characterization of Polymers and ...
    For CHNS or CHN determination the FlashSmart. Elemental Analyzer operates with dynamic flash combustion of the sample. Samples are weighed in tin containers and ...
  36. [36]
    [PDF] investigation of the mechanical and thermal properties of
    FTIR and elemental analysis showed that the copolymers have three ... CHN analysis is performed via combustion at 990 °C with the elemental analyzer.
  37. [37]
    The Infrared Spectra of Polymers, Part I: Introduction
    Jul 1, 2021 · The spectra of functional groups in polymers are similar to those in small molecules. The spectra of low-density and high-density polyethylene ...
  38. [38]
    (PDF) Quantification of copolymer composition (methyl acrylate and ...
    Nov 19, 2015 · Conditions for simple and sensitive FTIR quantitative determination of acrylonitrile, methyl acrylate, and itaconic acid in their copolymer ...
  39. [39]
    Determination of Ethylene-Propylene Copolymer Composition by ...
    Quantitative Analysis of Monomer Composition in Ethylene-Propylene Block Copolymers by FT-IR Spectroscopy. Applied Spectroscopy 1987, 41 (2) , 319-320 ...
  40. [40]
    Quantitative 13C NMR analysis of sequence distributions in poly ...
    Different 13C NMR methods of determining triad distributions in two poly(ethylene-co-1-hexene) copolymers are examined using high signal-to-noise ratio 13C ...Missing: ¹H ¹³C polymer monomer peaks dyad
  41. [41]
    [PDF] NMR analysis and triad sequence distributions of poly(3 ...
    In this work, we have re-examined the 1H and 13C NMR spectra of eight samples of PHBV with a large compositional range. We were able to achieve good spectral ...
  42. [42]
    1H NMR and 13C NMR investigations of sequence distribution and ...
    Aug 6, 2025 · 1 H NMR and 13 C NMR spectroscopies can provide detail and quantitative information about nature of linkages and sequence distribution to ...
  43. [43]
    Recent Advances and Applications of NMR Techniques in Plastic ...
    Mar 3, 2025 · To address the inherent sensitivity limitations of NMR, various hyperpolarization techniques were developed, such as chemically induced dynamic ...Missing: post- | Show results with:post-
  44. [44]
    Automated Ultra-Fast 13C NMR Analysis of Polyolefin Materials - PMC
    Jan 21, 2025 · In this paper, we introduce an instrument for automated fast determinations of the 13 C NMR microstructure on polyolefin materials.
  45. [45]
    NMR Analyses and Statistical Modeling of Biobased Polymer ... - MDPI
    This article provides a selective review of the developments in both the NMR analysis of biobased polymers and the statistical models that can be used to ...
  46. [46]
    High-Resolution 13C NMR Configurational Analysis of ...
    In this paper, we present the results of a comparative high-resolution 13C NMR microstructural analysis of an “atactic” and an “isotactic” polypropylene ...
  47. [47]
    Tacticity Changes during Controlled Degradation of Polypropylene
    Sep 21, 2021 · Causes of tacticity defects in both iPP and sPP have been studied extensively, predominantly by 13C NMR. These defects are related to the ...Missing: seminal | Show results with:seminal
  48. [48]
    Two-dimensional NMR studies of polypropylene tacticity
    The 1 H and 13 C NMR spectral assignments of polypropylene tacticity are revised using two-dimensional NMR techniques.
  49. [49]
    The Dimensions of Chain Molecules Containing Branches and Rings
    Zimm, Walter H. Stockmayer; The Dimensions of Chain Molecules Containing Branches and Rings. J. Chem. Phys. 1 December 1949; 17 (12): 1301–1314. https://doi ...
  50. [50]
    Determination of long chain branching in PE samples by GPC-MALS ...
    Branching index are calculated by GPC-MALS and GPC-VIS. Determined structure factor present a clear systematic trend with molecular weight.
  51. [51]
    Statistical mechanics of chain molecules : Flory, Paul J
    Jun 25, 2023 · Statistical mechanics of chain molecules. by: Flory, Paul J. Publication date: 1969. Topics: Molecules, Statistical mechanics. Publisher: New ...
  52. [52]
    Foundations of Rotational Isomeric State Theory and General ...
    Variability in the Persistence Length of an Atactic Polymer Due to Quenched Randomness, As Illustrated by Atactic Polystyrene. Macromolecules 2007, 40 (2) ...
  53. [53]
    Characterizing polymer structure with small-angle neutron scattering
    May 3, 2021 · In this tutorial, we provide an overview of SANS and a guide to interpreting SANS measurements of polymers that is aimed at new and prospective users.
  54. [54]
    Polymer crystallinity determinations by DSC - ScienceDirect.com
    Differential scanning calorimetry (DSC) and quantitative differential thermal analysis (DTA) are widely used to determine the crystallinity of semicrystalline ...<|separator|>
  55. [55]
    [PDF] THERMAL APPLICATIONS NOTE Polymer Heats of Fusion
    Fortunately, the heats of fusion values for 100 % crystalline polymers may be determined by indirect methods such as extrapolation using the Flory equation (2).
  56. [56]
    [PDF] X-Ray Diffraction Methods in Polymer Science
    : The special usefulness of x-ray diffraction in the study of solid substances lies in its ability to distinguish ordered from disordered states. Thus it is ...
  57. [57]
    Meaning and measurement of crystallinity in polymers: A Review
    Methods of determination of the degree of crystallinity using density, infrared, thermal, N.M.R. and X-ray measurements are examined in light of modern notions ...
  58. [58]
  59. [59]
    Polymer spherulites: A critical review - ScienceDirect
    Polymeric and non-polymeric materials often crystallize as spherulites when crystallized from viscous melts or solutions at large undercooling.<|control11|><|separator|>
  60. [60]
    A review of small-angle scattering models for random segmented ...
    Small-angle X-ray scattering data from several experimental and commercial poly(ether urethane) formulations was used to test various scattering models ...
  61. [61]
    Porod scattering study of coarsening in immiscible polymer blends
    The coarsening behavior of the dispersed phase in immiscible blends of PS/PMMA prepared by either melt-extrusion or pulverization was characterized using SAXS ...
  62. [62]
    Small-angle x-ray-scattering study of ordering kinetics in a block ...
    May 7, 1990 · The kinetics of microphase separation and ordering of a diblock copolymer onto a lattice has been studied using time-resolved high ...Missing: SAXS Porod
  63. [63]
    Understanding tapping-mode atomic force microscopy data on the ...
    In this paper, we focus on better understanding tapping-mode atomic force microscopy (AFM) data of soft block copolymer materials with regard to: (1) phase ...
  64. [64]
    Imaging Block Copolymer Crystallization in Real Time with the ...
    In the images shown, the MB- or SEB-rich domains are darkest, followed by the ethylene-rich domains, with the crystalline ethylene areas having the brightest ...
  65. [65]
    High-resolution three-dimensional structural determination of ...
    Jul 26, 2023 · Staining is a conventional protocol for the characterization of the 3D structures of block copolymers at the nanometer level; it helps enhance ...
  66. [66]
    Three‐dimensional structural analysis of a block copolymer by ...
    Feb 2, 2007 · It was confirmed that the lamellar structure had a domain spacing of 0.10 μm, which quantitatively agreed with the value obtained from the TEMT ...
  67. [67]
    Microphase Separation and Crystallization in All-Conjugated Poly(3 ...
    In situ temperature-resolved synchrotron small-angle X-ray scattering (SAXS) ... 30 Years of small-angle x-ray scattering. Guinier, Andre. Physics Today (1969) ...
  68. [68]
    [PDF] Overview of Glass Transition Analysis by Differential Scanning ...
    This note will discuss the available approaches for the analysis of a glass transition by Differential Scanning Calorimetry (DSC).
  69. [69]
    Differential Scanning Calorimetry - Polymer Science Learning Center
    Because of this change in heat capacity that occurs at the glass transition, we can use DSC to measure a polymer's glass transition temperature. You may ...
  70. [70]
    [PDF] exploring the sensitivity of thermal analysis techniques to the glass ...
    Most DSC experiments are performed at 5 or 10ºC/minute. A higher heating rate is beneficial in detecting Tg, however, because the heat flow signal associated ...
  71. [71]
    Melting point, crystallization, and glass transition in polymers - Linseis
    Therefore, the crystallization temperature is always below the thermodynamically controlled melting temperature. Glass transition. Non-crystalline materials ...
  72. [72]
    [PDF] Measurement of Glass Transition Temperatures by Dynamic ...
    This note will detail the use of dynamic mechanical analysis (DMA) and rheology to characterize the glass transition of a polymeric material. GLASS TRANSITIONS.Missing: WLF | Show results with:WLF
  73. [73]
    Glass Transition Temperature-Composition Relationship of ... - Nature
    The Fox equation is based on the assumption that cer- tain properties of a copolymer, e.g., specific volume, mo- lar cohesive energy, or chain stiffness are ...
  74. [74]
    Is the Molecular Weight Dependence of the Glass Transition ...
    The immense dependence of the glass transition temperature Tg on molecular weight M is one of the most fundamentally and practically important features of ...
  75. [75]
    Glass Transition Temperature (Tg) of Polymers - Protolabs
    Glass transition temperature (Tg) affects polymer moldability and characteristics such as tensile strength, modulus elasticity, and transparency.
  76. [76]
    Polymer Glass Transition Temperature - Material Properties, Impact
    Sep 21, 2023 · The Tg provides valuable insights into a polymer's behavior, such as its processing, mechanical, thermal, and overall performance ...
  77. [77]
    Thermogravimetric Analysis - an overview | ScienceDirect Topics
    Thermogravimetric analysis (TGA) is a thermal analysis technique that measures the mass of a material over time as its temperature changes at a constant heat ...
  78. [78]
    Thermogravimetric Analysis Integrated with Mathematical Methods ...
    Aug 11, 2025 · This review emphasized the role of mathematical models and correlations in thermogravimetric analysis to evaluate the thermal stability of ...<|separator|>
  79. [79]
    Flynn, J.H. and Wall, L.A. (1966) A Quick, Direct Method for the ...
    TGA data for the polymers were investigated by nonlinear fitting procedures that yielded activation energies and frequency factors for the combined chemical ...
  80. [80]
    Evolved Gas Analysis (EGA) | TGA-FTIR & TGA-MS - Measurlabs
    Evolved gas analysis (EGA) is a quantitative analysis method for monitoring the gases that are formed when a sample undergoes thermal decomposition.
  81. [81]
    On-Line Thermally Induced Evolved Gas Analysis: An Update—Part 2
    This second part reviews the latest applications of Evolved Gas Analysis performed by on-line coupling heating devices to infrared spectrometers (EGA-FTIR).
  82. [82]
    Thermal degradation and combustion properties of most popular ...
    In this work, the thermal degradation and combustion behaviours of the most popular synthetic biodegradable polymers in the market, poly(lactide acid) (PLA), ...
  83. [83]
    Polymer Degradation - an overview | ScienceDirect Topics
    The degradation of polyethylene occurs by random scission—a random breakage of the bonds within the polymer. When heated above 450 C (840°F), polyethylene ...
  84. [84]
    Thermal Stability and Decomposition Mechanism of PLA ... - MDPI
    Aug 22, 2021 · It was found that the PLA/lignin nanocomposites show better thermostability than neat PLA, while tannin filler has a small catalytic effect that can reduce the ...
  85. [85]
    D638 Standard Test Method for Tensile Properties of Plastics - ASTM
    Jul 21, 2022 · This test method is designed to produce tensile property data for the control and specification of plastic materials.
  86. [86]
    How to Measure the Mechanical Properties of Polymers - AZoM
    Aug 25, 2021 · Tests used to investigate polymer properties include standard configurations such as flexural, compressive and tensile in their pure form.
  87. [87]
    ASTM D638: The Definitive Guide To Plastic Tensile Testing - Instron
    ASTM D638 is the most common testing standard for determining the tensile properties of reinforced and non-reinforced plastics.
  88. [88]
    Standard Test Method for Compressive Properties of Rigid Plastics
    Aug 9, 2023 · This test method covers the determination of the mechanical properties of unreinforced and reinforced rigid plastics, including high-modulus composites.
  89. [89]
    What is Compression Testing? - Instron
    Learn how compression tests evaluate material strength under crushing loads using universal testing machines. Discover methods, fixtures, and applications.
  90. [90]
    D790 Standard Test Methods for Flexural Properties of Unreinforced ...
    Jul 24, 2017 · 1.1 These test methods are used to determine the flexural properties of unreinforced and reinforced plastics, including high modulus composites ...
  91. [91]
    ASTM D790 3-point flexure test plastics - ZwickRoell
    The 3-point flexure test to ASTM D790 is a traditional standardized characterization method for rigid and semi-rigid plastics.
  92. [92]
    Standard Test Method for Rubber Property—Durometer Hardness
    Jul 23, 2021 · ASTM D2240 is a standard test method for measuring rubber hardness using durometers, based on indentation, and covers twelve types of  ...<|separator|>
  93. [93]
    Shore Durometer Hardness Testing of Rubber and Plastics - MatWeb
    The hardness of plastics is most commonly measured by the Shore® (Durometer) test or Rockwell hardness test. Both methods measure the resistance of plastics ...
  94. [94]
    D785 Standard Test Method for Rockwell Hardness of Plastics and ...
    Apr 14, 2023 · ASTM D785 is a standard test method for Rockwell hardness of plastics and electrical insulating materials, using two procedures to measure ...
  95. [95]
    Rockwell Hardness ASTM D785, ISO 2039 - Intertek
    The Rockwell Hardness test measures hardness by the net increase in impression depth under load. Higher numbers on R, L, M, E, or K scales indicate harder ...
  96. [96]
    Hardness of Plastics - When to Use RockWell & Shore Scales
    Jul 11, 2025 · Various scales measure plastic hardness (Shore A for soft elastomers, Shore D for rigid plastics, Rockwell for harder engineering polymers), ...
  97. [97]
    Influence of molecular weight and molecular weight distribution on ...
    Generally, mechanical properties of polymers increase with molecular weight, but above a limit, the properties are usually unaffected.
  98. [98]
    Dynamic Mechanical Analysis - Menard - Major Reference Works
    Jun 30, 2023 · Dynamic mechanical analysis (DMA) is applying a stress or strain to a sample at controlled frequencies and analyzing the response to obtain phase angle and ...
  99. [99]
    Dynamic mechanical analysis in materials science: The Novice's Tale
    Oct 7, 2020 · This article discusses the new avenues for DMA in different fields and takes the reader from the fundamentals to its advanced applicability.
  100. [100]
    Dynamic Mechanical Analysis - an overview | ScienceDirect Topics
    The dynamic mechanical analysis (DMA) method is used to track mechanical changes in cross-linked polymers as a function of temperature, as well as to quantify ...
  101. [101]
    [PDF] principles and applications of the cox-merz rule, rn-14 - TA Instruments
    Deviations from the rule occur at high frequencies, and the oscillatory data can either over- or under estimate the steady state data.
  102. [102]
    Cox-Merz Rule - an overview | ScienceDirect Topics
    The Cox-Merz rule states that the shear-rate dependence of the steady-state viscosity equals the frequency dependence of the complex viscosity.2 Flow Curves Of... · 2.2 Polydispersity As The... · 5.3 Rheological Measurements
  103. [103]
    (PDF) From melt flow index to rheogram - ResearchGate
    Aug 5, 2025 · A knowledge of the complete flow curve or rheogram of a polymeric melt depicting the variation of the melt viscosity over industrially ...
  104. [104]
    NMR Spectroscopy in the Characterization of Polymers
    Jun 1, 1983 · Bovey found the use of 1H NMR to be a fertile field to characterize the microstructure of hydrocarbon polymers. He used 19F NMR for ...
  105. [105]
    Polymer Molecular Weights via DOSY NMR | Analytical Chemistry
    May 5, 2023 · Diffusion-ordered spectroscopy (DOSY) 1H nuclear magnetic resonance (1H NMR) has become a powerful tool to characterize the molecular weights of ...Results and Discussion · Conclusions · Supporting Information · References
  106. [106]
    Solid state nuclear magnetic resonance of polymers - ScienceDirect
    We present a review of solid-state NMR studies applied to characterization of polymers. •. An order parameter that condensates the elastic response within ...
  107. [107]
    NMR Spectroscopy | Laboratory Analysis Services - Measurlabs
    Both solid and liquid samples are suitable for NMR analysis. Solid samples need to be soluble in a deuterated NMR solvent before the analysis.
  108. [108]
    Infrared, Raman, and Near-Infrared Spectroscopic Evidence for the ...
    IR spectra were collected at a 4 cm-1 resolution on a Nicolet Magna 550 FT-IR ... stretching region (2800 -3000 cm-1) were the focus points. Chloroform ...
  109. [109]
    Introducing Raman Spectroscopy through Analysis of Microplastic ...
    Sep 22, 2025 · The summative evaluation, primarily based on the laboratory reports, demonstrated students' ability to connect spectral features to polymer ...
  110. [110]
    Spectroscopic Techniques for the Characterization of Polymer ...
    One major advantage of Raman scattering is to allow the analysis of thick polymer samples while only very thin films can be examined by infrared transmission ...
  111. [111]
    Characterization of Modified PVDF Membranes Using Fourier ...
    May 30, 2024 · This review aims to description of various Fourier Transform IR (FTIR)-based spectroscopic techniques suitable to characterize (i) different ...
  112. [112]
    Infrared Linear Dichroism for the Analysis of Molecular Orientation in ...
    Mar 21, 2022 · Infrared dichroism is particularly well suited for characterizing polymer chain orientation at a molecular level.
  113. [113]
    Some Applications of Vibrational Spectroscopy for the Analysis ... - NIH
    Jul 8, 2019 · Selected examples of application of infrared and Raman spectroscopies illustrate their potential for monitoring polymer processes, measuring ...
  114. [114]
    Fast and quantitative 2D and 3D orientation mapping using Raman ...
    Dec 5, 2019 · Confocal Raman microscopy is a powerful non-destructive technique for chemical mapping of organic and inorganic materials. Here we demonstrate ...
  115. [115]
    Polarised Raman Spectroscopy - Edinburgh Instruments
    Jan 23, 2020 · Polarised Raman spectroscopy provides information about molecular structure, including the symmetry of vibrational modes.
  116. [116]
    Applications of Micro-Fourier Transform Infrared Spectroscopy (FTIR ...
    Dec 17, 2015 · Fourier transform infrared spectroscopy (FTIR) can provide crucial information on the molecular structure of organic and inorganic ...<|separator|>
  117. [117]
    Review of the fundamental theories behind small angle X-ray ... - NIH
    In this paper, the fundamental concepts and equations necessary for performing small angle X-ray scattering (SAXS) experiments, molecular dynamics (MD) ...
  118. [118]
    X-ray scattering in polymers and micelles - ScienceDirect.com
    SAXS is sensitive to the differences in electron density within a material and hence is able to resolve density fluctuations within polymer materials. It is ...
  119. [119]
    Wide-Angle X-ray Scattering Study on Shear-Induced Crystallization ...
    Jan 6, 2011 · Polymorphism, preferred crystal orientation, crystallization kinetics, and disorder effect were investigated based on WAXS 2D whole pattern ...
  120. [120]
    [PDF] Understanding Polymers Through X-ray Scattering
    Apr 1, 2025 · In transmission WAXS mode, the detector is positioned close to the sample. In accordance with Bragg's Law of Diffraction, two parallel rays of ...
  121. [121]
    Current progresses of small-angle neutron scattering on soft-matters ...
    This Review is intended to be a relatively comprehensive source for applying SANS on soft-matters investigation, including the basic principles and theory.
  122. [122]
    An in Situ Small- and Wide-Angle X-ray Scattering Study Using ...
    Dec 11, 2019 · The aim of this work is to extend the use of simultaneous SAXS and WAXS methods to characterize crystalline structure and morphology, ...
  123. [123]
  124. [124]
    A new Fourier transformation method for SAXS of polymer lamellar ...
    With the revised procedure, the lamellar thickness and long period can be obtained readily from real scattering, whether for a lamellar two-phase system or a ...
  125. [125]
    How can Polymers be Characterized Using SEM? - AZoM
    Oct 18, 2023 · SEM provides a high-resolution, three-dimensional visualization platform for in-depth examination and characterization of micro and nanoscale polymer surfaces ...Missing: seminal | Show results with:seminal
  126. [126]
    Transmission Electron Microscopy Methodology to Analyze Polymer ...
    Mar 28, 2024 · Our methodology enables long-sought insights into the polymer structure, introducing a new tool for high resolution studies of polymer crystallinity.Missing: seminal | Show results with:seminal
  127. [127]
    Electron microscopy for polymer structures - Oxford Academic
    Feb 18, 2022 · This paper reviews recent advances and perspectives of electron microscopy and its application to polymer hierarchical structures.Missing: seminal | Show results with:seminal
  128. [128]
    Soft matter and nanomaterials characterization by cryogenic ...
    Dec 10, 2019 · Cryo-TEM reduces beam damage and allows for characterization in a native, frozen-hydrated state, providing direct visual representation of soft ...Missing: seminal | Show results with:seminal
  129. [129]
    Polymer Characterization with the Atomic Force Microscope
    Atomic force microscopy is a powerful characterization tool for polymer science, capable of revealing surface structures with superior spatial resolution.