Fact-checked by Grok 2 weeks ago

Protein tertiary structure

Protein tertiary structure refers to the complete three-dimensional conformation of a protein, encompassing its polypeptide backbone and all side chains, which arises from the folding of the chain into a specific spatial arrangement. This level of structure is essential for determining a protein's functional properties, as it positions key residues to enable interactions with other molecules, such as substrates in enzymes or ligands in receptors. Unlike primary structure, which is the linear sequence of , or secondary structure, which involves local folding patterns like alpha helices and beta sheets, tertiary structure integrates these elements into a global fold stabilized by non-covalent and covalent interactions. The formation of tertiary structure is primarily dictated by the protein's , with folding occurring spontaneously under physiological conditions, often assisted by molecular chaperones to prevent aggregation. Key stabilizing interactions include hydrophobic interactions, where nonpolar side chains cluster in the protein's interior away from ; hydrogen bonds between polar groups; ionic bonds or salt bridges between oppositely charged residues; van der Waals forces providing weak attractions between closely packed atoms; and disulfide bonds, which are covalent linkages between residues that lock distant parts of the chain together. These forces collectively minimize , driving the polypeptide to adopt its native, functional conformation, which can be reversible— as demonstrated by experiments where denatured proteins like refold correctly upon removal of disrupting agents. Environmental factors such as , , and can influence this folding process. Tertiary structure is crucial for protein and , with disruptions often leading to loss of function or pathological states. For instance, globular proteins like exhibit compact tertiary folds that facilitate oxygen transport, while fibrous proteins such as form elongated structures for tensile strength. Clinically, mutations altering the primary sequence can misfold the tertiary structure, contributing to diseases including (due to defective CFTR channel folding), Alzheimer's disease (via amyloid-beta aggregation), and disorders (from conformational changes in PrP protein). Experimental determination of tertiary structures relies on techniques like , (NMR) spectroscopy, and cryo-electron microscopy, which have revealed atomic-level details for thousands of proteins.

Overview

Definition

Protein tertiary structure refers to the three-dimensional spatial arrangement of a polypeptide chain, achieved through folding that brings distant residues into close proximity via interactions between their side chains, known as R-groups. This level of structure builds upon secondary elements, such as alpha helices and beta sheets, which serve as the foundational motifs that pack together to form the overall fold. The stability of tertiary structure is maintained by several key non-covalent and covalent interactions. Hydrophobic effects drive non-polar R-groups to cluster in the protein's interior, away from , while hydrogen bonds form between polar groups, ionic bonds (or bridges) occur between oppositely charged residues, disulfide bridges provide covalent links between residues, and van der Waals forces contribute weak attractions between closely packed atoms. These interactions collectively minimize the protein's energy state, enabling a compact conformation essential for function. In many proteins, secondary structure motifs fold into a compact globule, exemplified by globular proteins like , which adopt spherical shapes for solubility and enzymatic activity in aqueous environments. In contrast, fibrous proteins, such as , exhibit elongated, rope-like tertiary structures suited for structural support and tensile strength. The folding process is thermodynamically driven by the minimization of , described by the equation \Delta G = \Delta H - T\Delta S where \Delta G is the change in free energy, \Delta H is the enthalpy change, T is the temperature, and \Delta S is the entropy change; a negative \Delta G favors the folded state as the most stable conformation.

Comparison with other structural levels

Protein tertiary structure represents the three-dimensional folding of a single polypeptide chain, building directly upon the primary structure, which is defined as the linear sequence of connected by covalent bonds. This foundational sequence, first elucidated for insulin by in the early 1950s, dictates the potential for higher-order folding through the chemical properties of the amino acid side chains. In contrast to the one-dimensional nature of primary structure, tertiary structure achieves a compact, globular form that positions distant residues in space for functional interactions. Secondary structure encompasses local conformational motifs within the polypeptide chain, such as α-helices and β-sheets, primarily stabilized by hydrogen bonds between the backbone carbonyl oxygen and amide hydrogen atoms. Proposed by and Robert Corey in 1951, these elements represent repetitive patterns that emerge early in folding but do not capture the full spatial organization of the protein. Tertiary structure, however, integrates these secondary motifs into a global architecture, where interactions between non-adjacent side chains—such as hydrophobic clustering or occasional disulfide bridges—bridge separate secondary elements to define the overall shape. Quaternary structure arises from the non-covalent or covalent association of two or more polypeptide subunits, each of which typically possesses its own structure, to form a multi-chain complex essential for certain proteins like . Unlike assemblies, which involve inter-chain contacts across multiple folded units, structure is confined to the intramolecular folding of an individual chain, though that chain may later participate in interactions. This distinction highlights how structure provides the stable subunit scaffold necessary for higher-level oligomerization. The hierarchy of protein structures is often depicted in schematic diagrams progressing from the linear primary sequence (1D) through local secondary folds (2D-like elements) to the fully realized 3D conformation of a single chain, and culminating in multi-subunit arrangements (beyond 3D). These visualizations underscore the sequential dependency, where alterations at the primary level can propagate through secondary and levels to affect assembly.

Historical development

Early theories

In the and , protein structure was often interpreted through the lens of colloid chemistry, which viewed globular proteins as compact, spherical molecular with defined molecular weights, as demonstrated by Svedberg's ultracentrifugation studies that established proteins like as discrete molecules rather than indefinite aggregates. This perspective emphasized the compactness of globular proteins to maintain and stability in aqueous environments, influenced by colloidal principles that highlighted and effects. A key early conceptual advance came from Hsien Wu's theory of protein denaturation, which posited that native globular proteins adopt a highly compact configuration where nonpolar (hydrophobic) side chains are internalized, shielded from , while denaturation disrupts this arrangement and exposes these groups, leading to aggregation or changes. Dorothy Maud Wrinch proposed the cyclol theory in the late , suggesting that globular proteins form rigid, cage-like polyhedral structures through covalent cyclol bonds—resulting from dehydration reactions between bonds and side-chain carboxyl or hydroxyl groups—creating a fabric of cross-linked hexagons and pentagons that encapsulate the protein's interior. This model aimed to explain the stability and globular shape of proteins like insulin, predicting flat, basket-like architectures without relying on non-covalent interactions. However, the theory was largely disproven by the late due to inconsistencies with bond energy calculations and emerging data, as and Carl Niemann demonstrated that cyclol linkages would be thermodynamically unstable compared to standard bonds. Linus Pauling contributed foundational ideas in 1951 by proposing the and beta-pleated sheet as regular secondary structures for polypeptide chains, allowing for compact folding while satisfying hydrogen bonding and stereochemical constraints, thereby laying groundwork for understanding how linear sequences might achieve tertiary compactness. These early theories, while innovative, were constrained by the absence of atomic-resolution crystallographic data, limiting models to speculative geometries based on chemical and low-resolution fiber diffraction patterns, which could not yet distinguish precise side-chain arrangements or confirm hydrophobic burial.

Key milestones

In 1960, and colleagues determined the first high-resolution three-dimensional structure of a protein, , using at 2 Å resolution, revealing a compact globular fold dominated by an alpha-helical bundle comprising about 75% of the polypeptide chain with eight helices arranged around the . This breakthrough provided the initial empirical visualization of protein tertiary structure, demonstrating how the chain folds into a functional domain without beta-sheets and highlighting the role of hydrophobic interactions in stabilizing the core. The 1969 formulation of by Cyrus Levinthal underscored the computational improbability of random searching for the native tertiary structure, estimating that a 100-residue protein with approximately three possible conformations per residue could adopt around $10^{27} states, which at a rate of $10^{13} conformations per second would require over $10^{27} years to explore—far exceeding the age of the —thus implying guided folding pathways rather than exhaustive trials. Christian Anfinsen's experiments in the early 1960s, culminating in his 1972 , demonstrated through refolding of denatured A that the spontaneously regained full activity after removal of denaturants and reoxidation of its four disulfide bonds, with only one of 105 possible pairings yielding the native structure, establishing the thermodynamic hypothesis that the primary encodes all information necessary for achieving the lowest free-energy fold under physiological conditions. This principle was further articulated in Anfinsen's 1973 overview, emphasizing that structure emerges from -directed minimization of without requiring additional genetic instructions. During the 1980s and 1990s, advances in disulfide bond mapping, pioneered by Thomas Creighton's studies on bovine pancreatic trypsin inhibitor (BPTI), elucidated oxidative folding pathways by isolating productive intermediates with specific disulfide pairings—such as the initial 30-51 and 5-55 bonds—revealing that tertiary structure formation couples sequentially with covalent cross-linking to avoid kinetic traps . Concurrently, the discovery of molecular chaperones transformed understanding of assisted folding; R. John Ellis coined the term in 1987 to describe proteins like Rubisco-binding protein that prevent unproductive aggregation during assembly, while Sean Hemmingsen's 1988 identification of bacterial (now a chaperonin) as a homolog essential for oligomeric protein maturation highlighted their role in facilitating tertiary structure attainment in crowded cellular environments. These developments bridged Anfinsen's insights with mechanisms, showing chaperones such as and /GroES actively guide nascent chains toward native folds without specifying the final structure.

Formation and stability

Folding mechanisms

Protein folding mechanisms encompass the kinetic pathways through which a linear polypeptide chain assembles into its compact tertiary structure, driven by the interplay of conformational and intramolecular interactions. Small proteins, typically under 100 residues, often fold via a two-state model, characterized by a highly transition from a heterogeneous unfolded ensemble to the unique native state, with no detectable stable intermediates on the folding timescale. This process reflects rapid equilibration among unfolded conformations prior to barrier crossing, ensuring efficient folding without kinetic partitioning into off-pathway states. Larger or multi-domain proteins, however, generally exhibit multi-state folding kinetics, involving populated intermediates that bridge the unfolded and native states, allowing sequential assembly of structural modules. These pathways arise from the increased conformational complexity, where partial folding of subdomains precedes global compaction, reducing the risk of aggregation but extending overall folding times. The energy landscape theory conceptualizes these mechanisms as navigation of a rugged, funnel-shaped free energy surface, where the native state forms a deep global minimum and progressive structure formation narrows the accessible conformations, guiding the chain downhill despite local roughness that can create kinetic traps. This statistical view emphasizes evolutionary selection for minimally frustrated landscapes that minimize trapping and promote rapid folding to the thermodynamically stable native . A prominent intermediate in multi-state folding is the molten globule, a compact state with significant native-like secondary structure but fluctuating tertiary contacts and exposed hydrophobic regions, serving as a transient platform for subsequent tight packing. This lacks a fixed side-chain yet retains a globular , bridging early collapse and late-stage refinement in the . Stopped-flow fluorescence spectroscopy provides direct experimental evidence for these mechanisms, monitoring burial or environmental changes on millisecond timescales to resolve folding phases, such as submillisecond collapse to molten globule-like states followed by slower native rearrangements in proteins like . These techniques reveal chevron plots and relaxation that validate two-state in small proteins and accumulation in larger ones. In cellular environments, molecular chaperones briefly assist these pathways by preventing unproductive aggregates.

Physicochemical determinants

The tertiary structure of proteins is primarily stabilized by non-covalent and covalent interactions between side chains, which collectively minimize and drive the polypeptide chain to adopt its native conformation. These physicochemical determinants include the , hydrogen bonding, electrostatic interactions, bonds, and van der Waals forces, each contributing to the burial of nonpolar residues in the core and the precise packing of the structure. The serves as the dominant force, while the others provide specificity and additional stability, particularly in the protein interior where is excluded. The arises from the tendency of nonpolar side chains, such as those of , , and , to aggregate in the protein's interior, away from the aqueous environment. This burial releases structured water molecules surrounding hydrophobic groups, leading to an increase in solvent entropy that thermodynamically favors the . Originally proposed by Kauzmann in 1959 as a key driver of , this effect accounts for much of the change during collapse of the polypeptide chain, with estimates suggesting it contributes 20-30 kJ/mol per residue pair in the core. In globular proteins like , nonpolar residues occupy over 50% of the buried surface area, exemplifying how this entropic gain enforces compact tertiary structures. Hydrogen bonding between polar side chains, such as those of serine, threonine, and asparagine, supplements the backbone hydrogen bonds of secondary structures and enhances tertiary stability. These bonds form when a hydrogen atom attached to an electronegative atom (donor) interacts with another electronegative atom (acceptor), typically with energies of 10-40 kJ/mol, though their net contribution to folding is context-dependent due to competition with water solvation. In the protein interior, side-chain hydrogen bonds are stronger by 1-5 kJ/mol compared to solvent-exposed ones, as desolvation penalties are offset by the low environment. For instance, in ribonuclease A, side-chain hydrogen bonds between and aspartate residues help lock distant parts of the chain together, contributing to overall rigidity. Electrostatic interactions, particularly salt bridges between oppositely charged residues like aspartate or glutamate (negative) and or (positive), provide long-range stabilization that is sensitive to and . A involves both ionic attraction and hydrogen bonding components, with typical interaction energies of 5-15 kJ/mol in low-dielectric protein interiors, though effects can reduce this to near-neutral in some cases. These bridges are more prevalent on protein surfaces but can anchor core elements; their disruption often lowers melting temperatures by 2-10°C, as seen in engineered variants of barnase where removing a key Asp-Lys bridge destabilizes the structure. changes that protonate or deprotonate residues can modulate these interactions, influencing tertiary conformation in responsive proteins. Covalent disulfide bonds, formed by the oxidation of thiol groups between cysteine residues, impart exceptional stability to tertiary structures, especially in extracellular proteins exposed to oxidative environments. These S-S linkages, with bond energies around 200 kJ/mol, covalently cross-link distant parts of the chain, reducing loss upon unfolding and preventing misfolding. Disulfide bonds are rare in cytosolic proteins due to reducing conditions but are abundant in secreted ones, where they can constitute up to 10% of s. A classic example is insulin, an extracellular stabilized by three disulfide bonds: two intra-chain in the A subunit and one inter-chain linking A and B subunits, which are essential for its compact fold and ; mutation of these s leads to diabetes-associated instability. Van der Waals interactions and pi-stacking contribute to the tight packing of the hydrophobic core, where atoms approach within 3-4 to enable weak attractive forces. Van der Waals forces, encompassing dispersion and other non-specific attractions, each provide 1-5 kJ/mol but accumulate significantly in densely packed interiors, filling voids and optimizing surface complementarity; in , core packing efficiency reaches 75% via thousands of such contacts. Pi-stacking occurs between aromatic rings of , , or , involving parallel or offset overlaps stabilized by dispersion forces rather than , with energies up to 10 kJ/mol per pair. These interactions are common in protein cores, as in the beta-barrel of , where stacked aromatics enhance stability without covalent links.

Role of molecular chaperones

Molecular chaperones play a crucial role in protein tertiary structure formation by facilitating the correct folding of nascent or unfolded polypeptides, shielding hydrophobic regions to prevent off-pathway aggregation, and providing a controlled environment for conformational rearrangements. These ATP-dependent or independent proteins do not impart the final fold but instead mitigate kinetic traps during folding, ensuring efficient attainment of the native tertiary structure. The family of chaperones, including DnaK in and Hsp70/Hsc70 in eukaryotes, binds transiently to exposed hydrophobic segments of unfolded proteins via an ATP-regulated substrate-binding domain. This binding, which occurs in an ATP-dependent manner, stabilizes extended conformations and prevents intermolecular hydrophobic interactions that lead to aggregation, thereby allowing time for productive folding. ATP hydrolysis locks the substrate in a high-affinity state, while exchange promotes release, cycling the chaperone for reuse; co-chaperones like DnaJ stimulate activity to enhance efficiency. In , the /GroES chaperonin system provides an isolated, cage-like compartment for folding. GroEL forms a cylindrical double-ring structure with a central cavity that accommodates unfolded substrates, binding them to hydrophobic residues lining the inner wall. Upon ATP binding and GroES capping, the cavity expands and becomes hydrophilic, releasing the substrate into a sequestered where it can fold without interference from cellular crowding or aggregation; subsequent and GroES dissociation release the folded protein. This system is essential for approximately 10-15% of bacterial proteins that obligately require chaperonin assistance for tertiary structure attainment. Trigger factor serves as a ribosome-associated chaperone in , aiding co-translational folding of nascent chains emerging from the ribosomal exit tunnel. It docks via the ribosomal protein L23 and extends over the tunnel exit to form a cradle-like structure that confines the nascent polypeptide, preventing premature domain folding or aggregation by countering local compaction and unfolding stable secondary structures as needed. This ribosome-bound action ensures nascent chains adopt compatible conformations early in synthesis, reducing misfolding risks during tertiary structure development. In eukaryotic cells, and form a lectin-based chaperone system in the (ER) specialized for folding. These chaperones bind monoglucosylated N-linked glycans on nascent , retaining them in the ER to allow iterative cycles of folding and ; is membrane-anchored, while is soluble, but both recruit folding enzymes like . This glycan-mediated interaction promotes proper tertiary domain assembly and oligomerization, preventing export of misfolded . Dysfunction or mutations in molecular chaperones underscore their essential role, as evidenced in folding diseases; for instance, in caused by CFTR mutations like ΔF508, impaired chaperone interactions (e.g., reduced or binding) fail to rescue the misfolded protein from degradation, leading to loss of functional tertiary structure and activity. Similarly, mutations in family members have been linked to neurodegenerative disorders where overwhelms chaperone capacity.

Environmental factors

The cellular environment significantly influences the formation and stability of protein tertiary structures through various abiotic factors. in the , where protein concentrations can reach 200–300 mg/mL, promotes compact folding by effects, which entropically favor the native state over extended conformations. This congestion restricts the accessible volume for unfolded chains, accelerating folding rates and stabilizing tertiary interactions without altering the underlying energy landscape. pH and ionic strength modulate electrostatic interactions critical to tertiary structure, as changes in protonation states of residues like histidines and aspartates alter salt bridges and repulsion forces. In hemoglobin, the Bohr effect exemplifies this: at lower , protonation of specific residues stabilizes the deoxy (T-state) quaternary and tertiary conformation, reducing oxygen affinity and facilitating . Higher ionic strength screens these charges, weakening electrostatic contributions and potentially shifting equilibrium toward less stable conformers. Temperature elevations increase thermal motion, disrupting weak non-covalent interactions, while chemical denaturants like and (GdnHCl) specifically target bonds in the protein backbone and side chains, leading to unfolding. penetrates the hydration shell and forms direct bonds with groups, solvating the unfolded state more favorably than the native fold. , in contrast, disrupts via stronger ionic interactions but similarly destabilizes structure by weakening intramolecular bonding networks. The of the cellular compartment controls bond formation, which covalently stabilizes tertiary structures in many extracellular proteins. In the oxidizing of , enzymes like DsbA facilitate pairing, promoting correct folding of secreted proteins. Conversely, the reducing , maintained by thioredoxins and glutaredoxins, prevents aberrant formation, ensuring cytoplasmic proteins remain in reduced, non-covalent tertiary states. Recent studies highlight osmolytes such as N-oxide (TMAO) in counteracting environmental stress on tertiary structure. Under osmotic or , TMAO preferentially excludes from the protein surface, enhancing water-mediated interactions and stabilizing the native fold against denaturation. In and marine organisms, TMAO accumulation protects proteomes during abiotic challenges, with in situ analyses showing it stabilizes over 60% of proteins by reinforcing hydrophobic cores.

Experimental determination

X-ray crystallography

is a primary experimental technique for determining the three-dimensional tertiary structure of proteins at atomic resolution, relying on the of X-rays by ordered protein crystals to reconstruct maps. The process begins with the purification and of the protein, where a concentrated protein is subjected to various conditions such as , temperature, and precipitating agents to form well-ordered crystals, typically containing multiple copies of the protein molecule in a repeating . Once crystals are obtained, they are exposed to a beam of X-rays, which interact with the electrons in the protein atoms, producing a pattern of spots on a detector; the intensity of these spots provides information about the amplitudes of the structure factors, but the phases must be determined separately to compute the . The phase problem is solved using methods such as multiple isomorphous replacement (MIR), where heavy atoms like mercury or are introduced into the protein to create derivative crystals, allowing phase calculation from differences in intensities between native and derivative datasets, or molecular replacement (MR), which uses a known homologous structure as a search model to estimate initial phases. After phasing, the map is fitted with an atomic model, which is iteratively refined to minimize discrepancies between observed and calculated structure factors. The technique achieved its first major milestones in protein structure determination with the elucidation of in 1958 by and colleagues, who reported a 2 Å resolution model revealing the protein's predominantly helical fold and heme-binding pocket, marking the initial atomic-level view of a . This was followed in 1960 by Max Perutz's team solving the structure of at 5.5 Å resolution, later improved, which demonstrated the quaternary arrangement of its four subunits and laid the groundwork for understanding in proteins. These pioneering works, for which Kendrew and Perutz shared the 1962 , established as indispensable for tertiary structure analysis. Protein structures determined by typically achieve resolutions of 1-2 Å, enabling precise placement of atoms and identification of side-chain conformations in small to medium-sized proteins up to about 100 , though larger complexes can be studied with advancements in technology. A key advantage is its ability to provide high-resolution static snapshots of the protein in a crystalline environment, offering detailed insights into active sites, binding, and folding motifs that are crucial for functional studies. However, the method's disadvantages include the necessity for high-quality crystals, which can be challenging to obtain for flexible or proteins, and the provision of only frozen, non-dynamic views that may not fully represent solution-state conformations. Refinement of the atomic model is often performed using software like PHENIX, which automates coordinate, occupancy, and thermal parameter adjustments while incorporating geometric restraints to ensure stereochemical accuracy. This technique is frequently complemented by other methods for validating dynamic aspects of tertiary structure.

Nuclear magnetic resonance spectroscopy

Nuclear magnetic resonance (NMR) spectroscopy determines protein tertiary structures in solution by measuring nuclear spin interactions that report on atomic distances, bond angles, and chemical environments. Key techniques include (NOESY), which detects through-space dipole-dipole interactions between protons typically separated by less than 5 , providing distance restraints essential for modeling three-dimensional folds. Rotating-frame Overhauser Effect Spectroscopy (ROESY) complements NOESY for larger proteins by mitigating artifacts from molecular tumbling rates, yielding similar short-range distance information. Chemical shift indexing, such as the (CSI) method, assigns secondary and tertiary structural elements by comparing observed NMR chemical shifts of backbone atoms (e.g., ^{13}C^\alpha, ^{13}C^\beta, ^{15}N^\mathrm{H}, ^{1}H^\mathrm{N}) to random coil values, with deviations indicating \alpha-helices, \beta-sheets, or turns. These assignments integrate with distance restraints to refine tertiary topology using structure calculation software like CYANA or . Isotope labeling with ^{13}C and ^{15}N is crucial for overcoming spectral overlap in proteins, enabling multidimensional experiments (e.g., ^{1}H-^{15}N HSQC, ^{13}C-edited NOESY) that correlate resonances across nuclei for unambiguous and restraint collection. Uniform labeling incorporates these isotopes into all relevant sites during recombinant expression, while selective labeling targets specific residues to simplify spectra. This approach extends the practical size limit of solution NMR to proteins up to approximately 50 kDa, beyond which signal broadening from slow tumbling limits resolution, though deuteration (^2H) further reduces relaxation and enables studies of systems approaching 100 kDa. NMR uniquely captures dynamic aspects of tertiary structure, revealing conformational flexibility in loops, hinges, or domain interfaces through measurements of longitudinal (R_1) and transverse (R_2) relaxation rates of ^{15}N nuclei. Elevated R_2 rates indicate millisecond-to-microsecond motions, such as loop fluctuations that modulate active sites, while order parameters (S^2) from model-free analysis quantify picosecond-to-nanosecond dynamics, showing restricted motion in core regions versus higher amplitude in solvent-exposed loops. These data complement static structures by illustrating how tertiary architecture accommodates functional motions, as in or ligand binding. An early landmark in NMR-based tertiary structure determination was the solution structure of ubiquitin, a 76-residue protein, solved in the mid-1980s using 2D NOESY and sequential assignment strategies, which revealed its compact \alpha/\beta fold with a mixed parallel/antiparallel \beta-sheet and \alpha-helix. This work, building on resonance assignments from ^{1}H NMR, demonstrated NMR's feasibility for globular proteins and established ubiquitin as a benchmark for validation against X-ray crystallography. Recent advances in solid-state NMR (ssNMR) have expanded applications to larger systems and challenging targets like membrane proteins, where solution methods falter due to insolubility or aggregation. Techniques such as magic-angle spinning and enhance sensitivity, allowing atomic-resolution structures of proteins exceeding 50 kDa in bilayers, including interhelical distances via ^{13}C-^{13}C correlations and torsion angles from chemical shifts. For instance, ssNMR has elucidated the tertiary arrangement of bacterial transporters in native-like membranes, capturing -protein interactions absent in detergent-solubilized studies. These developments integrate and solid-state data for hybrid models, providing comprehensive views of tertiary structure in heterogeneous environments.

Cryo-electron microscopy

Cryo-electron microscopy (cryo-EM) is a powerful technique for determining the three-dimensional structures of proteins, particularly those forming large macromolecular complexes or embedded in membranes. The process begins with the preparation of protein samples, which are applied to a grid and flash-frozen in liquid to form a thin layer of vitreous ice, preserving the native conformation without crystallization. These frozen samples are then imaged using a transmission , where an electron beam passes through the specimen to generate two-dimensional projections. Through single-particle , thousands of such projections are computationally aligned and averaged to reconstruct a high-resolution three-dimensional model of the protein's structure. Significant advancements in cryo-EM resolution were achieved in the with the introduction of direct electron detectors, which capture images with reduced and higher sensitivity, enabling routine atomic-level resolutions below 3 . This "resolution revolution" was recognized by the 2017 awarded to , , and Richard Henderson for their foundational contributions to cryo-EM methodology, including vitreous ice embedding and image processing algorithms. Cryo-EM excels in visualizing protein tertiary structures that exhibit conformational heterogeneity, as it can classify and reconstruct multiple states from a single dataset, and it is particularly advantageous for membrane proteins, which often resist required by methods. For instance, in 2020, cryo-EM structures of the at resolutions around 3 revealed distinct prefusion and postfusion conformational states, providing critical insights into viral entry mechanisms and guiding vaccine design. As of 2025, the integration of has further enhanced cryo-EM workflows, particularly through AI-assisted particle picking, which automates the identification of protein particles in noisy micrographs, significantly improving throughput and accuracy for complex datasets. Cryo-EM maps can also be briefly integrated with data to produce hybrid models that refine local features.

Other biophysical methods

Dual polarization interferometry (DPI) is a surface-sensitive optical technique that measures the thickness, , and mass density of protein layers adsorbed onto waveguides in real time, providing insights into conformational changes and tertiary structure during adsorption or binding events. For instance, DPI has been used to detect structural rearrangements in proteins like β-lactoglobulin at oil-water interfaces, where changes in layer density indicate shifts from native to partially unfolded states. This method is particularly valuable for studying protein-membrane interactions, as demonstrated in analyses of foldamers binding to bilayers, where DPI quantifies conformational adaptations at the nanoscale. Fluorescence resonance (FRET) probes protein tertiary structure by measuring distances (typically 1-10 nm) between donor and acceptor fluorophores attached to specific residues, allowing inference of folding dynamics and domain orientations in solution. In seminal studies, folded and unfolded subpopulations of proteins like cold shock protein B were distinguished during , revealing heterogeneity in tertiary structure formation. FRET has also mapped conformational changes in multidomain proteins, such as voltage-gated calcium channels, by quantifying between labeled cytoplasmic domains. Circular dichroism (CD) spectroscopy assesses protein tertiary structure through near-UV (250-300 nm) spectral signatures arising from the asymmetric environments of aromatic residues (, , ) and bonds, complementing far-UV measurements of secondary structure.30317-1) For example, perturbations in near-UV CD bands indicate tertiary disruptions during or stability assessments, as seen in studies of engineered antibodies where spectral shifts correlate with folding integrity. The technique is widely applied to monitor unfolding transitions, providing quantitative data on the stability of tertiary interactions in solution. Small-angle X-ray scattering (SAXS) yields low-resolution (1-2 nm) envelopes of protein tertiary structure in solution by analyzing scattering patterns from hydrated samples, capturing overall shape, flexibility, and conformational ensembles without . It excels for flexible or multidomain proteins, such as those modeled via ensemble optimization methods, where SAXS data refines predictions of native tertiary arrangements by comparing scattering profiles to theoretical models. Similarly, (SANS) provides analogous low-resolution structural information but leverages neutron contrast variation (e.g., via deuteration) to isolate tertiary features in complexes or crowded environments. SANS has elucidated protein-detergent interactions, revealing how preserve native tertiary folds in proteins. These methods offer indirect probes of tertiary structure, often at lower than primary techniques, and typically require with high- data from NMR or cryo-EM for validation and atomic-level interpretation.

Computational prediction

Traditional algorithms

Traditional algorithms for protein tertiary structure prediction rely on physics-based simulations, , and fragment-based assembly to model three-dimensional conformations from sequences, predating -driven approaches. These methods emerged in the 1980s and 1990s, addressing the challenge of simulating folding pathways or inferring structures from evolutionary relatives. They typically achieve moderate success for proteins with detectable homologs but struggle with novel folds due to computational limitations. Ab initio methods simulate from first principles using empirical s to minimize . These approaches represent the protein as a collection of atoms or residues and apply physics-based potentials to explore conformational space, aiming to identify the native state as the global energy minimum. For instance, the CHARMM (Chemistry at HARvard Macromolecular Mechanics) models bonded and non-bonded interactions, including van der Waals, , and hydrogen bonding terms, to drive or minimization simulations of folding trajectories. Early implementations, such as those in the , successfully predicted small proteins (<100 residues) with root-mean-square deviation (RMSD) values around 5-10 Å from experimental structures, but required significant computational resources for larger systems. Homology modeling, also known as comparative modeling, constructs tertiary structures by aligning a target sequence to homologous proteins with known structures in the (PDB). The process involves identifying templates via sequence similarity searches (e.g., using ), threading the target sequence onto the template backbone to account for insertions/deletions, and refining side-chain placements and loops using energy minimization. Threading, a key variant for low-homology cases, evaluates sequence-structure compatibility with statistical potentials derived from known folds, as pioneered in the early 1990s. This method excels when sequence identity exceeds 30%, yielding models with RMSD <2 Å, but accuracy drops for distant homologs due to alignment errors. Fragment assembly methods, exemplified by the Rosetta algorithm, build structures by combining short segments (3-9 residues) from PDB fragments that match local sequence patterns. These fragments are assembled via Monte Carlo sampling, guided by a scoring function that includes statistical terms for backbone torsion angles, hydrogen bonding, and solvation, alongside physics-inspired energy evaluations. Rosetta iteratively perturbs the chain and accepts low-energy conformations, effectively sampling folding funnels to generate decoys for clustering into native-like models. For small proteins, this approach produced topologies correct within 5-6 Å RMSD in the early 2000s. The Critical Assessment of Structure Prediction (CASP) competitions, initiated in 1994, provided blind benchmarks for these algorithms through the 2010s. In CASP1-CASP5 (1990s), homology and ab initio methods achieved Global Distance Test Total Score (GDT-TS) values around 50 for easy targets with clear templates, while hard targets scored below 30. By CASP9-CASP11 (2010-2014), refinements improved easy-target GDT-TS to 70-80, but free-modeling categories for novel folds remained at 20-40, highlighting persistent gaps. A core challenge for these methods is the Levinthal paradox, which posits that exhaustive sampling of conformational space (∼10^{300} states for a 100-residue protein) would take longer than the universe's age, yet proteins fold in milliseconds. Solutions involve biased sampling techniques like , which explore energy landscapes with funnels leading to the native state, reducing the effective search space to ∼10^{10} semi-compact conformations via rapid collapse and multiple transition pathways. Lattice models demonstrated this resolution, showing folding times scale with chain length rather than exponentially.

AI and machine learning approaches

Artificial intelligence and machine learning have revolutionized protein tertiary structure prediction by leveraging large-scale datasets of known structures and evolutionary information to achieve unprecedented accuracy, often surpassing traditional methods that rely heavily on homology modeling. These approaches employ deep neural networks trained end-to-end on massive protein sequence and structure data, enabling predictions directly from amino acid sequences without requiring experimental templates in many cases. Seminal advancements, such as those from and other research groups, have democratized access to high-fidelity models, accelerating research across biology and medicine. AlphaFold, developed by DeepMind, represents a landmark in this field, with its second iteration () introduced in 2020 and detailed in 2021. The architecture features an Evoformer module that processes multiple sequence alignments (MSAs) and pairwise residue representations through attention-based layers to capture evolutionary and spatial relationships, followed by structure modules that iteratively refine atomic coordinates via invariant point attention. In the Critical Assessment of Structure Prediction () competition held in 2020, achieved median backbone root-mean-square deviation (RMSD) accuracies below 1 Å for many targets, outperforming all competitors and demonstrating near-atomic precision even for proteins without close homologs. This performance stemmed from training on over 100,000 experimentally determined structures from the (), combined with vast genomic sequence data. Building on similar principles, RoseTTAFold, released in 2021 by the Baker lab at the University of Washington, employs a three-track neural network that jointly models sequence, structural distance, and orientation information, allowing for efficient prediction and enabling de novo protein design through diffusion-based sampling of novel folds. Unlike earlier template-dependent methods, RoseTTAFold's end-to-end learning facilitates rapid generation of diverse structures, with accuracies comparable to on CASP14 benchmarks while requiring less computational resources, making it accessible for broader use in structure-based design. Meta AI's ESMFold, introduced in 2022, advances single-sequence prediction by integrating a protein language model () with a folding head that directly outputs 3D coordinates, bypassing the need for time-consuming MSAs. Trained on over 250 million protein sequences from metagenomic data, ESMFold achieves median GDT-TS scores of around 70 on diverse test sets, enabling predictions in seconds per protein—orders of magnitude faster than MSA-based systems—while maintaining high accuracy for monomeric structures. This approach highlights the power of transformer-based language models in capturing sequence patterns for folding without alignment dependencies. AlphaFold 3, unveiled by DeepMind in 2024, extends these capabilities to multimodal predictions, incorporating ligands, nucleic acids, and modified residues into joint complex modeling using a diffusion-based architecture that refines atomic positions progressively. It achieves up to 50% higher accuracy than prior tools on protein-ligand benchmarks, such as PoseBusters, facilitating predictions of biomolecular interactions critical for understanding cellular processes. Subsequent evaluations, including in CASP16 (2024), confirmed AlphaFold 3's leading performance, with high GDT-TS scores across diverse targets, further solidifying deep learning's role in the field. In October 2025, the OpenFold Consortium released OpenFold3, an open-source implementation inspired by AlphaFold 3, enhancing accessibility for non-commercial research and enabling custom fine-tuning for specialized applications. The collective impact of these AI methods has been profound: as of November 2025, AlphaFold has generated predictions for over 200 million protein structures, covering nearly all known sequences in and vastly expanding the 's utility beyond its ~200,000 experimental entries. These predictions have accelerated drug design by enabling virtual screening of protein-ligand interactions, with applications in identifying novel inhibitors for targets like kinases and proteases, reducing the time and cost of lead optimization in pharmaceutical pipelines.

Biological significance

Functional implications

The tertiary structure of proteins plays a crucial role in forming active sites that enable enzymatic catalysis by precisely positioning catalytic residues in three-dimensional space. In serine proteases, such as and , the tertiary fold brings together a catalytic triad—consisting of serine, histidine, and aspartate residues—into a configuration that facilitates nucleophilic attack on peptide bonds, allowing substrate-specific hydrolysis. This spatial arrangement, stabilized by hydrogen bonds and hydrophobic interactions, ensures efficient proton transfer and transition state stabilization, underscoring how tertiary structure directly dictates enzymatic function. Allostery represents another key functional implication of tertiary structure, where ligand binding at one site induces conformational changes that propagate through the protein to modulate activity at a distant site, enabling regulatory control and signal transduction. In G-proteins, such as heterotrimeric GTPases, GTP binding triggers a tertiary rearrangement in the α-subunit, shifting switch regions to disrupt interactions with the βγ-subunits and activate downstream effectors like adenylyl cyclase, thereby amplifying cellular signals in pathways like vision and olfaction. This mechanism highlights the dynamic nature of tertiary folds, which allow proteins to act as molecular switches for precise physiological responses. In multi-domain proteins, tertiary structure manifests as independent folds for each domain, promoting modularity that enhances functional versatility through combinatorial assembly and specialized interactions. Domains typically range from 40 to 350 amino acids and fold autonomously, enabling proteins to perform multiple tasks; for instance, the in signaling proteins binds C-terminal peptides via a conserved carboxylate-binding loop, facilitating targeted protein-protein associations in pathways like synaptic transmission. This modularity allows evolutionary tinkering at domain interfaces while preserving core functions, contributing to the complexity of eukaryotic proteomes where approximately 67% of proteins are multi-domain. Evolutionary conservation of tertiary structures across homologous proteins ensures the preservation of functional scaffolds, even when primary sequences diverge significantly, thereby maintaining core activities over long timescales. Structural analyses reveal that related proteins often retain similar folds—such as or motifs—despite low sequence identity, which supports conserved mechanisms like nucleotide binding in dehydrogenases; this conservation arises from purifying selection on structurally critical residues, linking evolutionary history directly to functional reliability. A prominent example of tertiary structure's functional role is seen in the variable regions of antibodies, where the immunoglobulin fold of VH and VL domains positions six complementarity-determining regions (CDRs) to form a paratope that binds antigens with high specificity and affinity. The tertiary arrangement creates a binding groove or pocket, with CDR-H3 providing the most variability for accommodating diverse epitopes, as in monoclonal antibodies targeting viral proteins; this structural plasticity, including elbow hinge flexibility, enables induced fit adjustments for effective immune recognition and neutralization.

Role in diseases

Disruptions in protein tertiary structure often underlie a range of diseases characterized by misfolding, aggregation, and loss of function, leading to cellular toxicity and organ dysfunction. These pathological conformations deviate from the native fold, promoting aberrant interactions that propagate disease states. In prion diseases, including Creutzfeldt-Jakob disease in humans and bovine spongiform encephalopathy in cattle, the cellular prion protein (PrPC) undergoes a conformational shift from its predominantly α-helical tertiary structure to the β-sheet-enriched PrPSc isoform. This misfolded PrPSc forms aggregates that template further conversion of PrPC, resulting in neuronal damage and spongiform encephalopathy. Alzheimer's disease features the accumulation of β-amyloid (Aβ) plaques, where soluble Aβ monomers misfold into β-sheet-dominated fibrillar structures that disrupt synaptic function and trigger . Familial mutations in the amyloid precursor protein accelerate this tertiary misfolding, exacerbating plaque formation and cognitive decline. Cystic fibrosis arises primarily from the ΔF508 mutation in the (CFTR), which destabilizes the nucleotide-binding domain 1 (NBD1) tertiary structure, causing global misfolding and retention in the . This prevents CFTR trafficking to the plasma membrane, impairing transport and leading to mucus accumulation in lungs and . Serpinopathies, such as alpha-1-antitrypsin deficiency, involve mutations like the Z variant in alpha-1-antitrypsin (A1AT) that trap the protein in a polymerogenic intermediate conformation, characterized by aberrant β-sheet insertion and loop displacement in its tertiary structure. These polymers accumulate in hepatocytes, causing liver damage, while reduced circulating A1AT levels predispose to . Therapeutic strategies targeting tertiary structure disruptions include chemical chaperones, such as 4-phenylbutyrate, which stabilize folding intermediates to enhance mutant protein trafficking in , and proteostasis regulators that modulate chaperone activity or degradation pathways to reduce aggregates in diseases. Molecular chaperones like can briefly mitigate misfolding in these contexts by assisting refolding, though their overload in disease states often exacerbates pathology.

References

  1. [1]
    Biochemistry, Tertiary Protein Structure - StatPearls - NCBI Bookshelf
    A protein's distinctive 3-dimensional configuration, or tertiary structure, arises from interactions between residues as the chain bends and folds in a 3- ...Introduction · Fundamentals · Clinical Significance
  2. [2]
    The Shape and Structure of Proteins - Molecular Biology of the Cell
    Stretches of polypeptide chain that form α helices and β sheets constitute the protein's secondary structure. The full three-dimensional organization of a ...Missing: globule | Show results with:globule
  3. [3]
    The Cyclol Theory and the Globular Proteins - Nature
    The Cyclol Theory and the Globular Proteins. D. M. Wrinch. Nature volume 139, pages 972–973 (1937)Cite this article.
  4. [4]
    How protein chemists learned about the hydrophobic factor
    Dec 31, 2008 · How protein chemists learned about the hydrophobic factor. Charles Tanford,. Corresponding Author. Charles Tanford. candj@dial.pipex.com.
  5. [5]
    A Three-Dimensional Model of the Myoglobin Molecule Obtained by X-Ray Analysis - Nature
    - **Title**: A Three-Dimensional Model of the Myoglobin Molecule Obtained by X-Ray Analysis
  6. [6]
    [PDF] Levinthal's Paradox - TheisticScience.com
    We feel that protein folding is speeded and guided by the rapid formation of local interactions which then determine the further folding of the peptide. This ...Missing: reference | Show results with:reference
  7. [7]
  8. [8]
    Principles that Govern the Folding of Protein Chains - Science
    David S. Eisenberg,. To What Extent is Anfinsen's Thermodynamic Hypothesis Consistent With the Formation and Polymorphism of Amyloid Fibrils?, Journal of ...
  9. [9]
    Proteins as molecular chaperones - Nature
    Ellis, J. Proteins as molecular chaperones. Nature 328, 378–379 (1987). https://doi.org/10.1038/328378a0. Download citation. Issue date: 30 July ...
  10. [10]
    Homologous plant and bacterial proteins chaperone oligomeric ...
    May 26, 1988 · The product of the Escherichia coli groEL gene is essential for cell viability and is required for the assembly of bacteriophage capsids.
  11. [11]
    Two-state models of protein folding kinetics - PMC - NIH
    A two-state kinetic model is justified if protein molecules rapidly equilibrate between different unfolded conformations prior to complete folding.
  12. [12]
    The nature of protein folding pathways - PNAS
    Experimental results show that, contrary to prior belief, proteins are multistate rather than two-state objects. They are composed of separately cooperative ...
  13. [13]
    [PDF] THEORY OF PROTEIN FOLDING: The Energy Landscape Perspective
    Mar 15, 2020 · ABSTRACT. The energy landscape theory of protein folding is a statistical description of a protein's potential surface.
  14. [14]
    Molten globule and protein folding - PubMed
    Author. O B Ptitsyn. Affiliation. 1 Institute of Protein Research, Russian Academy of Sciences, Pushchino, Moscow Region, Russia.
  15. [15]
    Submillisecond protein folding kinetics studied by ultrarapid mixing
    Addition of imidazole dramatically changes the kinetics. At 3 ms, a typical dead time for stopped-flow experiments, the fluorescence is already 90% quenched.
  16. [16]
    How the hydrophobic factor drives protein folding - PNAS
    Oct 17, 2016 · When Kauzmann published his classic 1959 paper (1) on a new hydrophobic factor that drives protein folding, he gave examples from the literature ...
  17. [17]
    Quantitative theory of hydrophobic effect as a driving force of protein ...
    Various studies suggest that the hydrophobic effect plays a major role in driving the folding of proteins. In the past, however, it has been challenging to ...
  18. [18]
    Contribution of hydrogen bonds to protein stability - PMC - NIH
    (1) Hydrogen bonds contribute favorably to protein stability. (2) The contribution of hydrogen bonds to protein stability is strongly context dependent.
  19. [19]
    Hydrogen Bonds: Simple after All? | Biochemistry - ACS Publications
    Apr 20, 2018 · Intermolecular hydrogen bonds have angles of ∼180°, whereas intramolecular hydrogen bonds of similar lengths generally have bent hydrogen bonds.Simple View of Hydrogen... · Hydrogen Bond Energetics... · Supporting Information<|separator|>
  20. [20]
    Protein stabilization by salt bridges: concepts, experimental ...
    Salt bridges in proteins are bonds between oppositely charged residues that are sufficiently close to each other to experience electrostatic attraction.
  21. [21]
    Evaluating the Strength of Salt Bridges: A Comparison of Current ...
    Apr 5, 2014 · Force fields differ in their treatment of electrostatic interactions, including the formation of salt bridges in proteins. Here we conducted ...
  22. [22]
    Review Protein disulfide engineering - ScienceDirect.com
    Jan 21, 2014 · Disulfide bonds are found predominantly in secreted extracellular proteins. The redox environment within the cytosol preserves cysteine ...
  23. [23]
    Insulin Biosynthesis, Secretion, Structure, and Structure-Activity ...
    Feb 1, 2014 · Insulin is a 51-residue protein, a 51-residue anabolic protein, with two chains (A and B) connected by disulfide bonds, and is a ligand that ...INSULIN BIOGENESIS AND... · BIOCHEMISTRY AND... · INSULIN STRUCTURE<|separator|>
  24. [24]
    The Energetic Origins of Pi–Pi Contacts in Proteins - ACS Publications
    We find that neutral pi–pi interactions in proteins are dominated by Pauli repulsion and London dispersion rather than repulsive quadrupole electrostatics.
  25. [25]
  26. [26]
    Protein folding assisted by the GroEL/GroES chaperonin system
    The chaperonin system GroEL/GroES assists in the folding of proteins in the bacterial cytosol. Recent applications of biophysical techniques for the structural ...Missing: seminal paper
  27. [27]
    Concerted Action of the Ribosome and the Associated Chaperone ...
    Oct 12, 2012 · Upon ribosome binding, the chaperone Trigger Factor counters folding also of longer nascent chains, to extents varying between different chain ...Article · Results · Tf Unfolds Nascent Chains
  28. [28]
    Beyond lectins: the calnexin/calreticulin chaperone system of the ...
    Feb 15, 2006 · Calnexin and calreticulin are related proteins that comprise an ER chaperone system that ensures the proper folding and quality control of ...
  29. [29]
    Mechanisms of CFTR Folding at the Endoplasmic Reticulum - Frontiers
    Dec 12, 2012 · ... mutations that disrupt protein folding. It is caused by ... cystic fibrosis mutation. Chem. Biol. 15, 62–69. Pubmed Abstract ...Abstract · Introduction · CFTR Folding · The Role of Chaperones in...
  30. [30]
    Volume Exclusion and Soft Interaction Effects on Protein Stability ...
    Hard interactions, also known as volume exclusion, are the result of impenetrable crowding agents occupying solvent space, removing volume otherwise accessible ...
  31. [31]
    [PDF] Estimation of Macromolecule Concentrations and Excluded Volume ...
    The use of solutions of purified macromolecules that emulate the crowding conditions inferred for cytoplasm is discussed. Keywords: proteins; cytoplasm; ...
  32. [32]
    Beyond the Excluded Volume Effects: Mechanistic Complexity ... - NIH
    The mere presence of high concentrations of crowding agents defines the appearance of the excluded volume effects, which are purely entropic hard-core ...
  33. [33]
    THE BOHR EFFECT - Annual Reviews
    This review examines recent experiments to determine which groups are responsible for the Bohr effect and how these ligand-linked processes are modulated by ...Missing: seminal papers
  34. [34]
    Bohr effect and ionic strength dependence of individual groups
    At the higher ionic strength 10 groups per tetramer contributed to the Bohr effect, whereas 28 groups were contributory at the lower ionic strength.Missing: protein tertiary seminal
  35. [35]
    Urea, but not guanidinium, destabilizes proteins by forming ... - PNAS
    Feb 24, 2009 · The mechanism by which urea and guanidinium destabilize protein structure is controversial. We tested the possibility that these denaturants ...Urea Blocks Amide Hx · Urea H-Bonds To The Peptide... · Urea-Peptide H-Bonding Seems...
  36. [36]
    A study of urea and GnHCl induced protein denaturation
    Urea and GnHCl are the well-known chaotropic agents used in molecular biology and biochemistry to denature the proteins by disrupting the hydrogen bonding ...
  37. [37]
    Disulfide Bond Formation in Prokaryotes: History, Diversity and Design
    The formation of structural disulfide bonds is essential for the function and stability of a great number of proteins, particularly those that are secreted.
  38. [38]
    Review Production of disulfide-bonded proteins in Escherichia coli
    In prokaryotes and eukaryotes, disulfide bond formation is compartmentalized to non-cytoplasmic locations such as the periplasm or ER, respectively. Proteins ...
  39. [39]
    In situ analysis of osmolyte mechanisms of proteome thermal ...
    Feb 29, 2024 · TMAO stabilized the largest number of proteins (67.2% of detected proteins) and showed the strongest stabilization, whereas glycerol at the same ...
  40. [40]
    Trimethylamine N-oxide is a new plant molecule that promotes ...
    May 19, 2021 · TMAO operates as a protective osmolyte in plants, promoting appropriate protein folding and as an activator of abiotic stress–induced gene expression.Results · Tmao Enhances Appropriate... · Tmao Promotes Abiotic...
  41. [41]
    x Ray crystallography - PMC - PubMed Central - NIH
    The aim of x ray crystallography is to obtain a three dimensional molecular structure from a crystal. A purified sample at high concentration is crystallised ...
  42. [42]
    [PDF] Macromolecular Structure Determination by X-ray Crystallography
    This article outlines the principles and the key methods involved in macromolec- ular structure determination by X-ray crystallography. X-radiation was ...
  43. [43]
    Introduction to phasing - PMC - NIH
    The use of heavy-atom substitution to solve the phase problem was invented very early on by small-molecule crystallographers, for example the isomorphous ...
  44. [44]
    [PDF] Phase Problem in X-ray Crystallography, and Its Solution
    MAD experiments are most commonly performed using protein in which all methionine residues have been modified to contain a selenium instead of a sulfur atom.
  45. [45]
    John Kendrew and myoglobin: Protein structure determination ... - NIH
    The essay reviews John Kendrew's pioneering work on the structure of myoglobin for which he shared the Nobel Prize for Chemistry in 1962.
  46. [46]
    Methods for Determining Atomic Structures - PDB-101
    X-ray crystallography can provide very detailed atomic information, showing every atom in a protein or nucleic acid along with atomic details of ligands, ...
  47. [47]
    X-Ray Crystallography vs. NMR Spectroscopy - News-Medical
    Oct 30, 2019 · Disadvantages of X-ray Crystallography and NMR Spectroscopy​​ The types of sample that can be analyzed are limited. In particular, membrane ...
  48. [48]
    PHENIX: a comprehensive Python-based system for ... - NIH
    Jan 22, 2010 · PHENIX has been developed to provide a comprehensive system for macromolecular crystallographic structure solution with an emphasis on the automation of all ...<|control11|><|separator|>
  49. [49]
    NMR Spectroscopy Principles, Interpreting an NMR Spectrum and ...
    Feb 22, 2024 · NOESY spectra contain peaks that correlate pairs of nearby nuclei (typically, separated by less than 5-6 Å). Differently to COSY, NOESY ...
  50. [50]
    a fast and simple method for the assignment of protein secondary ...
    The chemical shift index: a fast and simple method for the assignment of protein secondary structure through NMR spectroscopyMissing: tertiary | Show results with:tertiary
  51. [51]
    Secondary and Tertiary Structural Effects on Protein NMR Chemical ...
    Recent theoretical developments permit the prediction of 1 H, 13 C, 15 N, and 19 F nuclear magnetic resonance chemical shifts in proteins.
  52. [52]
    Specific isotopic labelling and reverse labelling for protein NMR ...
    Dec 16, 2022 · ... isotopes 13C and/or 15N. The standard approach is to uniformly label a protein with one or both of these nuclei such that all C and/or N ...
  53. [53]
    15 N/ 13 C Labeling of Proteins for Structural Investigations Using ...
    We have developed a method that has proven highly reliable for the deuteration and triple labeling (2H/15N/13C) of a broad range of proteins by recombinant ...
  54. [54]
    Extending the size limit of protein nuclear magnetic resonance - PNAS
    In addition, the NMR signals have been spread effectively into additional 15N and/or 13C dimensions, alleviating spectral degeneracy. These advances, together ...
  55. [55]
    An introduction to NMR-based approaches for measuring protein ...
    Here we provide an introduction to NMR-based approaches for studying protein dynamics, highlighting eight distinct methods with recent examples.
  56. [56]
    Characterization of backbone dynamics using solution NMR ...
    Nov 4, 2021 · This protocol consists of a step-by-step setup and analysis scheme for the measurements of 15N protein NMR backbone relaxation experiments using ...
  57. [57]
    Functional protein dynamics on uncharted time scales detected by ...
    Aug 14, 2019 · Commonly used NMR spin relaxation in solution provides valuable information on very fast and slow motions but is insensitive to the intermediate ...
  58. [58]
    The Quiet Renaissance of Protein NMR - PMC - NIH
    By the mid 1980's Wüthrich and others determined the structures of a number of small proteins. This was enabled by the development of two-dimensional NMR ...
  59. [59]
    (PDF) Ubiquitin: A Small Protein Folding Paradigm - ResearchGate
    Aug 6, 2025 · For the past twenty years, the small, 76-residue protein ubiquitin has been used as a model system to study protein structure, stability, ...
  60. [60]
    Solid-state NMR and Membrane Proteins - PMC - NIH
    Solid-state NMR has advanced sufficiently to enable the structures of ... A solid-state NMR index of helical membrane protein structure and topology.
  61. [61]
    Solid-state NMR structure determination of a membrane protein in E ...
    Nov 1, 2023 · This cellular ssNMR structure determination paves the way for analyzing the atomic structural details for membrane proteins in native cellular membranes.
  62. [62]
    Solid-state NMR spectroscopy based atomistic view of a membrane ...
    Aug 27, 2019 · We devise a methodology that combines hydrogen-deuterium (H/D) exchange and solid-state NMR (SSNMR) to follow membrane protein unfolding in lipid membranes at ...
  63. [63]
    Solid-state NMR of membrane proteins in situ - ScienceDirect.com
    Here, we review recent advances in nuclear magnetic resonance (NMR) that enable the examination of membrane protein structure and activity in situ, within ...
  64. [64]
    Single-particle cryo-electron microscopy | Nature Methods
    Dec 30, 2015 · The grid is then plunged into a cryogen such as liquid ethane, flash-freezing it and trapping the particles in a thin film of vitreous ice. In ...
  65. [65]
    Structure Determination by Single-Particle Cryo-Electron Microscopy
    We provide here a brief overview of single-particle Cryo-EM and show how Cryo-EM has revolutionized structural investigations of membrane proteins. We also show ...
  66. [66]
    The Nobel Prize in Chemistry 2017 - Popular information
    Using cryo-electron microscopy, researchers can now freeze biomolecules midmovement and portray them at atomic resolution. This technology has taken ...Missing: tertiary | Show results with:tertiary
  67. [67]
    Membranes under the Magnetic Lens: A Dive into the Diverse World ...
    By removing the need to form well-diffracting crystals, cryo-EM has eliminated a major bottleneck for determining the structure of membrane proteins. This has ...2.2. 1. Cryo-Em Overview · Figure 3 · Figure 10
  68. [68]
    Distinct conformational states of SARS-CoV-2 spike protein | Science
    Here, we report two cryo–electron microscopy structures derived from a preparation of the full-length S protein, representing its prefusion (2.9-angstrom ...
  69. [69]
    Artificial intelligence in cryo-EM protein particle picking
    Jan 16, 2025 · This review critically examines the evolution and current state of cryo-EM particle picking methods, with an emphasis on the impact of AI.Abstract · Introduction · Emergence of AI in particle... · comparative study of the AI...
  70. [70]
    combining Cryo-TEM, X-ray crystallography, and NMR - PMC - NIH
    The combination of Cryo-TEM and other methods such as X-ray crystallography, nuclear magnetic resonance spectroscopy, and molecular dynamics modeling
  71. [71]
    A comparison between dual polarization interferometry (DPI) and ...
    The measurements give information about both the thickness and refractive index of an adsorbed protein film, which can be used to calculate the adsorbed amount.
  72. [72]
    Structural rearrangement of β-lactoglobulin at different oil-water ...
    Aug 2, 2011 · DPI measurements suggest that β-Lg formed a thin dense layer at emulsion droplet surfaces. The effects of high temperature and the presence of ...
  73. [73]
    Differential membrane binding of α/β-peptide foldamers - NIH
    Dual polarisation interferometry was used to study the binding of each peptide to two different model membrane systems designed to mimic either the plasma ...
  74. [74]
    FRET Spectrometry: A New Tool for the Determination of Protein ...
    Förster resonance energy transfer (FRET) is an exquisitely sensitive method for detection of molecular interactions and conformational changes in living cells.Missing: seminal papers
  75. [75]
    Single-molecule protein folding: Diffusion fluorescence ... - PNAS
    Folded and denatured subpopulations were observed by fluorescence resonance energy transfer (FRET) measurements on freely diffusing single protein molecules.
  76. [76]
    Fluorescence Resonance Energy Transfer (FRET) Indicates That ...
    Here, we have taken the approach of measuring FRET between fluorescent proteins inserted into pairs of α1S cytoplasmic domains. Expression of these constructs ...
  77. [77]
    Circular dichroism spectroscopy in protein engineering and ...
    CD spectroscopy is a reliable and convenient method for evaluating higher order protein structure that has proven valuable in the development of protein ...Circular Dichroism... · 2. Cd Analysis Of New... · 4. Application To...
  78. [78]
    Tools and methods for circular dichroism spectroscopy of proteins
    Jun 15, 2021 · This tutorial review describes the background to the technique of CD spectroscopy and good practice methods for high quality data collection.
  79. [79]
    A practical guide to small angle X-ray scattering (SAXS) of flexible ...
    Sep 14, 2015 · Small-angle X-ray scattering (SAXS) is a biophysical method to study the overall shape and structural transitions of biological macromolecules in solution.
  80. [80]
    [PDF] Structural characterization of flexible proteins using SAXS
    A new approach, ensemble optimization method (EOM), is proposed to quantitatively characterize flexible proteins in solution using small-angle X-ray scattering ...
  81. [81]
    Small-angle neutron scattering contrast variation studies of ...
    Small-angle neutron scattering (SANS) has been a beneficial tool for studying the structure of biological macromolecules in solution for several decades.
  82. [82]
    Small-angle neutron scattering study of the structure of protein ...
    Small-angle neutron scattering (SANS) was used to study the structure of protein/sodium dodecylsulfate complexes. Two water soluble proteins, bovine serum ...
  83. [83]
    Applying Small Angle X-Ray Scattering (SAXS) To Protein Structure ...
    Here, we describe how to obtain protein structure predictions, compare them to experimental SAXS data and improve models to reflect experimental information ...
  84. [84]
    Ab initio protein structure prediction - ScienceDirect.com
    All protein structure prediction techniques comprise a representation of the protein, a force field commensurate with this representation, a technique for ...
  85. [85]
    Recent Progress of Protein Tertiary Structure Prediction - PMC
    Feb 13, 2024 · In this review, we start with an overview of the history of protein structure prediction, including template-based modeling (TBM) and template- ...
  86. [86]
    [PDF] Ab Initio Protein Structure Prediction
    Well-known examples of such all-atom physics-based force fields include AMBER (Weiner et al. 1984;. Cornell et al. 1995; Duan and Kollman 1998), CHARMM (Brooks ...
  87. [87]
    Ab Initio Protein Structure Prediction: Methods and challenges
    Dec 16, 2013 · This chapter presents the state of the art of the protein structure prediction (PSP) problem from a computational science perspective.
  88. [88]
    Advances in Protein Tertiary Structure Prediction
    Protein structure prediction methods have been mainly classified into three types: ab Initio folding, comparative (homology) modeling and threading.
  89. [89]
    Protein Structure Prediction Using Rosetta - ScienceDirect.com
    In the Rosetta method, short fragments of known proteins are assembled by a Monte Carlo strategy to yield native-like protein conformations.
  90. [90]
    [PDF] Protein Structure Prediction Using Rosetta
    1. Structure Prediction by Fragment Assembly. The fragment assembly approach has multiple benefits for de novo pro- tein structure prediction. First, and ...
  91. [91]
    Protein Structure Prediction Center
    CASP aims to establish the current state of the art in protein structure prediction and identify progress, using blind prediction to test methods.CASP13 · CASP_Commons · CASP16 · CASP15
  92. [92]
    CASP3-11 Results Published in E-Life - Institute for Protein Design
    Nov 4, 2015 · In the early 1990s, researchers in the field of protein structure prediction were challenged by the problem of how to impartially judge the ...
  93. [93]
    How does a protein fold? - Nature
    May 19, 1994 · To determine how the Levinthal paradox is resolved, we use a lattice Monte Carlo model in which the global minimum (native state) is known. The ...
  94. [94]
    The Levinthal paradox: yesterday and today - ScienceDirect
    The essential concept introduced by Levinthal is that the appropriate point of reference for protein folding is a random search problem. Taken literally, as it ...
  95. [95]
    A Universal Allosteric Mechanism for G Protein Activation - PMC
    G proteins play a central role in signal transduction and pharmacology. Signaling is initiated by cell-surface receptors, which promote GTP binding and the ...Missing: implications | Show results with:implications
  96. [96]
    The Nature of Protein Domain Evolution: Shaping the Interaction ...
    In this review, we aim to describe the basic concepts of protein domain evolution and illustrate recent developments in molecular evolution.
  97. [97]
    Proteome-wide structural analysis quantifies structural conservation ...
    Proteins related by evolution can have similar tertiary structures, even when the sequence similarities are statistically marginal or seemingly nonexistent ( ...
  98. [98]
    Antibody Structure and Function: The Basis for Engineering ...
    The intact antibody molecule shown in Figure 1 has three functional components, two Fragment antigen binding domains (Fabs) and the fragment crystallizable (Fc) ...
  99. [99]
    Toward a mechanism of prion misfolding and structural models of ...
    Physicochemical variables known to influence misfolding are reviewed to help elucidate the mechanism of prionogenesis and identify salient features of PrP(Sc), ...
  100. [100]
    New insights into structural determinants of prion protein folding and ...
    Prions are the etiological agent of fatal neurodegenerative diseases called prion diseases or transmissible spongiform encephalopathies.
  101. [101]
    Biophysical Studies of the Amyloid β-peptide: Interactions With Metal ...
    Sep 23, 2013 · Alzheimer's disease is the most common of the protein misfolding ("amyloid") diseases. The deposits in the brains of afflicted patients ...
  102. [102]
    Effects of Familial Alzheimer's Disease Mutations on the Folding ...
    Oct 6, 2022 · Familial Alzheimer's disease (FAD) mutations of the amyloid β-peptide (Aβ) are known to lead to early onset and more aggressive Alzheimer's disease.
  103. [103]
    Requirements for efficient correction of ΔF508 CFTR ... - PubMed - NIH
    Jan 20, 2012 · Misfolding of ΔF508 cystic fibrosis (CF) transmembrane conductance regulator (CFTR) underlies pathology in most CF patients.
  104. [104]
    Restoration of NBD1 thermal stability is necessary and ... - PubMed
    Jan 16, 2015 · The ΔF508 mutation has two major impacts on CFTR: reduced thermal stability of NBD1 and disruption of its interface with membrane-spanning domains (MSDs).
  105. [105]
    Deficiency Mutations of Alpha-1 Antitrypsin. Effects on Folding ...
    Misfolding, polymerization, and defective secretion of functional alpha-1 antitrypsin underlies the predisposition to severe liver and lung disease in alpha-1 ...<|control11|><|separator|>
  106. [106]
    Molecular pathogenesis of alpha-1-antitrypsin deficiency - PubMed
    The mechanism of α1-AT deficiency provides a paradigm for a wider group of conditions that we have termed the serpinopathies. We also examine the strategies ...<|control11|><|separator|>
  107. [107]
    Innovative strategies to treat protein misfolding in inborn errors of ...
    Apr 1, 2014 · Innovative strategies to treat protein misfolding in inborn errors of metabolism: pharmacological chaperones and proteostasis regulators.
  108. [108]
    Biological and chemical approaches to diseases of proteostasis ...
    We review how small molecules can enhance proteostasis by binding to and stabilizing specific proteins (pharmacologic chaperones) or by increasing the ...