Fact-checked by Grok 2 weeks ago

Degree of polymerization

The degree of polymerization (DP), also denoted as n or x, is defined as the number of repeating monomeric units in a , molecule, block, or chain within a . This measure quantifies the length of the polymer chain and is fundamental to understanding the size and structure of polymeric materials. In polymer samples, which are typically polydisperse—meaning they contain chains of varying lengths—DP is expressed as statistical averages, such as the number-average degree of polymerization (\overline{DP}_n), calculated from the total number of monomer units divided by the number of chains, or the weight-average degree of polymerization (\overline{DP}_w), which weights longer chains more heavily. The DP directly relates to the 's molecular weight, where the number-average molecular weight M_n is given by M_n = \overline{DP}_n \times M_0, with M_0 being the molecular weight of the unit. DP values for common polymers often range from thousands to tens of thousands, reflecting the extensive chain lengths achieved during synthesis. The DP profoundly influences the physical, mechanical, and rheological properties of polymers, making it a critical parameter in materials design. Higher DP generally enhances tensile strength, toughness, and viscosity by increasing chain entanglement and reducing chain mobility, though it can also raise the glass transition temperature and complicate processing due to higher melt viscosity. For instance, in elastomers, a DP exceeding 10,000 promotes rubbery behavior through extensive entanglements, enabling elasticity and recovery under deformation. DP is commonly determined experimentally via methods like end-group analysis using nuclear magnetic resonance (NMR) spectroscopy, which ratios signals from repeat units to chain ends.

Fundamentals

Definition

The degree of polymerization (), denoted as n for a single or \overline{DP} for an average value, is defined as the number of repeating units comprising the . This measure quantifies the length of the in terms of structural units rather than absolute size. For a given , the degree of polymerization is calculated as DP = \frac{M_p}{M_m}, where M_p is the total molecular weight of the and M_m is the molecular weight of the repeating unit. This relation assumes a linear without significant branching or end-group contributions, providing a direct link between length and molecular weight. In homopolymers, which consist of identical repeating units derived from a single type, the DP simply counts these uniform units along the chain. For instance, a simple linear homopolymer such as has the structure -(\mathrm{CH_2 - CH_2})_n-, where n represents the degree of polymerization as the number of ethylene-derived units. In contrast, for copolymers formed from multiple types, the DP denotes the total number of all units incorporated into the chain, regardless of their sequence or . The term "degree of polymerization" originated in the early through the work of , who in 1920 described as chains of identical basic molecules, stating that "the number of the latter in the is called its degree of polymerisation." This introduction formed a key part of Staudinger's macromolecular hypothesis, challenging prevailing views and establishing the foundational understanding of chain structures. In typical polydisperse systems, the degree of polymerization is reported as an average to reflect the distribution of chain lengths.

Significance

The degree of polymerization (DP) quantifies the average number of units in a , serving as a primary indicator of chain length that directly governs essential material behaviors including processability, , and overall end-use performance. Longer chains with higher DP enhance intermolecular interactions, leading to improved and utility in applications, but they also increase melt , which can hinder fabrication techniques like or molding, and diminish solubility in solvents due to reduced chain mobility. Conversely, shorter chains facilitate easier processing and higher solubility but often lack the robustness required for demanding uses. DP plays a pivotal role in classifying macromolecules, distinguishing oligomers—characterized by low DP values typically below 10–20 repeating units—from true polymers, which exhibit high DP exceeding 100 units and display the unique viscoelastic and structural of long-chain molecules. Oligomers, with their limited chain length, behave more like small molecules, offering and reactivity advantages in niche roles such as intermediates or additives, while polymers' extended chains enable the formation of solid, functional materials essential to modern . This threshold-based distinction underscores DP's importance in defining macromolecular identity and potential. In contexts, precise control of is economically critical, as it allows optimization of polymers for high-volume applications like plastics, where balanced chain lengths ensure durability and formability; fibers, requiring high DP for tensile strength in textiles; and adhesives, where tunable DP influences and flexibility. Annual global of such materials exceeds hundreds of millions of tons, with DP adjustment via conditions or additives directly impacting cost-efficiency and product quality in sectors from to automotive components. The theoretical underpinnings of DP's effects trace to Paul Flory's statistical theories of the , which modeled chains as random coils prone to entanglement at sufficient lengths, providing a for predicting how chain dimensionality influences bulk behavior and reinforcing DP's centrality in design.

Polymerization Mechanisms

Step-Growth Polymerization

Step-growth polymerization involves the sequential reaction of bifunctional or multifunctional monomers, where each step forms a between reactive functional groups, typically through condensation reactions that eliminate small molecules such as . This leads to the formation of dimers, trimers, and higher oligomers, with chain growth occurring randomly between any compatible functional groups present in the system, resulting in a broad distribution of chain lengths. Unlike other polymerization types, the degree of polymerization () in step-growth processes builds gradually, as molecules of all sizes can react with each other, emphasizing the statistical nature of chain extension and termination. The relationship between the number-average degree of polymerization, \overline{DP}_n, and the extent of reaction p (the fraction of functional groups that have reacted) is described by the Carothers equation: \overline{DP}_n = \frac{1}{1 - p} This equation arises from a probabilistic model of the polymerization process. Consider a linear step-growth system with bifunctional monomers; the probability that a given functional group at one end of a chain has not reacted with another chain (effectively terminating growth) is $1 - p. Thus, the average number of monomers per chain is the reciprocal of this probability, yielding the form above. derived this in his foundational work on polyfunctional condensations, establishing the theoretical basis for predicting chain lengths in such systems. A key limitation of is the challenge in achieving high due to constraints, where the reverse reaction hinders complete . For instance, at 99% (p = 0.99), \overline{DP}_n \approx 100, meaning chains are still relatively short on average, and molecular weights remain modest without strategies like removing by-products to shift the . This slow approach to high molecular weight underscores the need for nearly quantitative yields, often exceeding 99.5% for practical applications. Representative examples of step-growth polymerization include the synthesis of polyesters, such as (PET) formed from and via esterification, and polyamides, like Nylon 6,6 produced from and through amidation. In systems with monomers of higher functionality (f > 2), such as trifunctional acids or alcohols, branching occurs, leading to a gelation point at p = \frac{1}{f-1}, beyond which an infinite network forms, marking the transition to a crosslinked gel. This phenomenon, observed in phenolic resins or polymers, highlights how functionality influences the structural evolution and ultimate material properties.

Chain-Growth Polymerization

Chain-growth polymerization, also known as addition polymerization, involves the sequential addition of monomers to a growing polymer chain through active centers, typically radicals, ions, or coordination complexes. The degree of polymerization (DP) in this mechanism is primarily controlled by the relative rates of propagation and termination steps, as each propagation event adds a monomer unit while termination halts chain growth. Initiation creates the active center, often from an initiator that generates radicals or ions, but it does not directly limit DP; instead, the balance between rapid propagation and slower termination or transfer reactions determines the average chain length. In conventional chain-growth processes, termination dominates, leading to a statistical distribution of chain lengths, but specialized variants like living polymerization suppress termination, enabling precise control over high DP values. In free-radical , a common variant, the number-average degree of polymerization \overline{DP}_n is given by the kinetic equation \overline{DP}_n = \frac{k_p [M]}{(2 k_t R_i)^{1/2}}, where k_p is the rate constant, [M] is the concentration, k_t is the termination rate constant, and R_i is the rate (often proportional to the initiator concentration [I] via R_i = 2 f k_d [I], with f as the initiator and k_d as the decomposition rate constant). This equation arises from the steady-state approximation for radical concentration, where the rate of divided by the rate of termination yields the average number of units added per . Higher concentrations or lower initiator levels increase \overline{DP}_n by favoring over termination, typically achieving values from hundreds to thousands depending on conditions. The mode of termination significantly influences the DP distribution. In disproportionation, two growing abstract a from each other, yielding two dead chains with saturated ends; here, \overline{DP}_n equals the kinetic chain length \nu (monomers per radical lifetime). In , the radicals couple to form one longer chain with a single link, doubling the chain length such that \overline{DP}_n = 2\nu; this also narrows the molecular weight distribution compared to disproportionation. The prevalence of each mode varies by monomer—for instance, styrene favors , while leans toward —directly affecting the polydispersity and overall chain length control. Living polymerization, pioneered by anionic initiation without termination or transfer, allows for exceptionally high and predictable DP by maintaining active chain ends throughout the reaction. The DP is directly proportional to the monomer-to-initiator ratio, \overline{DP}_n = \frac{[M]_0}{[I]_0}, enabling values exceeding 1000 with narrow distributions (polydispersity index near 1). This kinetic control contrasts with conventional methods, as chains resume growth upon monomer addition, facilitating block copolymer synthesis. Representative examples illustrate these principles. In free-radical polymerization of styrene, \overline{DP}_n often reaches 500–2000 under typical conditions (e.g., 60°C, azoisobutyronitrile initiator), yielding with versatile properties for and insulation. For produced via Ziegler-Natta , a chain-growth mechanism using titanium-based catalysts, DP values commonly exceed 1000–10,000, resulting in with superior strength for pipes and containers. These processes highlight how mechanistic control tunes DP for industrial applications.

Molecular Weight Averages

Number-Average

The number-average degree of polymerization, denoted \overline{DP}_n, represents the average number of units per polymer chain, calculated as the total number of units across all chains divided by the total number of chains in a polydisperse sample. In mathematical terms, for a where N_i is the number of chains containing exactly i units, \overline{DP}_n is expressed as \overline{DP}_n = \frac{\sum_{i=1}^{\infty} i N_i}{\sum_{i=1}^{\infty} N_i}. This formula arises from the arithmetic mean of chain lengths, weighted equally by the number of chains regardless of their size. Equivalently, in terms of mole fractions x_i = N_i / \sum_{j=1}^{\infty} N_j, the expression simplifies to the summation \overline{DP}_n = \sum_{i=1}^{\infty} i x_i, highlighting its dependence on the frequency distribution of chain lengths in polydisperse systems. Experimental determination of \overline{DP}_n relies on techniques that effectively count the number of polymer chains. End-group is a primary , where the concentration of terminal functional groups is quantified to infer the chain ; for instance, ^1H NMR spectroscopy can detect hydroxyl end groups in polyesters by integrating signals from protons adjacent to these groups, allowing calculation of \overline{DP}_n via the ratio of end-group to repeat-unit signals. Another direct approach is membrane osmometry, which measures the \Pi of dilute polymer solutions according to \Pi = cRT / M_n + A_2 c^2 + \cdots, where c is concentration, R is the , T is , M_n is the number-average molecular weight, and A_2 is the second virial ; \overline{DP}_n is then obtained as M_n / M_0, with M_0 being the monomer molecular weight. These methods are particularly suited for samples with functional end groups or where are accessible. Despite their utility, these measurements have limitations inherent to the number-averaging approach. \overline{DP}_n is highly sensitive to low-molecular-weight species, as each short chain contributes equally to the denominator as longer ones, thereby skewing the average downward and underestimating the presence of high-DP chains in the . Additionally, end-group analysis can be inaccurate if side reactions produce cyclic structures without end groups or if impurities interfere with or spectroscopic signals. Osmometry, meanwhile, requires careful of conditions and is less reliable for very high-molecular-weight polymers where osmotic pressures become immeasurably small.

Weight-Average

The weight-average degree of polymerization, denoted as \overline{DP}_w, represents a measure that weights each chain by its mass contribution to the overall sample. It is calculated as \overline{DP}_w = \frac{\sum_{i} N_i (DP_i)^2}{\sum_{i} N_i DP_i}, where N_i is the number of chains with degree of polymerization DP_i. This formula arises from the weight fractions of the chains, w_i = \frac{N_i DP_i M_0}{\sum N_i DP_i M_0}, where M_0 is the molecular weight of the repeating unit, leading to \overline{DP}_w = \sum w_i DP_i. In practice, \overline{DP}_w is often determined indirectly through the weight-average molecular weight M_w, which is measured using techniques such as (GPC) or (SEC) coupled with light scattering detectors, or standalone . These methods yield M_w as an independent of calibration standards, with \overline{DP}_w = M_w / M_0. relies on the intensity of scattered light from polymer solutions, following the Debye equation Kc / R_\theta = 1/M_w + 2A_2 c, where K is an optical constant, c is concentration, R_\theta is the Rayleigh ratio, and A_2 is the second virial coefficient; extrapolation to infinite dilution provides M_w. Compared to the number-average degree of polymerization, \overline{DP}_w provides a more accurate reflection of the overall mass distribution in polydisperse polymers, as it emphasizes contributions from longer chains that dominate the sample's total mass. This makes \overline{DP}_w particularly relevant for physical properties influenced by chain length and entanglement, such as melt , where empirical relations show \eta \propto M_w^{3.4} for entangled linear polymers above the critical entanglement molecular weight. For illustration, consider a simple bimodal distribution with 10 chains of DP_1 = 10 and 1 chain of DP_2 = 100. The weight-average degree of polymerization is then \overline{DP}_w = \frac{10 \cdot 10^2 + 1 \cdot 100^2}{10 \cdot 10 + 1 \cdot 100} = \frac{1000 + 10000}{100 + 100} = \frac{11000}{200} = 55. This value highlights how the single long chain significantly influences the average, unlike the number-average of approximately 18.2./05%3A_Molecular_Weight_Averages)

Polydispersity

The polydispersity index (PDI), denoted as Đ, quantifies the breadth of the distribution of chain lengths in a sample and is calculated as the ratio of the weight-average degree of polymerization (\overline{DP}_w) to the number-average degree of polymerization (\overline{DP}_n): \PDI = \frac{\overline{DP}_w}{\overline{DP}_n}. A PDI value of 1 indicates a perfectly monodisperse with all chains of identical length, while values greater than 1 reflect increasing heterogeneity in chain lengths. The number-average and weight-average degrees of polymerization are defined in the sections on molecular weight averages. In at high , the PDI theoretically approaches 2 due to the statistical nature of the process, resulting in a broad distribution of chain lengths./03:_Kinetics_and_Thermodynamics_of_Polymerization/3.02:_Kinetics_of_Step-Growth_Polymerization) Conversely, living polymerization mechanisms, which avoid termination and , yield nearly monodisperse polymers with PDI values close to 1, enabling precise control over molecular architecture./02:_Synthetic_Methods_in_Polymer_Chemistry/2.05:_Living_Cationic_Polymerization) Several factors influence the PDI during . In chain-growth processes, termination reactions—such as radical recombination or —and events broaden the by prematurely ending at varying lengths. Monomer impurities, including inhibitors or reactive contaminants, can exacerbate this by inducing side reactions that disrupt uniform growth and increase polydispersity. PDI is typically measured using (GPC), a size-exclusion technique that separates chains by hydrodynamic volume. The method involves calibrating the system with narrow molecular weight standards and generating a distribution curve by plotting the logarithm of molecular weight against volume, from which \overline{DP}_n and \overline{DP}_w are computed to derive the PDI. A narrow PDI, often below 1.5, is critical for applications requiring high uniformity, such as block copolymers used in self-assembling nanostructures, where broad distributions can disrupt and ordered morphologies.

Physical Properties

Mechanical Properties

The degree of polymerization (DP) significantly influences the mechanical properties of polymers, particularly through its effect on chain entanglements, which enhance load-bearing capacity and deformation resistance. As DP increases, tensile strength and rise due to greater chain entanglement density, allowing for more effective stress distribution across the material. This improvement is most pronounced up to the critical entanglement threshold, beyond which further increases yield diminishing returns as the material reaches maximum entanglement. Entanglement effects also manifest in the rheological behavior that underpins mechanical performance, such as in melt processing and viscoelastic response under load. For unentangled polymers (low ), melt scales as \eta \propto \overline{[DP](/page/DP)}, reflecting simple dynamics where chains move independently. In entangled regimes (high ), follows \eta \propto \overline{[DP](/page/DP)}^{3.4}, as predicted by extensions of the Doi-Edwards reptation theory, indicating constrained chain motion that contributes to higher and strength in the solid state. At low DP, polymers exhibit due to insufficient chain interlock, leading to easy crack propagation and low under stress. In contrast, high DP promotes through energy dissipation via chain slippage and uncoiling. For example, in nylon-6,6, low-molecular-weight oligomers (DP < 50) are brittle and prone to fracture with minimal elongation, whereas high-DP variants (DP > 100) demonstrate superior tensile strength and impact resistance, enabling applications in load-bearing components. Under sustained mechanical stress, polymers undergo involving chain scission, which progressively reduces average and degrades overall mechanical integrity. This process, often initiated at weak points along the chain, leads to decreased entanglement density, lowered tensile strength, and eventual brittle failure, as observed in where scission correlates directly with reduced molecular weight during tensile loading.

Thermal Properties

The glass transition temperature (T_g) of amorphous polymers increases with the degree of polymerization (), reflecting reduced chain end mobility and enhanced cooperative segmental motion in longer chains. This dependence arises because chain ends contribute free volume and lower the barrier for relaxation, an that diminishes as DP grows. The seminal empirical relation, known as the Fox-Flory equation, quantifies this trend: T_g = T_g^\infty - \frac{K}{M_n}, where T_g^\infty represents the asymptotic T_g at infinite molecular weight, M_n is the number-average molecular weight (M_n = \overline{DP}_n \times M_0, with M_0 the monomer unit molecular weight), and K is a polymer-specific constant related to end-group contributions. The dependence is typically expressed using the number-average DP (or M_n) for T_g, as it reflects chain-end concentration. This equation holds for many linear polymers above a critical DP of approximately 100-200, beyond which T_g plateaus. A representative example is , where low-DP fractions exhibit a T_g of about 70°C, rising to 100°C for high-DP materials with \overline{DP} > 1000. This shift underscores how increasing DP stabilizes the glassy state against thermal softening, influencing applications like and coatings. Similar behavior occurs in other amorphous polymers, such as , though the exact K value varies with chain flexibility and side-group interactions. For semicrystalline polymers capable of ordered packing, the melting point (T_m) likewise increases asymptotically with DP, driven by the stabilization of lamellar crystals by longer chain segments that reduce surface free energy defects. In polyethylene, short-chain oligomers (low DP, akin to n-paraffins) melt at temperatures below 100°C, resembling waxes, whereas high-DP variants like high-density polyethylene (HDPE) achieve T_m values of 130-140°C, enabling high-temperature structural uses. This progression reflects a convergence toward the theoretical infinite-chain limit, with T_m leveling off as end-group influences wane. Thermal degradation kinetics are markedly influenced by DP, with lower-DP polymers showing accelerated due to a higher of end groups that serve as sites for volatile fragment release. These end groups, often containing weaker bonds (e.g., hydroperoxides or allylic structures), promote unzipping or random scission at lower temperatures, reducing overall . For instance, fractionated samples demonstrate that low-molecular-weight fractions volatilize rapidly at initial stages, contrasting with the higher onset temperatures (above 350°C) for high-DP chains where mid-chain degradation dominates. This sensitivity guides the design of heat-resistant materials, emphasizing end-group passivation strategies.

References

  1. [1]
    degree of polymerization (D01569) - IUPAC Gold Book
    degree of polymerization ... The number of monomeric units in a macromolecule or oligomer molecule, a block or a chain. Source: PAC, 1996, 68, 2287. (Glossary of ...Missing: authoritative | Show results with:authoritative
  2. [2]
  3. [3]
    [PDF] Mechanical Properties of Polymers
    Amorphous & Crystalline. Degree of Polymerization (DP): the number of monomer units in a molecule. Page 4. Classes of Polymers. • Thermoplastics: such as PE ...
  4. [4]
    Degree of Polymerization - an overview | ScienceDirect Topics
    It is calculated as the ratio of molecular weight of a polymer and molecular weight of the repeat unit. Number average DP and weight average DP are the two main ...
  5. [5]
    Polymers - MSU chemistry
    Polymers formed by a straightforward linking together of monomer units, with no loss or gain of material, are called addition polymers or chain-growth polymers.
  6. [6]
    Symmetric diblock copolymers - degree of polymerization
    There are several ways to define the degree of polymerization for a block copolymer. The most naive one is just the total number of monomers.<|control11|><|separator|>
  7. [7]
    Life and work of Hermann Staudinger - K 2025
    “So macromolecules represent chains of one and the same basic molecule. The number of the latter in the macromolecule is called its degree of polymerisation.” ( ...
  8. [8]
    Hermann Staudinger Foundation of Polymer Science - Landmark
    During this reaction, which he called "polymerization," individual repeating units are joined together by covalent bonds. This new concept, referred to as " ...
  9. [9]
    [PDF] Principles of Polymerization
    ... Polymers and Polymerizations / 1. 1-1a Polymer Composition and Structure / 2 ... Degree of Polymerization / 274. 3-9b. Thermodynamics of Polymerization ...
  10. [10]
    Statistical theory of chain configuration and physical properties of ...
    Aug 6, 2025 · The earliest model of swelling of polymer networks, due to Flory and Rehner [87] , combined the Flory-Huggins theory for polymer/solvent ...
  11. [11]
    Step-Growth Polymerization - an overview | ScienceDirect Topics
    Step-growth polymerization, also known as polycondensation, is defined as a method for synthesizing polymers by exploiting the reactivity of functional ...
  12. [12]
    Polymers and polyfunctionality - Transactions of the Faraday Society ...
    Volume 32, 1936 From the journal: Transactions of the Faraday Society Polymers and polyfunctionality Wallace H. Carothers Abstract
  13. [13]
    Principles of Polymer Chemistry - Paul J. Flory - Google Books
    Principles of Polymer Chemistry · Contents · Common terms and phrases · References to this book · About the author (1953) · Bibliographic information ...
  14. [14]
    [PDF] Free-Radical Polymerization of Olefin Monomers
    Flory showed that a free-radical polymerization reaction, like other radical ... a specific ceiling temperature, and by the same token, Equation (11-56).
  15. [15]
    [PDF] Mechanism and Kinetics of Free Radical Chain Polymerization
    Jan 29, 2001 · Most anionic reactions have no inherent termination step. G. Odian, Principles of Polymerization, 3rd Ed., 1991, p 274. Define the rate of ...
  16. [16]
    [PDF] STATISTICS OF LINEAR POLYCONDENSATION
    At 90% conversion (p = 0.90) the number average degree of polymerization is only 10 ! Equivalent to number average molecular weight of ≈ 1000 g/mole. • At 95% ...
  17. [17]
    Polymerization and Structure of Opposing Polymer Brushes Studied ...
    Dec 8, 2021 · ... average chain length must be described by the averaged degree of polymerization. The number-averaged degree of polymerization DPn is defined ...
  18. [18]
    Synthesis of a Biodegradable PLA: NMR Signal Deconvolution and ...
    Dec 16, 2021 · End-group analysis through 1H NMR spectroscopy is a useful technique for the estimation of the number-average molecular weight in polymers (8−11) ...
  19. [19]
    Overview of Methods for the Direct Molar Mass Determination ... - NIH
    Mn can be determined by using the colligative properties of a polymer solution, e.g., by using osmometry, by the end group method, or by light scattering after ...
  20. [20]
    [PDF] Characterisation and Polymerisation Studies of Energetic Binders
    However, Mn is sensitive to low molecular weight species as they ... number average molecular weight of the polymer and its hydroxyl functionality.
  21. [21]
    Molecular-weight Determination by Light Scattering.
    A gram–Charlier analysis of scattering to describe nonideal polymer conformations. Macromolecules 2024, 57 (20) , 9518-9535.
  22. [22]
    Measurement of Molecular Weight by using GPC method - Shimadzu
    Methods used to measure the molecular weight of polymers include the light scattering method, osmotic pressure method, and viscosity method.
  23. [23]
    3 types of calculated molecular weight data | Malvern Panalytical
    Sep 21, 2023 · A GPC/SEC system with a static light scattering is the ideal way to calculate molecular weight. Since the molecular weight of the sample is ...
  24. [24]
    Molecular Polymer Weight and Zero-Shear Viscosity - AZoM
    May 7, 2019 · There is a strong link between what is called the weight-average molecular weight of a polymer and a parameter called the zero-shear viscosity.
  25. [25]
    Polydispersity Index: How Accurately Does It Measure the Breadth of ...
    Polydispersity index (PDI) is used as a measure of the breadth of the molecular weight distribution. PDI is defined as M w /M n where M w and M n are the ...Missing: step- | Show results with:step-
  26. [26]
    Tailoring polymer dispersity and shape of molecular weight ... - NIH
    The width and shape of molecular weight distributions can significantly affect the properties of polymeric materials and thus are key parameters to control.
  27. [27]
    Atom Transfer Radical Polymerization | Chemical Reviews
    Prolonged reaction times leading to nearly complete monomer conversion may not increase the polydispersity of the final polymer but will induce loss of end ...
  28. [28]
    Employing MALDI-MS on Poly(alkylthiophenes): Analysis of ...
    The polydispersity index calculated from the MALDI spectra were the same or ... Monomer purity has also been found to affect the end-group distribution.
  29. [29]
    Effects of Polydispersity on the Order−Disorder Transition in Block ...
    The effects of PDI on the ODT have potential implications on the processability of block copolymers. In general, processing is best carried out above the ...
  30. [30]
    Synthesis, characterization and properties of amphiphilic block ...
    Jun 6, 2012 · The polymerization resulted in copolymers with polydispersity index below 1.5. The self-assembly behavior of the triblock copolymers was studied ...
  31. [31]
    Entanglements and Fracture in Polymers | Chemical Reviews
    Polymer chains entangle when they are sufficiently long, dense, and mobile, comprising the microstructure of polymers. Entangled polymer chains cannot pass ...
  32. [32]
    [PDF] The melt viscosity-molecular weight relationship for linear polymers
    Jan 24, 2020 · A. Melt Viscosity-Molecular Weight Relationship. Viscosities at 25 °C for all samples are listed together with molecular weights in ...
  33. [33]
    Effect of Molecular Structure Change on the Melt Rheological ...
    Dec 4, 2018 · Sufficiently entangled polymer melts show a strong viscosity (η) increase with their molar mass (M), η ∼ M3, compared to unentangled polymers (η ...
  34. [34]
    [PDF] Experiment 5 Tensile Properties of Various Polymeric Materials
    If the testing temperature is below Tg, then materials with low molecular weight are extremely brittle. For semicrystalline polymers, the same trends apply, ...
  35. [35]
    [PDF] HOW MOLECULAR STRUCTURE AFFECTS MECHANICAL ...
    For the uncrosslinked SI material, it was shown that notched tensile strength is a strong function of both temperature and molecular weight, whereas stiffness.
  36. [36]
    [PDF] Mechanical Shear Degradation of Polymers in Solution: A Review.
    This Report reviews the literature on mechanical degradation of polymer solutions up to June 1975 and includes work which has been reported since the.
  37. [37]
    Second-Order Transition Temperatures and Related Properties of ...
    Dilatometric and visco metric data on fractionated polystyrenes containing diethylbenzene end groups are presented over wide temperature ranges.Missing: original | Show results with:original
  38. [38]
    The Melting Temperatures of the n–Paraffins and the Convergence ...
    The highest experimentally observed melting temperature for polyethylene is 138.7 °C, and for poly-methylene, 141.4 °C [6, 7]. Both of these were extended-chain ...Missing: oligomers | Show results with:oligomers
  39. [39]
    Thermal Degradation of Fractionated High and Low Molecular ... - NIH
    The low molecular weight whole polymer shows a very rapid initial degradation rate, over 0.13 percent volatilization per minute, which drops very rapidly to ...
  40. [40]
    The influence of the functional end groups on the properties of ...
    This review summarizes the various structures of PLA end groups as well as the methods of their modification and characterization.