Fact-checked by Grok 2 weeks ago

Biochemistry

Biochemistry is the branch of that studies the chemical substances and processes occurring in living organisms, integrating principles from and chemistry to understand life at the molecular level. It examines the structure, properties, and interactions of key biomolecules, including proteins, nucleic acids (such as and ), carbohydrates, and , which form the foundation of cellular function and organization. The field aims to explain biological phenomena through chemical mechanisms, revealing how these molecules enable processes like , , and genetic information transfer. The discipline of biochemistry emerged in the late 19th and early 20th centuries from chemical analyses of biological tissues, marking a shift from descriptive to molecular explanations of . Pioneering work, such as Eduard Buchner's 1897 demonstration of cell-free fermentation, established enzymes as catalysts for biochemical reactions, laying the groundwork for modern enzymology. By the early 1900s, advancements in isolating biological molecules and the advent of journals like Zeitschrift für Physiologische Chemie (founded 1877) solidified biochemistry as a distinct field, blending organic, inorganic, and physical chemistry with biology. This evolution accelerated in the mid-20th century with discoveries in molecular biology, including the structure of DNA in 1953, which illuminated the chemical basis of heredity. Biochemistry encompasses several core areas, including metabolism, which investigates how organisms convert food into energy and building blocks through interconnected pathways like and the ; enzymology, focusing on enzymes that accelerate reactions and are classified into six main categories such as oxidoreductases and hydrolases; and , exploring how nucleic acids store and transmit genetic information. Additional subfields include structural biochemistry, which analyzes biomolecular architectures using techniques like , and , which studies energy flow in cells. These areas highlight biochemistry's role in elucidating dynamic cellular processes, from protein synthesis to . The importance of biochemistry lies in its applications across diverse sectors, driving innovations in through targeting metabolic pathways and disease mechanisms, as well as in via that enhance nutrient efficiency. In health sciences, it informs by revealing how drugs interact with biomolecules, supports in assessing chemical toxicity, and advances through techniques like for development. Furthermore, biochemical insights into and improve dietary recommendations and , while contributions to address pollutant impacts on ecosystems. Overall, biochemistry provides the molecular framework for understanding and manipulating life processes, underpinning progress in human , , and .

History

Origins and Early Discoveries

The roots of biochemistry emerged from longstanding philosophical debates in the 18th and early 19th centuries between , which posited that living organisms were governed by a non-physical "vital force" distinct from ordinary chemical processes, and , which advocated that life could be explained through physical and chemical laws alone. This tension shaped early efforts to study biological phenomena chemically, as vitalists argued that compounds could only arise in , while mechanists sought to bridge the gap between inorganic and . A pivotal moment came in 1828 when German chemist synthesized —an found in —from inorganic , demonstrating that complex biological molecules could be produced in a without vital intervention. This experiment challenged vitalist doctrines by showing continuity between inorganic and organic realms, paving the way for viewing life as a series of chemical transformations, though its immediate impact on vitalism's decline was more symbolic than revolutionary. Building on such insights, French chemist conducted groundbreaking respiration experiments in the 1770s, measuring oxygen consumption and production in animals and humans, which framed breathing as a form of slow akin to chemical oxidation. These studies established as a quantifiable chemical process, linking physiological functions to elemental reactions and influencing later biochemical inquiries into energy transformation. In the 1840s, German chemist advanced these ideas through his work in physiological or "animal" chemistry, analyzing the chemical composition of foods and bodily fluids to elucidate metabolic pathways in and . Liebig's experiments demonstrated that animal heat and work arose from the oxidation of and in foodstuffs, treating organisms as chemical engines and emphasizing the role of nitrogenous compounds in tissue repair. Toward the late , Emil Fischer's structural elucidations of sugars like glucose and proteins, including formations, provided foundational models for biomolecular architecture, revealing life's building blocks as intricate yet chemically analyzable entities. A culminating discovery occurred in 1897 when Eduard Buchner extracted a press-juice that fermented into and without intact cells, proving that enzymatic processes could operate extracorporeally and solidifying biochemistry's focus on isolated chemical mechanisms in . These milestones collectively shifted perceptions, portraying vital processes as governed by chemistry rather than mystical forces.

Development of Key Concepts and Techniques

The elucidation of the by Hans Krebs in 1937 marked a pivotal advancement in understanding , demonstrating how is oxidized through a series of enzymatic reactions involving intermediates to generate energy. Krebs' work, building on earlier observations of , integrated with and earned him the 1953 Nobel Prize in Physiology or Medicine, shared with Fritz Lipmann for discoveries on . This cycle became a cornerstone for subsequent research, highlighting the interconnectedness of biochemical processes. In the 1940s and 1950s, advanced protein by proposing the alpha-helix and beta-sheet configurations based on data and quantum mechanical principles, revolutionizing the understanding of polypeptide folding. Pauling's models emphasized hydrogen bonding's role in secondary structures, influencing later studies on protein function and earning recognition in his 1954 for chemical bond research. Concurrently, and Francis Crick's 1953 double-helix model of DNA provided a structural basis for genetic information storage and replication, integrating biochemical and crystallographic evidence from Rosalind Franklin's work. Their discovery, published in , laid the foundation for and was honored with the 1962 Nobel Prize in Physiology or Medicine, shared with . Technological innovations in the mid-20th century further propelled biochemical analysis. The development of in the 1950s, pioneered by for protein separation based on charge and size, enabled precise purification and characterization of biomolecules, earning Arne Tiselius the 1948 Nobel Prize in Chemistry for foundational electrophoretic methods. By the 1970s, (NMR) spectroscopy emerged as a non-destructive tool for determining molecular structures in solution, with Kurt Wüthrich's applications to proteins in the 1980s resolving three-dimensional folds and contributing to his 2002 . These techniques democratized structural biochemistry, allowing detailed studies of enzyme-substrate interactions and conformations. The 2010s introduced CRISPR-Cas9 as a transformative tool for biochemical manipulation, with and Emmanuelle Charpentier's 2012 demonstration of RNA-guided DNA cleavage enabling precise gene editing and studies. Their innovation, recognized with the 2020 shared with Charpentier, facilitated biochemical investigations into gene regulation and protein expression. Nobel recognitions underscored these shifts: the 1975 Chemistry Prize to for his studies on the of enzyme-catalyzed reactions, the 1980 Chemistry Prize to , , and for their contributions to the biochemistry of nucleic acids, including and sequencing, the 2023 or Medicine Prize to and for discoveries concerning base modifications that enabled the development of effective mRNA vaccines, and the 2024 Chemistry Prize to David Baker, , and John Jumper for computational and . Post-2000, biochemistry evolved toward integrative approaches like systems biochemistry, which models network-level interactions using computational tools to predict cellular responses. , advanced through high-throughput and NMR, profiles small-molecule metabolites to map dynamic pathways, as exemplified in global human projects initiated around 2007. A landmark in predictive modeling came with DeepMind's in 2020, achieving near-experimental accuracy in via , transforming structural and . These developments, incorporating genomic data, have unified biochemistry with , enabling holistic views of cellular function.

Chemical Foundations

Essential Elements and Atoms

Living organisms are primarily composed of a limited set of chemical elements, with six elements—oxygen (O), carbon (C), (H), (N), calcium (Ca), and (P)—accounting for approximately 99% of the mass in the . The elements carbon (C), (H), oxygen (O), (N), (P), and (S), collectively known as , form the foundational building blocks of biological molecules, enabling the complexity and functionality of life. Oxygen dominates by mass at approximately 65%, largely due to its prevalence in , which constitutes 60-70% of body weight, while carbon makes up about 18%, serving as the structural backbone for compounds. Hydrogen and follow at roughly 10% and 3%, respectively, contributing to , structures, and key biomolecules like proteins and nucleic acids. Calcium, at approximately 1.5%, is essential for mineralization and cellular signaling. Phosphorus and are present in smaller amounts, at about 1% and 0.25%, yet play critical roles in energy transfer molecules such as ATP and in like and . The roles of these elements are tightly linked to their chemical properties. Carbon's versatility stems from its ability to form four stable covalent bonds, arranged in a tetrahedral geometry that allows for diverse three-dimensional structures in biomolecules. This bonding capacity, combined with moderate (2.55 on the Pauling scale), enables carbon to create stable chains and rings essential for life's molecular diversity. ( 2.20) readily forms nonpolar bonds with carbon but polar bonds with more electronegative atoms like oxygen (3.44) and (3.04), facilitating hydrogen bonding crucial for molecular interactions. (2.19) and (2.58) contribute to high-energy bonds and disulfide bridges, respectively, due to their ability to form multiple oxidation states and polar linkages. Besides these major elements, trace elements constitute less than 1% of body mass but are indispensable for specific functions. For instance, iron (), at under 0.01% of total mass, is central to , where it binds oxygen reversibly through changes between Fe²⁺ and Fe³⁺ states, enabling efficient transport in blood. Magnesium (), comprising about 0.05%, forms the core of in , coordinating with nitrogenous ligands to absorb light for . These elements' low abundance belies their catalytic and structural importance, often as cofactors in enzymes. Isotopic variations of these elements also hold biological significance. (¹⁴C), a radioactive with a of 5,730 years, is incorporated into biomolecules during life via atmospheric CO₂ fixation and decays post-mortem, allowing to determine the age of organic remains up to about 50,000 years old. This technique has revolutionized biochemical studies of ancient ecosystems and molecular turnover rates, providing insights into evolutionary timelines without altering the primary elemental composition.
ElementApproximate % by Mass in Human BodyKey Biological Role
Oxygen (O)65Component of water and organic molecules; enables respiration
Carbon (C)18Backbone of organic structures
Hydrogen (H)10In water and C-H bonds for energy storage
Nitrogen (N)3In amino acids and nucleic acids
Calcium (Ca)1.5Structural component in bones and teeth; cellular signaling
Phosphorus (P)1In ATP, DNA, and phospholipids
Sulfur (S)0.25In cysteine, methionine, and coenzymes
Iron (Fe)<0.01Oxygen transport in hemoglobin
Magnesium (Mg)0.05Central ion in chlorophyll; enzyme cofactor

Role of Water and Acid-Base Chemistry

Water serves as the universal solvent in biological systems primarily due to its polarity, arising from the electronegative oxygen atom pulling electron density away from the hydrogen atoms, creating partial negative and positive charges, respectively. This polarity enables water molecules to form hydrogen bonds with each other and with polar solutes, such as ions and hydrophilic biomolecules, facilitating their dissolution and interactions essential for cellular processes. Hydrogen bonding also contributes to water's high specific heat capacity, approximately 4.18 J/g·°C, which allows it to absorb or release large amounts of heat with minimal temperature change, thereby stabilizing cellular temperatures during metabolic activities. In aqueous environments, water undergoes autoionization, where two water molecules react to produce hydronium (H₃O⁺) and hydroxide (OH⁻) ions, represented by the equilibrium equation: \mathrm{2H_2O \rightleftharpoons H_3O^+ + OH^-} The equilibrium constant for this process, known as the ion product of water K_w, equals $1.0 \times 10^{-14} at 25°C, indicating that in pure water, the concentrations of H⁺ (or H₃O⁺) and OH⁻ are each $1.0 \times 10^{-7} M, resulting in a neutral pH of 7. Acid-base chemistry in biological systems is governed by the pH scale, defined as pH = −log₁₀[H⁺], which quantifies the hydrogen ion concentration and determines the acidity or basicity of solutions. Cellular processes require tight pH control, achieved through buffer systems that resist changes in pH upon addition of acids or bases. The describes the pH of such buffers: pH = pKₐ + log₁₀([A⁻]/[HA]), where pKₐ is the negative logarithm of the , [A⁻] is the conjugate base concentration, and [HA] is the acid concentration. In physiological contexts, the bicarbonate buffer system plays a critical role in maintaining blood pH around 7.4, with carbonic acid (H₂CO₃) having a pKₐ of 6.1, allowing it to effectively buffer against CO₂-derived acids from respiration. Intracellularly, the phosphate buffer system is prominent, particularly in the cytosol where the second dissociation of phosphoric acid (HPO₄²⁻/H₂PO₄⁻) has a pKₐ of 7.2, closely matching the typical cytosolic pH range of 7.2–7.4 and enabling stable conditions for enzymatic reactions and ion transport.

Biomolecules

Carbohydrates: Structure and Function

Carbohydrates are polyhydroxy aldehydes or ketones, or compounds that yield such units upon hydrolysis, consisting primarily of carbon, hydrogen, and oxygen in a 1:2:1 ratio. They serve as essential biomolecules in living organisms, functioning as rapid energy sources through oxidation and as structural components in cell walls and exoskeletons. Classified by polymerization degree, carbohydrates include monosaccharides (single units), disaccharides (two units), and polysaccharides (many units linked by glycosidic bonds). Their structures determine solubility, digestibility, and biological roles, with variations in linkage types (α or β) influencing helical or linear conformations. Monosaccharides are the building blocks of carbohydrates, simple sugars with 3 to 7 carbon atoms and the general formula \ce{(CH2O)_n}, where n typically ranges from 3 to 7. , an aldohexose with molecular formula \ce{C6H12O6}, exemplifies this class and is ubiquitous in biology. In its open-chain form, glucose is depicted in a Fischer projection as a straight chain with an aldehyde group at C1, hydroxyl groups on C2–C5 (configuring as D-glucose with the C5 OH on the right), and a CH2OH at C6. However, in aqueous solution, over 99% of glucose exists in cyclic forms via intramolecular hemiacetal formation, where the C5 hydroxyl attacks the C1 carbonyl, predominantly yielding a six-membered (β-D-glucopyranose or α-D-glucopyranose, differing at the anomeric C1). This cyclization introduces a new chiral center at C1, enabling α (axial OH) and β (equatorial OH) anomers, which interconvert via mutarotation. Disaccharides form when two monosaccharides join via a glycosidic bond, an acetal linkage from dehydration of hemiacetal and hydroxyl groups, releasing water. Sucrose, a non-reducing disaccharide, comprises α-D-glucose and β-D-fructose connected by an α-1,2-glycosidic bond, making it the primary transport sugar in plants and a key dietary source. Lactose, a reducing disaccharide in mammalian milk, links β-D-galactose to D-glucose through a β-1,4-glycosidic bond, providing energy for infants and serving as a precursor for other galactosides. These linkages dictate enzymatic digestibility; for instance, lactase hydrolyzes the β-1,4 bond in lactose. Polysaccharides are long chains of monosaccharides (often hundreds to thousands of units) polymerized via glycosidic bonds, enabling diverse functions based on linkage stereochemistry. , the plant energy storage polysaccharide, includes (linear α-1,4-linked D-glucose, forming a left-handed helix) and (branched with α-1,6 linkages every 24–30 residues), allowing compact storage and rapid mobilization. , the analogous animal storage form in liver and muscle, is more branched (α-1,6 every 8–12 residues) than amylopectin, facilitating quicker glucose release during energy demands. In contrast, provides structural support in plant cell walls as linear β-1,4-linked D-glucose chains, which hydrogen-bond into rigid microfibrils resistant to hydrolysis. , a nitrogen-containing structural polysaccharide, consists of β-1,4-linked N-acetyl-D-glucosamine units, forming tough fibers in arthropod exoskeletons and fungal cell walls. Carbohydrates fulfill critical biological functions as energy providers, stores, and structural elements. Glucose oxidation via cellular respiration yields approximately 30 ATP molecules per molecule, underscoring its role as a universal fuel. Storage polysaccharides like glycogen and starch maintain glucose homeostasis, with glycogenolysis releasing glucose-1-phosphate for metabolic use. Structurally, cellulose imparts tensile strength to plants, enabling upright growth, while chitin reinforces invertebrate cuticles against mechanical stress. Glycoconjugates extend carbohydrate functionality by covalently attaching oligosaccharides to proteins or lipids, forming that mediate cell adhesion, signaling, and immune recognition. In N-linked glycosylation, oligosaccharides attach to asparagine residues in Asn-X-Ser/Thr motifs via an N-acetylglucosamine intermediate, occurring co-translationally in the endoplasmic reticulum. O-linked glycosylation, attaching to serine or threonine hydroxyls, proceeds in the Golgi via direct galactosamine or other sugar transfer, contributing to mucin-like protections and protein stability. These modifications influence glycoprotein folding, trafficking, and interactions with .

Lipids: Structure and Function

Lipids are a diverse class of hydrophobic biomolecules primarily composed of carbon, hydrogen, and oxygen, essential for cellular structure, energy storage, and signaling in living organisms. Unlike other biomolecules, lipids are defined more by their solubility in nonpolar solvents than by a single structural motif, encompassing simple lipids like and complex ones like and . Their amphipathic properties—possessing both hydrophilic and hydrophobic regions—enable critical roles in forming barriers and facilitating molecular interactions. Fatty acids serve as the foundational building blocks of most lipids, consisting of a hydrocarbon chain attached to a carboxylic acid group. They are classified as saturated if the chain lacks double bonds, such as palmitic acid (C16:0), a 16-carbon straight-chain molecule common in animal fats, or unsaturated if containing one or more double bonds, like oleic acid (C18:1). Saturated fatty acids typically have even-numbered carbon chains ranging from 14 to 24 atoms, contributing to the rigidity of lipid structures. Unsaturated variants introduce kinks that affect packing and fluidity. Triglycerides, or triacylglycerols, are formed by esterifying three fatty acid molecules to a glycerol backbone, creating neutral, nonpolar storage lipids. These molecules aggregate into droplets in adipose tissue, serving as the primary energy reserve in animals and plants due to their high caloric density—approximately 9 kcal per gram. For instance, in human adipocytes, triglycerides composed of saturated and monounsaturated fatty acids like palmitic and oleic acid predominate, enabling efficient long-term energy storage. Phospholipids are key amphipathic lipids that constitute the bulk of cell membranes, featuring a glycerol spine linked to two hydrophobic fatty acid tails, a phosphate group, and a polar head, as in phosphatidylcholine (glycerol + two fatty acids + phosphate + choline). This structure allows phospholipids to self-assemble into bilayers, with hydrophilic heads facing aqueous environments and hydrophobic tails sequestered inward, forming a semipermeable barrier that regulates cellular transport and maintains compartmentalization. The diversity in fatty acid chain length (C14–C26) and saturation further modulates bilayer properties, such as thickness and permeability. Steroids, including cholesterol, possess a characteristic four-fused-ring system (three six-membered and one five-membered) with a hydroxyl group at one end and a flexible hydrocarbon tail, rendering them rigid and planar. Cholesterol, abundant in eukaryotic membranes, intercalates between phospholipid tails to modulate fluidity—preventing excessive rigidity at low temperatures and excessive disorder at high ones—thus maintaining optimal membrane integrity. Sphingolipids, another membrane component, feature a sphingoid base (e.g., sphingosine) backbone amide-linked to a fatty acid, forming ceramides that can attach polar heads like phosphocholine in sphingomyelin. These lipids cluster in ordered membrane domains, contributing to signaling platforms and cellular recognition. In terms of function, lipids excel in energy provision through β-oxidation of fatty acids, where each cycle cleaves two carbons as acetyl-CoA, yielding 4 ATP equivalents per round via NADH and FADH₂ oxidation in the electron transport chain. For example, complete β-oxidation of (C16:0) generates 8 acetyl-CoA units and nets approximately 106 ATP molecules, far exceeding glucose oxidation yields per carbon atom and underscoring lipids' role in sustained energy during fasting. Membrane lipids like and ensure structural stability and fluidity, integrating briefly with proteins to form functional complexes such as lipid rafts. Additionally, lipids act as precursors for signaling molecules; (C20:4 n-6), an unsaturated fatty acid, is released from membrane to produce eicosanoids like prostaglandins, which mediate inflammation and homeostasis. Steroids from further serve as hormone precursors, such as cortisol and sex hormones, regulating diverse physiological processes.

Proteins: Structure and Function

Proteins are linear polymers composed of amino acid monomers linked by peptide bonds, serving as the primary functional units in cellular processes due to their diverse structures and roles. The 20 standard amino acids found in proteins share a common backbone structure consisting of a central α-carbon atom bonded to a hydrogen atom, a carboxyl group (-COOH), an amino group (-NH₂), and a variable side chain (R group). In aqueous environments at physiological pH, amino acids predominantly exist as zwitterions, where the carboxyl group is deprotonated to -COO⁻ and the amino group is protonated to -NH₃⁺, resulting in a net neutral charge but with separated positive and negative charges. The properties of the R group determine the amino acid's classification: nonpolar (hydrophobic) side chains, such as those in alanine (methyl group) and leucine (isobutyl group), promote interactions in non-aqueous environments; polar uncharged side chains, like those in serine (hydroxymethyl) and asparagine (amide), enable hydrogen bonding; and charged side chains include acidic ones, such as aspartic acid (carboxymethyl, negatively charged at pH 7) and basic ones, such as lysine (butylamine, positively charged at pH 7). The primary structure of a protein is its linear sequence of amino acids, dictated by the genetic code from messenger RNA and determining all higher levels of organization. Secondary structure arises from local hydrogen bonding between the backbone atoms, forming regular motifs such as the α-helix, proposed by and in 1951, where the polypeptide chain coils into a right-handed spiral stabilized by hydrogen bonds between the carbonyl oxygen of one residue and the amide hydrogen four residues ahead, and the β-sheet, also described by and Corey, consisting of extended strands aligned either parallel or antiparallel with hydrogen bonds between adjacent strands. Tertiary structure represents the overall three-dimensional fold of a single polypeptide chain, driven by interactions among side chains and the backbone, including a hydrophobic core where nonpolar residues cluster away from water to minimize free energy, as demonstrated in studies of protein folding, and covalent disulfide bonds formed between the thiol groups of cysteine residues to stabilize the fold, particularly in extracellular proteins. Quaternary structure occurs in proteins composed of multiple polypeptide subunits, assembled through noncovalent interactions and sometimes disulfide bonds; for example, human consists of two α and two β subunits arranged in a tetrahedral configuration, enabling cooperative oxygen binding. Proteins exhibit a wide array of functions shaped by their structures. As enzymes, they catalyze biochemical reactions by providing an active site that lowers activation energy through precise substrate binding and orientation, exemplified by ribonuclease A, whose folding was shown to be thermodynamically driven by Anfinsen's experiments in the 1960s. In transport roles, proteins such as hemoglobin facilitate the movement of oxygen across membranes via its heme-binding pockets, while ion channels like the potassium channel form selective pores lined by polar residues to allow passive diffusion. Defense functions are fulfilled by antibodies, which are Y-shaped immunoglobulins with variable regions that bind specific antigens through complementary shapes, triggering immune responses. Structural proteins provide mechanical support; collagen, the most abundant protein in animals, forms a triple helix from three left-handed polyproline II-like chains wound into a right-handed superhelix, rich in glycine, proline, and hydroxyproline, imparting tensile strength to connective tissues. Post-translational modifications expand protein diversity and functionality beyond the primary sequence encoded by nucleic acids. Phosphorylation involves the addition of a phosphate group to serine, threonine, or tyrosine residues by , introducing negative charge that can alter protein conformation, activity, or interactions, as seen in . Ubiquitination attaches , a small 76-amino-acid protein, to lysine residues via a cascade of E1, E2, and E3 enzymes, often marking proteins for proteasomal degradation but also regulating non-degradative processes like trafficking.

Nucleic Acids: Structure and Function

Nucleic acids are essential biomolecules that serve as the primary carriers of genetic information in living organisms, composed of repeating nucleotide units. Each nucleotide consists of a nitrogenous base attached to a five-carbon sugar (pentose) and at least one phosphate group, with nucleotides linked by phosphodiester bonds between the sugar of one nucleotide and the phosphate of the next, forming a directional sugar-phosphate backbone. The nitrogenous bases fall into two classes: purines (adenine [A] and guanine [G]) and pyrimidines (cytosine [C], thymine [T] in DNA, or uracil [U] in RNA). In deoxyribonucleic acid (DNA), the sugar is 2'-deoxyribose, lacking a hydroxyl group at the 2' carbon position, whereas ribonucleic acid (RNA) contains ribose with a hydroxyl group at that site. The structure of DNA is a right-handed double helix, in which two antiparallel polynucleotide strands wind around a common axis, stabilized by hydrogen bonding between complementary bases: A pairs with T via two hydrogen bonds, and G pairs with C via three. This Watson-Crick base pairing ensures specificity in genetic information storage and dictates the sequence of one strand based on the other. The most prevalent conformation in vivo is the B-form, characterized by approximately 10.5 base pairs per helical turn, a pitch (axial rise per turn) of 3.4 nm, a diameter of about 2 nm, and nearly perpendicular base pairs to the helix axis, resulting in major and minor grooves that facilitate protein interactions. These grooves expose the edges of the bases, allowing regulatory proteins to bind and access the genetic code without unwinding the helix. In contrast, RNA is generally single-stranded, enabling it to fold into complex secondary and tertiary structures through intramolecular base pairing, which influences its diverse roles. The three major types of RNA include messenger RNA (), which is linear and conveys genetic instructions from DNA; transfer RNA (), which adopts a characteristic cloverleaf secondary structure with stem-loops, including an anticodon loop for base-pairing with mRNA; and ribosomal RNA (), which forms intricate folded structures as a core component of ribosomes. These RNA structures enable specific functions, such as tRNA's role in recognizing codons during protein assembly. Nucleic acids primarily function in the storage, transmission, and expression of hereditary information. DNA maintains genetic continuity through semiconservative replication, in which each parental strand serves as a template for a new complementary strand, producing two identical daughter molecules. This mechanism, demonstrated experimentally using density-labeled DNA in Escherichia coli, ensures faithful copying of the genome across cell divisions. In gene expression, DNA is transcribed into RNA transcripts, which are then translated into proteins, with RNA serving as an intermediary to direct the synthesis of functional polypeptides from the genetic blueprint.

Enzymes and Catalysis

Enzyme Structure and Mechanism

Enzymes are primarily proteins, though some RNA molecules also exhibit catalytic activity, that accelerate biochemical reactions by lowering activation energies through specific structural features. The core of an enzyme's catalytic function resides in its active site, a specialized region typically formed by a pocket or cleft on the protein surface where substrates bind and undergo transformation. This site is composed of amino acid residues precisely positioned to interact with the substrate, often involving hydrogen bonds, electrostatic interactions, and van der Waals forces to achieve specificity. The classical model for substrate binding, proposed by Emil Fischer in 1894, describes the active site as a rigid, preformed structure complementary in shape, charge, and hydrophobic properties to the substrate, analogous to a lock and key fitting precisely to ensure specificity. This lock-and-key hypothesis explained the high selectivity of enzymes like glycosidases for particular sugar configurations but failed to account for cases where substrates induce structural adjustments in the enzyme. In 1958, Daniel Koshland introduced the induced fit model, positing that substrate binding triggers conformational changes in the enzyme, reshaping the active site for optimal alignment and catalysis, thereby enhancing specificity and efficiency beyond mere geometric complementarity. This dynamic process is exemplified in , where glucose binding causes a large hinge motion that closes the active site, excluding water and positioning catalytic residues. Amino acid residues in the active site, such as those in , play critical roles; for instance, the catalytic triad consisting of aspartate (Asp), histidine (His), and serine (Ser) residues facilitates nucleophilic attack on peptide bonds. In , Asp102 orients His57 for proton abstraction from Ser195, enabling the serine hydroxyl to act as a nucleophile in forming a covalent acyl-enzyme intermediate. Enzyme mechanisms often involve acid-base catalysis, where residues donate or accept protons to stabilize transition states. In ribonuclease A, His12 and His119 act as general bases and acids, respectively, to facilitate the hydrolysis of RNA phosphodiester bonds via a 2'-3'-cyclic phosphate intermediate. Another common strategy employs covalent intermediates, as seen in ping-pong mechanisms where the enzyme alternates between substrates. Nucleoside diphosphate kinase (NDPK), for example, uses a phosphohistidine intermediate: the enzyme first transfers phosphate from to His122, forming a phosphoenzyme, then transfers it to , enabling sequential phosphoryl group exchange. Many enzymes require cofactors to achieve full catalytic competence, as their protein scaffolds alone lack necessary chemical groups. Metal ions like Zn^{2+} serve as Lewis acids, polarizing substrates or stabilizing charged intermediates; in carbonic anhydrase, Zn^{2+} coordinated to three histidine residues deprotonates a bound water molecule, generating a Zn^{2+}-OH^- nucleophile that attacks CO_2 to form bicarbonate (HCO_3^-). \text{Zn}^{2+}-\text{OH}^- + \text{CO}_2 \rightarrow \text{Zn}^{2+}-\text{OCO}_2^- + \text{H}^+ This reaction proceeds at near-diffusion-limited rates, underscoring the ion's role in enhancing nucleophilicity. Coenzymes, organic cofactors derived from vitamins, participate directly in catalysis; nicotinamide adenine dinucleotide (NAD^+), composed of nicotinamide and adenine nucleotides linked by a pyrophosphate bond, accepts a hydride ion (H^-) from substrates in dehydrogenation reactions. For example, NAD^+-dependent enzymes, such as alcohol dehydrogenase, oxidize alcohols to aldehydes or ketones, thereby reducing NAD^+ to NADH. \ce{R-CH2OH + NAD^+ ⇌ R-CHO + NADH + H^+} The nicotinamide ring's C4 position stereospecifically accepts the pro-R hydride, ensuring reaction specificity. Beyond the active site, allosteric sites—distinct binding pockets—allow regulatory molecules to induce conformational changes that modulate catalysis, a phenomenon first formalized by Monod, Wyman, and Changeux in 1965. Effector binding at these sites shifts the enzyme between tense (T, low-activity) and relaxed (R, high-activity) states, as in hemoglobin's cooperative oxygen binding, but applied to enzymes like aspartate transcarbamoylase where CTP binding stabilizes the T state to inhibit activity. Such changes propagate through the protein via altered hydrogen bonding networks or rigid-body motions, fine-tuning substrate affinity without directly competing at the active site.00391-7)

Enzyme Kinetics and Regulation

Enzyme kinetics quantifies the rates of enzymatic reactions and how they depend on substrate concentration, providing a framework for understanding catalytic efficiency. The foundational model is the Michaelis-Menten equation, which assumes a simple reversible binding of substrate to enzyme followed by product formation, yielding the initial reaction velocity v as v = \frac{V_{\max} [S]}{K_m + [S]}, where V_{\max} is the maximum velocity achieved at saturating substrate concentration [S], and K_m is the Michaelis constant representing the substrate concentration at which v = \frac{1}{2} V_{\max}. This equation, derived from steady-state assumptions, highlights hyperbolic saturation kinetics typical of many enzymes and enables determination of kinetic parameters through nonlinear regression or linear transformations. A common linearization, the Lineweaver-Burk plot, transforms the Michaelis-Menten equation into \frac{1}{v} = \frac{K_m}{V_{\max}} \cdot \frac{1}{[S]} + \frac{1}{V_{\max}}, plotting \frac{1}{v} versus \frac{1}{[S]} to yield a straight line with slope \frac{K_m}{V_{\max}}, y-intercept \frac{1}{V_{\max}}, and x-intercept -\frac{1}{K_m}. This graphical method facilitates parameter estimation and analysis of deviations, such as those caused by inhibitors, though it can amplify errors at low substrate concentrations. Enzyme inhibition modulates kinetics by reducing activity, classified by effects on K_m and V_{\max} in Lineweaver-Burk plots. Competitive inhibition occurs when an inhibitor binds reversibly to the active site, competing with substrate and increasing apparent K_m (reduced substrate affinity) while V_{\max} remains unchanged, as higher [S] can outcompete the inhibitor; the modified equation is v = \frac{V_{\max} [S]}{K_m (1 + \frac{[I]}{K_i}) + [S]}, where [I] is inhibitor concentration and K_i is the inhibition constant. Noncompetitive inhibition involves binding to a site distinct from the active site, unaffected by substrate presence, decreasing V_{\max} (fewer functional enzymes) but leaving K_m unchanged, with the equation v = \frac{V_{\max} [S]}{(K_m + [S])(1 + \frac{[I]}{K_i})}. Uncompetitive inhibition binds only to the enzyme-substrate complex, lowering both K_m (apparent increased affinity) and V_{\max}, yielding parallel lines in Lineweaver-Burk plots, described by v = \frac{V_{\max} [S]}{K_m + [S](1 + \frac{[I]}{K_i})}. Beyond inhibition, enzymes are regulated physiologically to fine-tune metabolic flux. Allosteric regulation involves effector binding at sites remote from the active site, inducing conformational changes that alter activity; for instance, feedback inhibition in (ATCase), a key enzyme in pyrimidine biosynthesis, is allosterically inhibited by (CTP), the pathway's end product, reducing activity to prevent overproduction, while (ATP) acts as an activator.65038-6/fulltext) This heterotropic regulation exemplifies the , where effectors shift equilibrium between tense (low-affinity) and relaxed (high-affinity) states. Covalent modification provides rapid, reversible control, often via phosphorylation that adds a negatively charged phosphate group to serine, threonine, or tyrosine residues, altering charge and conformation. A classic example is glycogen phosphorylase, where phosphorylation by phosphorylase kinase converts the inactive b form to the active a form, enhancing glycogen breakdown in response to hormonal signals like epinephrine.74379-8/fulltext) Zymogens represent irreversible activation through proteolytic cleavage, ensuring enzymes like proteases remain inactive until needed; trypsinogen, secreted by the pancreas, is converted to active trypsin by enterokinase in the intestine, cleaving a peptide bond to expose the active site and initiate protein digestion. Environmental factors also influence kinetics: temperature affects rate via Arrhenius kinetics, with most enzymes exhibiting optimal activity around 37°C in humans before denaturation reduces V_{\max}, while extremes inactivate via unfolding. pH optima, typically near neutrality for cytoplasmic enzymes (e.g., pH 7 for ), arise from protonation states of active-site residues; deviations alter K_m and V_{\max} by ionizing key groups, as seen in 's acidic optimum (pH 2) for gastric function.

Metabolism

Catabolic Pathways

Catabolic pathways in biochemistry encompass the degradative processes that break down complex biomolecules, such as carbohydrates, lipids, and proteins, into simpler molecules, thereby releasing energy in the form of and reducing equivalents like and . These pathways converge on central routes, including , the , β-oxidation, and amino acid degradation, which funnel substrates into energy-yielding reactions while producing CO₂ as a byproduct. Primarily occurring in the cytosol and mitochondria, these processes provide substrates for , enabling cells to meet energetic demands under varying conditions. Glycolysis, also known as the Embden-Meyerhof-Parnas pathway, is a universal 10-step anaerobic process that converts one molecule of glucose into two molecules of pyruvate, occurring in the cytosol of nearly all cells. The pathway begins with the phosphorylation of glucose by or , consuming ATP, followed by isomerization to fructose-6-phosphate and another ATP-dependent phosphorylation to fructose-1,6-bisphosphate. Cleavage by yields dihydroxyacetone phosphate and , which interconvert via ; the latter then undergoes oxidation by to 1,3-bisphosphoglycerate, producing NADH. Subsequent steps involve substrate-level phosphorylation to generate ATP via and , resulting in a net yield of 2 ATP, 2 NADH, and 2 pyruvate per glucose molecule. The overall reaction is: \text{glucose} + 2 \text{ADP} + 2 \text{P}_\text{i} + 2 \text{NAD}^+ \rightarrow 2 \text{pyruvate} + 2 \text{ATP} + 2 \text{NADH} + 2 \text{H}^+ + 2 \text{H}_2\text{O} Under aerobic conditions, pyruvate enters the mitochondria for further oxidation, linking glycolysis to the TCA cycle. The TCA cycle, or Krebs cycle, serves as the central hub for oxidizing acetyl-CoA derived from carbohydrates, fats, and proteins, completing the catabolism of pyruvate from glycolysis and other sources. This eight-step cyclic pathway occurs in the mitochondrial matrix and begins with the condensation of acetyl-CoA (two carbons) with oxaloacetate (four carbons) by citrate synthase to form citrate, followed by isomerization to isocitrate. Oxidative decarboxylation by isocitrate dehydrogenase produces α-ketoglutarate, NADH, and CO₂; α-ketoglutarate dehydrogenase then yields succinyl-CoA, another NADH, and CO₂. Succinyl-CoA synthetase generates GTP (or ATP) and succinate, which is dehydrogenated to fumarate by succinate dehydrogenase, producing FADH₂. Fumarase hydrates fumarate to malate, and malate dehydrogenase oxidizes it back to oxaloacetate, yielding a third NADH. Per acetyl-CoA, the cycle produces 3 NADH, 1 FADH₂, 1 GTP, and 2 CO₂, with no net accumulation of intermediates due to the regenerative nature of the cycle. β-Oxidation is the primary catabolic route for fatty acids, sequentially cleaving two-carbon units as acetyl-CoA from the carboxyl end of activated acyl-CoA chains in the mitochondrial matrix. The process initiates with activation of free fatty acids to acyl-CoA by acyl-CoA synthetase in the cytosol or outer mitochondrial membrane, requiring ATP and producing AMP and pyrophosphate; the acyl group is then transported into the matrix via the carnitine shuttle. Each cycle of β-oxidation comprises four enzymatic steps: dehydrogenation by acyl-CoA dehydrogenase to form a trans-enoyl-CoA and FADH₂; hydration by enoyl-CoA hydratase to L-3-hydroxyacyl-CoA; oxidation by 3-hydroxyacyl-CoA dehydrogenase to 3-ketoacyl-CoA and NADH; and thiolysis by β-ketothiolase to acetyl-CoA and a shortened acyl-CoA. For a saturated even-chain fatty acid like palmitate (16 carbons), seven cycles yield eight acetyl-CoA, 7 FADH₂, and 7 NADH, which feed into the TCA cycle and electron transport chain for maximal energy extraction. Amino acid catabolism involves the initial removal of the α-amino group through transamination, transferring it to α-ketoglutarate to form glutamate and the corresponding α-keto acid, primarily catalyzed by aminotransferases like or . These α-keto acids, such as from alanine or from aspartate, then enter central catabolic pathways: glucogenic amino acids feed into gluconeogenesis or the TCA cycle at points like α-ketoglutarate, succinyl-CoA, or fumarate, while ketogenic ones like produce or acetoacetate for the TCA cycle or ketogenesis. The ammonia from deamination is detoxified via the , ensuring nitrogen homeostasis during protein breakdown. This integration allows amino acids to contribute to energy production, with yields varying by residue but ultimately converging on the TCA cycle for oxidation.

Anabolic Pathways

Anabolic pathways in biochemistry encompass the energy-requiring processes that synthesize complex biomolecules from simpler precursors, essential for growth, repair, and maintenance of cellular structures. These pathways contrast with by building macromolecules such as , , , and , often drawing on intermediates from central metabolism like the . In anabolic reactions, and reducing equivalents like drive the formation of carbon-carbon bonds and other linkages, ensuring the production of molecules critical for cellular function. Gluconeogenesis represents a primary anabolic route for carbohydrate synthesis, generating glucose from non-carbohydrate precursors such as lactate, glycerol, and amino acids, primarily in the liver and kidneys to maintain blood glucose levels during fasting. This pathway largely reverses glycolysis but circumvents its three irreversible steps—those catalyzed by , , and —through specialized enzymes. The conversion of pyruvate to oxaloacetate is mediated by , a biotin-dependent enzyme that requires ATP and occurs in the mitochondria, while then decarboxylates oxaloacetate to phosphoenolpyruvate using GTP, predominantly in the cytosol. Additional bypasses include and , which hydrolyze their respective phosphate esters without energy input. Overall, synthesizing one glucose molecule from two pyruvates demands six high-energy phosphate bonds (four ATP and two GTP equivalents), highlighting the energetic cost of this reversal. Fatty acid synthesis constructs long-chain fatty acids from acetyl-CoA units, occurring in the cytosol and serving as a key step in lipid anabolism for membrane formation and energy storage. The process initiates with the carboxylation of acetyl-CoA to malonyl-CoA by , an ATP-dependent reaction that provides the two-carbon donor for chain elongation while preventing futile cycling with β-oxidation. Malonyl-CoA is then transferred to the complex, a multifunctional enzyme that iteratively adds two-carbon units from malonyl-CoA to the growing acyl chain, releasing CO₂ in each condensation step. Each elongation cycle involves β-ketoacyl reduction, dehydration, and enoyl reduction, all powered by as the electron donor, ultimately yielding after seven cycles. This NADPH dependence underscores the pathway's reliance on reductive biosynthesis, with the coordinating all activities in a single polypeptide in mammals.01125-X) Nucleotide synthesis occurs via de novo and salvage pathways, enabling the production of purine and pyrimidine nucleotides for DNA, RNA, and cofactor assembly. In de novo purine biosynthesis, the pathway assembles the purine ring stepwise on 5-phosphoribosyl-1-pyrophosphate (PRPP), starting with the activation of PRPP by glutamine phosphoribosyl pyrophosphate amidotransferase to form 5-phosphoribosylamine, followed by additions of glycine, formate, aspartate, and CO₂ to yield inosine monophosphate (IMP). Pyrimidine de novo synthesis first constructs the ring as orotate from carbamoyl phosphate and aspartate, then attaches it to PRPP via orotate phosphoribosyltransferase to form orotidine monophosphate, which is decarboxylated to uridine monophosphate (UMP). These pathways require ATP for multiple steps and integrate one-carbon units from folate metabolism. Salvage pathways, in contrast, recycle free bases or nucleosides—such as adenine via adenine phosphoribosyltransferase with PRPP to AMP, or hypoxanthine to IMP—conserving energy and precursors from nucleic acid turnover. Amino acid biosynthesis draws heavily from TCA cycle intermediates as carbon skeletons, allowing cells to produce non-essential amino acids from central metabolic pools. For instance, glutamate is synthesized by the reductive amination of α-ketoglutarate using , which transfers an amino group from ammonia (or glutamine via ) in an NADPH-dependent reaction, serving as a precursor for glutamine, proline, and arginine. Aspartate derives from oxaloacetate through transamination with glutamate, feeding into asparagine, lysine, methionine, threonine, and isoleucine synthesis. Other TCA-derived amino acids include alanine from pyruvate (a glycolysis-TCA link) and the branched-chain group from α-ketoglutarate and oxaloacetate branches. These pathways are amphibolic, linking anabolism to TCA flux and nitrogen assimilation, with glutamate acting as a key nitrogen donor across multiple routes.

Metabolic Integration and Regulation

Metabolic pathways in eukaryotic cells are organized into distinct subcellular compartments, which facilitates efficient integration and prevents futile cycles between catabolism and anabolism. Glycolysis, the initial breakdown of glucose to pyruvate, occurs primarily in the cytosol, allowing rapid response to energy demands without the need for organelle transport. In contrast, the tricarboxylic acid (TCA) cycle takes place in the mitochondrial matrix, where it links carbohydrate, lipid, and protein catabolism to oxidative phosphorylation for ATP production. Beta-oxidation of fatty acids is compartmentalized between peroxisomes, which handle the initial shortening of very-long-chain fatty acids, and mitochondria for complete oxidation, ensuring specialized handling of hydrophobic substrates. This spatial separation is maintained by membrane transporters, such as the mitochondrial pyruvate carrier, which shuttles metabolites between cytosol and mitochondria to coordinate flux across pathways. Hormonal signals further integrate metabolic pathways by modulating enzyme activities in response to systemic energy needs, promoting homeostasis through reciprocal actions on anabolism and catabolism. Insulin, secreted in response to elevated blood glucose, promotes anabolic processes by activating glycogen synthase through dephosphorylation via the IRS-PI3K-Akt pathway, thereby enhancing glucose storage as glycogen in liver and muscle. This action inhibits gluconeogenesis and lipolysis while stimulating glycolysis and lipogenesis, directing nutrients toward storage. Conversely, glucagon, released during low glucose states, drives catabolic pathways by elevating cyclic AMP levels via G-protein-coupled receptors, which activates protein kinase A to promote hepatic glycogenolysis and gluconeogenesis, increasing blood glucose availability. Glucagon also stimulates fatty acid oxidation and amino acid catabolism, counteracting insulin's effects to mobilize energy reserves during fasting or stress. At the cellular level, reciprocal regulation ensures that catabolic and anabolic pathways do not operate simultaneously, with key enzymes like (PFK-1) serving as control points for glycolytic flux. PFK-1 is allosterically inhibited by high ATP and citrate, signals of energy abundance and active , respectively, which prevents unnecessary glucose breakdown when cellular needs are met. In energy-deficient states, PFK-1 is activated by AMP, indicating low ATP levels, and by , a potent regulator produced by that overrides ATP inhibition to favor glycolysis over gluconeogenesis. This feed-forward and feedback mechanism integrates cytosolic glycolysis with mitochondrial oxidation, allowing cells to adjust flux dynamically to maintain redox and energy balance. Metabolic flux analysis provides a quantitative framework for understanding how perturbations in one pathway affect overall integration, emphasizing distributed control rather than single rate-limiting steps. Flux control coefficients, introduced by Kacser and Burns, measure the fractional change in steady-state flux through a pathway in response to a fractional change in enzyme activity, revealing that control is often shared among multiple steps. For instance, in glycolysis, enzymes like hexokinase and PFK-1 may have higher coefficients under varying conditions, guiding how hormonal or compartmental signals redistribute flux for homeostasis. This approach highlights the robustness of integrated networks, where compensatory adjustments in enzyme levels or activities maintain steady-state metabolism despite external changes.

Bioenergetics

Thermodynamic Principles

In biochemistry, the Gibbs free energy change, denoted as ΔG, serves as a fundamental criterion for determining the spontaneity of chemical reactions under constant temperature and pressure. It is defined by the equation \Delta G = \Delta H - T \Delta S where ΔH represents the change in enthalpy, T is the absolute temperature in Kelvin, and ΔS is the change in entropy. This relationship integrates the energetic (enthalpic) and disorder-related (entropic) contributions to a process, allowing biochemists to predict whether a reaction will proceed without external energy input. For standard conditions, particularly in aqueous solutions at pH 7 (denoted as ΔG°'), the standard Gibbs free energy change relates to the equilibrium constant K_eq via \Delta G^\circ = -RT \ln K_\text{eq} where R is the gas constant (8.314 J/mol·K). A reaction is spontaneous if ΔG < 0, indicating that the products are more stable than the reactants and the process favors forward progression toward equilibrium. Biochemical reactions are classified as exergonic (ΔG < 0, energy-releasing and spontaneous) or endergonic (ΔG > 0, energy-requiring and non-spontaneous). In living systems, endergonic processes essential for biosynthesis or transport are rarely isolated; instead, they are coupled to exergonic reactions to achieve an overall negative ΔG. A classic example is the hydrolysis of adenosine triphosphate (ATP) to adenosine diphosphate (ADP) and inorganic phosphate (Pi), with ΔG°' ≈ -30.5 kJ/mol under standard biochemical conditions (pH 7); the actual ΔG under physiological conditions is more negative, approximately -50 kJ/mol, which drives unfavorable syntheses such as the formation of glutamine from glutamate and ammonia. This coupling mechanism underscores ATP's role as a universal energy currency in cells, enabling the thermodynamic feasibility of otherwise improbable reactions. Many biological reactions operate far from , characterized by equilibrium constants K_eq that are either extremely large (>>1) or small (<<1), corresponding to large negative or positive ΔG° values, respectively. These near-irreversible reactions (large negative ΔG°) ensure directional flux in metabolic pathways, preventing wasteful back-reactions and maintaining cellular . profoundly influences biochemical processes through its on ΔG, as the TΔS term amplifies entropic contributions at higher temperatures. Reaction rates in biology often exhibit a Q_{10} , where rates approximately double (Q_{10} ≈ 2) for every 10°C increase, reflecting the Arrhenius temperature dependence integrated into thermodynamic feasibility. This sensitivity ensures that physiological temperatures optimize both spontaneity and in enzymatic .

Energy Carriers and Redox Processes

(ATP) serves as the primary energy carrier in cellular processes, consisting of an base linked to a sugar and three groups attached via phosphoanhydride bonds. The high-energy phosphoanhydride bonds between the groups store significant potential energy due to electrostatic repulsion among the negatively charged phosphates, making thermodynamically favorable with a standard free energy change of ΔG°' ≈ -30.5 kJ/mol. This hydrolysis reaction cleaves the terminal phosphoanhydride bond, yielding (ADP) and inorganic (Pi), thereby releasing energy to drive endergonic reactions in . In redox processes, coenzymes such as (NAD⁺/NADH) and (FAD/FADH₂) act as electron carriers, facilitating between metabolic pathways. The of these carriers determines their ability to donate or accept electrons, quantified by the : E = E^{\circ'} + \frac{RT}{nF} \ln \left( \frac{[\text{ox}]}{[\text{red}]} \right) where E is the actual reduction potential, E^{\circ'} is the standard reduction potential at 7, R is the , T is in , n is the number of electrons transferred, F is the , and [\text{ox}] and [\text{red}] are the concentrations of the oxidized and reduced forms, respectively. For NAD⁺/NADH, the standard reduction potential E^{\circ'} is -0.32 V, indicating a strong , while for free FAD/FADH₂ it is approximately -0.22 V, though this value varies when bound to enzymes. These potentials enable NADH and FADH₂ to donate electrons to the (ETC), powering . The electron transport chain comprises four membrane-bound protein complexes (I–IV) embedded in the inner mitochondrial membrane, which sequentially accept electrons from NADH and FADH₂ to generate a proton gradient. Complex I (NADH dehydrogenase) oxidizes NADH and transfers electrons to ubiquinone while pumping protons (H⁺) across the membrane; Complex II (succinate dehydrogenase) handles FADH₂-derived electrons without proton pumping; Complex III (cytochrome bc₁) passes electrons to cytochrome c and pumps additional protons; and Complex IV (cytochrome c oxidase) reduces oxygen to water, further contributing to the proton gradient. This vectorial proton translocation establishes an electrochemical gradient (proton-motive force) across the membrane, with approximately 10 H⁺ translocated per NADH oxidized. The chemiosmotic theory, proposed by Peter Mitchell in 1961, posits that this proton gradient directly drives ATP synthesis without requiring high-energy chemical intermediates. , a rotary complex consisting of the membrane-embedded F₀ subunit (proton channel) and the peripheral F₁ subunit (catalytic head), harnesses the proton-motive force as protons flow back into the matrix, inducing conformational changes that phosphorylate to ATP. This mechanism integrates redox-driven proton pumping with efficient energy conservation, underpinning aerobic respiration.

Interdisciplinary Connections

Biochemistry intersects with molecular biology through the chemical processes that govern the flow of genetic information, as encapsulated in the central dogma proposed by Francis Crick, which states that genetic information flows from DNA to RNA to proteins, with no reverse transfer from proteins to nucleic acids. This unidirectional pathway relies on biochemical reactions involving nucleotide polymerization and hydrolysis, ensuring the fidelity and efficiency of information transfer in cells. In prokaryotes and eukaryotes, these processes are catalyzed by enzymes that utilize energy from nucleotide triphosphates, highlighting biochemistry's role in enabling molecular biology's core mechanisms. Transcription, the synthesis of RNA from a DNA template, begins when binds to promoter sequences, such as the in eukaryotes, approximately 25-30 upstream of the transcription start site. In eukaryotes, , responsible for mRNA synthesis, forms a pre-initiation complex with general transcription factors like TFIID, which recognizes the promoter and recruits the polymerase, initiating RNA chain elongation through formation. This biochemical mechanism ensures precise , with the promoter's core elements dictating the start site and regulatory sequences modulating the rate. Translation converts the of mRNA into a polypeptide chain at ribosomes, which are ribonucleoprotein complexes that decode the via the codon table, where each three- codon specifies an or stop signal. The process involves transfer RNAs (tRNAs) carrying to the ribosome's A site, where activity—catalyzed by —forms peptide bonds, driven by GTP hydrolysis for translocation. This biochemical orchestration achieves high efficiency, with ribosomes synthesizing proteins at rates up to 20 per second in bacteria, underscoring the interplay between chemistry and . DNA replication maintains genetic integrity by duplicating the genome semi-conservatively, with DNA polymerases adding nucleotides to the 3' end of a growing strand in a 5' to 3' direction, using deoxynucleoside triphosphates as substrates. On the lagging strand, discontinuous synthesis produces Okazaki fragments, short RNA-primed DNA segments (typically 100-200 nucleotides in eukaryotes) that are later joined by DNA ligase after primer removal. Proofreading fidelity is enhanced by the polymerase's 3' to 5' exonuclease activity, which excises mismatched nucleotides, reducing error rates to about 1 in 10^7 base pairs incorporated. Gene regulation at the molecular level integrates biochemical signals to control expression, as exemplified by the in , where the protein, encoded by the lacI gene, binds the operator sequence to block transcription in the absence of , but binding induces a conformational change, releasing the and allowing access. Enhancers, distal DNA elements often thousands of base pairs from promoters, boost transcription by looping to interact with promoter-bound factors, recruiting co-activators like to enhance activity. Epigenetic modifications, such as at CpG islands by DNA methyltransferases, add methyl groups to residues, recruiting repressive proteins like methyl-CpG-binding domain proteins that compact and inhibit binding, thereby silencing genes biochemically without altering the sequence. RNA processing in eukaryotes matures pre-mRNA through splicing, where the —a complex of snRNPs—recognizes intron-exon boundaries and catalyzes two reactions to excise introns and ligate exons, ensuring accurate sequence assembly. Capping occurs co-transcriptionally at the 5' end, adding a 7-methylguanosine cap via guanylyltransferase and methyltransferases, which protects against exonucleases and facilitates binding during . Polyadenylation at the 3' end involves cleavage at a poly(A) signal (AAUAAA) followed by addition of 200-250 residues by , stabilizing the mRNA, promoting nuclear export, and enhancing translational efficiency through cap-tail synergy. Biochemistry intersects with through advanced techniques that elucidate the three-dimensional architectures of biomolecules, essential for understanding their functions. X-ray crystallography remains a cornerstone method, providing atomic-resolution structures of proteins and complexes, though it is constrained by resolution limits typically around 1-3 due to crystal quality and radiation damage. A key challenge in this technique is the phase problem, where diffraction experiments yield only amplitude information, requiring methods like multiple isomorphous replacement or molecular replacement to determine phases and reconstruct maps. Complementing crystallography, (cryo-EM) has revolutionized the field by enabling high-resolution imaging of large macromolecular assemblies without crystallization, such as the bacterial at 1.55 resolution, revealing dynamic conformational states inaccessible to other methods. Biophysical principles underpin these structural insights by governing molecular interactions at the atomic level. Van der Waals forces, arising from transient interactions, contribute to the stability of close-packed atomic surfaces in protein interfaces and ligand binding. Electrostatic interactions, including salt bridges and hydrogen bonds between charged residues, play a critical role in protein-DNA binding, facilitating sequence-specific recognition in transcription factors like the . of biomolecules, vital for cellular processes, follows Fick's first law, where the J is given by J = -D \nabla C with D as the and \nabla C the concentration gradient, influencing reaction rates in crowded cellular environments. Protein folding exemplifies the integration of structural and biophysical perspectives, where the Levinthal paradox highlights the improbability of random conformational searches reaching the native state within biological timescales, necessitating guided pathways. Molecular chaperones like Hsp70 assist by binding hydrophobic regions of nascent or misfolded polypeptides, preventing aggregation and promoting refolding through ATP-dependent cycles. Misfolding can lead to pathogenic aggregates, such as amyloid-β fibrils in Alzheimer's disease, which disrupt cellular proteostasis and contribute to neurodegeneration. In membrane biophysics, lipid rafts—cholesterol- and sphingolipid-enriched domains—organize channels and receptors, modulating their localization and activity in . gating, the conformational switch between open and closed states, is influenced by tension and composition, enabling selective permeation crucial for neuronal signaling and . These principles reveal how physical forces dictate biochemical function, with techniques like cryo-EM capturing transient gating intermediates.

References

  1. [1]
    Biological/Biochemistry - American Chemical Society
    Biochemistry emerged as a separate discipline when scientists combined biology with organic, inorganic, and physical chemistry. They began to study areas such ...Missing: key | Show results with:key
  2. [2]
    What is Biochemistry? - Michigan Technological University
    Biochemistry is the study of the chemicals and chemistry of living organisms. Biochemists study biomolecules (such as proteins, RNA, DNA, sugars, and lipids)
  3. [3]
    Biochemistry: Background Information - Library Guides
    Jan 17, 2025 · Biochemistry is the study of chemical processes within and relating to living organisms. Biochemistry is closely related to molecular biology.
  4. [4]
    Biochemistry - History | Vassar College
    The discipline of biochemistry evolved from chemical studies on biological tissue in the late 19 th and early 20 th centuries.
  5. [5]
    Biochemistry, Proteins Enzymes - StatPearls - NCBI Bookshelf - NIH
    There are six main categories of enzymes: oxidoreductases, transferases, hydrolases, lyases, isomerases, and ligases. Each category carries out a general type ...
  6. [6]
    A Critique Of Vitalism And Its Implications For Integrative Medicine
    In this sense, vitalism has operated throughout the history of ideas as a “motor force” against which mechanism, reductionism, have had to defend themselves.
  7. [7]
    Vitalism - William Bechtel
    Vitalism developed as a contrast to this mechanistic view. Over the next three centuries, numerous figures opposed the extension of Cartesian mechanism to ...
  8. [8]
    History of Clinical Chemistry - PMC - NIH
    Jun 10, 2002 · In 1828, Friedrich Wöhler (1800-82) found that urea, an 'organic' substance, could be synthesized in vitro without any 'vital force' or living ...Missing: debunked | Show results with:debunked
  9. [9]
    Antoine Lavoisier and the study of respiration: 200 years old - PubMed
    Antoine Lavoisier has been called the father of modern chemistry. From a medical point of view, he introduced the study of respiration and metabolism.Missing: 1770s | Show results with:1770s
  10. [10]
    Antoine-Laurent de Lavoisier (1743-1794) and the birth of ...
    Aug 6, 2025 · In the 18th century, the French chemist Antoine-Laurent de Lavoisier conducted breathing experiments on human and animal respiration.
  11. [11]
    Justus Liebig and Animal Chemistry - ResearchGate
    Aug 9, 2025 · In 1840, Justus von Liebeg (the founder of organic chemistry) discovered that phosphate, nitrogen and potassium salts, and not organic ...
  12. [12]
    Justus Liebig (our Eponym) — JLU
    From the late 1830s onwards, Liebig shifted his research focus to physiological chemistry – the research of plant and animal metabolism. His treatise ...
  13. [13]
    Emil Fischer – Biographical - NobelPrize.org
    His greatest success was his synthesis of glucose, fructose and mannose in 1890, starting from glycerol. This monumental work on the sugars, carried out between ...
  14. [14]
    [PDF] ALCOHOLIC FERMENTATION WITHOUT YEAST CELLS*
    ALCOHOLIC FERMENTATION. WITHOUT YEAST CELLS*. Eduard Buchner. Until now it has not been possible to separate fermenting activity from living yeast cells; the ...
  15. [15]
    Centenary of the Award of a Nobel Prize to Eduard Buchner, the ...
    Sep 4, 2007 · The discovery of cell-free alcohol fermentation: In 1897 Eduard Buchner laid the foundation stone for modern in vitro enzymology from his ...
  16. [16]
    The Elements of Life: A Biocentric Tour of the Periodic Table - PMC
    Six macronutrients (CHNOPS) account for >99% of all elements in the human body (Fraústo_da_Silva & Williams, 2001). Life also requires other elements, many of ...
  17. [17]
    The Chemical Components of a Cell - Molecular Biology of ... - NCBI
    The four covalent bonds that can form around a carbon atom, for example, are arranged as if pointing to the four corners of a regular tetrahedron. The precise ...
  18. [18]
    2.3: Carbon - Biology LibreTexts
    Apr 9, 2022 · The carbon atom has unique properties that allow it to form covalent bonds to as many as four different atoms, making this versatile element ...Missing: electronegativity | Show results with:electronegativity
  19. [19]
    Biological molecules
    Briefly, atoms such as carbon or oxygen are said to be reduced if they form covalent bonds with an atom with lesser electronegativity, such as hydrogen.<|control11|><|separator|>
  20. [20]
    Review on iron and its importance for human health - PMC
    For example, iron is a very important component of the hemoglobin molecule; erythropoietin, a molecule secreted by the kidneys, promotes the formation of RBCs ...
  21. [21]
    Physiological Essence of Magnesium in Plants and Its Widespread ...
    Apr 25, 2022 · Mg is the fundamental component of chlorophyll (Chl) pigments in the light-capturing complex of chloroplasts and, hence, is involved in ...
  22. [22]
    Radiocarbon Dating - American Chemical Society
    Oct 10, 2016 · In a system where carbon-14 is readily exchanged throughout the cycle, the ratio of carbon-14 to other carbon isotopes should be the same in a ...Missing: variations | Show results with:variations
  23. [23]
    Radiocarbon Dating - Chemistry LibreTexts
    Jan 29, 2023 · Although 12C is definitely essential to life, its unstable sister isotope 14C has become of extreme importance to the science world. ...
  24. [24]
    Polarity of Water - Chemistry Tutorial - The University of Arizona
    Jan 28, 2003 · Water is polar due to uneven electron density, with partial negative charge near oxygen and partial positive charges near hydrogen atoms.
  25. [25]
    How Water's Properties Are Encoded in Its Molecular Structure and ...
    Water's properties are encoded by its molecular structure, including hydrogen bonding, which creates open tetrahedral cage-like structuring.
  26. [26]
    14.1 Unique Properties of Water - Maricopa Open Digital Press
    Water has the highest specific heat capacity of any liquid. Water's high heat capacity is another property caused by hydrogen bonding among the water molecules.
  27. [27]
  28. [28]
    pH, Acids, Bases and Buffers | Introduction to Biology
    pH is the negative, base-10 logarithm of the hydrogen ion (H+) concentration of the solution. As an example, a pH 4 solution has an H+ concentration that is ...
  29. [29]
    Henderson Hasselbalch equation - GlobalRPH
    Oct 18, 2017 · Henderson-Hasselbalch equation describes the relationship of pH as a measure of acidity with the acid dissociation constant (pKa)
  30. [30]
    Henderson-Hasselbalch Equation - an overview - ScienceDirect.com
    The pK's of the major buffer systems cover a wide pH range, from 6.1 for the bicarbonate buffer system, to 9.0 for the ammonia buffer system (see Fig. 83-2). A ...
  31. [31]
    [PDF] Chapter 9: Phosphate transfer reactions - Organic Chemistry
    Phosphoric acid is triprotic, meaning that it has three acidic protons available to donate, with pKa values of 2.1, 7.2, and 12.3, respectively. O. OH. OH. OP.
  32. [32]
    Intracellular pH – Advantages and pitfalls of surface-enhanced ...
    Tumor microenvironment is characterized by slightly acidic values outside the cells (pH around 6.5–7.0) and slightly alkaline intracellular pH (ca. 7.2–7.4, see ...
  33. [33]
    Physiology, Carbohydrates - StatPearls - NCBI Bookshelf - NIH
    Carbohydrates are one of the three macronutrients in the human diet, along with protein and fat. These molecules contain carbon, hydrogen, and oxygen atoms.
  34. [34]
    Biological Functions of Glycans - Essentials of Glycobiology - NCBI
    This chapter provides an overview of the biological functions of glycans in three broad categories: structural roles in, on, and outside cells; energy ...
  35. [35]
    Carbohydrates
    They contain from 3 to 7 carbons and have the general formula of (CH2O)n where n ranges from 3 to 7 (5 or 6 being the most common).Missing: review | Show results with:review
  36. [36]
    Structure & Reactivity in Chemistry: IB4 - IMF: Carbohydrates - csbsju
    In starch and glycogen, which are energy storage polysaccharides, the linkage is alpha 1-4 while in cellulose, the most abundant biomolecule, the linkage is ...
  37. [37]
    Monosaccharide Diversity - Essentials of Glycobiology - NCBI - NIH
    Formation of Hemiacetals. Monosaccharides can also be represented as Haworth projections in which both five- and six-membered cyclic structures are depicted as ...
  38. [38]
    [PDF] Chapter 7 Carbohydrates: Nomenclature Monosaccharides
    Sugars can form cyclic hemiacetals and hemiketals. Because the alcohol ... Fischer Projection of ring structure. • Fischer projections of glucose anomers.
  39. [39]
    Chapter 25 Notes - Carbohydrates
    the two stereoisomers at the hemiacetal (anomeric) carbon. alpha anomer: OH group is down (Haworth); beta anomer: OH group is up (Haworth) · anomers are ...
  40. [40]
    Carbohydrates – Biology - UH Pressbooks
    Common disaccharides include lactose, maltose, and sucrose ([link]). Lactose is a disaccharide consisting of the monomers glucose and galactose. It is found ...
  41. [41]
    Carbohydrates - OpenEd CUNY
    In sucrose, a glycosidic linkage forms between carbon 1 in glucose and carbon 2 in fructose. Common disaccharides include lactose, maltose, and sucrose (Figure) ...
  42. [42]
    Disaccharides - EdTech Books
    The resulting linkage is called a glycosidic linkage. There are three important disaccharides that we will discuss: sucrose, lactose, and maltose. In all ...
  43. [43]
    CH103 - Chapter 8: The Major Macromolecules - Chemistry
    In the D-family, the alpha and beta bonds have the same orientation defined ... Starch, glycogen, cellulose, and chitin are examples of polysaccharides.
  44. [44]
    Carbohydrates - OpenEd CUNY
    Glycosidic bonds (or glycosidic linkages) can be an alpha or beta type. ... Starch, glycogen, cellulose, and chitin are primary examples of polysaccharides.
  45. [45]
    Carbohydrate Background - RockEDU Science Outreach
    Carbohydrates play important roles in energy storage and metabolism as well as cell structure, cell-to-cell communication, and cellular-level defenses.<|control11|><|separator|>
  46. [46]
    Complex Carbohydrates and Glycoconjugates: Structure, Functions ...
    Nov 12, 2021 · The functional studues of carbohydrates concern molecular recognition such as carbohydrate–lectin or glycoside–enzyme interactions, cell ...
  47. [47]
    Biochemistry, Lipids - StatPearls - NCBI Bookshelf - NIH
    May 1, 2023 · Lipids are an essential component of the cell membrane. The structure is typically made of a glycerol backbone, 2 fatty acid tails (hydrophobic) ...
  48. [48]
    Lipid classification, structures and tools - PMC - PubMed Central
    Lipids are a diverse and ubiquitous group of compounds which have many key biological functions, such as acting as structural components of cell membranes, ...
  49. [49]
    Mammalian lipids: structure, synthesis and function - PubMed Central
    This review summarises the biosynthesis of the lipids of the mammalian cell; phospholipids, sphingolipids and cholesterol and how lipid diversity is achieved.
  50. [50]
    Biochemistry, Fatty Acid Oxidation - StatPearls - NCBI Bookshelf - NIH
    Mitochondrial beta-oxidation yields 4 ATP equivalents per round of oxidation ... Fatty acid beta-oxidation is regulated by the cell's energy requirements.Introduction · Fundamentals · Cellular Level · Molecular Level
  51. [51]
    Physiology, Proteins - StatPearls - NCBI Bookshelf
    Proteins serve as structural support, biochemical catalysts, hormones, enzymes, building blocks, and initiators of cellular death. Proteins can be further ...
  52. [52]
    The Amino Acids
    The 20 Standard Amino Acids. NAME, STRUCTURE (AT NEUTRAL pH). Nonpolar (Hydrophobic) R Groups. Glycine (Gly). Alanine (Ala). Valine (Val). Leucine (Leu).
  53. [53]
    Biochemistry, Essential Amino Acids - StatPearls - NCBI Bookshelf
    Apr 30, 2024 · Proteins are made up of 20 amino acids. Each amino acid has an α-carboxyl group, a primary α-amino group, and a side chain called the R group ( ...
  54. [54]
    Biochemistry, Primary Protein Structure - StatPearls - NCBI Bookshelf
    Oct 31, 2022 · Each protein at least contains a primary, secondary, and tertiary structure. Only some proteins have a quaternary structure as well.
  55. [55]
    The discovery of the α-helix and β-sheet, the principal structural ...
    PNAS papers by Linus Pauling, Robert Corey, and Herman Branson in the spring of 1951 proposed the α-helix and the β-sheet, now known to form the backbones ...
  56. [56]
    Biochemistry, Tertiary Protein Structure - StatPearls - NCBI Bookshelf
    These disulfide bonds are often called bridges because they create strong links between distinct regions of the same polypeptide or 2 separate polypeptides.Introduction · Fundamentals · Clinical Significance
  57. [57]
    Hemoglobin: Structure, Function and Allostery - PMC
    The two αβ dimers (named α1β1 and α2β2) are arranged around a 2-fold axis of symmetry resulting in a large central water cavity in the T or unliganded or ...
  58. [58]
    COLLAGEN STRUCTURE AND STABILITY - PMC - NIH
    Their structure was a right-handed triple helix of three staggered, left-handed PPII helices with all peptide bonds in the trans conformation and two hydrogen ...
  59. [59]
    Post-Translational Modification of Cellular Proteins by Ubiquitin and ...
    Ubiquitination of endogenous proteins is one of the key regulatory steps that guides protein degradation through regulation of proteasome activity.Introduction · Role of Ubiquitination... · Proliferating Cell Nuclear... · p27(KIP1)
  60. [60]
    The Structure and Function of DNA - Molecular Biology of the Cell
    The nucleotides are covalently linked together in a chain through the sugars and phosphates, which thus form a “backbone” of alternating sugar-phosphate-sugar- ...
  61. [61]
    A Structure for Deoxyribose Nucleic Acid - Nature
    The determination in 1953 of the structure of deoxyribonucleic acid (DNA), with its two entwined helices and paired organic bases, was a tour de force in ...
  62. [62]
    From DNA to RNA - Molecular Biology of the Cell - NCBI Bookshelf
    RNA can fold into specific structures. RNA is largely single-stranded, but it often contains short stretches of nucleotides that can form conventional base- ...
  63. [63]
    Biochemistry, RNA Structure - StatPearls - NCBI Bookshelf - NIH
    Jul 29, 2023 · mRNA, rRNA, and tRNA are the three main types of RNA involved in protein synthesis. RNA also serves as the primary genetic material for viruses.
  64. [64]
    The replication of DNA in Escherichia coli* | PNAS
    In this paper reference will be made to the apparent molecular weight of DNA samples determined by means of density-gradient centrifugation.
  65. [65]
    Lock-Key Model - an overview | ScienceDirect Topics
    The textbook lock-and-key model for enzyme catalysis was first introduced by the Nobel laureate organic chemist Emil Fischer in 1894 (Fischer, 1894). The ...
  66. [66]
    [PDF] What Made Emil Fischer Use this Analogy? - LSU School of Medicine
    In 1894, when Fischer first used the lock-and-key analogy to illustrate enzyme specifity, he was 42. He lived for another. 25 years, during which time he ...
  67. [67]
    The Key–Lock Theory and the Induced Fit Theory - Koshland - 1995
    Jan 3, 1995 · The new theory proposed by DE Koshland, Jr. in 1958 allows one to explain regulation and cooperative effects, and adds some new specificity principles as well.
  68. [68]
    A REFINED MECHANISM OF SERINE PROTEASE ACTION
    The mechanism of serine proteases prominently illustrates how charged amino acid residues and proton transfer events facilitate enzyme catalysis.Missing: paper | Show results with:paper
  69. [69]
    An alternate geometry for the catalytic triad of serine proteases
    Unconventional serine proteases: Variations on the catalytic Ser/His/Asp triad configuration. Protein Science 2008, 17 (12) , 2023-2037. https://doi.org ...Missing: original | Show results with:original
  70. [70]
    Chapter 7: Catalytic Mechanisms of Enzymes - Chemistry
    Acid-Base Catalysis is involved in any reaction mechanism that requires the transfer of a proton from one molecule to another. It is very common to see this ...
  71. [71]
    Mechanism of phosphoryl transfer by nucleoside diphosphate kinase
    It involves a covalent intermediate with a histidine residue phosphorylated in the active site. ... ping-pong enzyme such as NDP kinase. Previously, we have ...Materials And Methods · Fluorometric Binding Studies... · Results And Discussion
  72. [72]
    Carbonic Anhydrase: Evolution of the Zinc Binding Site by Nature ...
    Zinc-bound hydroxide attacks the carbonyl carbon of CO2 to form zinc-bound bicarbonate. The initial mode of bicarbonate binding (a) may reflect the structure of ...
  73. [73]
    Elucidating the role of metal ions in carbonic anhydrase catalysis - NIH
    Sep 11, 2020 · Our study provides evidence that the metal ions in metalloenzymes have a crucial impact on the catalytic mechanism beyond their primary chemical properties.
  74. [74]
    Nicotinamide Adenine Dinucleotide (NAD) - Chemistry LibreTexts
    Jul 4, 2022 · The NAD+ coenzyme is involved with many types of oxidation reactions where alcohols are converted to ketones or aldehydes. It is also involved ...
  75. [75]
    NAD+ metabolism: pathophysiologic mechanisms and therapeutic ...
    Oct 7, 2020 · Beyond its vital role as a coenzyme in energy metabolism, the important role of NAD+ has expanded to be a co-substrate for various enzymes ...
  76. [76]
    Translation of the 1913 Michaelis-Menten Paper - PMC - NIH
    Nearly 100 years ago Michaelis and Menten published their now classic paper (Michaelis, L., and Menten, M. L. (1913) Die Kinetik der Invertinwirkung, ...
  77. [77]
    The Determination of Enzyme Dissociation Constants
    Hans Lineweaver · Dean Burk. ACS Legacy ... Smart citations by scite.ai include citation statements extracted from the full text of the citing article.
  78. [78]
    The origin and use of the terms competitive and non ... - PubMed
    The terms competition and competitive were in use for appropriate types of interaction in human and animal behaviour from the seventeenth century.
  79. [79]
    7.1: An Overview of Metabolic Pathways - Catabolism
    May 6, 2022 · CATBOLISM: Catabolic reactions involve the breakdown of carbohydrates, lipids, proteins, and nucleic acids to produce smaller molecules and ...
  80. [80]
    Physiology, Krebs Cycle - StatPearls - NCBI Bookshelf
    Nov 23, 2022 · Each molecule of acetyl-CoA entering the TCA cycle yields 12 ATP molecules. ... Acetyl-CoA undergoes oxidation to CO2 in 8 steps, and the energy ...
  81. [81]
    Metabolism | Essays in Biochemistry - Portland Press
    Aug 24, 2020 · Metabolism consists of a series of reactions that occur within cells of living organisms to sustain life.Missing: enzymology | Show results with:enzymology
  82. [82]
    Biochemistry, Glycolysis - StatPearls - NCBI Bookshelf - NIH
    During glycolysis, glucose ultimately breaks down into pyruvate and energy; a total of 2 ATP is derived in the process (Glucose + 2 NAD+ + 2 ADP + 2 Pi --> 2 ...
  83. [83]
    Amino acid metabolism in health and disease - Nature
    Sep 13, 2023 · The oxidation pathway begins with aminotransferase-mediated deamination and transfers the amino group to alpha-ketoglutaric acid to form ...
  84. [84]
    Amino Acid Catabolism: An Overlooked Area of Metabolism - PMC
    Jul 29, 2023 · (B) Amino acid catabolism in the liver involves several degrading processes, including the removal and transfer of the amino group, deamination ...
  85. [85]
    Physiology, Gluconeogenesis - StatPearls - NCBI Bookshelf - NIH
    Nov 13, 2023 · Pyruvate carboxylase requires ATP and the coenzyme biotin for activation. This conversion is the first step that reverses the nonequilibrium ...
  86. [86]
    Gluconeogenesis Flux in Metabolic Disease - Annual Reviews
    Abstract. Gluconeogenesis is a critical biosynthetic process that helps maintain whole- body glucose homeostasis and becomes altered in certain medical ...
  87. [87]
    Gluconeogenesis - an overview | ScienceDirect Topics
    Gluconeogenesis Energy Cost​​ In sum, synthesis of one molecule of glucose from two of pyruvate is coupled with the conversion of six molecules of ATP to ADP. In ...
  88. [88]
    Fatty Acid Biosynthesis Revisited: Structure Elucidation and ...
    Typically, fatty acid biosynthesis begins with acetyl-CoA, carboxylation produces the malonyl-CoA building blocks that are subsequently condensed and reduced in ...
  89. [89]
    Fatty Acid Synthesis - an overview | ScienceDirect Topics
    The core of fatty acid synthesis is the production of palmitate from acetyl-CoA and malonyl-CoA in a reaction that requires NADPH and is catalyzed by fatty acid ...
  90. [90]
    Purine and Pyrimidine Nucleotide Synthesis and Metabolism - PMC
    Using 5-phosphoribosyl-1-pyrophosphate (PRPP), the de novo pathway enzymes build purine and pyrimidine nucleotides from “scratch” using simple molecules such as ...
  91. [91]
    Amino Acid Metabolism - PMC - PubMed Central - NIH
    The carbon skeleton of amino acids can be converted into TCA cycle intermediates that can be used either to generate ATP through oxidative phosphorylation or ...
  92. [92]
    Alpha-Ketoglutarate: Physiological Functions and Applications - PMC
    Jan 1, 2016 · In addition, glutamate, released from nerve fibers in bone tissue, is synthesized by the reductive amination of AKG in peri-vein hepatocytes ( ...
  93. [93]
    Principles and functions of metabolic compartmentalization - PMC
    Oct 20, 2022 · The metabolic state within compartments is integrated through multiple levels in an organism by the activity of signal transduction pathways.
  94. [94]
    Current perspective on the role of insulin and glucagon in the ...
    May 9, 2019 · In this review, we begin by evaluating the principal differences between insulin and glucagon with regard to their mechanism and control of their secretion.
  95. [95]
    New Insights Into the Role and Mechanism of Glycogen Synthase ...
    Insulin causes stable activation of glycogen synthase by promoting dephosphorylation of multiple sites in the enzyme. A model linking this action to the mitogen ...
  96. [96]
    Structural basis for allosteric regulation of human ... - PubMed Central
    Mar 16, 2024 · ATP and fructose-2,6-bisphosphate regulate skeletal muscle 6-phosphofructo-1-kinase by altering its quaternary structure. IUBMB Life 60, 526 ...
  97. [97]
    Structural basis for allosteric regulation of human ... - Nature
    Aug 25, 2024 · ATP and fructose-2,6-bisphosphate regulate skeletal muscle 6-phosphofructo-1-kinase by altering its quaternary structure. IUBMB Life 60, 526 ...
  98. [98]
    Control-pattern Analysis of Metabolic Pathways. Flux and ... - PubMed
    This paper describes a non-algebraic diagrammatic method which generates the mathematical expressions for flux or concentration-control coefficients in terms of ...
  99. [99]
    Free energy - CHEM 245 Biochemistry
    The change in Gibbs free energy, ΔG, for any process which occurs at constant temperature is given by the equation ΔG = ΔH − TΔS. In what follows, we ...
  100. [100]
    Energy and enzymes | Biological Principles
    Gibbs Free Energy. Gibbs free energy is a measure of the amount of work that ... ΔG = ΔH – TΔS. where H = enthalpy (a measure of energy content that is ...Missing: biochemistry | Show results with:biochemistry
  101. [101]
    [PDF] Biological Chemistry I: Biochemical Transformations II
    ΔGº' = +13.8 kj/mol. ATP + H2O ADP + Pi. ΔGº' = -30.5 kj/mol. Conclusion: the thermodynamic lability of ATP is used frequently to drive reactions to the.
  102. [102]
    [PDF] Notes LECTURE 5 Biochemistry 9.17.04
    Sep 17, 2004 · In the world of thermodynamics, exergonic reactions are considered. “spontaneous.” But in reality, despite being energetically favorable, some ...
  103. [103]
    [PDF] 7.014 Handout
    Keq >> 1. ΔGo' has large negative value, therefore reaction ⇒ is irreversible. Keq << 1. ΔGo' has large positive value, therefore reaction ⇒ cannot occur. 2 ...Missing: near- | Show results with:near-
  104. [104]
    [PDF] Effects of size and temperature on metabolic rate
    Sep 21, 2001 · Temperature governs metabolism through its effects on rates of biochemical reactions. Reaction kinetics vary with temperature according to the ...
  105. [105]
    Physiology, Adenosine Triphosphate - StatPearls - NCBI Bookshelf
    Through metabolic processes, ATP becomes hydrolyzed into ADP, or further to AMP, and free inorganic phosphate groups. The process of ATP hydrolysis to ADP ...Missing: phosphoanhydride | Show results with:phosphoanhydride
  106. [106]
    4.4 ATP: Adenosine Triphosphate – Human Biology
    The bonds that connect the phosphates (phosphoanhydride bonds) have high-energy content. The energy released from ATP hydrolysis into ADP + Pi performs cellular ...
  107. [107]
    What is the redox potential of a cell? - Bionumbers book
    ... oxidation of NADH (produced, for example, in the TCA cycle). The NAD+/NADH pair has a redox potential of E = -0.32 V and it is oxidized by oxygen to give ...
  108. [108]
    Redox Reactions - Oxidation & Reduction
    Half reactions: Reduction or oxidation? ; (1), fumarate + 2H+ + 2 e– succinate, E°' = 0.030 V ; (2), FAD + 2H+ + 2 e– FADH · E°' = -0.180 V ...
  109. [109]
    Biochemistry, Electron Transport Chain - StatPearls - NCBI Bookshelf
    Sep 4, 2023 · The electron transport chain is a series of four protein complexes that couple redox reactions, creating an electrochemical gradient that leads to the creation ...
  110. [110]
    [PDF] Peter Mitchell - Nobel Lecture
    I shall presently explain the difference between the physiological and the biochemical levels at which the chemiosmotic theory has helped to promote useful ...
  111. [111]
    Central Dogma of Molecular Biology - Nature
    Article; Published: 08 August 1970. Central Dogma of Molecular Biology. FRANCIS CRICK. Nature volume 227, pages 561–563 (1970)Cite this article.
  112. [112]
    Eukaryotic RNA Polymerases and General Transcription Factors
    The promoters of many genes transcribed by polymerase II contain a sequence similar to TATAA 25 to 30 nucleotides upstream of the transcription start site. This ...
  113. [113]
    From RNA to Protein - Molecular Biology of the Cell - NCBI Bookshelf
    They join together on an mRNA molecule, usually near its 5′ end, to initiate the synthesis of a protein. The mRNA is then pulled through the ribosome; as its ...
  114. [114]
    Fidelity of DNA replication—a matter of proofreading - PMC
    The fidelity of DNA replication relies on nucleotide selectivity of replicative DNA polymerase, exonucleolytic proofreading, and postreplicative DNA mismatch ...
  115. [115]
    DNA Replication—A Matter of Fidelity - ScienceDirect.com
    Jun 2, 2016 · Proofreading by DNA polymerases is the second mechanism that ensures high fidelity during DNA replication, and this can be divided into two ...Main Text · Fidelity Of Dna Polymerases · Mismatch Repair
  116. [116]
    Genetic regulatory mechanisms in the synthesis of proteins - PubMed
    Genetic regulatory mechanisms in the synthesis of proteins. J Mol Biol. 1961 Jun:3:318-56. doi: 10.1016/s0022-2836(61)80072-7. Authors. F JACOB, J MONOD. PMID ...
  117. [117]
    Enhancers: five essential questions - PMC - PubMed Central - NIH
    Enhancers are DNA-regulatory elements that activate transcription of a gene or genes to higher levels than would be the case in their absence. These elements ...
  118. [118]
    DNA Methylation and Its Basic Function | Neuropsychopharmacology
    Jul 11, 2012 · DNA methylation regulates gene expression by recruiting proteins involved in gene repression or by inhibiting the binding of transcription factor(s) to DNA.
  119. [119]
    Integrating mRNA Processing with Transcription - ScienceDirect.com
    The messenger RNA processing reactions of capping, splicing, and polyadenylation occur cotranscriptionally. They not only influence one another's efficiency ...Main Text · Mrna Processing Reactions · Pol Ii Transcriptional...
  120. [120]
    Roles of mRNA poly(A) tails in regulation of eukaryotic gene ...
    Mar 13, 2023 · The cap binds directly to translation factors to promote initiation. Synergy between the 5′ cap and the 3′ poly(A) tail further stimulates this ...
  121. [121]
    x Ray crystallography - PMC - PubMed Central - NIH
    The upper resolution limit required. This will usually be determined by the quality of the crystal itself and in practice most data sets are collected to the ...
  122. [122]
    [PDF] Phase Problem in X-ray Crystallography, and Its Solution
    This limits the resolution of the diffraction pattern to typically 2 or 3 A˚ . Since these molecules may also contain many thousands of atoms, a different ...Missing: biochemistry | Show results with:biochemistry
  123. [123]
    The translating bacterial ribosome at 1.55 Å resolution generated by ...
    Feb 25, 2023 · The ribosome has been a suitable target for single particle cryo-EM due to its large size, globular shape, and high RNA content, which generates ...
  124. [124]
    Molecular Interactions (Noncovalent Interactions) - Loren Williams
    Jun 10, 2024 · Molecular interactions refer to the attractive or repulsive forces between molecules and non-bonded atoms.Missing: biophysical | Show results with:biophysical
  125. [125]
    Dynamics of Ionic Interactions at Protein–Nucleic Acid Interfaces
    Aug 26, 2020 · Electrostatic interactions via ion pairs (salt bridges) of nucleic acid phosphates and protein side chains are crucial for proteins to bind to DNA or RNA.
  126. [126]
    4.5 Mechanisms of Protein Folding | BS1005 / CM1051: Biochemistry I
    How proteins fold to achieve a stable state is still an active area of research. Levinthal's paradox states that proteins cannot fold by sampling all ...
  127. [127]
    Protein Folding in the Cytoplasm and the Heat Shock Response - PMC
    Chaperones such as Hsp70 prevent protein misfolding and aggregation. They are up-regulated when cells are stressed but decline during aging.Protein Folding In The... · Ribosome-Associated... · The Chaperonins
  128. [128]
    The Levinthal Problem in Amyloid Aggregation: Sampling of a ... - NIH
    Here we report a method to efficiently sample this space and generate an ensemble of growth trajectories.Modeling Fibril Growth As A... · Markov State Model Of Fibril... · Net Fibril Growth Rates And...Missing: chaperones Hsp70
  129. [129]
    Lipid microdomains and the regulation of ion channel function - PMC
    Many types of ion channel localize to cholesterol and sphingolipid-enriched regions of the plasma membrane known as lipid microdomains or 'rafts'.Missing: biophysics | Show results with:biophysics