Fact-checked by Grok 2 weeks ago

Cross-link

A cross-link is a or short sequence of bonds that connects two or more chains or molecular segments, typically forming a small from which at least four chains emanate, as defined in macromolecular chemistry. This structural feature is fundamental in both synthetic and natural materials, enhancing stability and altering physical properties through the creation of interconnected networks. In , cross-linking transforms linear or branched polymers into rigid, three-dimensional structures by forming covalent bonds between chains, which significantly increases the material's hardness, , and resistance to deformation. For instance, thermosetting polymers like resins and polyurethanes rely on extensive cross-linking to achieve and , enabling applications in adhesives, foams, and composites. The degree of cross-linking directly influences properties such as elasticity and , with higher cross-link density leading to insoluble, infusible networks that do not melt upon heating. In biological systems, cross-links play a in maintaining and function, often occurring naturally through bonds or enzymatic processes, or artificially via chemical reagents that target functional groups like amines, sulfhydryls, or carboxyls. These linkages stabilize tertiary and quaternary protein conformations, capture transient molecular interactions, and are essential in processes such as fiber formation in connective tissues. Cross-linking techniques are also widely used in for , such as linking enzymes to antibodies or immobilizing proteins on surfaces for diagnostic assays.

Fundamentals

Definition and Formation

A cross-link is a or short sequence of bonds that connects two or more chains or molecular segments, typically forming a small region from which at least four chains emanate, thereby enhancing the structural integrity and mechanical of the . This forms a network that restricts chain mobility, distinguishing cross-linked systems from uncross-linked polymers, which primarily rely on physical entanglements—topological constraints where long chains become intertwined without chemical bonds—for their viscoelastic behavior. Cross-links form through various chemical mechanisms, including radical polymerization, where free radicals initiate bonding between chains; condensation reactions, involving the elimination of small molecules to link functional groups; and ionic interactions, which can create temporary or permanent bridges via electrostatic forces. In radical processes, initiators generate reactive sites that propagate to form covalent ties, often using multifunctional monomers to branch and connect chains. typically occurs in step-growth polymerizations with monomers bearing multiple reactive sites, such as diols and triacids, leading to a three-dimensional network. Ionic cross-linking, while often reversible, arises from coordination between charged groups on polymer segments. These mechanisms can produce either covalent (permanent) or non-covalent (dynamic) cross-links, depending on the bonding type. A seminal example of cross-link formation is the of , discovered by in 1839, where heating with creates covalent sulfur bridges between diene units in adjacent chains, transforming the sticky material into a durable . This process exemplifies radical-induced cross-linking, as sulfur radicals facilitate the addition across double bonds, establishing a foundational application of the concept in . Cross-links are broadly categorized into chemical and physical types based on the nature of the bonds involved, with hybrid variants combining elements of both. Chemical cross-links form through covalent bonds, which provide permanent and mechanically robust connections between molecular chains. These bonds, such as linkages (–S–S–) or groups (–COO–), exhibit high strength and thermal stability due to their high bond dissociation energies, typically ranging from 200 to 400 kJ/mol. For instance, bonds are formed by the oxidation of groups and are known for their durability in various chemical environments. linkages, often created via reactions, contribute to network rigidity and resistance to deformation. Unlike weaker interactions, covalent cross-links do not readily dissociate under ambient conditions, ensuring long-term structural integrity. Physical cross-links, in contrast, arise from non-covalent interactions, including hydrogen bonding, ionic associations, and crystalline domains, which allow for reversibility and adaptability. Hydrogen bonds, with energies around 10–40 kJ/mol, form between electronegative atoms like oxygen or and , enabling dynamic association and dissociation influenced by environmental factors such as temperature or . Ionic bonds involve electrostatic attractions between charged groups, providing moderate strength (up to 80 kJ/mol) but susceptibility to disruption by solvents or ions. Crystalline domains act as physical junctions where aligned segments create ordered regions that restrict chain mobility, yet these can melt or reorganize under or stress. This reversibility stems from the lower energy barriers for bond breaking and reforming compared to covalent types. Hybrid cross-links integrate covalent and non-covalent mechanisms, such as ionically reinforced covalent networks, to balance permanence with tunability. In these systems, a primary covalent framework maintains overall stability, while secondary ionic interactions enhance or enable self-healing by dissipating energy through reversible . For example, networks with covalent backbones supplemented by ionic clusters exhibit improved properties without sacrificing adaptability. A key distinction between these types lies in their stability profiles: chemical cross-links offer both thermodynamic stability (favoring the bonded state at ) and kinetic stability (high for reversal), whereas physical cross-links are primarily kinetically stable under specific conditions but thermodynamically prone to , allowing stimulus-responsive behavior.

Applications in

Synthetic

Cross-links in synthetic polymers are covalent bonds that connect polymer chains, transforming linear or branched macromolecules into three-dimensional networks that enhance integrity and durability. These networks are essential for creating thermosetting materials, where the cross-linking renders the polymer insoluble and infusible upon heating, unlike thermoplastics. In applications such as rubbers and plastics, cross-links improve elasticity by preventing chain slippage under , increase tensile strength through load distribution across the network, and boost thermal by restricting segmental motion at elevated temperatures. The historical development of cross-linked synthetic polymers began in the with modifications to , marking a pivotal shift from thermoplastic-like behavior to robust elastomers. In 1839, discovered , a process involving heating with to form polysulfide cross-links between chains, which dramatically extended the material's service temperature range and elasticity while eliminating tackiness. This breakthrough laid the foundation for production, particularly during when natural supplies were scarce, leading to the synthesis of cross-linkable elastomers like rubber. By the early , the advent of fully synthetic thermosets, such as phenol-formaldehyde resins () in 1907, expanded cross-linking to rigid plastics via condensation reactions, evolving into modern high-performance materials like epoxies and polyurethanes. A classic example is vulcanized rubber, where sulfur atoms bridge isoprene units, typically at 1-3% concentration, to yield a material with superior rebound resilience and to abrasion, as seen in tires and seals. In contrast, epoxy resins form cross-links through nucleophilic ring-opening reactions between epoxide groups and amine hardeners, producing a densely networked structure with hydroxyl side groups that confer and chemical . These cross-links elevate the by up to several orders of magnitude compared to uncured resins, reduce in solvents due to the infinite molecular weight of the network, and shift the temperature (Tg) higher—often from below to over 100°C—enabling use in structural composites and coatings. Physical cross-links in systems arise from non-covalent interactions that temporarily connect chains, imparting a dynamic that contrasts with the permanent bonds of covalent cross-links. These interactions enable reversible and , allowing the to respond adaptively to external stimuli such as or . Unlike the rigid networks formed in synthetic through chemical means, physical cross-links contribute to viscoelastic in materials like elastomers and gels. Key mechanisms of physical cross-linking include hydrogen bonding, ionic clustering, and chain entanglements. In , hydrogen bonding between groups along the backbone creates transient junctions that enhance elasticity and , as evidenced by studies showing interchain hydrogen bonds significantly influencing properties. Ionic clustering occurs in ionomers, where oppositely charged groups aggregate into multipolar domains that act as physical cross-links, increasing and strength by restricting chain mobility while maintaining reversibility through ion hopping. Physical entanglements in thermoplastics involve the topological of long chains, which transfers stress between chains and provides pseudo-cross-linking without chemical bonds, particularly effective in high-molecular-weight linear polymers. The primary advantages of physical cross-links stem from their reversibility, enabling reprocessability and self-healing capabilities that are not feasible with covalent networks. For instance, materials can be melted or reshaped at elevated temperatures as interactions dissociate, facilitating in applications like elastomers. Self-healing occurs through reformation of bonds at damage sites, restoring mechanical integrity without external intervention, as demonstrated in dynamic networks where chain repairs fractures. Representative examples include supramolecular polymers, where multiple hydrogen bonds or host-guest interactions form linear or networked structures with tunable dynamics, enabling applications in healable coatings and adhesives. Block micelles also exemplify physical cross-links, with amphiphilic chains self-assembling into core-shell structures stabilized by hydrophobic interactions and hydrogen bonding, providing stable yet responsive nanostructures for . Despite these benefits, physical cross-links exhibit limitations, particularly lower strength under high , as the transient of interactions leads to and reduced load-bearing compared to covalent alternatives. This vulnerability can result in material failure in demanding environments, necessitating hybrid designs for enhanced .

Characterization Methods

Degree of Crosslinking

The degree of crosslinking, commonly quantified as cross-link density (ν), represents the average number of cross-links per unit volume or per unit mass in a network, typically expressed in units of mol/m³ or mol/cm³. This metric captures the extent to which polymer chains are interconnected, directly determining the network's structural integrity and transition from a viscous melt to an elastic solid. In crosslinked systems, ν is inversely related to the average molecular weight between cross-links (M_c), often approximated as ν = ρ / M_c, where ρ is the . The importance of cross-link density lies in its profound impact on key material properties. Higher ν restricts chain mobility, reducing the equilibrium swelling ratio by limiting solvent uptake and enhancing resistance to dissolution, while also improving mechanical strength, elasticity, and thermal stability at the expense of ductility. Conversely, lower ν promotes greater flexibility and higher swelling, which can accelerate degradation through increased permeability to environmental agents like water or oxidants. These effects are critical in tailoring polymers for specific uses, such as elastomers or hydrogels, where balanced density ensures optimal performance without brittleness or excessive softness. Theoretical models, such as the Flory-Rehner equation, provide a foundational approach to understanding and predicting cross-link density through equilibrium swelling behavior. Derived from thermodynamic principles, the equation balances the elastic retraction of the network against the of mixing: \ln(1 - \phi) + \phi + \chi \phi^2 + \nu V_1 \left( \phi^{1/3} - \frac{\phi}{2} \right) = 0 where ϕ is the polymer volume fraction at equilibrium, χ is the Flory-Huggins interaction parameter, V_1 is the molar volume, and ν is the cross-link density. This model, originally developed for lightly crosslinked rubbers, remains widely used to estimate ν from experimental swelling data despite assumptions like affine deformation. Several synthesis parameters govern the achievable degree of crosslinking. The monomer-to-crosslinker ratio directly controls ν, as higher crosslinker content increases interconnection points during . Extended curing time allows more complete reaction progression, elevating until a plateau is reached, while concentration accelerates and rates, influencing the final network uniformity and extent. Physical cross-links, such as those from hydrogen bonding or ionic interactions, can contribute to effective alongside chemical ones, modulating overall behavior without altering covalent .

Experimental Techniques

Swelling tests, including gel fraction analysis, are fundamental experimental techniques for assessing the presence and extent of cross-links in by measuring the insoluble fraction after extraction. In this method, a polymer sample is immersed in a suitable , such as acetone or , to dissolve uncross-linked chains while leaving the cross-linked network intact; the gel fraction is then calculated as the ratio of the dry weight after extraction to the initial weight, providing a direct indicator of cross-link . This approach is particularly effective for thermoset polymers and elastomers, where higher gel fractions correlate with greater cross-linking, and it is standardized under ASTM D2765 for determining insoluble content in cross-linked plastics through procedures involving extraction and weighing. Rheological measurements, such as dynamic shear , evaluate cross-link extent by analyzing the storage (G'), which reflects the response of the polymer network. In frequency sweep experiments, the plateau value of G' at low frequencies indicates the cross-link , as denser networks exhibit higher due to restricted mobility; this is commonly applied to hydrogels and rubbers to monitor gelation during curing. These metrics build on degree of crosslinking concepts by providing practical quantification through viscoelastic behavior. Spectroscopic methods like () spectroscopy identify cross-links by probing molecular structure and dynamics, with techniques such as solid-state 1H measuring transverse relaxation times (T2) that shorten in highly cross-linked systems due to reduced . Advanced methods include electron spin (ESR) spectroscopy for detecting cross-links, which monitors free concentrations and their evolution during initiation and propagation in processes like peroxide-induced cross-linking of . (DMA) assesses viscoelastic properties by applying oscillatory stress and measuring storage (E') and loss (E'') as functions of or , revealing the rubbery plateau proportional to cross-link in cured polymers. Recent advancements enable in-situ monitoring of cross-linking via , which tracks vibrational changes in chemical bonds during real-time , such as in resins where peak shifts indicate cross-link formation post-2010 developments in confocal setups. Calibration against standards like ASTM D2765 ensures accuracy across these techniques, with gel fraction results often validated against rheological data for comprehensive cross-link assessment.

Biological Contexts

In Proteins

Cross-links in proteins refer to covalent or non-covalent bonds that connect different parts of a polypeptide chain or multiple chains, playing crucial roles in maintaining structural integrity and enabling biological functions. These interactions are essential for proper , stability, and activity, particularly in extracellular and secreted proteins exposed to oxidative environments. Covalent cross-links in proteins include bonds, formed between the thiol groups of residues (-S-S-), and isopeptide bonds, which link the carboxyl group of an aspartic or to the ε-amino group of a . bonds are the most common covalent type, providing rigidity to protein structures by linking distant regions of the chain. Isopeptide bonds, often intramolecular, form spontaneously during folding in hydrophobic environments and enhance resistance to and , as seen in Gram-positive bacterial surface proteins. Non-covalent cross-links, such as salt bridges, involve electrostatic interactions between oppositely charged side chains (e.g., and aspartate), contributing to secondary and tertiary structure stabilization without forming permanent bonds. Disulfide bond formation typically occurs through oxidative coupling of cysteine thiols, catalyzed by enzymes like (PDI) in the (ER), where the oxidizing environment promotes pairing to achieve native configurations. This process is vital for secretory proteins, ensuring correct folding before export. In contrast, isopeptide bonds often form autocatalytically without dedicated enzymes, relying on proximity during folding. Salt bridges, being non-covalent, form dynamically based on local and ionic conditions, adjusting to support protein dynamics. These cross-links stabilize and structures, particularly in enzymes and antibodies, by reducing conformational and locking functional domains in place. In enzymes, disulfide bonds maintain geometry for , while in antibodies, interchain disulfides link heavy and light chains, enabling . For instance, human insulin features three disulfide bonds—one intra-chain in the A chain (CysA6-CysA11) and two interchain (CysA7-CysB7, CysA20-CysB19)—that are essential for its folded structure and . Recent studies since 2015 have highlighted (AGEs) as non-enzymatic covalent cross-links formed between reducing sugars and protein residues, accumulating with age and contributing to tissue stiffening and dysfunction. AGEs, such as pentosidine, create intra- and intermolecular bridges in long-lived proteins like , impairing elasticity and promoting , which accelerates aging-related pathologies. Aberrant cross-links are implicated in diseases like Alzheimer's, where dityrosine and transglutaminase-mediated bonds stabilize amyloid-β plaques, enhancing their aggregation and neurotoxicity. These pathological cross-links in plaques resist degradation, perpetuating neuronal damage and cognitive decline.

In Nucleic Acids

Cross-links in nucleic acids primarily affect DNA and RNA, forming covalent bonds between nucleotides that disrupt the double helix structure. These lesions include interstrand cross-links (ICLs), which covalently link bases on opposite strands, and intrastrand cross-links, which connect adjacent bases within the same strand. ICLs, such as those induced by psoralen in phototherapy, occur at sites like 5'-TA-3' sequences and require activation by long-wavelength UVA light to form monoadducts that subsequently cross-link strands. In contrast, intrastrand cross-links, exemplified by cisplatin's 1,2-GG or 1,2-AG adducts, primarily target adjacent guanines on the same strand, comprising about 90% of cisplatin's DNA lesions. UV radiation induces intrastrand pyrimidine dimers, such as cyclobutane pyrimidine dimers (CPDs) between adjacent thymines or cytosines, representing a major form of solar damage to DNA. Chemical agents like nitrogen mustards (e.g., mechlorethamine) and mitomycin C form ICLs at GpC or CpG sites through alkylation, with nitrogen mustards yielding about 5% ICLs upon reaction with guanine N7 positions. These cross-links profoundly impair function by blocking essential processes. ICLs and intrastrand lesions prevent strand separation, halting forks and progression during transcription, which triggers arrest or in affected cells. If unrepaired, they lead to double-strand breaks during replication or via error-prone bypass mechanisms, contributing to genomic . For instance, persistent from UV exposure can cause C-to-T transitions if bypassed inaccurately. Repair of these lesions involves specialized pathways to restore genome integrity. Intrastrand cross-links, including UV-induced and cisplatin adducts, are primarily excised via the (NER) pathway, which recognizes helical distortions and removes oligonucleotides containing the damage using endonucleases like XPF-ERCC1 and XPG. ICLs require a more coordinated response, often initiated during S-phase replication; the (FA) pathway plays a central role, monoubiquitinating FANCD2-FANCI to recruit nucleases for unhooking the cross-link, followed by translesion synthesis and to fill gaps. Defects in FA genes hypersensitize cells to ICL agents, as seen in patients. Repair enzymes may form transient cross-links with DNA intermediates, linking to protein-associated mechanisms briefly noted in broader contexts. In , cross-linking agents are pivotal in due to their toward proliferating cells. , a cornerstone chemotherapeutic, exploits intrastrand and minor formation to treat testicular, ovarian, and cancers, with over 23% of clinical trials featuring it as of 2018. , approved by the FDA in 1974 for gastric and pancreatic carcinomas, induces ICLs at CpG sites to inhibit tumor growth, often in combination regimens. Psoralen-based photochemotherapy (PUVA) leverages ICLs for treatment and certain lymphomas, where UVA-activated targets hyperproliferative skin cells. Nitrogen mustards, like , form ICLs in regimens for leukemias and , underscoring the therapeutic exploitation of cross-link repair vulnerabilities.

In Structural Biomolecules

In structural biomolecules, cross-links play a crucial role in providing mechanical strength and rigidity to extracellular matrices in plants and animals. , a complex polyphenolic polymer abundant in plant walls, features (such as β-O-4 aryl ) and ester bonds that interconnect phenylpropanoid units, forming a branched network that imparts hydrophobicity and resistance to compression. These cross-links integrate with hemicelluloses and , creating a composite that reinforces secondary walls during . In contrast to intracellular linkages in peptides or , lignin's polyphenolic cross-links are extracellular and contribute to long-term structural integrity in vascular tissues. The formation of lignin's cross-links occurs through oxidative of monolignols—such as p-coumaryl, coniferyl, and sinapyl alcohols—catalyzed primarily by class III peroxidases in the presence of . This process takes place in the during , where radical intermediates couple combinatorially to yield the irregular and bonds, ensuring adaptability to environmental stresses. Peroxidases localize to the , binding to growing polymer chains to direct deposition and enhance cross-linking efficiency. Beyond , cross-links in animal structural biomolecules like exemplify similar principles of extracellular stabilization. In , lysyl oxidase () mediates the oxidative of and hydroxylysine residues, generating reactive aldehydes that form stable aliphatic cross-links, such as aldimine or pyridinoline bonds. These interconnections between tropocollagen molecules increase diameter and tensile strength, essential for the biomechanical properties of tissues like tendons and . Lignin's cross-linked structure confers ecological significance by rendering plant biomass highly resistant to microbial degradation, thereby slowing carbon cycling in terrestrial ecosystems and preserving . This recalcitrance poses substantial challenges for production, as enzymatic breakdown requires harsh pretreatments to disrupt and bonds, with post-2020 research highlighting gaps in developing efficient ligninolytic enzymes like laccases and peroxidases for sustainable . Ongoing efforts focus on fungal and bacterial systems to overcome these barriers, aiming to improve conversion yields without compromising integrity.

References

  1. [1]
    crosslinking (CT07136) - IUPAC Gold Book
    Reaction involving sites or groups on existing macromolecules or an interaction between existing macromolecules that results in the formation of a small region ...
  2. [2]
    CROSS-LINK Definition & Meaning - Merriam-Webster
    Oct 6, 2025 · a crosswise connecting part (such as an atom or group) that connects parallel chains in a complex chemical molecule (such as a polymer)
  3. [3]
    8.26: Cross-Linking - Chemistry LibreTexts
    Jun 25, 2023 · Cross-linking is the formation of covalent bonds between polymer chains, creating a rigid, interconnected network. This results in a higher ...
  4. [4]
    Overview of Crosslinking and Protein Modification
    Crosslinking is the process of chemically joining two or more molecules by a covalent bond. Modification involves attaching or cleaving chemical groups to alter ...
  5. [5]
    Cross-Link - an overview | ScienceDirect Topics
    Cross links are defined as covalent linkages formed between acceptors and a matrix, typically involving proteins, that connect different molecular structures ...
  6. [6]
    What is Crosslinking? | Beyond Chemistry - Stahl
    Crosslinking is joining polymers by coating bonds, connecting polymer chains with covalent or ionic bonds, increasing size and durability.
  7. [7]
    Polymer Chains Entanglement - an overview | ScienceDirect Topics
    Polymer chain entanglements refer to the interactions between long polymer chains that become intertwined or knotted, influencing the physical properties ...
  8. [8]
    Crosslinking Polymers: Types, Effects, Applications & Trends
    Jun 13, 2024 · Crosslinking involves the formation of covalent bonding between adjacent polymer backbone. It creates an interconnected three-dimensional structure.
  9. [9]
    Mechanisms of Polymerization - an overview | ScienceDirect Topics
    The mechanism of radical polymerization follows three steps namely, initiation, propagation, and termination. During initiation, a molecule called a radical ...Missing: cross- | Show results with:cross-
  10. [10]
    Intramolecularly Cross-Linked Polymers: From Structure to Function ...
    May 22, 2020 · A cross-link is defined as a small region in a macromolecule from which at least four chains emanate that was formed by reactions involving ...
  11. [11]
    (PDF) Cross-linking of polymers by various radiations: Mechanisms ...
    be formed by using a monomer with three or more functional groups as a component of monomers. Post-cross-linking between polymer chains is usually carried out ...
  12. [12]
    Charles Goodyear | Rubber, Vulcanization, Inventor - Britannica
    In 1839 he accidentally dropped some India rubber mixed with sulfur on a hot stove and so discovered vulcanization. He was granted his first patent in 1844 but ...
  13. [13]
    Vulcanization - an overview | ScienceDirect Topics
    The vulcanization process established by Goodyear in 1839 is still currently in use. Sulphur vulcanization takes place in three stages. (1). Induction period.
  14. [14]
    Polymers - MSU chemistry
    If instead, the chains of rubber molecules are slightly cross-linked by sulfur atoms, a process called vulcanization which was discovered by Charles Goodyear ...
  15. [15]
    Effects of Cross-Link Density and Distribution on Static and Dynamic ...
    Simulation results show that the introduction of cross-links slows the chain dynamics down and thus leads to a slight increase in the glass transition ...
  16. [16]
    [PDF] 3.5 POLYMERS - CSUN
    Jul 23, 2002 · Thermoset Plastics: Thermoset plastics are extensively cross-linked and do not deform or soften upon heating. As a result, thermoset ...
  17. [17]
    Recent development of biodegradable synthetic rubbers and ... - NIH
    The industrialization of natural rubber began following the discovery of the vulcanization process by Charles Goodyear in 1839. The vulcanization process ...Missing: sulfur | Show results with:sulfur
  18. [18]
    [PDF] Lecture 2, Early History
    Continued reaction builds up a densely cross-linked network. This is Bakelite, a thermosetting polymer. Once the reaction is complete, the material cannot.
  19. [19]
    A Case Study on Natural Rubber” | ACS Applied Materials & Interfaces
    Invented by Charles Goodyear, chem. crosslinking of rubbers by sulfur vulcanization is the only method by which modern automobile tires are manufd. The ...
  20. [20]
    Cross-linking of Epoxy Resins - American Chemical Society
    The network build-up is dependent on reactivities of functional groups, functionality of monomers, composition of the system, and reaction paths as a function ...
  21. [21]
    Soft Materials by Design: Unconventional Polymer Networks Give ...
    ... physical crosslinks in polymer networks are relatively weak, transient and reversible. Typical examples of weak physical crosslinks include hydrogen bond ...
  22. [22]
    Overlooked Impact of Interchain H-Bonding between Cross-Links on ...
    The impact of interchain H-bonding on the mechanical properties of the polyurethane thermoset was unprecedently emphasized and supported by evidence.Missing: ionomers | Show results with:ionomers
  23. [23]
    The role of ionic clusters in the determination of the properties of ...
    Jan 5, 2021 · The mechanical properties of the ionomers are determined by their crystalline structure and cluster formation. Clusters increase strength and modulus, while ...
  24. [24]
    Tough Materials Through Ionic Interactions - PMC - NIH
    ... ionic clusters at high ion contents. Since the dynamic crosslinker forms ionic clusters that act as physical crosslinks, they reduce the chain mobility and ...
  25. [25]
    Interplay between entanglement and crosslinking in determining ...
    In polymer physics, the concept of entanglement refers to the topological constraints between long polymer chains that are closely packed together. Both theory ...
  26. [26]
    Entanglements
    The structural conformation that leads to transfer of a mechanical stress from one chain to another is termed an entanglement.
  27. [27]
    Dynamic cross-linked topological network reconciles the ... - Science
    Mar 19, 2025 · This work represents a molecular strategy that combines electronic effect and topology network design to modulate materials' properties.Missing: ionically | Show results with:ionically
  28. [28]
    Design of self-healing and self-restoring materials utilizing ... - Nature
    Feb 18, 2022 · This review describes an overview of the strategies used to prepare self-healing and self-restoring materials utilizing reversible and movable crosslinks.
  29. [29]
    The Versatility of Polymeric Materials as Self-Healing Agents ... - NIH
    A summary of physical cross-linking and chemical cross-linking use in the self-healing polymer mechanism. Self-healing hydrogels can be formed through physical ...
  30. [30]
    Supramolecular Polymers: Historical Development, Preparation ...
    A variety of noncovalent interactions (e.g., multiple hydrogen bonding, metal coordination, host–guest interactions, and aromatic stacking) have been employed ...
  31. [31]
    Highly robust supramolecular polymer networks crosslinked by a ...
    Apr 9, 2024 · Various supramolecular polymers have been constructed by using different recognition motifs, such as hydrogen bonding, π‒π stacking, host–guest ...
  32. [32]
    Core-Crosslinked Polymeric Micelles: Principles, Preparation ... - NIH
    These crosslinks can e.g. be based on covalent bonds, hydrogen bonding or π-π-stacking. However, although the pharmacokinetic properties of PM tend to ...
  33. [33]
    Effects of Crosslinking on the Mechanical Properties Drug Release ...
    Aug 29, 2013 · In this report, we explore the potential of modulating the water content, mechanical properties, and drug release profiles of protein films.
  34. [34]
    [PDF] Crosslink Density of Rubbers - IUPAC
    Crosslink density (ν) is a measure of crosslinked points per unit volume, normally expressed in mol/cm3.
  35. [35]
    Crosslinking Density - an overview | ScienceDirect Topics
    The density of crosslinks in a polymer can be experimentally obtained by the equation d = E ' r 3 R ( T g + 40 ) , where d = crosslinking density per unit ...
  36. [36]
    Understanding the role of cross-link density in the segmental ...
    Aug 8, 2022 · Cross-linking is known to play a pivotal role in the relaxation dynamics and mechanical properties of thermoset polymers, which are commonly ...INTRODUCTION · Cross-linked polymer model · Mechanical properties of cross...
  37. [37]
    Effect of the Cross-Linking Density on the Swelling and Rheological ...
    Flory–Rehner Theory​​ The number of cross-links per unit volume in a polymer network is defined as cross-linking density [37]. The polymer volume fraction, an ...
  38. [38]
    Competing Effects of Molecular Additives and Cross-Link Density on ...
    Nov 9, 2023 · We find that increasing the additive mass percentage m progressively decreases both the glass-transition temperature T g and the fragility of glass formation.
  39. [39]
    Influence of curing modes on crosslink density in polymer structures
    Results of this study suggest that polymerization with PD resulted in a lower crosslink density and gave rise to polymers with an increased susceptibility to ...
  40. [40]
    [PDF] Influence of monomer structure and catalyst concentration on ...
    Jul 31, 2024 · Taking into consideration the molecular struc- ture and the reaction pattern, there are five factors that influence the bond exchange rate: (i) ...
  41. [41]
    Curing behavior and impact of crosslinking agent variation in ...
    In general, as more crosslinking agent was added, the hardness increased, but this result can be explained by the difference between “crosslinking density” and ...
  42. [42]
    D2765 Standard Test Methods for Determination of Gel Content and ...
    Feb 19, 2024 · ASTM D2765 standard determines gel content and swell ratio of crosslinked ethylene plastics by extracting with solvents, using methods A, B, ...
  43. [43]
    NMR Studies on the Phase-Resolved Evolution of Cross-Link ...
    Dec 3, 2020 · Nuclear magnetic resonance (NMR) spectroscopy has grown to be a viable tool for probing molecular structure and dynamics. However, a rather ...
  44. [44]
    ESR Study of Peroxide-Induced Cross-Linking of High Density ...
    Peroxide-initiated cross-linking of high-density polyethylene (HDPE) was studied using an on-line electron spin resonance (ESR) technique.
  45. [45]
    Introduction to Dynamic Mechanical Analysis and its Application to ...
    DMA allows users to characterize the viscoelastic properties of the material such as storage modulus, loss modulus and tan δ. These properties help understand ...Missing: cross- | Show results with:cross-
  46. [46]
    In situ EPR and Raman spectroscopy in resin curing
    Jan 19, 2022 · Monitoring of polymerization by in situ EPR spectroscopy. The polymerization of styrene/BADGE-MA was monitored by in situ EPR spectroscopy ...
  47. [47]
    Comparison of techniques for determining crosslinking in silane ...
    In the determination of degree of crosslinking, the approach which has been widely used is based on ASTM D2765 (standard test methods for determination of gel ...
  48. [48]
    From structure to redox: the diverse functional roles of disulfides and ...
    Disulfide bonds have classically been shown to stabilize proteins by maintaining overall structure via intermolecular and intra-domain covalent bonds between ...
  49. [49]
    Intramolecular isopeptide bonds: protein crosslinks built for stress?
    These intramolecular isopeptide bonds form autocatalytically during protein folding, as the reacting groups are brought together in a hydrophobic environment.
  50. [50]
    Intramolecular Isopeptide Bonds Give Thermodynamic and ... - NIH
    Further investigations have shown that similar isopeptide bonds can be found as internal cross-links in other proteins. Examination of previously determined ...
  51. [51]
    A Highly Conserved Salt Bridge Stabilizes the Kinked Conformation ...
    Salt bridges, commonly formed by a combination of non-covalent interactions (hydrogen bonding and electrostatic interactions) between acidic and alkaline amino ...
  52. [52]
    Disulfide bonds in ER protein folding and homeostasis - PMC - NIH
    Disulfide bond formation in protein folding and oligomerization. Due to their covalent nature, disulfide bonds can have profound effects on the folding pathway ...
  53. [53]
    Native Disulfide Bond Formation in Proteins - PMC - NIH
    Native disulfide bond formation is critical for the proper folding of many proteins. Recent studies using newly identified protein oxidants, folding catalysts, ...
  54. [54]
    Control of blood proteins by functional disulfide bonds - PMC
    Protein disulfide bonds are the links between the sulfur atoms of 2 cysteine amino acids (the cystine residue) that form as proteins mature in the cell. These ...
  55. [55]
    Role of disulfide bonds in the structure and activity of human insulin
    Insulin contains two inter-chain disulfide bonds between the A and B chains (A7-B7 and A20-B19), and one intra-chain linkage in the A chain (A6-A11).
  56. [56]
    Advanced Glycation End-Products (AGEs): Formation, Chemistry ...
    Apr 12, 2022 · Advanced glycation end-products (AGEs) constitute a non-homogenous, chemically diverse group of compounds formed either exogeneously or endogeneously on the ...
  57. [57]
    The role of advanced glycation end products in aging and metabolic ...
    Sep 4, 2019 · Accumulation of advanced glycation end products (AGEs) on nucleotides, lipids and peptides/proteins are an inevitable component of the aging process.
  58. [58]
    A central role for dityrosine crosslinking of Amyloid-β in Alzheimer's ...
    Dec 18, 2013 · Examination of samples from AD patients revealed the presence of dityrosine linkages within amyloid plaques and its colocalisation with Aβ in ...
  59. [59]
    Tissue Transglutaminase, Protein Cross-linking and Alzheimer's ...
    The major proteinaceous component of the plaques is the extensively cross-linked β-amyloid (Aβ) with non-amyloid components comprising the core of the plaques ...
  60. [60]
    The evolving role of DNA inter-strand crosslinks in chemotherapy - NIH
    Mitomycin C ICLs cause minimal distortion to DNA. Psoralen ICLs cause moderate DNA unwinding. Cisplatin ICLs cause severe DNA bending and unwinding. Nitrogen ...
  61. [61]
    Mechanism of the formation of DNA–protein cross-links by antitumor ...
    Numerous studies show that cisplatin forms in DNA ∼90% intrastrand cross-links (CLs) between neighboring purine bases (1,2-GG or 1,2-AG intrastrand CLs) and ...
  62. [62]
    All You Need Is Light. Photorepair of UV-Induced Pyrimidine Dimers
    ... (UV-A). The direct absorption of UV by DNA leads mainly to the formation of pyrimidine dimers between adjacent pyrimidines in a DNA strand. The most frequent ...
  63. [63]
    Preparation of Stable Nitrogen Mustard DNA Inter Strand Crosslink ...
    Nitrogen mustards (NMs) react with two bases on opposite strands of a DNA duplex to form a covalent linkage, yielding adducts called DNA interstrand crosslinks ...<|separator|>
  64. [64]
    Cellular response to DNA interstrand crosslinks: the Fanconi anemia ...
    Apr 19, 2016 · Interstrand crosslinks (ICLs) are a highly toxic form of DNA damage. ICLs can interfere with vital biological processes requiring separation ...Origin Of Icls · The Fanconi Anemia Pathway · Phosphorylation
  65. [65]
    Mechanisms of interstrand DNA crosslink repair and human disorders
    May 1, 2016 · In this review, we briefly summarize the recent studies of ICL associated disorders and repair mechanism, with emphasis in the first incision of ICLs.
  66. [66]
    Regulation of DNA cross-link repair by the Fanconi anemia/BRCA ...
    The Fanconi anemia (FA) pathway is essential for the repair of DNA interstrand cross-links (ICLs), and a germline defect in the pathway results in FA, a cancer ...
  67. [67]
    Mitomycin: ten years after approval for marketing - PubMed
    Mitomycin was approved for marketing by the Food and Drug Administration in 1974 for use in gastric and pancreatic carcinomas when combined with other ...
  68. [68]
    Lignin Biosynthesis and Structure - PMC - NIH
    Lignin is able to readily copolymerize alternative units that derive from incomplete monolignol biosynthesis in plants with pathway perturbations.
  69. [69]
    Peroxidases Bound to the Growing Lignin Polymer Produce Natural ...
    Monolignols are oxidized in the cell wall by oxidative enzymes (peroxidases and/or laccases) to radicals, which then couple with the growing lignin polymer.
  70. [70]
    Lignin polymerization: how do plants manage the chemistry so well?
    Oct 22, 2018 · Lignin polymerization occurs via a combinatorial radical coupling process that is highly flexible in nature and allows plants to incorporate numerous lignin ...
  71. [71]
    The Role of the Lysyl Oxidases in Tissue Repair and Remodeling
    Lysyl oxidase (LOX) and four lysyl oxidase-like proteins are a group of enzymes capable of catalyzing cross-linking reaction of collagen and elastin.
  72. [72]
  73. [73]
    Recent Advancements and Challenges in Lignin Valorization
    The enzyme depolymerization of the lignin takes place in step 2, generating certain lignin-derived aromatics. The aromatics are hydrolyzed in step 3 through the ...
  74. [74]