A fibril is a small filament or fiber, typically threadlike in appearance, that constitutes a basic structural element in biological tissues and cellular components. These nanoscale structures, often measuring 5–100 nm in diameter depending on context, provide mechanical support, enable elasticity, and facilitate organization in everything from extracellular matrices to intracellular assemblies across diverse organisms.[1][2]In connective tissues, collagen fibrils represent a primary example, forming the major load-bearing elements of the extracellular matrix in multicellular animals ranging from echinoderms to vertebrates; their hierarchical assembly from triple-helical collagen molecules imparts tensile strength and resistance to deformation essential for tissue integrity.[3] Myofibrils, another prominent type, are cylindrical organelles within striated muscle cells composed of repeating sarcomere units—the most ordered macromolecular arrays in eukaryotic cells—that integrate actin and myosin filaments to drive contraction and force generation.[4][5]Beyond physiological roles, fibrils also feature in pathological processes, such as amyloid fibrils formed by misfolded proteins adopting a cross-β sheet architecture; these elongated aggregates, consisting of laterally associated protofilaments, underlie over 50 human diseases including Alzheimer's and type 2 diabetes by disrupting cellular function through toxicity and deposition.[6][7] Fibril formation generally involves self-assembly mechanisms influenced by environmental factors like pH and ionic strength, with implications for both normal development—such as wound healing via fibronectin fibrillogenesis—and therapeutic targeting of aberrant structures.[8][9]
Structure and Formation
Hierarchical Organization
Fibrils are elongated, nanoscale fibrous structures typically ranging from 3 to 300 nm in diameter depending on type, composed of biopolymers such as proteins or polysaccharides that self-assemble into higher-order architectures.[10][11] For example, cellulose microfibrils are 3-6 nm in diameter, amyloid fibrils 5-15 nm, and collagen fibrils 20-200 nm.[12][13][14] These structures exhibit high aspect ratios, often exceeding 100:1, with lengths extending into the micrometer scale, enabling them to serve as fundamental building blocks in biological materials.[11]The hierarchical organization of fibrils spans multiple scales, beginning at the primary level with linear sequences of amino acids in proteins or monosaccharides in polysaccharides. At the secondary level, these sequences fold into motifs such as α-helices or β-sheets in proteins and extended chains in polysaccharides. Tertiary structures emerge as protofilaments or subfibrils form through lateral associations of these motifs, while the quaternary level involves bundling of protofilaments into mature fibrils and further aggregation into macroscopic fibers.[15][16] This multi-level assembly imparts crystallinity and molecular orientation, which contribute to the material's anisotropy by aligning polymer chains along the fibril axis.[17]A representative example of this assembly is observed in protein fibrils, where staggered molecular packing creates periodic banding patterns, such as the 67 nm D-periodicity arising from the quarter-staggered arrangement of tropocollagen molecules.[18] These patterns result from the offset alignment of molecular units, enhancing structural regularity. Such hierarchical features influence mechanical properties, including stiffness and toughness, by distributing loads across scales.[17]Visualization of fibril hierarchy relies on techniques like atomic force microscopy (AFM), which resolves surface topography and periodicity at the nanoscale, and X-ray diffraction, which reveals internal crystalline order and banding patterns through scattering patterns.[19][20] AFM provides direct imaging of fibril contours and twists, while X-ray methods quantify long-range order in aligned samples.[21]
Biogenesis Processes
Fibril biogenesis involves self-assembly processes characterized by distinct nucleation, growth, and maturation phases, where molecular precursors organize into ordered fibrous structures through non-covalent interactions. Nucleation initiates assembly by forming stable oligomeric seeds from monomers, such as tropocollagen for collagen fibrils or β-sheet-rich cores for amyloid fibrils, overcoming an energy barrier via entropy-driven aggregation. Growth proceeds by sequential monomer addition to fibril ends, elongating the structure, while maturation stabilizes the fibril through lateral association and polymorphism refinement, often resulting in hierarchical architectures. These phases are primarily driven by hydrophobic interactions that bury non-polar residues, hydrogen bonding that aligns backbones (e.g., in cross-β sheets of amyloids or β-1,4-glucan chains of cellulose), and electrostatic forces that modulate charge complementarity between subunits.[22][23][24]Environmental conditions significantly influence fibril dimensions and kinetics during biogenesis. Factors such as pH, temperature, and ionic strength modulate intermolecular forces; for instance, neutral pH in the range of 6.9 to 8.0 promotes fibrillogenesis in collagen by optimizing electrostatic attractions,[25] while physiological temperatures around 37°C facilitate nucleation and assembly. Higher ionic strength screens charges, favoring lateral growth and thicker fibrils, as seen in collagen assemblies where divalent cations enhance diameter control. In cellulose microfibrils, plasma membrane dynamics indirectly respond to cellular environmental cues like microtubule orientation, guiding fibril deposition.[24]Enzymatic processes are crucial for precursor activation and polymerization in fibril formation. In collagen biogenesis, procollagen peptidases, including N- and C-proteinases like BMP-1/tolloid-like enzymes, cleave N- and C-terminal propeptides extracellularly, enabling tropocollagen molecules to initiate self-assembly into fibrils.[23] For cellulose microfibrils in plants and bacteria, glycosyltransferases within cellulose synthase complexes (e.g., CESA proteins forming rosette-shaped units in the plasma membrane) polymerize UDP-glucose into linear β-1,4-glucan chains, which spontaneously crystallize into microfibrils via interchain hydrogen bonding. These enzymatic steps ensure precise chain length and orientation, distinguishing biogenesis from purely physical aggregation.[26]In vitro fibrillogenesis typically occurs spontaneously under controlled conditions, mimicking self-assembly without cellular templates, as in collagen solutions neutralized to form D-periodic fibrils or amyloid peptides aggregating at acidic pH. In contrast, in vivo processes are templated by cellular machinery, such as fibronectin and integrins directing collagen nucleation at plasma membranes or bacterial secretion systems guiding curli amyloid assembly, ensuring spatially regulated and hierarchical outcomes. This templated approach integrates fibril formation with tissue architecture, differing from the isotropic growth often observed in lab settings.[23][24][22]Fibril polymerization motifs exhibit evolutionary conservation across kingdoms, reflecting ancient mechanisms for structural biomaterials, including in bacteria (e.g., curli amyloids and bacterial cellulose) and fungi (e.g., chitin fibrils). The cross-β architecture in amyloids, from bacterial curli to eukaryotic prions, shares self-templating nucleation-growth dynamics, suggesting a primordial role in prebiotic replication and biofilm formation. Similarly, cellulose synthase-mediated glucan polymerization is preserved from bacterial linear complexes to plant rosette assemblies, while collagen-like triple helices in metazoans echo simpler fibrous motifs in prokaryotes, highlighting convergent evolution in force-driven self-organization for mechanical support.[27][26][22]
Mechanical Properties
Intrinsic Mechanics
Isolated fibrils exhibit a range of tensile properties depending on their molecular composition, with stiff variants like cellulose microfibrils displaying Young's moduli in the range of 29-88 GPa, as measured in bacterial and microfibrillated forms using atomic force microscopy (AFM) and network analysis techniques.[28] In contrast, protein-based fibrils such as type I collagen show moduli from 3.75 to 11.5 GPa in dry conditions, determined via AFM nanoindentation on rat tailtendon samples, reflecting their semi-crystalline structure.[29] Ultimate tensile strengths for collagen fibrils vary from 47 to 580 MPa, with extensibility reaching up to 81% strain at failure in isolated mouse tailtendon fibrils tested under tension.[30][31] Elastin-like fibrils, characterized by their amorphous domains, demonstrate lower moduli around 1 MPa but high extensibility exceeding 100% strain, enabling reversible deformation in soft tissues.[32]Deformation in isolated fibrils occurs through distinct mechanisms that balance elasticity and failure. At low strains, molecular chains slide relative to one another, particularly in collagen where tropocollagen molecules exhibit intermolecular slippage at low cross-link densities, leading to ductile behavior.[33] Higher strains induce uncoiling of helical structures, such as in collagen's triple helix, transitioning to energy dissipation via viscoelastic relaxation in amorphous regions, where type I collagen fibrils show a fast relaxation time of 7 seconds and a slow one of 102 seconds, modeled using a two-time-constant Maxwell-Weichert framework.[34] Fracture typically initiates at defects, with brittle failure in highly cross-linked collagen (cross-link density β ≥ 40) resulting from tropocollagen rupture, while lower cross-linking promotes sliding and strain softening up to 103% elongation.[33]Size effects govern fibril strength through statistical scaling laws, where thinner fibrils exhibit higher tensile strength due to reduced defect probability, following Weibull statistics for failure. In bamboo fibers as a model for fibrillar systems, strength decreases with increasing diameter (196-584 μm) and length (20-60 mm), with a modified Weibull distribution incorporating diameter variation improving predictions of the scale parameter (403-599 MPa) and shape parameter (2-6).[35] This weak-link scaling implies that smaller cross-sections sample fewer flaws, elevating the characteristic strength σ₀, as validated in simulations and tensile tests where longer gauges reduce average fracture stress.[35] Micromechanical modeling, such as finite element representations of fibril bundles, further elucidates these effects by simulating defect distributions and predicting failure probabilities under uniaxial load.[36]Experimental characterization of intrinsic mechanics relies on high-resolution techniques to probe individual fibrils. AFM-based nanoindentation applies localized forces (up to 2 μN) to measure reduced moduli via the Oliver-Pharr method, revealing anisotropic responses in collagen with axial stiffness exceeding transverse by factors of 10.[29] Pulling tests using AFM cantilevers or MEMS devices enable direct tensile loading, quantifying stress-strain curves and viscoelastic creep in sea cucumber collagen fibrils at strains up to 20%.[34]Buckling analysis from AFM topography images assesses tensile moduli without direct pulling, as demonstrated on nanometer-scale collagen where release from substrates induces measurable deflection.[37] These methods, often combined with scanning electron microscopy for visualization, provide data on isolated fibrils while minimizing substrate interactions.Thermal and environmental factors significantly modulate fibril stiffness. Elevated temperatures from 25°C to 45°C reduce the Young's modulus of type I collagen fibrils by destabilizing intermolecular bonds, as observed in torsion pendulum tests.[38]Hydration softens fibrils dramatically; immersion in aqueous media decreases collagen modulus by up to three orders of magnitude compared to dehydrated states, attributed to water bridges weakening peptide interactions, measured via AFM force-volume analysis on bovine Achilles tendon samples.[39] Dehydration, conversely, increases stiffness by 30% over days, highlighting water's role in modulating viscoelastic energy dissipation in amorphous domains.[29]
Role in Biomaterials
Fibrils serve as critical reinforcement phases within biomaterial matrices, facilitating efficient load transfer that enhances overall toughness through mechanisms such as crack deflection and fibril pull-out. In composite structures like mineralized collagen, staggered fibril arrangements allow stresses to distribute across the matrix-fibril interface, where pull-out dissipates energy during fracture by requiring additional work to extract embedded fibrils. This reinforcement is evident in hierarchical designs where fibrils bridge cracks, preventing rapid propagation and increasing fracture resistance by up to several fold compared to unreinforced matrices. Similarly, crack deflection at fibril boundaries redirects propagating flaws, promoting tortuous paths that absorb more energy and improve ductility in otherwise brittle composites.[40][41][42]The alignment of fibrils introduces anisotropy and directionality to biomaterials, enabling tailored directional strength that aligns with physiological loading demands, as seen in tendon-like composites. Aligned fibril orientations create preferential stiffness along the fiberaxis, with tensile moduli often exceeding 1 GPa in the longitudinal direction while remaining compliant transversely, thus optimizing load-bearing without bulk rigidity. This directional reinforcement arises from fibril-matrix shear interactions that transmit forces unidirectionally, mimicking natural tissues where misalignment reduces peak strength by 50% or more. In engineered scaffolds, such anisotropy supports cell alignment and tissueintegration, enhancing functional performance under uniaxial stresses.[43][44][45]Fibrils contribute to fatigue resistance and durability in biomaterials under cyclic loading, exhibiting hysteresis loops that enable energy storage and dissipation without permanent deformation. In elastic fibril networks like those inspired by resilin, near-complete recovery (over 95%) occurs after thousands of cycles due to reversible conformational changes, with hysteresis providing damping to mitigate fatigue failure. This cyclic resilience stems from interfibrillar sliding and matrix yielding, which distribute strains and prevent localized crack initiation, allowing biomaterials to withstand millions of load cycles at strains up to 300%. Such properties are vital for dynamic applications, where unrecovered energy loss below 10% per cycle ensures long-term integrity.[46][47][48]In bone biomaterials, fibril mechanics at the collagen-mineral interface drive hardness by coupling organic flexibility with inorganic rigidity, where mineral platelets within fibrils increase compressive strength to approximately 150 MPa while maintaining toughness. Conversely, in plant cell wall biomaterials, fibril networks confer flexibility through microfibril reorientation and sliding, enabling reversible extension under tensile loads up to 10% strain without rupture. These examples highlight how fibril contributions scale to bulk properties, balancing stiffness and compliance.[49][50][51][52][53]Finite element analysis (FEA) models of fibril-matrix interactions provide predictive insights into bulk biomaterial properties by simulating load transfer and deformation at multiple scales. These models represent fibrils as orthotropic elements embedded in viscoelastic matrices, revealing stress concentrations at interfaces that inform design for enhanced toughness, with predictions matching experimental moduli within 10-20%. By varying fibril volume fractions and alignments, FEA elucidates how mineral content modulates overall stiffness, aiding the optimization of synthetic composites for biomedical use.[54][55][56]
Fibrils in Animals
Collagen Fibrils
Collagen represents the predominant fibril-forming protein in animal extracellular matrices, with 28 distinct types identified in humans, each characterized by unique molecular compositions and tissue distributions.[57] Among these, the fibril-forming types I, II, and III are the most abundant and critical for structural integrity; type I predominates in skin, bone, and tendons, providing robust tensile support; type II forms the primary scaffold in cartilage; and type III contributes to the flexibility of blood vessels, skin, and hollow organs.[58] These types assemble into fibrils that vary in diameter from 50 to 200 nm, enabling tissue-specific mechanical adaptations.[59]The assembly of collagen fibrils begins with tropocollagen molecules, rigid triple-helical structures approximately 300 nm in length and 1.5 nm in diameter, which self-assemble extracellularly through a quarter-stagger arrangement.[60] In this process, molecules overlap with an axial stagger of about 67 nm, creating a characteristic D-period banding pattern visible under electronmicroscopy, where gap and overlap zones alternate along the fibril length.[61] This hierarchical organization, influenced by cellular secretion and environmental cues during biogenesis, results in staggered fibrils that pack into larger fibers, optimizing load distribution in tissues.[62]In tendons, collagen fibrils, primarily type I, confer exceptional tensile strength, with ultimate stresses reaching up to 100 MPa, essential for transmitting forces from muscle to bone.[63] Their resilience arises from hydration-dependent swelling, where water interactions with proteoglycans and the fibrillar matrix allow reversible deformation and energy dissipation under cyclic loading.[64] Disruptions in fibril assembly, such as mutations in collagen genes, underlie disorders like Ehlers-Danlos syndrome, where abnormal fibril diameters lead to tissue fragility and hyperelasticity.[59]Fibril maturation involves cross-linking that enhances stability and stiffness; enzymatic cross-links, catalyzed by lysyl oxidase, form early during development by oxidizing lysine residues to create covalent bonds between tropocollagen molecules.[65] Non-enzymatic advanced glycation end-products (AGEs) accumulate over time, particularly with aging or diabetes, further stiffening fibrils by forming additional intermolecular bridges that reduce flexibility but increase resistance to enzymatic degradation.[66]
Cytoskeletal and Elastic Fibrils
Cytoskeletal fibrils play essential roles in cellular motility and structural support within animal cells, primarily through the actin-myosinsystem. Actinmicrofilaments, approximately 7 nm in diameter, consist of polymerized globular actin (G-actin) subunits that form dynamic thin filaments, while myosin thick filaments, about 15 nm in diameter, interact with these actin filaments to generate force.[67] These components assemble into sarcomeres, the basic contractile units of muscle cells, where the sliding filament theory describes muscle contraction: myosin heads bind to actin, undergo a power stroke powered by ATP hydrolysis, and pull the actin filaments toward the sarcomere center, shortening the muscle fiber.[67] This mechanism enables rapid and controlled motility, such as in skeletal and cardiac muscle.[67]Elastic fibrils contribute to tissue resilience and reversible deformation, distinct from rigid structural elements. Elastin microfibrils, assembled on scaffolds of fibrillin proteins, form extensible networks in connective tissues that allow up to 150% strain before yielding, enabling repeated stretching and recoil without permanent damage.[68] These structures provide long-range elasticity in dynamic environments like skin and blood vessels.[68] Complementing this, keratin intermediate filaments, approximately 10 nm in diameter, form rope-like bundles in epithelial cells, conferring mechanical resilience against shear and tensile stresses to maintain tissue integrity during movement and injury.[69]Specialized elastic proteins exemplify extreme adaptability in animal systems. Resilin, composed of beta-turn-rich polypeptide chains with glycine-proline repeats, functions in insect cuticles as a highly compliant elastomer, exhibiting a Young's modulus of approximately 1 MPa that supports energy storage and release in jumping and wing movements.[70] Similarly, dragline spider silk derives its exceptional toughness from beta-sheet nanocrystals embedded in an amorphous matrix, achieving tensile strengths up to 1 GPa while balancing strength and extensibility for prey capture and locomotion.[71]These fibrils exhibit dynamic functions critical for cellular and tissue adaptability. Cytoskeletal remodeling occurs through rapid actin polymerization at the filament plus end and depolymerization at the minus end, driven by actin-binding proteins like Arp2/3 and cofilin, allowing cells to reorganize their architecture in response to signals for migration or division.[72] In vascular tissues, elastic recoil in arteries relies on elastin networks that store and release energy during the cardiac cycle, damping pressure fluctuations and maintaining blood flow.[73]Disruptions in elastic fibril components lead to significant pathologies. Mutations in the fibrillin-1 gene impair microfibril assembly, causing Marfan syndrome, a connective tissue disorder characterized by weakened aortic walls and reduced elasticity, increasing the risk of aneurysms and dissections.[74]
Fibrils in Plants
Cellulose Microfibrils
Cellulosemicrofibrils are the primary structural components of plant cell walls, consisting of parallel β-1,4-linked glucan chains that assemble into crystalline bundles.[75] Each microfibril is estimated to comprise 18 to 24 glucan chains in a core-shell structure, with recent studies suggesting a 24-chain model for wood cellulose in seed plants; they have widths typically 3-5 nm and lengths extending up to several microns across the cell wall.[75][76] This arrangement forms a highly ordered Iβ crystalline polymorph, where hydrogen bonding between chains creates regions of high crystallinity, typically 40-70% overall, providing exceptional tensile strength.[77] The surface chains are more accessible and less ordered, allowing interactions with hemicelluloses like xyloglucan to form a composite matrix.[77]The number of glucan chains per microfibril remains a topic of debate, with models supporting 18 chains based on scatteringdata and synthase complexarchitecture, while structural analyses indicate 24 chains in a core-shell configuration for certain plants.[77][76]Synthesis of cellulose microfibrils occurs at the plasma membrane through rosette-shaped cellulose synthase complexes (CSCs), which are hexameric assemblies of cellulose synthase A (CESA) proteins.[78] These rosette terminal complexes extrude nascent glucan chains simultaneously, polymerizing up to 1,000 glucose units per minute per chain as the complex moves bidirectionally along the membrane.[79] Cortical microtubules guide this process by directing CSC trajectories via the cellulose synthase-interacting protein 1 (CSI1), ensuring microfibril deposition aligns with microtubule orientation and influences cellshape.[78] Disruption of microtubule-CSC linkage, as seen in CSI1 mutants, leads to misaligned microfibrils and reduced cellulose content.[79]In primary cell walls of growing cells, cellulose microfibrils provide flexibility by forming a transverse or oblique network that resists turgor pressure while permitting expansion.[80] Their orientation, controlled by microtubules, dictates the direction of cell elongation, with transverse alignment promoting longitudinal growth.[81] In secondary cell walls, microfibrils are deposited in thicker, more parallel arrays, enhancing rigidity and mechanical support for vascular tissues like xylem.[81] This reinforcement is crucial for load-bearing functions in mature plants, where microfibril alignment correlates with tissue stiffness.[82]While structural cellulose microfibrils predominate in land plants, algae exhibit variations such as paramylon and chrysolaminarin, which are β-1,3-glucans serving primarily as storage polysaccharides rather than structural elements.[83] Paramylon, found in euglenoids like Euglena gracilis, accumulates in membrane-bound granules as a linear β-1,3-glucan, distinct from the β-1,4-linked chains of true cellulose microfibrils.[84] Chrysolaminarin, a soluble β-1,3/1,6-glucan in diatoms and brown algae, functions similarly for energy storage under nutrient-rich conditions, contrasting with the insoluble, crystalline role of cellulose in cell wall architecture.[83]Environmental adaptations involve modifications to microfibril properties for enhanced drought resistance, such as increased cellulose content and crystallinity in tolerant varieties.[85] In upland rice, the DROT1 gene promotes thicker sclerenchyma cell walls with higher microfibril crystallinity under water deficit, improving mechanical support and water retention without compromising growth.[85] This adaptation strengthens vascular bundles, reducing wilting and enhancing survival, as evidenced by greater biomass in DROT1-overexpressing lines.[85]
Starch and Wood Fibrils
Starch in plants is stored in semi-crystalline granules composed mainly of amylopectin branched chains and amylose linear polymers of α-1,4-linked D-glucose units, serving as a compact form for energy storage in tissues such as tubers and seeds.[86] Amylose, typically 15-35% of granule mass, can adopt a left-handed single helix conformation, enabling dense packing and efficient enzymatic mobilization during germination or stress responses.[87] V-amylose forms helical inclusion complexes with ligands like fatty acids, which can occur in native starch and influence digestibility.[88]In wood, lignocellulosic composites integrate cellulose microfibrils within a matrix of hemicelluloses and lignin, forming multilayered secondary cell walls designated as S1, S2, and S3 layers that confer anisotropic mechanical properties.[89] The S2 layer, the thickest, features tightly wound helical microfibrils at low angles (5-30°) relative to the cell axis, embedded in a hemicellulose network that cross-links with lignin for enhanced rigidity and hydrophobicity.[90] This hierarchical organization, with lignin comprising 20-30% of the dry mass, seals the polysaccharide scaffold post-deposition, preventing microbial degradation while allowing radial expansion during growth.[91]Mechanically, the helical winding of microfibrils in wood's S2 layer provides superior compression resistance, with longitudinal strengths reaching up to 50 MPa in mature tissues, enabling trees to withstand wind loads and self-weight.[92] In starch granules, hydration induces reversible swelling, particularly in tubers like potatoes, where water uptake expands the amorphous regions significantly at temperatures below gelatinization (around 50-60°C), facilitating controlled energy release.[93]Biosynthesis of amylose occurs via granule-bound starchsynthase (GBSS), which processively elongates α-1,4-glucan chains within developing granules using ADP-glucose as the substrate, ensuring linear molecules integrate into the semi-crystalline matrix.[94] Lignification in wood follows cellulose and hemicellulose deposition, mediated by class III peroxidases that oxidize monolignols (coniferyl, sinapyl, and p-coumaryl alcohols) in the cell wallapoplast, coupling them into a branched polymer that infiltrates microfibril interstices.[95]
Applications and Biomimicry
Biomimetic Designs
Biomimetic designs draw inspiration from the hierarchical organization and mechanical properties of natural fibrils to engineer advanced materials with enhanced functionality in tissue engineering, robotics, and surface technologies. These synthetic constructs replicate the nanoscale architecture of biological fibrils, such as their aligned structures and periodic banding, to achieve superior performance in load-bearing, adhesion, and environmental responsiveness. By mimicking the self-assembly and orientation seen in collagen, silk, and cellulose fibrils, researchers have developed materials that promote cell interactions, provide tunable mechanics, and exhibit specialized surface behaviors.Collagen-mimetic electrospun nanofiber scaffolds are widely used in tissue engineering to replicate the native extracellular matrix, particularly the characteristic 67-nm D-banding pattern that facilitates cell adhesion and proliferation. These scaffolds, produced from type I collagen via electrospinning, form fibers that mimic the hierarchical assembly of natural collagen fibrils, enhancing biocompatibility and guiding tissue regeneration in applications like skin and vascular grafts. The D-banding promotes integrin-mediated cell attachment, improving outcomes in wound healing models compared to non-banded synthetic fibers.Recombinant silk-elastin-like proteins have been engineered into hydrogels that emulate the elasticity and toughness of natural silk and elastin fibrils, offering tunable mechanical properties in the range of 0.1-10 MPa for soft robotics and biomedical devices. These protein-based materials self-assemble into fibrillar networks upon stimuli like temperature or light, enabling reversible deformation and shape morphing essential for actuators in soft robots. For instance, silk-elastin copolymers cross-linked via enzymatic sites exhibit viscoelastic behavior that supports dynamic actuation while maintaining biocompatibility for tissue interfacing.Cellulose-inspired bacterial nanocellulose (BNC) materials serve as effective wound dressings due to their aligned nanofibril networks, which provide robust barrier properties against microbial penetration while allowing moisture vapor transmission. The ribbon-like, aligned fibrils in BNC form a porous yet impermeable scaffold that absorbs exudates and maintains a hydrated environment conducive to epithelialization, outperforming traditional dressings in chronic wound management. Functionalization of these aligned structures further enhances antibacterial activity without compromising the mechanical integrity derived from the fibrillar orientation.Self-cleaning surfaces biomimetic of aligned hydrophobic fibrils, such as those in plant waxes, achieve the lotus effect by reducing contact anglehysteresis to promote droplet rolling and contaminant removal. Hierarchical nanofibril arrays on substrates like modified cellulose films yield superhydrophobic properties with water contact angles exceeding 150°, minimizing adhesion of dirt and bacteria through the Cassie-Baxter state stabilized by fibril alignment. These designs, inspired by tubular wax fibrils on lotus leaves, have been applied in antifouling coatings for medical devices and textiles.Fabrication techniques like 3D printing and self-assembly of peptide amphiphiles enable the creation of hierarchical fibril mimics with precise control over nanoscale to macroscale organization. Peptide amphiphiles, with hydrophobic tails and bioactive heads, self-assemble into nanofibers that bundle into fibrils under physiological conditions, forming scaffolds that replicate collagen hierarchy for tissue engineering. Integrating these with 3D printing allows extrusion of aligned fibrillar hydrogels, tuning porosity and mechanics for customized implants while preserving bioactivity.
Pathological and Therapeutic Contexts
Amyloid fibrils, characterized by a cross-β sheet structure, play a central role in various neurodegenerative diseases, including Alzheimer's disease where they form β-amyloid (Aβ) plaques that contribute to neuronal toxicity through pore-forming mechanisms in cell membranes.[96] In Alzheimer's, brain-derived Aβ fibrils exhibit polymorphic structures with a conserved cross-β core, leading to synaptic dysfunction and cell death.[97] Similarly, prion fibrils in diseases like Creutzfeldt-Jakob syndrome feature a cross-β architecture, particularly in the sequence region 160–220, enabling self-propagation and neurotoxicity.[98]Fibrin fibrils are essential for normal blood clotting but become pathological in thrombotic disorders; thrombin cleaves fibrinopeptide A from fibrinogen, initiating polymerization into double-stranded fibrils that assemble into branched meshes, and excessive formation contributes to thrombus development by incorporating platelets and leukocytes.[99] In conditions like deep veinthrombosis, altered fibrin network density and stability impair fibrinolysis, increasing occlusion risk.[100]Collagen fibril abnormalities, such as irregular diameters observed in the gravis form of Ehlers-Danlos syndrome, result in connective tissue fragility and vascular complications.[101]Therapeutic strategies targeting pathological fibrils include stabilizers like tafamidis, which binds transthyretin tetramers with high affinity to prevent dissociation, misfolding, and amyloid fibril formation in transthyretin amyloidosis, thereby slowing neurologic progression.[102] Acoramidis, another transthyretin stabilizer, was approved by the U.S. Food and Drug Administration in 2024 for transthyretin cardiac amyloidosis (ATTR-CM).[103] For AL amyloidosis, anselamimab, an investigational fibril depleter, showed clinical benefits in phase 3 trials as of July 2025.[104] In Alzheimer's, lecanemab-irmb (anti-Aβ antibody) demonstrated continued benefits in reducing amyloid plaques and slowing cognitive decline through four years of treatment as of July 2025.[105] For regenerative applications, dense fibrillar collagen-based hydrogels mimic the osteoid matrix by providing oriented scaffolds that support osteoblast mineralization and bone formation.[106] Conversely, anticoagulants and fibrinolytics disrupt fibrin fibril networks in thrombosis to restore vascular flow.Diagnostic approaches leverage specific imaging for fibril detection; thioflavin-T binds the cross-β grooves of amyloid fibrils, inducing enhanced fluorescence for histological identification in tissues.[107]Ultrasound, particularly high-resolution duplex methods, enables real-time 3D assessment of fibrin clot volume and thrombolysis efficacy in thrombotic lesions.[108]Emerging post-2020 research includes in vivo CRISPR-Cas9 editing, such as NTLA-2001 (nexiguran ziclumeran), which in phase 1/2 trials for hereditary transthyretin amyloidosis with polyneuropathy achieved a sustained mean serum transthyretin reduction of approximately 90% through month 24 as of September 2025; however, phase 3 trials were placed on clinical hold in October 2025 due to serious adverse events including liver toxicity.[109][110]