Fact-checked by Grok 2 weeks ago

Molecule

A molecule is an electrically entity consisting of more than one . Rigorously, it corresponds to a on the that is deep enough to confine at least one vibrational , ensuring through characteristic chemical bonds. Molecules represent the fundamental units of chemical substances that exhibit distinct physical and chemical properties, distinguishing them from individual atoms or larger aggregates like ionic lattices. Molecules form through the sharing of s between atoms via covalent bonds in which atoms share electron pairs to achieve stable electron configurations. The three-dimensional structure of a molecule arises from the spatial arrangement of these bonds, governed by principles such as the valence-shell , which minimizes electron repulsion to predict geometries like tetrahedral (e.g., in , CH₄) or bent (e.g., in , H₂O). This structure directly influences molecular properties, including polarity, reactivity, and interactions with other molecules, which are crucial for processes ranging from chemical reactions to biological functions. Molecules are classified into types based on composition and complexity, such as diatomic molecules (e.g., O₂ or N₂, consisting of two identical atoms) for elemental gases, and polyatomic molecules for more complex entities like those in compounds. Molecules of elements contain only one kind of atom, while molecules of compounds incorporate two or more different elements, as seen in water (H₂O), where two hydrogen atoms bond with one oxygen atom. Beyond simple molecules, macromolecules—large polymers like proteins or DNA—play essential roles in biology, providing structural support, catalyzing reactions, and storing genetic information. In chemistry, molecules underpin the composition of gases, liquids, and molecular solids, enabling the diversity of materials and driving advancements in fields from materials science to pharmaceuticals.

Etymology and History

Etymology

The term "molecule" originates from the Latin word molecula, a form of moles, meaning "" or "barrier," implying a small mass or particle. This etymological root entered scientific discourse through molécule, which first appeared in the late around 1678, initially denoting an extremely minute particle without a strictly chemical . By the , chemists began adopting the term more formally to describe the fundamental units of , marking its transition from philosophical to empirical usage in emerging chemical theories. A pivotal milestone in the term's evolution occurred in 1811 when Italian chemist Amedeo Avogadro employed "molecule" (molécule in French) in his seminal paper to distinguish it from "atom." Avogadro used the word to refer to the smallest particle of a substance that retains its chemical properties, often composed of multiple atoms, thereby resolving ambiguities in gas volume laws and atomic combinations. This application built upon earlier ideas like John Dalton's atomic theory (1808) by distinguishing molecules from atoms, resolving ambiguities in gas laws and influencing later chemical developments. The modern etymological significance of "molecule" also reflects a conceptual bridge from ancient ideas of indivisible particles, such as ' atoms in the 5th century BCE, to 19th-century scientific precision, emphasizing bound aggregates rather than elemental units. This linguistic evolution underscores the term's role in formalizing the particulate nature of matter beyond mere philosophical speculation.

Historical Development

The concept of the molecule emerged in the early as part of John Dalton's , outlined in his 1808 publication A New System of Chemical Philosophy, where he distinguished between atoms as the fundamental units of elements and compound atoms—later recognized as molecules—formed by the union of atoms in fixed proportions. This framework provided the first systematic basis for understanding chemical composition, positing that molecules retain their integrity during chemical reactions unless decomposed by specific means. Building on Dalton's ideas, proposed in 1811 that equal volumes of gases at the same temperature and pressure contain equal numbers of molecules, introducing a critical distinction between atoms and molecules for elements that form diatomic gases like oxygen and . Avogadro's hypothesis clarified the role of molecules in gaseous reactions and laid groundwork for determining relative molecular masses, though it faced initial resistance and confusion with atomic weights. By the mid-19th century, revitalized Avogadro's work through his 1858 pamphlet Sunto di un corso di filosofia chimica, which resolved ambiguities in distinguishing atomic from molecular weights by applying Avogadro's principle to vapor densities and chemical formulas. Cannizzaro's approach enabled chemists to assign consistent molecular weights to compounds, such as establishing as H₂O rather than HO, and facilitated the development of the periodic table by in 1869. This clarification marked a pivotal advancement in molecular theory, shifting focus from empirical formulas to structural insights and standardizing molecular weights as twice the vapor density relative to hydrogen. In the late 19th century, structural theories advanced with August Kekulé's 1865 proposal of the cyclic hexagonal structure for , depicted as alternating single and double bonds between six carbon atoms, which explained the molecule's stability and isomerism despite its C₆H₆ formula. This model, introduced in Bulletin de la Société Chimique de Paris, represented a foundational step in by emphasizing valence and connectivity, influencing the representation of molecular architectures. Concurrently, emerged as a tool for probing molecular composition; for instance, the observed in 1885 for spectra provided early empirical data on energy levels, initially for atomic but extending to molecular 's band spectra in subsequent studies, revealing vibrational and rotational structures. The advent of in the early 20th century, pioneered by in 1912 and refined by William Henry and William Lawrence Bragg, allowed direct visualization of atomic arrangements in crystals, such as in simple salts, confirming molecular packing and bond lengths for the first time. The 20th century integrated into molecular theory, beginning with Gilbert N. Lewis's paper "The Atom and the Molecule," which introduced the electron-pair model for , where shared electron pairs form stable linkages, as in the H–H molecule with two electrons between protons. This cubic atom visualization evolved into the , providing a qualitative basis for molecular stability. In the , formalized these concepts; and London's 1927 valence bond treatment of the hydrogen molecule used wave mechanics to explain formation as a between ionic and covalent states, yielding a of approximately 4.7 , close to experimental values. This work, published in Zeitschrift für Physik, marked the quantum mechanical foundation of molecular bonding, enabling predictions of molecular geometries and energies that supplanted classical models. By the mid-20th century, methods determined the first protein structures, such as in 1958 by , revealing helical motifs and confirming molecular complexity at the atomic level.

Definition and Fundamentals

Core Definition

A molecule is defined as an electrically neutral entity consisting of more than one , where the atoms are bound together by chemical bonds, forming the smallest unit of a that retains its composition and chemical properties. This definition emphasizes that a molecule must correspond to a stable depression on the , deep enough to confine at least one vibrational state, ensuring its existence as a entity under normal conditions. Key characteristics of molecules include their electrical neutrality, discrete spatial arrangement of atoms, and relative stability, which allow them to maintain integrity in typical chemical environments. Molecules can be classified as diatomic, such as gas (H₂), which consists of two identical atoms, or polyatomic, like (H₂O), involving multiple atoms of different elements. They may be homonuclear, where all atoms are of the same element (e.g., O₂), or heteronuclear, involving different elements (e.g., CO₂). These examples illustrate how molecules serve as the fundamental building blocks of matter in covalent compounds. Molecules are distinguished from ions, which are charged species formed by the gain or loss of electrons, either as single atoms or groups of atoms; unlike molecules, ions do not maintain electrical neutrality. Single atoms, such as helium (He), represent uncombined elements and lack the multi-atomic structure of molecules. Even large entities like deoxyribonucleic acid (DNA), considered a giant molecule or macromolecule, fit the molecular definition as they comprise covalently bonded atoms in a stable, discrete arrangement, though they can contain up to tens of billions of atoms. Typically, molecules range from 2 to about 10⁶ atoms, but there is no strict upper limit on size, as evidenced by synthetic and biological macromolecules.

Prevalence and Composition

Molecules constitute the primary structural units of all on Earth, as well as the vast majority of gases and liquids, with notable exceptions including pure metals and certain ionic compounds that form lattice structures rather than discrete molecular entities. In biological contexts, these molecules are predominantly built from four elements—carbon (C), hydrogen (H), oxygen (O), and nitrogen (N)—which together comprise about 96% of the dry mass of living organisms, enabling the formation of complex structures like proteins, carbohydrates, lipids, and nucleic acids. Beyond biology, molecules in the atmosphere are chiefly diatomic, such as N₂ and O₂, while silicon (Si) plays a central role in mineral composition, forming silicate networks that account for over 90% of the Earth's crust by mass, second only to oxygen in abundance. In terms of prevalence, diatomic molecules dominate the composition of Earth's gaseous environments; for instance, dry air consists of approximately 78% N₂ and 21% O₂ by volume, underscoring their role in atmospheric stability and respiration. Polyatomic molecules, by contrast, are far more common in liquids and many solids, including (H₂O), which covers 71% of Earth's surface, and polymers that form the basis of and hydrocarbons. This distribution highlights molecules' versatility across phases of matter, with monatomic forms like (e.g., Ar at 0.93% of air) representing minor fractions in natural settings. While many molecules are heteronuclear compounds incorporating multiple elements, homonuclear elemental molecules consisting of a single element are also common. Examples include diatomic gases such as N₂ and O₂, and polyatomic species like white phosphorus (P₄) and elemental sulfur (S₈). More exotic elemental forms exist, such as fullerenes—cage-like carbon structures exemplified by buckminsterfullerene (C₆₀), first identified in 1985 through laser vaporization experiments on graphite. Furthermore, compounds of noble gases, long considered chemically inert, include xenon difluoride (XeF₂), synthesized in 1962 by direct reaction of xenon and fluorine under controlled conditions, marking an early breakthrough in noble gas chemistry. These examples illustrate the diversity of molecular formation, often requiring specific conditions or reactivity to stabilize.

Chemical Bonding

Covalent Bonding

A is a formed by the sharing of one or more pairs of electrons between two atoms, allowing each atom to achieve a more stable . These bonds are classified by the number of shared electron pairs: a single covalent bond involves two electrons (one pair), a shares four electrons (two pairs), and a shares six electrons (three pairs). Covalent bonds can be nonpolar, where electrons are shared equally between atoms of similar , or polar, where sharing is unequal due to differences in , typically with a difference less than 1.7 on the Pauling scale. Covalent bonds form through the overlap of atomic orbitals from adjacent atoms, creating regions of high between the nuclei. The strongest type, a (σ) bond, results from head-on overlap of orbitals, such as s-s or s-p, providing direct linkage along the axis. Pi (π) bonds, found in double and triple bonds, arise from sideways overlap of p orbitals, adding strength but restricting rotation. The Shell Electron Pair Repulsion ( predicts by assuming electron pairs around a central atom repel each other to minimize ; for example, in (CH₄), four bonding pairs arrange in a tetrahedral shape with bond angles of approximately 109.5°. The strength of covalent bonds is quantified by , the energy required to break the bond homolytically into radicals, typically ranging from 150 to 1000 kJ/ depending on the atoms involved. For instance, the C-H bond in has a dissociation energy of approximately 410 kJ/, reflecting its stability. Bond strength increases with (single < double < triple) and decreases with greater atomic size or lower electronegativity differences, as electron sharing becomes less effective. Representative examples illustrate these principles: the hydrogen molecule (H₂) features a nonpolar single sigma bond formed by 1s orbital overlap. In dioxygen (O₂), a double bond consists of one sigma bond from end-on p-orbital overlap and one pi bond from side-on overlap, resulting in a bond energy of about 498 kJ/mol. Organic molecules, such as alkanes, rely on chains of single C-C and C-H sigma bonds, enabling diverse structures while maintaining molecular integrity.

Non-Covalent Interactions

Non-covalent interactions are intermolecular or intramolecular forces that arise between molecules or within a molecule without the sharing of electrons, distinguishing them from that form the primary structural framework of molecules. These interactions, often weaker and more reversible, play crucial roles in determining molecular conformations, assembly, and properties in biological and chemical systems. Unlike ionic lattices, where strong electrostatic attractions dominate extended crystalline structures, non-covalent interactions in molecules typically operate over longer distances and contribute to dynamic stability rather than rigid frameworks. The main types of non-covalent interactions include electrostatic interactions (such as ion-ion and ion-dipole attractions), hydrogen bonding, van der Waals forces, and pi-pi stacking. Hydrogen bonding occurs when a hydrogen atom covalently bonded to an electronegative atom (such as oxygen or nitrogen) interacts electrostatically with another electronegative atom, forming a directional link with typical energies around 20 kJ/mol in simple systems like the water dimer. For instance, in the water dimer (H₂O)₂, the interaction energy is approximately -21 kJ/mol, stabilizing dimeric structures through partial charge attractions. Van der Waals forces encompass dipole-dipole interactions between polar molecules and London dispersion forces arising from transient induced dipoles in nonpolar molecules, with overall energies ranging from 1 to 40 kJ/mol across non-covalent types. These forces are exemplified in noble gas dimers, such as Ar₂, where London dispersion dominates the weak binding of about 1 kJ/mol, highlighting purely nonpolar attractions. Pi-pi stacking involves the overlap of electron clouds from aromatic rings, driven by dispersion and electrostatic components, with interaction energies typically 5-15 kJ/mol depending on orientation and distance. These interactions generally operate over distances of 0.2 to 1 nm, longer than the 0.1-0.2 nm typical for covalent bonds, allowing for flexibility in molecular arrangements. In biological contexts, hydrogen bonds are essential for stabilizing protein folding, where networks of such interactions (each ~20 kJ/mol) maintain secondary structures like alpha helices and beta sheets. Similarly, in DNA, hydrogen bonding facilitates base pairing—adenine-thymine via two H-bonds and guanine-cytosine via three—ensuring the double helix's specificity and stability without forming covalent links between strands. Van der Waals and pi-pi interactions further support these processes, such as pi-pi stacking in nucleobase dimers contributing to DNA's helical stacking. Overall, non-covalent interactions enable the reversible assembly and functional dynamics of molecular systems, with their cumulative effects often rivaling covalent strengths in large assemblies.

Molecular Representation

Empirical and Molecular Formulas

The empirical formula of a compound represents the simplest whole-number ratio of atoms of each element present in the substance. For instance, has the empirical formula CH_2O, indicating a 1:2:1 ratio of carbon, hydrogen, and oxygen atoms, respectively. This formula provides the relative proportions derived from elemental analysis but does not specify the actual number of atoms in a single . In contrast, the molecular formula indicates the exact number of atoms of each element in one molecule of the compound. For glucose, the molecular formula is C_6H_{12}O_6, which is six times the empirical formula since its molar mass is approximately 180 g/mol, compared to 30 g/mol for CH_2O. The molecular formula can be derived from the empirical formula by multiplying its subscripts by a whole-number factor n, determined as n = \frac{M_m}{M_e}, where M_m is the measured molar mass of the compound and M_e is the molar mass of the empirical formula. Historically, such formulas aided early chemists in estimating molecular weights through methods like vapor density measurements. Empirical formulas for organic compounds containing carbon, hydrogen, and oxygen are commonly determined via combustion analysis, where a known mass of the sample is burned in excess oxygen, producing carbon dioxide and water whose masses are measured to calculate the percentages of each element. These mass percentages are then converted to moles and simplified to the lowest whole-number ratio. To obtain the molecular formula, high-resolution mass spectrometry is used to measure the precise molecular mass, allowing identification of the correct multiple of the empirical formula based on the mass-to-charge ratio of the molecular ion. A key limitation of both empirical and molecular formulas is their inability to distinguish between isomers, which are compounds with the same formula but different atomic arrangements. For example, the formula C_3H_6 applies to both (a straight-chain alkene) and (a cyclic alkane), requiring additional structural analysis to differentiate them.

Structural and Graphical Formulas

Structural formulas depict the arrangement and connectivity of atoms within a molecule, explicitly showing the bonds between them to convey how atoms are linked, beyond the mere composition indicated by molecular formulas. For water, the structural formula H–O–H illustrates the two covalent bonds between the oxygen atom and the two hydrogen atoms, clarifying that the hydrogens are attached to the oxygen rather than to each other. Condensed structural formulas offer a more compact notation by grouping atoms and using subscripts or parentheses to represent chains or branches, such as CH₃COOH for acetic acid, where the CH₃ group is bonded to the COOH carboxyl unit. Graphical representations build on these by visualizing electron distribution and skeletal frameworks. Lewis dot structures, also known as electron dot diagrams, represent valence electrons as dots surrounding atomic symbols, with pairs of dots indicating either shared bonds or lone pairs on atoms; a single bond is shown as two electrons (one pair). In organic chemistry, skeletal formulas simplify depiction of carbon-based molecules by drawing lines for carbon-carbon bonds, implying a carbon atom at each line end or junction, while hydrogens are omitted but understood to fill carbon's tetravalency, and other heteroatoms are explicitly labeled. These notations are essential for distinguishing constitutional isomers, molecules with identical molecular formulas but different atom connectivities, which lead to distinct properties. For example, both ethanol and dimethyl ether have the formula C₂H₆O, but ethanol's structure (CH₃CH₂OH) shows a carbon-carbon bond with a terminal hydroxyl group (C–C–O–H), whereas dimethyl ether (CH₃OCH₃) features an oxygen bridging two methyl groups (C–O–C), as revealed by their structural formulas. Advanced graphical tools like ball-and-stick models enhance visualization by portraying atoms as colored spheres connected by rods representing bonds, allowing perception of spatial relationships in three dimensions without implying atomic sizes.

Structural Properties

Size and Dimensions

Bond lengths represent the fundamental scale of molecular structure, quantifying the internuclear distances in covalent bonds. These lengths typically range from about 70 pm for strong bonds like H-H to over 200 pm for weaker single bonds involving larger atoms, such as I-I at 266 pm. For example, the H-H bond in the diatomic hydrogen molecule measures 74 pm, while the C-C single bond in ethane is 154 pm. Such values are determined primarily through rotational spectroscopy, where the moments of inertia derived from microwave spectra yield precise equilibrium bond distances via the relation B = \frac{h}{8\pi^2 c \mu r^2}, with B as the rotational constant, \mu the reduced mass, and r the bond length. Molecular dimensions extend beyond covalent bonds to encompass the overall spatial extent, often defined using van der Waals radii, which approximate the effective size of atoms in non-bonded interactions. The van der Waals radius of hydrogen is 120 pm, leading to an effective diameter for the H₂ molecule of approximately 3.1 Å when accounting for the bond length and atomic contributions. For larger molecules, dimensions scale roughly with the number of atoms, following a near-linear increase in end-to-end distance for chain-like structures or a cubic root scaling for compact volumes, as seen in polymers where size grows as N^{1/2} to N^{1/3} (with N as atom count) depending on density. For small rigid molecules, dimensions are fixed, while for flexible macromolecules, they vary with conformation. This scaling underscores how simple diatomics occupy volumes on the order of 0.1 nm³, while polyatomic molecules expand accordingly. Molecules are classified by size into small molecules (typically under 1 nm, including diatomics like N₂ at ~0.3 nm) and macromolecules (often exceeding 10 nm, such as large proteins). Small molecules, with fewer than ~50 atoms, fit within angstrom scales, whereas macromolecules like or proteins can span tens of nanometers due to their polymeric nature. Conformational factors significantly influence effective dimensions, particularly in flexible macromolecules. In proteins, the unfolded state adopts a random coil configuration, with a root-mean-square end-to-end distance scaling approximately as 0.2 nm × N^{0.6}, reaching about 11 nm for a 300-residue chain. Upon folding into a compact native structure, the dimensions shrink to 3–6 nm in diameter due to intramolecular interactions, significantly reducing the effective hydrodynamic volume, often by a factor of 5-10 or more (corresponding to >80% reduction based on the cube of linear dimensions) and altering transport and reactivity properties.

Geometry and Shape

The three-dimensional of a molecule refers to the spatial arrangement of its atoms, which is primarily determined by the electronic structure and bonding interactions around central atoms. This arrangement influences the molecule's physical and chemical properties, such as reactivity, polarity, and . Factors like repulsions and orbital overlaps dictate the bond angles and overall , enabling predictions of molecular configurations from simple models. The Valence Shell Electron Pair Repulsion (VSEPR) model provides a foundational approach to predicting by considering the repulsion between pairs in the valence shell of a central atom. Developed by Ronald J. Gillespie and Ronald S. Nyholm, this theory posits that pairs—whether or lone pairs—arrange themselves to minimize mutual repulsion, leading to specific geometries. For instance, in (CH₄), classified as AX₄ under VSEPR notation (where A is the central atom and X represents a pair), the four pairs adopt a tetrahedral arrangement with bond angles of approximately 109.5°. This model extends to more complex cases, such as (H₂O, AX₂E₂), where two lone pairs distort the tetrahedral electron geometry into a bent molecular with a bond angle of about 104.5°. The VSEPR theory's simplicity makes it widely applicable for main-group elements, though it has limitations for transition metals or cases involving d-orbitals. Hybridization theory complements VSEPR by explaining the contributions to through the mixing of s, p, and sometimes d orbitals to form hybrid orbitals with equivalent energy and spatial orientation. introduced this concept in his valence bond framework, proposing that carbon in utilizes sp³ hybrid orbitals—formed from one s and three p orbitals—to achieve the tetrahedral with 109.5° , aligning with experimental observations. In (C₂H₄), each carbon atom employs sp² hybridization, resulting in a trigonal planar arrangement around each carbon with 120° bond and a that restricts rotation. Similarly, in (C₂H₂), sp hybridization yields a linear with 180° . For aromatic compounds like (C₆H₆), the sp² hybridization of each carbon atom ensures a planar, hexagonal ring with trigonal planar coordination, stabilizing the delocalized π-system. These hybridizations arise from the need to maximize orbital overlap for stronger bonds, directly linking electronic configuration to observed shapes. Chirality in molecules arises when a molecule is non-superimposable on its , often due to the presence of a chiral center, such as a carbon atom bonded to four different substituents. This phenomenon, first rationalized by and Joseph Achille Le Bel through their tetrahedral carbon model, leads to stereoisomers known as enantiomers that exhibit identical physical properties but rotate plane-polarized light in opposite directions. A classic example is (2-hydroxypropanoic acid), where the central carbon bearing a , hydroxyl, methyl, and carboxyl group creates two enantiomers: (R)-lactic acid and (S)-lactic acid. These enantiomers are crucial in biochemistry, as enzymes often distinguish between them, influencing processes like . Beyond tetrahedral carbons, chirality can also stem from axial or helical elements, but asymmetric centers remain the most common source in organic molecules. Conformational analysis examines how molecules adopt different spatial arrangements through rotation about single bonds, without breaking bonds, resulting in conformers that interconvert rapidly at room temperature. In ethane (C₂H₆), the staggered conformation—where hydrogen atoms on adjacent carbons are maximally separated by a 60° dihedral angle—represents the energy minimum, while the eclipsed conformation, with overlapping hydrogens at 0° dihedral, is a high-energy transition state separated by a rotational barrier of approximately 2.9 kcal/mol. This barrier, attributed to torsional strain from electron repulsion in eclipsed positions, was first evidenced through entropy measurements by John D. Kemp and Kenneth S. Pitzer, confirming hindered rotation around the C-C bond. Such analysis is vital for understanding reactivity in larger molecules, like butane, where gauche and anti conformers affect stability due to steric interactions.

Analytical Methods

Spectroscopy Techniques

Spectroscopy techniques play a crucial role in elucidating molecular and composition by probing interactions between molecules and , revealing information about vibrational, electronic, and nuclear spin states. These methods allow scientists to identify functional groups, assess connectivity, and infer environmental effects without destroying the sample, making them indispensable for , inorganic, and biochemical analysis. Common techniques exploit distinct physical principles: (IR) and focus on vibrational transitions, (NMR) on nuclear spins, and ultraviolet-visible (UV-Vis) on electronic excitations. Infrared (IR) spectroscopy measures the absorption of IR by molecules, corresponding to changes in during vibrational modes, which provides a for functional groups. For instance, the O-H stretching vibration in alcohols and typically appears as a broad band between 3200 and 3600 cm⁻¹ due to , while C=O stretches in carbonyl compounds occur around 1650–1750 cm⁻¹. This technique is widely used for qualitative identification, as each exhibits characteristic absorption frequencies, enabling rapid screening in synthetic chemistry and . Fourier-transform IR (FTIR) enhances resolution and speed, allowing analysis of complex mixtures by deconvoluting overlapping bands. Nuclear magnetic resonance (NMR) exploits the magnetic properties of atomic nuclei, such as ¹H and ¹³C, to determine the chemical environment of atoms through s and coupling patterns. The in ¹H NMR, measured in parts per million () relative to , reflects deshielding effects from electronegative atoms or ; for example, protons in methyl groups attached to alkanes appear at 0.9–1.8 , while those in aldehydes are downfield at 9–10 . Multidimensional NMR variants, like COSY and HSQC, map atom connectivity and correlations, providing detailed structural insights for biomolecules and pharmaceuticals. High-field spectrometers improve sensitivity, enabling studies of dilute samples in solution or solid states. Ultraviolet-visible (UV-Vis) spectroscopy detects electronic transitions in molecules, particularly those involving π to π* or n to π* excitations, which are sensitive to conjugation and chromophores. In conjugated systems like dyes, extended π bonds lower the energy gap, shifting absorption maxima to longer wavelengths (bathochromic shift); for example, absorbs at 255 nm, but polyenes can extend into the visible region, imparting color. This method quantifies concentrations via Beer's law and assesses purity or reaction progress, though it offers limited structural detail compared to vibrational techniques. Applications in biochemistry include monitoring through aromatic residue absorptions around 280 nm. Raman spectroscopy complements IR by detecting of light, arising from changes in molecular rather than , thus highlighting symmetric vibrations often inactive in IR. For instance, the C=C stretch in alkenes produces a strong Raman band around 1600–1680 cm⁻¹, while polar bonds like O-H show weaker signals. Surface-enhanced Raman scattering (SERS) amplifies signals for trace detection, making it valuable for studies and . The orthogonality of selection rules allows combined IR-Raman analysis to achieve near-complete vibrational spectra, enhancing structural elucidation in polymers and biological tissues.

Diffraction and Imaging

Diffraction and techniques enable the direct experimental determination of molecular structures by analyzing the patterns produced when waves—such as X-rays or —scatter off atomic electron clouds. These methods provide spatial information complementary to spectroscopic approaches, revealing lengths, angles, and overall geometries with high precision. Unlike theoretical predictions, yields empirical maps or reconstructed images that validate or refine molecular models. X-ray crystallography remains the gold standard for resolving atomic structures in crystalline samples, where a beam of X-rays interacts with the periodic lattice to produce diffraction spots. These patterns are analyzed via Fourier transform to generate electron density maps, allowing the placement of atoms within the molecule. A seminal application was the elucidation of the DNA double helix in 1953, where fiber diffraction data from oriented DNA samples, collected by Rosalind Franklin and Maurice Wilkins, provided key helical parameters and base-pairing insights at approximately 2 Å resolution, enabling James Watson and Francis Crick to propose the iconic B-form structure. This breakthrough demonstrated how X-ray methods can uncover macromolecular architectures essential for biological function. Modern refinements, including synchrotron sources, routinely achieve resolutions below 1 Å for small molecules, though protein crystals often limit to 1.5–2.5 Å due to disorder. Electron diffraction, particularly in the gas phase, offers a route to structure determination for volatile molecules without the need for . Here, a collimated beam scatters off an ensemble of randomly oriented gaseous molecules, yielding a from the diffraction intensity that encodes interatomic distances. Pioneered by Hermann Mark and Richard Wierl in 1930, the technique first quantified s in small organics like (CCl₄), revealing C–Cl distances of about 1.76 and validating tetrahedral . It excels for non-crystalline systems, providing accuracies to 0.01 , though it is best suited for molecules up to ~20 atoms due to vibrational averaging in the gas phase. Cryogenic electron microscopy (cryo-EM) has emerged as a powerful for biomolecules in near-native states, especially large protein complexes that resist . Samples are flash-frozen in vitreous to preserve , then imaged with low-dose electrons to minimize damage; thousands of 2D projections are computationally aligned and reconstructed into a map. Advancements in direct detectors and phase plates since the have pushed resolutions to near-atomic levels (2–4 ), as exemplified by the 2013 cryo-EM structure of the 70S ribosome at 3.3 , which resolved helices and protein side chains to reveal functional dynamics. This method has since enabled structures of over biomolecular assemblies, often surpassing for flexible targets. Despite their strengths, diffraction and imaging techniques share key limitations rooted in sample preparation and physics. All require some degree of molecular ordering—crystalline arrays for , gaseous dispersion for , or vitrified ensembles for cryo-EM—excluding truly amorphous or dynamic systems without averaging artifacts. is fundamentally capped by and sample quality; methods theoretically approach 0.5 but practically stall at 1 for imperfect crystals due to and mosaicity, while cryo-EM struggles below 2 for small proteins lacking stabilizing scaffolds. These constraints often necessitate hybrid approaches with computational modeling to interpret incomplete densities.

Theoretical Foundations

Quantum Mechanical Models

The quantum mechanical description of molecules begins with the time-independent , which governs the behavior of the multi-particle wavefunction \Psi(\mathbf{r}_1, \mathbf{r}_2, \dots, \mathbf{R}_1, \mathbf{R}_2, \dots), where \mathbf{r}_i are coordinates and \mathbf{R}_I are coordinates. The operator \hat{H} includes terms for all electrons and nuclei, as well as Coulombic potential energies between all charged particles: \hat{H} = -\sum_i \frac{\hbar^2}{2m_e} \nabla_i^2 - \sum_I \frac{\hbar^2}{2M_I} \nabla_I^2 + \sum_{i<j} \frac{e^2}{r_{ij}} + \sum_{I<J} \frac{Z_I Z_J e^2}{R_{IJ}} - \sum_{i,I} \frac{Z_I e^2}{r_{iI}}, yielding \hat{H} \Psi = E \Psi. This equation is analytically solvable only for the simplest molecular system, the hydrogen molecular ion H_2^+, which consists of two protons and one electron. In prolate spheroidal coordinates, the electronic Schrödinger equation for H_2^+ admits exact solutions, revealing bound states with bonding and antibonding characteristics that explain the stability of the ion at equilibrium bond lengths around 1.06 Å. For more complex molecules, the multi-electron wavefunction becomes intractable due to electron correlation and the need to satisfy the Pauli exclusion principle via antisymmetrization, necessitating approximations. A foundational approximation is the Born-Oppenheimer (BO) separation, which exploits the mass disparity between electrons (m_e) and nuclei (M_I \gg m_e) to decouple electronic and nuclear motions. In the BO framework, the nuclear wavefunction is treated parametrically, fixing nuclear positions \mathbf{R} to solve the electronic Schrödinger equation first: \left[ -\sum_i \frac{\hbar^2}{2m_e} \nabla_i^2 + V_{ee} + V_{eN}(\mathbf{R}) \right] \Psi_{el}(\mathbf{r}; \mathbf{R}) = E_{el}(\mathbf{R}) \Psi_{el}(\mathbf{r}; \mathbf{R}), where V_{ee} and V_{eN} are electron-electron and electron-nuclear potentials. The resulting E_{el}(\mathbf{R}) forms the for subsequent nuclear motion. This approximation, introduced in , enables the quantitative prediction of molecular geometries and vibrational spectra, with corrections for non-adiabatic effects typically small for ground states. Building on the BO approximation, (MO) theory describes the electronic wavefunction as a of single-particle molecular orbitals, \psi_k(\mathbf{r}), each approximated via the (LCAO) method: \psi_k = \sum_\mu c_{\mu k} \phi_\mu, where \phi_\mu are atomic basis functions centered on nuclei. The coefficients c_{\mu k} are variationally optimized to minimize the energy in the Hartree-Fock self-consistent approach. For the molecule H_2, the simplest diatomic, the ground-state MO is a bonding \sigma_g orbital from the in-phase LCAO of 1s atomic orbitals, lowering the energy by about 4.7 eV relative to separated atoms, while the antibonding \sigma_u^* orbital raises it; occupancy of the bonding MO yields a of 1. This framework, developed by Hund, Mulliken, and others in the late 1920s and early 1930s, captures distributions and is foundational for understanding molecular spectra and reactivity. In contrast, valence bond (VB) theory constructs the wavefunction from localized bonding pairs, emphasizing atomic-like orbitals that overlap to form covalent bonds, with spin alignment per the Pauli principle. For H_2, Heitler and London in derived the ground-state wavefunction as a symmetric of Heitler-London functions, \Psi = \frac{1}{\sqrt{2(1+S^2)}} [\phi_A(1)\phi_B(2) + \phi_B(1)\phi_A(2)] [\alpha(1)\beta(2) - \beta(1)\alpha(2)]/\sqrt{2}, where \phi_{A,B} are 1s orbitals on protons A and B, and S is the overlap integral; this yields a of approximately 3.14 eV, refined later by ionic terms. VB theory excels in incorporating for conjugated systems: in , the \pi-electron structure is a of two equivalent Kekulé forms, delocalizing the six electrons over the ring and stabilizing the molecule by about 36 kcal/mol relative to a single Kekulé structure, as quantified by Pauling in 1931. This concept rationalizes aromatic stability without invoking extended orbitals, though VB and MO theories converge in hybrid descriptions for complex molecules.

Computational Approaches

Ab initio methods form the cornerstone of computational for predicting molecular properties from first principles, without empirical parameters. These approaches solve the approximately by expanding the molecular wavefunction in a basis of orbitals. The Hartree-Fock (HF) method is the foundational technique, approximating the many- wavefunction as a single and optimizing orbitals self-consistently to minimize energy, though it neglects electron correlation beyond mean-field interactions. Post-Hartree-Fock methods address this limitation; for instance, second-order Møller-Plesset (MP2) incorporates electron correlation by treating the difference between the exact and the Hartree-Fock reference as a , yielding improved energies for small molecules like or . While accurate, ab initio methods scale poorly with system size, limiting routine applications to molecules with fewer than 100 atoms due to computational cost. Density functional theory (DFT) offers a more efficient alternative for larger systems, reformulating the many-electron problem in terms of the rather than the wavefunction, as per the Hohenberg-Kohn theorems. like B3LYP combine exact Hartree-Fock exchange with exchange-correlation approximations from the generalized gradient approximation, providing reliable geometries and energies for organic molecules and complexes at a fraction of cost—for example, B3LYP optimizes bond lengths in to within 0.01 of experiment. Widely adopted in software such as Gaussian and , DFT enables simulations of systems up to thousands of atoms, though it struggles with strong correlation or dispersion in cases like van der Waals complexes. Molecular dynamics (MD) simulations extend computational approaches to time-dependent behavior, evolving atomic positions according to classical Newtonian mechanics using force fields derived from quantum calculations or . This method captures dynamic processes like conformational changes; for , MD trajectories reveal pathways from unfolded to native states, as demonstrated in simulations of the 36-residue villin headpiece folding on microsecond timescales. sampling techniques, such as replica exchange, overcome energy barriers to access rare events, providing insights into folding funnels and intermediate states in proteins like chignolin. Post-2020 advances have integrated artificial intelligence (AI) to accelerate predictions, particularly for biomolecular structures. AlphaFold, developed by DeepMind, uses deep learning on evolutionary data to predict protein tertiary structures with atomic accuracy, achieving a median backbone RMSD of 0.96 Å in CASP14—surpassing traditional methods for cases without homologs. This AI-driven approach has revolutionized structural biology, enabling rapid modeling of complexes and influencing drug design, while extensions like AlphaFold-Multimer handle protein-protein interactions. In May 2024, AlphaFold 3 was released, extending predictions to joint structures of proteins with DNA, RNA, ligands, and ions, with improved accuracy for biomolecular complexes.

References

  1. [1]
    molecule (M04002) - IUPAC
    An electrically neutral entity consisting of more than one atom ( ). Rigorously, a molecule, in which must correspond to a depression on the potential energy ...
  2. [2]
    Chapter 4 - Atoms, Molecules, and Ions - CHE 110 - Textbook
    Aug 18, 2025 · A molecule is the smallest part of a substance that has the physical and chemical properties of that substance. In some respects, a molecule is ...
  3. [3]
    Molecular Structure & Bonding - MSU chemistry
    The three dimensional shape or configuration of a molecule is an important characteristic. This shape is dependent on the preferred spatial orientation of ...
  4. [4]
    Atoms, Molecules, and Compounds - University of Hawaii at Manoa
    Molecules of compounds have atoms of two or more different elements. For example, water (H2O) has three atoms, two hydrogen (H) atoms and one oxygen (O) atom.
  5. [5]
    Molecules - Chembook
    Molecules need a combination of atoms to be complete. The minimum is two atoms. Two atoms that are bound together via a chemical bond make a molecule.
  6. [6]
    Molecule - Etymology, Origin & Meaning
    Apparently coined in "On Macro-molecules, with the Determinations of the Form of Some of Them," by Anglo-Irish physicist G. Johnstone Stoney (1826–1911).
  7. [7]
  8. [8]
    Amedeo Avogadro | Biography, Law, Discoveries, & Facts - Britannica
    To distinguish between atoms and molecules of different kinds, Avogadro adopted terms including molécule intégrante (the molecule of a compound), molécule ...
  9. [9]
    Amedeo Avogadro - Science History Institute
    In 1811 Avogadro hypothesized that equal volumes of gases at the same temperature and pressure contain equal numbers of molecules. From this hypothesis it ...
  10. [10]
    Were molecules called atoms in the 19th century?
    Jan 17, 2024 · According to Online Etymology Dictionary: Molecule it can be traced back to Réné Descartes. However the modern use of the word was given by ...
  11. [11]
    John Dalton and the London atomists: William and Bryan Higgins ...
    Aug 20, 2014 · In 1808 John Dalton published A New System of Chemical Philosophy, which described principles such as the uniqueness of atoms of the same ...
  12. [12]
    Amedeo Avogadro - Le Moyne
    Our hypothesis, supposing it well-founded, puts us in a position to confirm or rectify his results from precise data, and, above all, to assign the magnitude of ...
  13. [13]
    Stanislao Cannizzaro | Science History Institute
    In 1858 Cannizzaro outlined a course in theoretical chemistry for students at the University of Genoa—where he had to teach without benefit of a laboratory.Missing: source | Show results with:source
  14. [14]
    Stanislao Cannizzaro - Linda Hall Library
    Jul 13, 2017 · In his 1858 paper, called "Sunto de un corso di filosofia chimica" (“Sketch of a course in chemical philosophy”), Cannizzaro attempted to ...Missing: source | Show results with:source
  15. [15]
    [PDF] The 150th Anniversary of the Kekul Benzene Structure
    Sep 24, 2014 · The first page of the 1865 article.[4]. Angewandte. Chemie. 47. Angew ... That first short French paper on benzene theory[4, 10, 14] is.
  16. [16]
    Chemistry's visual origins - Nature
    May 5, 2010 · In 1865, Kekulé proposed that benzene's structure was a hexagonal ring of six tetravalent carbon atoms, each able to form four bonds: one to a ...Missing: original | Show results with:original
  17. [17]
    Spectroscopy and Quantum Mechanics - MIT
    Finally in 1885, J.J. Balmer showed that the wavelengths of the visible spectral lines of atomic hydrogen, now known as the Balmer series, could be represented ...
  18. [18]
    The birth of X-ray crystallography - Nature
    Nov 7, 2012 · A century ago this week, physicist Lawrence Bragg announced an equation that revolutionized fields from mineralogy to biology.
  19. [19]
    THE ATOM AND THE MOLECULE. - ACS Publications
    Diverse self-assembled molecular architectures promoted by C–H···O and C–H···Cl hydrogen bonds in a triad of α-diketone, α-Ketoimine, and an imidorhenium ...
  20. [20]
    A brief history of macromolecular crystallography, illustrated by a ...
    We attempt to construct a genealogy tree of the principal lineages of protein crystallography, leading from the founding members to the present generation.
  21. [21]
    Atoms, molecules and ions | CPD article - RSC Education
    Jun 20, 2019 · Particles can be atoms, molecules or ions. · Atoms are single neutral particles. · Molecules are neutral particles made of two or more atoms ...
  22. [22]
    [PDF] Introduc)on to Biological Macromolecules
    – Smallest molecule: H2. – Largest molecules: DNAs (tens of billions of atoms!) • Biological macromolecule: a large and complex molecule with specific ...<|control11|><|separator|>
  23. [23]
    Mineral resource of the month: silicon - USGS Publications Warehouse
    The second most abundant element in Earth's crust and more than 25 percent of the crust by weight, silicon is one of the most useful elements to humans.
  24. [24]
    The Atmosphere: Getting a Handle on Carbon Dioxide
    Oct 9, 2019 · By volume, the dry air in Earth's atmosphere is about 78.08 percent nitrogen, 20.95 percent oxygen, and 0.93 percent argon. A brew of trace ...Missing: N2 | Show results with:N2
  25. [25]
    C60: Buckminsterfullerene - Nature
    Nov 14, 1985 · During experiments aimed at understanding the mechanisms by which long-chain carbon molecules are formed in interstellar space and ...
  26. [26]
    Xenon Difluoride and the Nature of the Xenon-Fluorine Bond - Science
    Xenon reacts with fluorine to form XeF2 which can be isolated before it reacts with fluorine to form XeF4. The linear configuration of XeF2 with the 2.00-A ...Missing: XeF2 | Show results with:XeF2
  27. [27]
    Molecular Interactions (Noncovalent Interactions) - Loren Williams
    Jun 10, 2024 · The enthalpy of a molecular interaction between two non-bonded atoms typically ranges from 1 to 10 kcal/mole (4 to 42 kJ/mole), where the lower ...
  28. [28]
    [PDF] The water dimer II: Theoretical investigations - ScienceDirect.com
    CCpol-8s dimer potential. They obtained an accurate water dimer interaction energy, De=-21.0. kJ/mol, with this new flexible CCpol-8sf potential, compared to ...
  29. [29]
    [PDF] Noncovalent Interactions in Extended Systems Described by the ...
    Noncovalent interactions, such as hydrogen-bonding, π-π, and van der Waals, consist of electrostatic and dispersion forces. The strongest noncovalent ...
  30. [30]
    [PDF] Non-Covalent Short Range Interactions
    These short range forces arise from the overlap of electron wave functions. Interactions of dipolar nature are classified further into strong hydrogen bonds and ...Missing: stacking | Show results with:stacking
  31. [31]
    [PDF] Study Of Non-Covalent Interactions In The Building Blocks Of ...
    Jun 5, 2023 · These interactions allow for the precise pairing of complementary nucleotides, which is critical for DNA replication and the transmission of ...
  32. [32]
    3.2 Determining Empirical and Molecular Formulas - Chemistry 2e
    Feb 14, 2019 · The elemental makeup of a compound defines its chemical identity, and chemical formulas are the most succinct way of representing this ...
  33. [33]
    empirical formula (E02063) - IUPAC Gold Book
    Formed by juxtaposition of the atomic symbols with their appropriate subscripts to give the simplest possible formula expressing the composition of a compound.
  34. [34]
    6.8: Calculating Empirical Formulas for Compounds
    Jul 28, 2025 · An empirical formula tells us the relative ratios of different atoms in a compound. The ratios hold true on the molar level as well.
  35. [35]
    Empirical and Molecular Formulas - LabXchange
    In the early 1800s, French chemist Joseph Gay-Lussac, whom we met in Unit 2, developed the process of combustion analysis as a way to calculate the ratio of ...
  36. [36]
    3.8: Determine Empirical Formula from Combustion Analysis
    Feb 11, 2023 · In this section, we will explore how to derive the chemical formulas of unknown substances from experimental mass measurements.
  37. [37]
    Determination of the Molecular Formula by High Resolution Mass ...
    May 30, 2020 · High resolution mass spectrometry can determine molecular formulas by distinguishing between the masses of compounds based on the isotopic ...
  38. [38]
    Chemical Formulas - FSU Chemistry & Biochemistry
    Another way to represent water would be its structural formula: This formula indicates that both of the hydrogens are attached to the oxygen but not each other ...
  39. [39]
    Lewis dot structure definition
    A way of representing atoms or molecules by showing electrons as dots surrounding the element symbol. One bond is represented as two electrons.
  40. [40]
    Illustrated Glossary of Organic Chemistry - Skeletal formula (skeletal ...
    Skeletal formula (skeletal structure; line-angle formula): A representation of molecular structure in which covalent bonds are shown as lines.
  41. [41]
    Compare-Contrast-Connect: Chemical Structures—Visualizing the ...
    Ball and stick models are three-dimensional models where atoms are represented by spheres of different colors and bonds are represented by sticks between the ...
  42. [42]
    List of Experimental Diatomic Bond Lengths - CCCBDB
    List of Experimental Diatomic Bond Lengths ; 0.917 · 1.564 · 1.361 · 1.267 ...
  43. [43]
    Experimental data for H 2 (Hydrogen diatomic) - CCCBDB
    Calculated geometries for H2 (Hydrogen diatomic). Bond descriptions. Examples: C-C single bond, C=C, double bond, C#C triple bond, C:C aromatic ...Missing: covalent | Show results with:covalent
  44. [44]
  45. [45]
    Van Der Waals Radius of the elements - Photographic Periodic Table
    Van Der Waals radius varies; Hydrogen is 120 pm, Zinc is 139 pm, Helium is 140 pm, Oxygen is 152 pm, and Carbon is 170 pm.Missing: typical | Show results with:typical
  46. [46]
    van der Waals Radii of Free and Bonded Atoms from Hydrogen (Z ...
    Aug 29, 2024 · In this work, we systematically review different definitions and methodologies to establish the free and bonded vdW radii for hydrogen.
  47. [47]
    Linear scaling methods for electronic structure calculations and ...
    Linear scaling methods using semi-empirical Hamiltonians allow one to perform simulations involving up to a thousand atoms on small workstations.
  48. [48]
    How big is the “average” protein? - Bionumbers book
    The typical protein length is about 300 amino acids in prokaryotes and 400 in eukaryotes, with a diameter of 3-6 nm.
  49. [49]
    Cellular and molecular scales - Nexus Wiki - ComPADRE
    mostly hydrogen, carbon, nitrogen, oxygen, sodium, potassium, and calcium are typically about 0.1 nm in ...Missing: classification diatomic macromolecules
  50. [50]
    The relative scale of biological molecules and structures - Nature
    Cells range from 1 μm to hundreds of μm. A DNA double helix is about 10 nm wide, while a nucleus is about 10 μm.
  51. [51]
    Size and Sequence and the Volume Change of Protein Folding - PMC
    Mar 29, 2011 · Our approach presents the advantage that the overall fold of the protein remains the same, while the size changes significantly, and the ...<|control11|><|separator|>
  52. [52]
    Under‐folded proteins: Conformational ensembles and their roles in ...
    Conformational ensembles of under‐folded proteins can be classified as transient (folding and misfolding intermediates) and permanent (IDPs and stable ...
  53. [53]
    Inorganic stereochemistry - Quarterly Reviews, Chemical Society ...
    Inorganic stereochemistry. R. J. Gillespie and R. S. Nyholm, Q. Rev. Chem. Soc., 1957, 11, 339 DOI: 10.1039/QR9571100339. To request permission to reproduce ...
  54. [54]
    Full article: THE EARLY HISTORY OF STEREOCHEMISTRY
    The purpose of this chapter is to describe the observations and reasoning that led Pasteur, van't Hoff, and Le Bel to make the epochal discoveries.Missing: origin | Show results with:origin
  55. [55]
    Quantum Monte Carlo methods for the solution of the Schrödinger ...
    This chapter discusses quantum Monte Carlo methods for the solution of the schrödinger equation for molecular systems.
  56. [56]
    Analytical solution of the Schrödinger equation for the hydrogen ...
    Aug 6, 2015 · The full analytical solution of the Schrödinger equation for the hydrogen molecular ion H_2^+ (special case of the quantum tree-body problem with the Coulomb ...
  57. [57]
    Accurate solutions of the Schrödinger and Dirac equations of H2+ ...
    Jun 5, 2012 · This review summarizes the studies of the exact solutions of the Schrödinger and Dirac equations of H 2 + in non-relativistic, relativistic, non ...
  58. [58]
    [PDF] On the Quantum Theory of Molecules
    In order to determine the eigenfunctions and thereby the transition probabilities only to the zeroth approximation, the energy calculation must be carried out ...
  59. [59]
    Electronic Population Analysis on LCAO–MO Molecular Wave ...
    LCAO molecular orbital overlap populations give in general much more flexible and widely useful measures of the non‐Coulombic parts of covalent bond energies.
  60. [60]
    [PDF] Robert S. Mulliken - Nobel Lecture
    In the case of MO's, the so-called LCAO approximations are rather gener- ally familiar. However, when people have talked about the electronic struc- tures of ...
  61. [61]
    [PDF] Valence bond theory. A tribute to the pioneers of 1927-1935 - iupac
    Heitler and London's 1927 paper on the hydrogen moleculae has been described (ref. 4) as. “the greatest single contribution to the clarification of the ...
  62. [62]
    39. The Fifth Paper - Linus Pauling and The Nature of the Chemical ...
    The most public demonstration of the power of Pauling's resonance ideas came when he used them to solve one of the oldest problems in organic chemistry.
  63. [63]
    Ab initio quantum chemistry: Methodology and applications - PNAS
    May 10, 2005 · Hartree–Fock theory is the simplest method of this type, involving optimization of a single determinant; however, as mentioned above, its ...
  64. [64]
    [PDF] Møller–Plesset perturbation theory - SMU
    how the Hartree–Fock (HF) method can be corrected for electron pair ... MP2 is the least costly ab initio method to correct. HF results for correlation ...
  65. [65]
    Thirty years of density functional theory in computational chemistry
    This review benchmarks a total of 200 density functionals on a molecular database (MGCDB84) of nearly 5000 data points.
  66. [66]
    Molecular dynamics simulations of protein folding from the transition ...
    The simulation results indicate that the tight packing of the side ... References. References. 1. A R Fersht Curr Opin Struct Biol 5, 79–84 (1995). Go ...
  67. [67]
    Protein-Folding Dynamics: Overview of Molecular Simulation ...
    May 5, 2007 · In this review, we present algorithms for MD and their extensions and applications to protein-folding studies, using all-atom models with explicit and implicit ...
  68. [68]
    Highly accurate protein structure prediction with AlphaFold - Nature
    Jul 15, 2021 · AlphaFold greatly improves the accuracy of structure prediction by incorporating novel neural network architectures and training procedures ...