Fact-checked by Grok 2 weeks ago

Protein secondary structure

Protein secondary structure refers to the local conformation of the polypeptide backbone in a protein, characterized by repeating patterns such as alpha-helices and beta-sheets, which are stabilized by hydrogen bonds between the carbonyl oxygen and hydrogen atoms of the peptide backbone. These structures represent the first level of folding beyond the primary sequence and are crucial for determining the overall three-dimensional architecture and function of proteins. The concept of secondary structure was first proposed in 1951 by , Robert Corey, and Herman Branson, who described the alpha-helix and beta-sheet as fundamental motifs based on stereochemical constraints and hydrogen bonding patterns. The most common secondary structural elements include the alpha-helix, a right-handed coiled structure with approximately 3.6 amino acid residues per turn and a pitch of 5.4 Å, where hydrogen bonds form between the carbonyl group of residue i and the amide nitrogen of residue i+4. In this motif, the phi (Φ) and psi (Ψ) dihedral angles are typically around -57° and -47°, respectively, allowing side chains to project outward from the helix axis. Alpha-helices are prevalent in both globular and fibrous proteins, such as keratin, and contribute to the stability of transmembrane segments. Beta-pleated sheets consist of two or more beta-strands—extended polypeptide segments—aligned either in or antiparallel orientations and linked by bonds between backbone atoms of adjacent strands. In antiparallel beta-sheets, the strands run in opposite directions with angles of approximately Φ = -139° and Ψ = 135°, while parallel sheets have angles around Φ = -119° and Ψ = 113°; both configurations result in a pleated appearance with side chains alternating above and below the plane. These sheets often form the core of globular proteins and can assemble into more complex forms like beta-barrels in proteins, such as porins. In addition to helices and sheets, proteins feature irregular secondary elements like beta-turns and loops, which connect the regular motifs and allow the chain to reverse direction or adopt flexible conformations without extensive hydrogen bonding networks. These regions, often involving specific like or , facilitate the packing of secondary structures into higher-order folds. Disruptions in secondary structure, such as excessive beta-sheet formation, are implicated in protein misfolding diseases including Alzheimer's and disorders.

Definition and Fundamentals

Definition

Protein secondary structure refers to the local conformation of the polypeptide backbone in a protein, characterized by regular, repeating patterns stabilized primarily by hydrogen bonds between the carbonyl oxygen and amide hydrogen atoms within the backbone, independent of side-chain interactions. This level of structure emerges from the primary sequence and serves as a fundamental building block for the protein's overall three-dimensional architecture. The concept of secondary structure was pioneered by and Robert Corey, who in 1951 proposed the as a coiled configuration where hydrogen bonds form between residues separated by three intervening along the chain, and the beta pleated sheet as a layered arrangement of extended strands linked laterally by interchain hydrogen bonds. These structures, along with less regular elements such as turns and loops, allow the polypeptide to fold compactly while maintaining stability through non-covalent interactions. In the alpha helix, the backbone adopts a right-handed spiral with 3.6 residues per turn and a of approximately 5.4 , enabling efficient packing in both soluble and proteins. sheets, in contrast, feature extended chains in a pattern, either parallel (strands running in the same direction) or antiparallel (opposite directions), forming the core of many globular proteins like enzymes. Turns and loops, often involving 3-5 residues, connect these motifs and frequently occur at the protein surface, facilitating flexibility and interactions with other molecules. Secondary structure elements are essential for protein function, as they dictate folding pathways, stability, and the positioning of functional groups, with disruptions leading to diseases such as .

Historical Background

The early investigations into protein secondary structure relied heavily on diffraction studies of fibrous proteins. In the 1930s, William T. Astbury and his collaborators at the analyzed keratin fibers from and , identifying distinct diffraction patterns they termed the "alpha" and "beta" forms, corresponding to unstretched and stretched states, respectively. These observations suggested regular, repeating structural features in proteins but lacked atomic-level details due to the limitations of the technology at the time. A breakthrough came in 1951 when , Robert B. Corey, and Herman R. Branson at the proposed specific atomic models for protein secondary structures using wire-and-cardboard model-building techniques informed by known covalent bond lengths, angles, and van der Waals radii. In April, they described the α-helix, a right-handed coil stabilized by intra-chain hydrogen bonds between the carbonyl oxygen of residue i and the amide hydrogen of residue i+4, with 3.7 residues per turn and a pitch of 5.4 Å. Just a month later, in May, and Corey introduced the β-pleated sheet, a layered configuration where adjacent polypeptide chains form inter-chain hydrogen bonds, creating extended, pleated structures observed in the beta form of silk fibroin and . These models were derived without full crystallographic data for entire proteins, relying instead on stereochemical feasibility and consistency with Astbury's diffraction patterns. The concept of secondary structure as a distinct level of protein organization was formalized in 1952 by Danish biochemist Kaj Ulrik Linderstrøm-Lang during his lectures at Stanford University, where he distinguished it from primary (amino acid sequence), tertiary (overall fold), and later quaternary (multi-subunit assembly) structures. This framework gained empirical validation in the late 1950s with the first X-ray crystal structures of proteins, such as myoglobin solved by John C. Kendrew in 1958, which prominently featured α-helices as predicted by Pauling. These developments laid the foundation for understanding how local hydrogen-bonding patterns dictate protein folding and function.

Types of Secondary Structures

Alpha Helix

The alpha helix is a prevalent in protein secondary structure, consisting of a right-handed helical coil formed by the polypeptide backbone, where each backbone group (N-H) forms a with the (C=O) of the four residues earlier in the sequence. This intra-chain hydrogen bonding pattern stabilizes the structure, with bond lengths typically around 2.8–3.0 between the donor and acceptor atoms. The configuration ensures that all residues are stereochemically equivalent, with side chains projecting outward from the axis. Proposed in 1951 by , Robert Corey, and Herman Branson through model-building informed by X-ray diffraction data from and simple peptides, the was one of the first regular secondary structures predicted for proteins, predating experimental confirmation. Their work identified it as a 3.7-residue helix (later refined to 3.6), emphasizing its compatibility with the planar peptide bonds and van der Waals radii of atoms in the backbone. This prediction was validated shortly after by of proteins like , where alpha helices comprise about 75% of the structure. Geometrically, the alpha helix features 3.6 residues per complete turn, a helical (advance along the axis per turn) of 5.4 , and a translation of 1.5 per residue, resulting in a tightly coiled approximately 1–2 in depending on side-chain bulk. The characteristic phi () and psi () backbone dihedral angles are approximately -57° and -47°, respectively, placing the structure within the allowed region of the for non-glycine residues. These parameters arise from optimizing geometry and minimizing steric clashes, with slight variations observed in real proteins due to side-chain interactions or environmental factors. In soluble proteins, alpha helices account for roughly 30% of all residues, serving as scaffolds for tertiary structure formation through packing against other helices or sheets, often mediated by hydrophobic interactions between nonpolar s. Certain , such as , , and , have high helix-forming propensities due to their ability to stabilize the core via van der Waals contacts, while disrupts helices by introducing kinks owing to its cyclic . Helices also contribute to functional roles, such as in DNA-binding proteins like motifs or in proteins where amphipathic helices span bilayers. For instance, in , alpha helices form the oxygen-binding pockets, enabling cooperative function. The stability of alpha helices is influenced by both local sequence and global context; isolated helices in short peptides are marginally stable in but are reinforced in proteins by capping interactions at the ends (e.g., or serine forming additional hydrogen bonds) and like salt bridges between charged side chains (i, i+4 positions). Thermal denaturation studies show helix melting temperatures around 50–60°C for model peptides, underscoring the cooperative nature of unfolding. Variants like the pi-helix (4.4 residues per turn) are rarer, comprising less than 1% of helical structures, as they accommodate suboptimal hydrogen bonding geometries.

Beta Sheet

The β-sheet, also known as the β-pleated sheet, is a prevalent form of regular secondary structure in proteins, composed of two or more β-strands—extended segments of polypeptide chain—aligned adjacently and stabilized by a network of hydrogen bonds between their backbone carbonyl oxygen and amide hydrogen atoms. This configuration allows for efficient packing of the polypeptide backbone, with the side chains projecting alternately above and below the plane of the sheet. The structure was originally proposed by and Robert B. Corey in 1951 as a "pleated sheet" layer of polypeptide chains, where the pleats arise from the zigzag arrangement of the peptide planes, enabling optimal hydrogen bonding between adjacent chains in an extended conformation. β-Strands in a sheet typically span 5 to 10 residues, adopting a nearly fully extended backbone with characteristic Ramachandran angles of φ ≈ −140° and ψ ≈ +130°; these angles position the backbone for interstrand hydrogen bonding while minimizing steric clashes. The hydrogen bonds form a ladder-like pattern across strands, with typical N–H···O distances of about 2.9 and near-linear geometry (N–H···O angles close to 180°). In practice, β-sheets are rarely flat; they exhibit a right-handed twist of approximately 30° per residue due to favorable side-chain orientations and backbone rigidity, which enhances and allows the sheet to curve into motifs like β-barrels or β-propellers. β-Sheets occur in two primary topologies: antiparallel, where adjacent strands run in opposite directions ( to ), and , where strands run in the same direction. In antiparallel sheets, bonds alternate directly between paired carbonyl and groups, resulting in more uniform and stronger bonds compared to the slightly distorted, wider-spaced bonds in sheets; this makes antiparallel arrangements generally more stable and common in isolated sheets. sheets, often embedded within larger mixed topologies, tend to require longer connecting loops and are frequently buried in protein cores to shield their less optimal bonding from . preferences differ markedly between the two: antiparallel strands favor hydrophobic residues like and for tight packing, while strands show a propensity for and aspartate, which can form additional side-chain bonds to compensate for backbone irregularities. Distortions such as β-bulges—insertions of extra residues that disrupt the regular hydrogen-bonding pattern—commonly occur to accommodate sequence variations or functional needs, allowing sheets to bend or adjust without losing overall integrity. In fibrous proteins like from , antiparallel β-sheets dominate, with stacked crystalline layers of Gly-Ala repeats providing exceptional tensile strength (up to 1 GPa) due to the dense hydrogen-bond network and intersheet van der Waals interactions. In globular proteins, β-sheets often form the core of structural domains, as seen in the immunoglobulin fold (PDB ID 1icf), where antiparallel sheets create a β-sandwich stabilized by a hydrophobic interface, or in (PDB ID 1tim), featuring a parallel β-sheet barrel surrounded by α-helices. These motifs underscore the β-sheet's role in mediating , stability, and interactions; aberrant β-sheet aggregation, as in , is implicated in diseases like Alzheimer's, where cross-β structures propagate via templated misfolding.

Turns and Loops

Turns and loops represent irregular regions of protein secondary structure that lack the repetitive hydrogen bonding patterns of alpha helices and beta sheets, instead serving primarily to connect these regular elements and enable the overall three-dimensional folding of the polypeptide chain. These motifs are essential for reversing the direction of the backbone, accommodating spatial constraints, and contributing to protein and function, often comprising hydrophilic residues exposed to . In typical globular proteins, turns and loops account for approximately 20-30% of residues, with their flexibility allowing dynamic conformational changes critical for enzymatic activity and molecular recognition. Beta-turns, the most common type of turn, involve four consecutive residues (i to i+3) where the chain direction reverses sharply, defined by a Cα(i) to Cα(i+3) distance of less than 7 Å and often stabilized by a between the carbonyl oxygen of residue i and the hydrogen of residue i+3. First proposed by Venkatachalam in 1968 through stereochemical modeling of units, beta-turns were as a distinct secondary structure alongside helices and sheets, with initial classifications into types I, II, and III based on backbone angles (φ, ψ). Subsequent refinements by Richardson in 1981 expanded this to eight canonical types (I, I', II, II', VIa, VIb, IV, VIII), characterized by specific φ and ψ values; for example, type I features φ2 ≈ -60°, ψ2 ≈ -30°, φ3 ≈ -90°, ψ3 ≈ 0°, while type II has a cis-like preference at position 3. Type I and type IV are the most prevalent, occurring in about 38% and 32% of beta-turns, respectively, and they frequently feature , , or at key positions due to their ability to adopt strained conformations. Other turn types include gamma-turns, which span three residues with a Cα(i) to Cα(i+2) distance under 7 and a from i to i+2, often involving classic (φ ≈ 70°, ψ ≈ -60°) or inverse (φ ≈ -70°, ψ ≈ 60°) variants stabilized by residues like or serine. Pi-turns, encompassing five residues, are rarer and feature a 4→1 , while alpha-turns (five residues) and wider motifs like beta-hairpin loops bridge strands in beta sheets. These tight turns are crucial for compact folding, with statistical analyses showing they cluster at protein surfaces and interfaces, influencing stability through side-chain interactions. Loops, in contrast, are longer irregular segments (typically 5-30 residues) that connect distant secondary structure elements without strict hydrogen bonding patterns, often adopting variable conformations classified as "" in secondary structure assignment schemes like DSSP. A prominent subclass, omega (ω) loops, consists of 6 or more residues forming a rigid, loop-shaped structure with ends separated by 5-10 in space despite sequence contiguity up to 18 residues apart, as defined by Leszczynski and Rose in through analysis of 67 high-resolution protein structures revealing 270 such motifs. These loops, frequently surface-exposed and hydrophilic, function as independent folding units and are enriched in functional sites, such as active centers in enzymes like where they contribute to substrate binding. Overall, turns and loops enhance protein versatility, with mutations in these regions often linked to diseases like due to disrupted folding or interactions.

Classification Systems

DSSP Classification

The Dictionary of Secondary Structure of Proteins (DSSP) is an algorithm for assigning secondary structure elements to the residues in a protein based on its three-dimensional atomic coordinates, rather than predicting structure from sequence. Developed by Wolfgang Kabsch and Chris Sander in 1983, DSSP analyzes the pattern of hydrogen bonds within the protein backbone to identify structural motifs, approximating intuitive notions of secondary structure through objective criteria. It processes PDB files or equivalent coordinate data to classify residues into one of originally eight states, expanded to nine in DSSP 4, focusing on hydrogen-bonded and geometrical features such as backbone dihedral angles and bond distances. This method has become the de facto standard for secondary structure annotation in , with over 400 citations annually and integration into major databases like the (PDB). At its core, the DSSP algorithm detects hydrogen bonds using an electrostatic model that considers the donor-acceptor distance (typically <3.0 Å for the hydrogen to acceptor) and the angle between the donor-hydrogen and hydrogen-acceptor vectors (deviation <30° from linearity). These bonds define primary elements like alpha-helices (sequential i to i+4 bonds) and beta-sheets (antiparallel or parallel bridges between strands). Residues not involved in such bonds are classified based on local geometry, such as turns or bends. The assignment is residue-specific, allowing for nuanced descriptions beyond binary helix/sheet categories, and it handles irregularities like distorted helices or isolated bridges. This hydrogen-bond-centric approach ensures consistency across protein structures, though it can be sensitive to coordinate resolution and refinement quality. DSSP classifies residues into the following nine secondary structure types (eight original plus one new in DSSP 4), each denoted by a single-letter code:
CodeStructure TypeDescription
Hα-helixRight-handed coil with 3.6 residues per turn, stabilized by i to i+4 hydrogen bonds.
G3₁₀-helixTighter helix with 3.0 residues per turn, i to i+3 bonds, often at helix ends.
Iπ-helixWider helix with 4.4 residues per turn, i to i+5 bonds, less common.
EExtended strandPart of a β-sheet, involved in extended hydrogen-bonded ladders.
Bβ-bridgeIsolated β-pair without full sheet formation.
THydrogen-bonded turnShort motif (e.g., type I or II) with non-helical hydrogen bonds.
SBendCurved backbone without hydrogen bonds, based on dihedral angles.
-LoopIrregular coil with no defined hydrogen bonds or geometry.
Pκ-helix (poly-proline II)Left-handed extended helix with approximately 3 residues per turn, common in unstructured regions and transmembrane proteins.
These assignments enable quantitative analysis of secondary structure content, such as the percentage of helical residues in a protein, and facilitate comparisons across homologs. For example, in myoglobin, DSSP identifies ~75% α-helical content, reflecting its classic globin fold. The original DSSP implementation has evolved, with the 2025 release of DSSP 4 introducing FAIR (Findable, Accessible, Interoperable, Reusable) principles for better data annotation. Key enhancements include detection of left-handed κ-helices (common in transmembrane proteins), improved handling of disulfide bridges, and compatibility with modern formats like mmCIF. This version also refines hydrogen bond calculations for higher accuracy in low-resolution structures and integrates with PDB-REDO for automated re-refinement. DSSP 4 maintains backward compatibility while expanding to nine structure types, enhancing its utility in large-scale structural genomics. Despite alternatives like STRIDE or KAKSI, DSSP remains predominant due to its simplicity and reproducibility.

Other Assignment Methods

In addition to the DSSP algorithm, several other computational methods have been developed to assign protein secondary structures from atomic coordinates, employing diverse criteria such as dihedral angles, backbone geometry, and knowledge-based potentials to address limitations in hydrogen-bond-centric approaches. These alternatives often aim to improve consistency in identifying irregular or edge elements like helix caps and strand distortions, though they can yield varying assignments for the same structure due to differing definitions. The STRIDE algorithm uses a combination of hydrogen bond patterns and empirical phi/psi dihedral angle propensities derived from known protein structures to classify residues into alpha-helices, beta-strands, or coils, providing smoother transitions at secondary structure boundaries compared to DSSP. It incorporates a spline-fitting procedure for beta-strands to better capture twisted conformations, achieving higher agreement with manual assignments in benchmark tests on globular proteins. DEFINE relies solely on Cα atom coordinates, comparing inter-residue distances and angles to idealized geometries of and without considering hydrogen bonds, which makes it computationally efficient for low-resolution models. This method identifies structural motifs by masking expected distance patterns, such as 5.4 Å for adjacent Cα in , and has been influential in early automated analyses of . P-SEA assigns secondary structures using only the Cα trace, applying pattern recognition on local curvature and torsion angles to delineate helices and strands, often resulting in fewer assigned helices but more extended strands than hydrogen-bond-based methods. It excels in handling distorted elements by prioritizing backbone linearity, with applications in fold recognition where precise atomic data is unavailable. More recent tools like KAKSI focus on phi/psi dihedral angles and Cα distances to emphasize linear helices while minimizing assignments to curved or kinked variants, reducing over-assignment of irregular structures observed in older methods. Similarly, SEGNO employs geometric criteria including residue distances, bond angles, and virtual torsion angles to classify elements, incorporating evolutionary conservation signals for enhanced accuracy in divergent protein families.
MethodPrimary CriteriaKey AdvantagesOriginal Reference
STRIDEHydrogen bonds + dihedral anglesBetter edge detectionFrishman & Argos (1995)
DEFINECα distances and anglesEfficiency for coarse modelsRichards & Kundrot (1988)
P-SEACα trace curvature/torsionHandles distortionsLabesse et al. (1997)
KAKSIDihedrals + Cα distancesLinear element focusMartin et al. (2005)
SEGNOGeometry + evolutionary signalsReflects physical propertiesSonego et al. (2005)

Experimental Determination

X-ray Crystallography

X-ray crystallography serves as the primary experimental method for determining the three-dimensional structures of proteins at atomic resolution, enabling the precise identification of secondary structure elements such as and . The technique involves directing a beam of X-rays at a protein crystal, where the X-rays scatter off the electrons of the atoms, producing a diffraction pattern that encodes information about atomic positions. By solving the phase problem—often through methods like multiple isomorphous replacement or molecular replacement—this pattern is transformed into an electron density map, into which the protein's amino acid sequence is fitted to model the structure. Secondary structures appear as characteristic density patterns: manifest as rod-like densities with 3.6 residues per turn, while show extended, pleated strands with hydrogen-bonded alignments. Historically, X-ray crystallography played a pivotal role in confirming the existence of regular secondary structures proposed by and in 1951 through model building. The first protein structure solved, sperm whale myoglobin at 2 Å resolution in 1960, revealed a predominantly alpha-helical fold comprising eight helices, validating the alpha helix geometry with its 5.4 Å pitch and intra-chain hydrogen bonds. Similarly, the refinement of hemoglobin to 2.8 Å resolution in the late 1960s by demonstrated both alpha helices and the absence of beta sheets in its quaternary structure, establishing X-ray as the gold standard for structural validation. These early achievements, which accounted for the initial entries in the established in 1971, underscored the method's ability to resolve secondary structural motifs at resolutions below 3 Å, where side-chain densities become discernible. In practice, achieving high-quality crystals is crucial, as disordered regions like flexible loops may exhibit poor electron density, leading to incomplete models of secondary structures such as turns. Resolutions of 1.5–2.5 Å are ideal for unambiguous secondary structure assignment, allowing clear visualization of backbone hydrogen bonding patterns that define helices and sheets; coarser resolutions around 3 Å suffice for overall folds but may blur subtle features. For instance, the structure of hen egg-white lysozyme at 2 Å resolution in 1965 highlighted a mix of alpha helices, beta sheets, and connecting loops, illustrating how the technique captures the architectural diversity of secondary elements. Modern synchrotron sources have accelerated data collection, enabling structures like protein kinase A at 2.0 Å, which displays conserved beta strands and regulatory helices critical for function. Despite its precision, X-ray crystallography has limitations for secondary structure determination, particularly for dynamic or membrane proteins where crystallization is challenging, potentially distorting native conformations of loops and turns. The method provides a static snapshot of the crystal lattice, which may not reflect solution-state flexibility, and regions with multiple conformations can appear as averaged, smeared densities, complicating assignment of irregular secondary elements. Complementary techniques like NMR are often needed for validation in such cases. As of 2020, X-ray structures comprise over 85% of the Protein Data Bank entries, yet ongoing advancements in phasing algorithms continue to mitigate these issues for more comprehensive secondary structure insights.

Nuclear Magnetic Resonance Spectroscopy

Nuclear magnetic resonance (NMR) spectroscopy determines protein secondary structure in solution, providing atomic-level insights under near-physiological conditions, unlike X-ray crystallography which requires crystals. This technique relies on measuring interactions between nuclear spins, yielding parameters such as chemical shifts, nuclear Overhauser effects (NOEs), and scalar couplings (J-couplings) that report on local conformational features. Seminal advancements by Kurt Wüthrich in the 1980s established NMR as a key method for protein structure elucidation, earning him the 2002 Nobel Prize in Chemistry. Chemical shifts, the resonant frequencies of nuclei influenced by their electronic environment, serve as primary indicators of secondary structure. In α-helices, Cα-H protons exhibit upfield shifts (negative secondary chemical shifts, Δδ ≈ -0.4 ppm), while β-sheets show downfield shifts (Δδ ≈ +0.4 ppm); similar trends occur for Cα carbons (Δδ ≈ -3 ppm for helices, +1 ppm for sheets). The (CSI), introduced by Wishart and Sykes in 1994, assigns structure types by thresholding deviations from random coil values across multiple nuclei (e.g., ¹Hα, ¹³Cα, ¹³CO), achieving 90-95% accuracy for helix and sheet identification in proteins up to 25 kDa. Databases like the (BMRB) facilitate validation and prediction using empirical correlations. NOEs provide distance restraints (<6 Å) between nearby protons, revealing spatial proximities diagnostic of secondary elements. Sequential NOE patterns distinguish structures: α-helices display strong dNN(i,i+1) (~2.8 Å) and medium dαN(i,i+3) (~3.5 Å) connectivities, reflecting amide proton hydrogen bonding, whereas β-sheets exhibit strong dαN(i,i+1) (~2.2 Å) and weak dNN(i,i+1) (>3 Å).90034-2) These are observed in 2D/3D NOESY spectra, with automation tools like ATNOS/CANDID accelerating assignment for larger proteins. For example, in bovine pancreatic , Wüthrich's group used NOEs to confirm β-sheet and helical segments in 1984.90034-2) J-couplings, mediated through bonds, report dihedral angles via the Karplus equation (³J ≈ A cos²θ + B cosθ + C, where θ is the torsion ). The ³JHNα coupling is ~4-5 Hz in α-helices (φ ≈ -60°) and ~8-9 Hz in β-sheets (φ ≈ -140°), enabling secondary structure assignment from HNHA spectra.72016-3) This complements NOEs; for instance, low ³J values confirm helical regions in . Combined with chemical shifts and NOEs, these parameters yield robust secondary structure models, often refined using tools like for φ/ψ prediction from shifts alone (accuracy >90% for backbone). For turns and loops, irregular elements are identified by the absence of regular NOE patterns and unique deviations, such as positive ¹Hα shifts in type I β-turns. NMR's solution-state advantage reveals dynamic aspects, like transient helices in , though it is limited to proteins <50 kDa without advanced labeling. Overall, these methods integrate into full structure calculations via software like CYANA or , with secondary structure serving as an initial scaffold.

Cryo-electron Microscopy

Cryo-electron microscopy (cryo-EM) has emerged as a pivotal technique for determining the three-dimensional structures of proteins and macromolecular assemblies, particularly those that are difficult to crystallize. Developed through foundational contributions recognized by the awarded to , , and Richard Henderson, cryo-EM involves flash-freezing biological samples in vitreous ice to preserve native states and imaging them under beams to generate projections. These projections are computationally reconstructed into density maps, enabling visualization of molecular architectures at resolutions ranging from low (worse than 10 ) to near-atomic (better than 3 ). Unlike , cryo-EM does not require crystals, making it ideal for flexible or heterogeneous proteins, and it has facilitated over 10,000 protein structures deposited in the by 2023. In cryo-EM density maps, protein secondary structures become discernible at intermediate resolutions of approximately 5–10 Å, where α-helices manifest as elongated cylindrical densities about 5 Å in diameter, and β-sheets appear as broad, flat planes with characteristic strand separations of 4.5–5 Å. At higher resolutions below 4 Å, individual amino acid side chains and backbone atoms can be resolved, allowing precise assignment of secondary elements through density fitting. This resolution-dependent visualization stems from the technique's ability to average thousands of particle images, reducing noise and revealing structural motifs that inform overall folding patterns. For instance, in the structure of the ribosome, cryo-EM maps at 3.5 Å resolution clearly delineate helical and sheet regions critical for function. Assignment of secondary structures from cryo-EM maps traditionally involved manual interpretation or rigid-body fitting of known models, but computational methods have advanced automation, especially using . A seminal tool, Emap2sec, employs a to classify voxels in maps of 6–10 resolution, achieving over 90% accuracy in identifying α-helices, β-sheets, and coils by training on simulated densities from PDB structures. Subsequent developments, such as DeepSSETracer, extend this to medium-resolution maps (5–8 ) via a architecture for segmenting secondary elements, enabling tracing of helices and sheets in complex assemblies like viral capsids. More recent frameworks like ModelAngelo integrate secondary structure prediction with , supporting resolutions down to 3 and reducing manual intervention in large-scale studies. These methods prioritize in density gradients, enhancing reliability for heterogeneous samples. The application of cryo-EM to secondary structure determination has profound implications for understanding protein dynamics and interactions, as seen in the elucidation of topologies where helical bundles are key functional units. By combining cryo-EM with predictive tools like , hybrid approaches refine secondary assignments in low-resolution regions, accelerating for targets like ion channels. Ongoing challenges include handling conformational heterogeneity, but improvements in detector technology and algorithms continue to lower resolution barriers, making cryo-EM indispensable for .

Computational Prediction

Early Physicochemical Methods

The early physicochemical methods for predicting protein secondary structure emerged in the 1970s, building on the foundational structural models proposed by Pauling and Corey in 1951, who hypothesized and beta-sheets based on polypeptide backbone geometry and hydrogen bonding patterns. These methods relied on empirical statistical analysis of propensities derived from known protein structures, incorporating physicochemical properties such as hydrophobicity, side-chain bulkiness, and conformational preferences to infer local folding tendencies from primary alone. Unlike later computational approaches, they emphasized rule-based heuristics and information-theoretic principles, achieving modest accuracies of 50-65% in three-state classifications (, sheet, ), which represented a significant advance over random guessing (33%) but highlighted limitations in capturing long-range interactions. A seminal contribution was the Chou-Fasman method, introduced in 1974, which classified into categories like helix-formers (e.g., , ), helix-breakers (e.g., , ), sheet-formers (e.g., , ), and sheet-breakers (e.g., ) based on their observed frequencies in secondary structures from a database of 15-20 proteins. The algorithm scanned the sequence for nucleation sites—clusters of at least four helix-formers within six residues or three sheet-formers within five—then extended these regions bidirectionally until average propensities fell below 1.00, resolving overlaps via rules prioritizing sheets over helices. It also predicted turns using propensities. Applied to test sets, this method yielded approximately 50-60% accuracy, with stronger performance for helices (63%) than sheets (56%), though it often overpredicted structures due to simplistic thresholds. Concurrently, the Lim method (1974) adopted a more theoretical physicochemical framework, emphasizing side-chain interactions like charge repulsion and steric hindrance to predict alpha-helices and beta-sheets without direct statistical propensities.90374-6) It used complex rules to identify potential structural segments, such as avoiding charged residues in hydrophobic cores for helices or favoring alternating patterns for sheets, drawing from 29 non-homologous proteins. This approach achieved around 56% accuracy on independent datasets, performing comparably to Chou-Fasman but with better specificity for beta-structures in some cases. The method, developed by , Osguthorpe, and Robson in 1978, marked a second-generation advancement by incorporating contextual information through a sliding window of 13-17 residues, treating as a joint probability problem informed by information theory.90297-8) It computed conditional probabilities for each residue's conformation based on pairwise interactions with neighbors, using log-odds scores from a of 30 proteins to assign the state (, sheet, or ) maximizing the overall likelihood. This method improved accuracy to 60-65%, particularly for (70%), by accounting for sequence dependencies absent in single-residue models, though it still struggled with turns and regions. Subsequent refinements, like GOR III (1996), extended this to multiple sequence alignments, but the original laid the groundwork for statistical mechanics-inspired predictions.90297-8) These early methods, while pioneering, were limited by small training datasets (often <50 proteins) and neglect of evolutionary information, leading to segment-based rather than residue-level predictions with three-state Q3 accuracies rarely exceeding 63%. They established the paradigm of propensity-based inference, influencing later and approaches by demonstrating the value of physicochemical descriptors in decoding sequence-to-structure relationships.

Modern Machine Learning Approaches

Modern approaches, particularly , have significantly advanced protein secondary structure prediction (PSSP) by leveraging large-scale datasets, evolutionary information from multiple sequence alignments (MSAs), and sophisticated neural architectures to capture complex sequence-structure relationships. Unlike early physicochemical methods that relied on empirical rules or basic statistical models, contemporary techniques employ convolutional neural networks (CNNs), recurrent neural networks (RNNs), and transformer-based models to achieve three-state (Q3) accuracies exceeding 84% and eight-state (Q8) accuracies around 70-75% on benchmark datasets like CB513 and CASP targets. These methods integrate profile-based inputs, such as position-specific scoring matrices (PSSMs), and increasingly incorporate pre-trained protein language models (PLMs) like ESM-1b or ProtTrans to encode contextual features without extensive MSAs, enabling faster and more accurate predictions even for single sequences. CNNs excel at extracting local structural motifs by treating protein sequences as one-dimensional signals, often combined with bidirectional (BiLSTM) units to model sequential dependencies. For instance, the SPIDER3 framework (2017) uses deep CNNs and BiLSTMs on PSSM inputs to predict secondary structures, achieving 83.9% Q3 accuracy on the TS115 dataset and outperforming prior tools like PSIPRED by incorporating deep residual connections for better gradient flow. Similarly, SPOT-1D (2018) employs a cascaded CNN-BiLSTM architecture with , reaching 87.16% Q3 on the TEST2016 set through multitask prediction of secondary structure alongside solvent accessibility and dihedral angles, demonstrating how joint learning enhances individual task performance. RNN variants, such as the CSI-LSTM model (2021), further refine predictions by processing NMR-derived data alongside sequence profiles, improving accuracy for disordered regions. Transformer architectures represent a paradigm shift by using self-attention mechanisms to capture long-range interactions across entire sequences, often pre-trained on massive protein corpora. NetSurfP-3.0 (2022) integrates ESM-1b embeddings into a transformer-based network, predicting secondary structure with 84.6% Q3 accuracy on CB513 while also estimating disorder and accessibility, and it reduces reliance on MSAs for broader applicability. The influence of end-to-end 3D structure predictors like AlphaFold2 (2021) has indirectly elevated PSSP, as its Evoformer module—comprising transformer blocks—derives secondary structures with near-perfect fidelity (over 90% Q3 in many cases) by jointly modeling residue contacts and distances from evolutionary couplings. Recent innovations, such as knowledge distillation in hybrid models (2025), combine PLMs with lightweight neural networks to boost efficiency, achieving comparable accuracies to larger models on CASP14 datasets while minimizing computational demands. These approaches prioritize conceptual integration of biophysical priors, like hydrogen bonding patterns, into learned representations, paving the way for real-time applications in protein engineering.

Applications

Protein Folding and Stability

Protein secondary structures, primarily alpha-helices and beta-sheets, serve as fundamental building blocks in by enabling the rapid formation of local, ordered motifs that guide the collapse of the polypeptide chain into its native conformation. During the folding process, these structures emerge early through intra- and inter-chain hydrogen bonding, acting as sites that reduce the entropic barrier to folding. This nucleation-condensation mechanism unifies various folding models, where the features partially formed native-like secondary elements stabilized by a mix of short-range backbone interactions and emerging long-range contacts, with phi-values indicating 50-70% native-like structure in the folding . The stability of alpha-helices arises predominantly from medium-range hydrogen bonds between the carbonyl oxygen of residue i and the amide hydrogen of residue i+4, supplemented by van der Waals interactions and side-chain contributions, which collectively provide 0.3-1 kcal/mol per residue to the of folding. These bonds bury polar groups, minimizing unfavorable interactions with and enhancing conformational rigidity, as evidenced by the high helical propensity of residues like and . In contrast, beta-sheets rely on long-range inter-strand hydrogen bonds and hydrophobic packing, yielding higher long-range contact densities (4.5-5.3 contacts per residue) and greater overall stabilization, with correlations between long-range order and folding rates (r = -0.78). Additional secondary forces, such as C-H···O s and n→π* interactions, fine-tune stability in these motifs; for instance, n→π* interactions stabilize over 70% of alpha-helical residues with 0.3-0.7 kcal/mol contributions, while C-H···O bonds affect about 35% of beta-sheet residues at 1-2 kcal/mol each. The cumulative effect of these interactions marginally contributes to net stability (average ~0.95 kJ/mol per backbone at ), but disruptions, such as mutations altering bond geometry, can significantly lower folding rates and thermal stability, as seen in variants of proteins like inhibitor 2 spanning 2.4-2300 s⁻¹ folding rates.

Biotechnology and Drug Design

Protein secondary structure plays a pivotal role in by enabling the rational engineering of proteins with enhanced stability, functionality, and therapeutic potential. In , computational methods that predict and design secondary structural elements, such as alpha-helices and beta-sheets, allow for the creation of novel scaffolds for biotechnological applications. For instance, design of self-assembling helical protein filaments has been achieved using Rosetta-based algorithms, which optimize helical coiled-coil motifs to form stable nanostructures suitable for and . Similarly, beta-barrel proteins have been designed from scratch, incorporating antiparallel beta-strands to create fluorescent reporters with improved photostability for cellular imaging in biotechnology workflows. These approaches rely on accurate secondary structure to ensure the folded proteins maintain their intended conformations under physiological conditions, thereby expanding the toolkit for and biosensors. In , secondary structure elements are targeted to disrupt pathological protein-protein interactions (PPIs) or aggregation events, often through the development of mimetics or . Alpha-helix mimetics, which replicate the spatial arrangement of side chains in helical segments, have emerged as potent inhibitors of helix-mediated PPIs. Seminal work introduced scaffolds as first-generation mimetics, capable of emulating i, i+3, i+4, and i+7 residue positions to selectively inhibit the /hDM2 interaction, restoring activity in cancer cells with low nanomolar potency. These non-peptidic compounds have since been optimized for broader applications, including pan-inhibitors of the that block anti-apoptotic signaling in models, demonstrating cellular efficacy and oral bioavailability in preclinical studies. More recent advancements incorporate constrained peptides and foldamers, such as those using hydrogen-bond surrogates, to target diverse PPIs like HIF-1α/p300, offering a modular platform for structure-based optimization. Beta-sheet structures, implicated in amyloid diseases, are addressed in drug design via beta-sheet breakers that prevent fibril formation. In Alzheimer's disease, short peptides derived from the amyloid-beta (Aβ) sequence, such as LVFF (residues 17-20), are conjugated with nicotinic acid to enhance solubility and beta-sheet disruption, shifting Aβ(1-42) aggregates toward random coil conformations and inhibiting neurotoxicity in vitro. This approach, informed by cryo-EM structures of Aβ fibrils revealing parallel beta-sheet cores, has led to candidates like NA-16KLVF19 that exhibit protease resistance and cognitive benefits in animal models. Beyond neurodegeneration, engineered beta-sheet miniproteins serve as scaffolds in vaccine design, where stabilized sheets present epitopes to elicit immune responses, as seen in nanoparticle platforms displaying helical and sheet motifs for neutralizing antibodies against viruses like influenza. Overall, integrating secondary structure with like deep learning-enhanced models has accelerated these applications, from therapeutic proteins mimicking IL-2 signaling to custom inhibitors, underscoring the translational impact in and precision medicine.

References

  1. [1]
    Biochemistry, Secondary Protein Structure - StatPearls - NCBI - NIH
    In this structure, 2 different regions of a polypeptide chain lie side by side and are bound by hydrogen bonds. They make up the core of many globular proteins.Introduction · Fundamentals · Clinical Significance
  2. [2]
    The structure of proteins: Two hydrogen-bonded helical ... - PNAS
    Two hydrogen-bonded helical structures for a polypeptide chain have been found in which the residues are stereochemically equivalent.Missing: original | Show results with:original
  3. [3]
    [PDF] Secondary Structure - Rose-Hulman
    Secondary Structure (α-helix and β-sheet). Secondary structural elements are repeated regular structures within proteins. In these repeated structures ...Missing: key | Show results with:key
  4. [4]
    Protein Structure | Learn Science at Scitable - Nature
    Known as alpha helices and beta sheets, these stable folding patterns make up the secondary structure of a protein. Most proteins contain multiple helices ...
  5. [5]
    The Pleated Sheet, A New Layer Configuration of Polypeptide Chains
    The Pleated Sheet, A New Layer Configuration of Polypeptide Chains. Linus Pauling and Robert B. CoreyAuthors Info & Affiliations. May 15, 1951. 37 (5) 251-256.
  6. [6]
    Secondary Structure of Proteins and Nucleic Acids - NCBI - NIH
    The secondary structural features of proteins can be grouped into three broad classes: helical features, extended strands, and turns or loops.Missing: definition | Show results with:definition
  7. [7]
    The discovery of the α-helix and β-sheet, the principal structural ...
    PNAS papers by Linus Pauling, Robert Corey, and Herman Branson in the spring of 1951 proposed the α-helix and the β-sheet, now known to form the backbones ...
  8. [8]
    Terminology of Protein Structure - Nature
    This terminology, introduced by Linderstrøm-Lang in his Lane Memorial Lectures1, served well in the few years since it was invented, and has thoroughly ...
  9. [9]
    The discovery of the α-helix and β-sheet, the principal structural ...
    PNAS papers by Linus Pauling, Robert Corey, and Herman Branson in the spring of 1951 proposed the α-helix and the β-sheet, now known to form the backbones ...
  10. [10]
    CHEM 245 - α helix
    Jan 1, 2018 · Characteristics of the α helix · Right-handed · 3.6 residues per turn · Pitch = 5.4 Å · H-bond distances (amide N to acyl O) ~ 2.8 Å · In proteins, ...
  11. [11]
    Analysis of forces that determine helix formation in α-proteins - NIH
    A model for prediction of α-helical regions in amino acid sequences has been tested on the mainly-α protein structure class.
  12. [12]
    The Membrane- and Soluble-Protein Helix-Helix Interactome - NIH
    The α-helix is by far the most common regular secondary structure element. In water-soluble proteins approximately 35% of all protein residues are in the α- ...
  13. [13]
    PPS 97' - Beta-Sheet Geometry
    Nov 4, 1996 · Typical values are phi = -140 degrees and psi = 130 degrees. In contrast, alpha-helical residues have both phi and psi negative.
  14. [14]
    Antiparallel and parallel β-strands differ in amino acid residue ...
    Nov 1, 1979 · Parallel β P ) and antiparallel (β A ) arrangements of strands in a sheet differ in the hydrogen bond pattern between strands.
  15. [15]
    On the strength of β-sheet crystallites of Bombyx mori silk fibroin
    Jul 6, 2014 · Structurally, silk fibroin is composed of β-sheet crystallite domains that are embedded in a matrix of amorphous domains (figure 1). The fibroin ...Introduction · Models and methods · Results · Discussion
  16. [16]
    Hierarchical Structure of Proteins - PDB-101
    There are two common secondary structures: alpha helices and beta strands. ... In this structure, four beta strands form the sheet with neighboring strands ...
  17. [17]
    A Perspective on the (Rise and Fall of) Protein β-Turns - PMC
    Oct 14, 2022 · The β-turn is the third defined secondary structure after the α-helix and the β-sheet. The β-turns were described more than 50 years ago and account for more ...
  18. [18]
    A Perspective on the (Rise and Fall of) Protein β-Turns - MDPI
    The β-turn is the third defined secondary structure after the α-helix and the β-sheet. The β-turns were described more than 50 years ago and account for more ...<|control11|><|separator|>
  19. [19]
    Stereochemical criteria for polypeptides and proteins. V ...
    First published: October 1968 ... The two nonhelical conformations are very similar, and in both the cases the peptide chain turns around, reversing the direction ...
  20. [20]
  21. [21]
    Prediction of Tight Turns and Their Types in Proteins - ScienceDirect
    Tight turns are generally categorized as δ-turn, γ-turn, β-turn, α-turn, and π-turn, which are formed by two-, three-, four-, five-, and six-amino-acid residues
  22. [22]
    Loops in Globular Proteins: A Novel Category of Secondary Structure
    A systematic study was made of 67 proteins of known structure revealing 270 omega loops. ... Citations. Cite as. Jacquelyn F. Leszczynski,; George D. Rose. ,.
  23. [23]
    [PDF] Dictionary of Protein Secondary Structure: Pattern Recognition of ...
    Our goal is to approximate the intuitive notion of secondary structure by an objective algorithm. An algorithm for extracting structural features from the ...
  24. [24]
    Protein Secondary Structure Classification Revisited
    The most widely used program for secondary structure assignment remains DSSP, introduced in 1983, with currently more than 400 citations per year.
  25. [25]
    Protein secondary structure assignment revisited: a detailed ...
    Sep 15, 2005 · We have developed a method for secondary structure assignment, called KAKSI. Assignments made by KAKSI are compared with assignments given by DSSP, STRIDE, ...
  26. [26]
    DSSP 4: FAIR annotation of protein secondary structure - Hekkelman
    Jul 17, 2025 · The DSSP algorithm is based on the detection of backbone hydrogen bonds that form primary structural elements: hydrogen bonded turns and bridges ...
  27. [27]
    PDBe adopts the new DSSP for protein secondary structure annotation
    Oct 30, 2025 · Developed by the PDB-REDO team at the Netherlands Cancer Institute, DSSP analyzes atomic coordinates to consistently identify helices, strands, ...
  28. [28]
    Knowledge‐based protein secondary structure assignment
    We have developed an automatic algorithm STRIDE for protein secondary structure assignment from atomic coordinates based on the combined use of hydrogen bond ...Missing: original | Show results with:original
  29. [29]
    STRIDE: a web server for secondary structure assignment from ... - NIH
    STRIDE is a software tool for secondary structure assignment from atomic resolution protein structures. It implements a knowledge-based algorithm that makes ...
  30. [30]
    Identification of structural motifs from protein coordinate data - PubMed
    Identification of structural motifs from protein coordinate data: secondary structure and first-level supersecondary structure ... F M Richards , C E Kundrot ...
  31. [31]
    Protein Secondary Structure - an overview | ScienceDirect Topics
    DEFINE-S (Richards and Kundrot, 1988) uses only Cα coordinates, compares their distances with the distances in ideal SSEs and also provides information ...
  32. [32]
    P-SEA: a new efficient assignment of secondary structure from Cα ...
    Accurate assignment of the secondary structure elements (SSE) (helix, strand, coil) is an essential step for the analysis and modelling of protein structure.
  33. [33]
    Secondary structure assignment that accurately reflects physical and ...
    Dec 1, 2005 · Secondary structure is used in hierarchical classification of protein structures, identification of protein features, such as helix caps and ...
  34. [34]
    x Ray crystallography - PMC - PubMed Central - NIH
    x Ray crystallography is currently the most favoured technique for structure determination of proteins and biological macromolecules.
  35. [35]
    Methods for Determining Atomic Structures - PDB-101
    Several methods are currently used to determine the structure of a protein, including X-ray crystallography, NMR spectroscopy, and electron microscopy.Missing: papers | Show results with:papers
  36. [36]
    Review A Glimpse of Structural Biology through X-Ray Crystallography
    Nov 20, 2014 · As an extension of X-ray crystallography, electron crystallography was successfully applied to determine the structure of membrane proteins ...Kinases And Anticancer Drugs · Membrane Protein Structure... · Membrane Transport Proteins
  37. [37]
    Heterogeneity and Inaccuracy in Protein Structures Solved by X-Ray ...
    In protein structure determination by X-ray crystallography, the observed diffraction pattern results from the scattering of X-rays by an ensemble of ...
  38. [38]
  39. [39]
    [PDF] Nobel Lecture - Kurt Wüthrich
    Standard presentation of 1H NMR data leading to sequence-specific resonance as- signments and the identification of regular secondary structures in proteins.
  40. [40]
    Press release: The 2017 Nobel Prize in Chemistry - NobelPrize.org
    Oct 4, 2017 · The Nobel Prize in Chemistry 2017 is awarded to Jacques Dubochet, Joachim Frank and Richard Henderson for the development of cryo-electron microscopy.<|control11|><|separator|>
  41. [41]
    Cryo-electron microscopy wins chemistry Nobel - Nature
    Oct 12, 2017 · Jacques Dubochet, Joachim Frank and Richard Henderson were awarded the prize on 4 October for their work in developing cryo-electron microscopy (cryo-EM).
  42. [42]
    Protein secondary structure detection in intermediate-resolution cryo ...
    Jul 29, 2019 · Emap2sec uses a three-dimensional deep convolutional neural network to assign secondary structure to each grid point in an EM map. We tested ...Missing: details | Show results with:details
  43. [43]
    Protein Secondary Structure Detection in Intermediate Resolution ...
    We have developed a computational method, Emap2sec, which identifies the secondary structures of proteins (α helices, β sheets, and other structures) in an EM ...
  44. [44]
    A Tool for Segmentation of Secondary Structures in 3D Cryo-EM ...
    Nov 2, 2021 · We have developed an effective software bundle, DeepSSETracer, for the detection of protein secondary structure from cryo-EM component maps in medium ...
  45. [45]
    Automated model building and protein identification in cryo-EM maps
    Feb 26, 2024 · Here we present ModelAngelo, a machine-learning approach for automated atomic model building in cryo-EM maps.
  46. [46]
    Determining Protein Secondary Structures in Heterogeneous ...
    Jun 8, 2024 · In this work, we present a U-net-based segmentation model, cryoSSESeg, for detecting protein secondary structures in medium-resolution cryo-EM ...Introduction · Materials and Method · Results and Discussion · Author Information
  47. [47]
    Review Predictive modeling and cryo-EM: A synergistic approach to ...
    Feb 20, 2024 · This review examines the synergy between these two techniques, as well as their limitations, in the structural and functional characterization of large ...
  48. [48]
    Modeling cryo-EM structures in alternative states with AlphaFold2 ...
    Oct 30, 2025 · Cryo-electron microscopy (cryo-EM) is one of the most widely used and rapidly developing techniques for biomolecular structure determination.Results · Discussion · Methods<|control11|><|separator|>
  49. [49]
    GOR V server for protein secondary structure prediction
    Summary: We have created the GOR V web server for protein secondary structure prediction. The GOR V algorithm combines information theory, Bayesian statistics ...
  50. [50]
    [PDF] 1 Secondary structure prediction - cs.Princeton
    Nov 16, 2000 · Early secondary structure prediction methods (such as Chou-Fasman and GOR, out- lined below) had a 3-state accuracy of 50–60%. (They initially ...
  51. [51]
    Sixty-five years of the long march in protein secondary structure ...
    Protein secondary structure prediction began in 1951 when Pauling and Corey predicted helical and sheet conformations for protein polypeptide backbone even ...
  52. [52]
    Deep learning for protein secondary structure prediction - NIH
    This paper aims to provide a comprehensive review of the trends and challenges of deep neural networks for protein secondary structure prediction (PSSP).Missing: seminal | Show results with:seminal
  53. [53]
  54. [54]
  55. [55]
    Transition-state structure as a unifying basis in protein-folding ...
    The rate of folding depends primarily on the interactions that directly stabilize the nucleus, especially those in native-like secondary structure and those ...
  56. [56]
    Inter-residue interactions in protein folding and stability - ScienceDirect
    In this review, we provide a comprehensive account of the role of inter-residue interactions in the structure, folding and stability of proteins.<|control11|><|separator|>
  57. [57]
    Secondary Forces in Protein Folding - PMC - PubMed Central
    The dominant contributors to protein folding include the hydrophobic effect and conventional hydrogen bonding, along with Coulombic interactions and van der ...
  58. [58]
    Hydrogen-bonding classes in proteins and their contribution to ... - NIH
    This paper proposes to assess hydrogen-bonding contributions to the protein stability, using a set of model proteins for which both X-ray structures and ...Theoretical Premise · Results · Fig. 2
  59. [59]
    Advances in protein structure prediction and design - Nature
    Aug 15, 2019 · In this Review, we describe current approaches for protein structure prediction and design and highlight a selection of the successful applications they have ...
  60. [60]
    Helix mimetics: Recent developments - PubMed
    Jun 3, 2015 · This review explores recent significant developments in the area of small molecule α-helix mimetics and provides a critical overview of success stories.
  61. [61]
    Helix mimetics: Recent developments - ScienceDirect
    The development of protein–protein interaction (PPIs) inhibitors represents a challenging goal in chemical biology and drug discovery.
  62. [62]
    Novel Design of Neuropeptide-Based Drugs with β-Sheet Breaking ...
    Mar 5, 2022 · This study offers a unique insight into the design of neuropeptide-based drugs with β-sheet breaker potential in the amyloid-beta cascade for Alzheimer's ...