Fact-checked by Grok 2 weeks ago

Enzyme

Enzymes are biological catalysts, primarily proteins but also including molecules known as ribozymes, that accelerate the rate of chemical reactions in living organisms by lowering the required for those reactions without being consumed or permanently altered in the process. These macromolecules possess a complex three-dimensional structure, consisting of one or more polypeptide chains folded into specific shapes, with an —a specialized region that binds to molecules to form an enzyme-substrate complex and facilitate . The binding mechanism often follows the lock-and-key model, where the active site precisely matches the substrate, or the more dynamic induced fit model, in which the enzyme adjusts its conformation upon substrate binding to optimize the reaction. Enzymes operate under physiological conditions, enabling reactions to occur at rates compatible with life, and their activity is influenced by factors such as temperature, pH, and the presence of cofactors or inhibitors. Enzymes are systematically classified by the Nomenclature Committee of the International Union of Biochemistry and (IUBMB) into six major classes based on the type of reaction they catalyze: oxidoreductases (EC 1, electron transfer), transferases (EC 2, group transfer), hydrolases ( 3, hydrolysis), lyases (EC 4, addition or removal of groups to form double bonds), isomerases (EC 5, isomerization), and ligases (EC 6, bond formation coupled to ). Each enzyme is assigned a unique EC number (e.g., EC 1.1.1.1 for ) for precise identification, and most are named with the suffix "-ase" to indicate their catalytic function, such as amylase for . The discovery of enzymes traces back to the , with the term "enzyme" coined by Wilhelm Kühne in 1877 to describe non-cellular catalysts, building on earlier observations like Anselme Payen's isolation of in 1833; the field advanced dramatically in the 1980s with the identification of ribozymes by and , earning them the 1989 . In biological systems, enzymes are indispensable, with thousands present in cells to regulate , , , and other processes essential for sustaining life. Their specificity and efficiency underpin applications, from industrial production to medical diagnostics, highlighting their profound impact across disciplines.

Etymology and History

Etymology

The term "enzyme" originates from the Greek phrase en zymē, meaning "in " or "in leaven," and was coined in 1877 by German physiologist Wilhelm Kühne to describe the active agents responsible for processes observed in . This nomenclature reflected the era's focus on as a primary model for studying catalytic activity in biological systems, distinguishing these agents from the broader category of ferments. In the early 19th century, before the adoption of "enzyme," such catalytic substances were commonly termed "ferments," a concept advanced by chemists including , who interpreted as a purely rather than a process tied exclusively to living organisms. Liebig and contemporaries like used "ferment" to encompass both living and non-living agents that accelerated reactions, though debates persisted over whether these were vital forces or chemical catalysts. By the mid-19th century, ferments were often subdivided into "organized ferments," which were believed to require intact living cells (such as ), and "unorganized ferments," which were thought to act independently. This distinction shifted dramatically in 1897 when Eduard Buchner demonstrated that cell-free extracts could still perform alcoholic , proving that the active principles—now termed enzymes—were soluble substances extractable from cells, independent of vital life processes. Buchner's discovery solidified "enzyme" as the preferred term for these non-cellular catalysts, marking a terminological from organism-bound ferments to biochemical entities.

Historical Development

The recognition of enzymatic processes dates back to ancient civilizations, with the earliest known fermented beverages dating to around 7000 BCE in at the site. Fermentation was subsequently utilized in regions such as and by approximately 4000 BCE for food and beverage production, unknowingly harnessing enzymatic activity in processes like and bread-making. In ancient Greece, philosophers such as documented observations of fermentation, noting in his works how substances like must (grape juice) transformed into wine through a resembling , attributing it to natural changes rather than vital forces. A pivotal advancement occurred in 1833 when French chemists Anselme Payen and Jean-François Persoz isolated diastase from germinating barley malt, marking the first preparation of an enzyme and demonstrating its ability to hydrolyze starch into sugars. This discovery laid the groundwork for enzymology as a scientific discipline, shifting focus from vague biological processes to isolable catalysts. In 1897, Eduard Buchner conducted cell-free fermentation experiments using yeast extracts, proving that enzymes could catalyze reactions without intact living cells and refuting the doctrine of vitalism; for this, he received the 1907 Nobel Prize in Chemistry. The early 20th century saw further purification and mechanistic insights, exemplified by James B. Sumner's 1926 crystallization of from jack bean, confirming enzymes as proteins and earning him the 1946 (shared with John H. Northrop and Wendell M. Stanley for related work). In 1913, and developed the foundational model of , describing the hyperbolic relationship between concentration and through their equation, which revolutionized quantitative studies of . Later discoveries expanded the scope of enzymology beyond proteins. In the 1980s, and independently demonstrated that molecules could act as catalysts, termed ribozymes, challenging the protein-centric view of enzymes and earning them the 1989 . Building on this, in the 1990s, pioneered techniques, randomly mutating enzyme genes and selecting variants for desired functions, with her first successful application in 1993; this method transformed enzyme engineering and led to her 2018 . More recently, the 2024 recognized advancements in and structure prediction, with David Baker awarded for computational methods to create novel enzymes and and John Jumper for AI-based prediction tools like , which have revolutionized enzyme engineering and understanding.

Classification and Nomenclature

Enzyme Commission System

The Enzyme Commission (EC) System was established in 1956 by the International Union of Biochemistry (IUB), now known as the International Union of Biochemistry and Molecular Biology (IUBMB), to provide a standardized framework for classifying enzymes based on the reactions they catalyze and to prevent inconsistencies in nomenclature. This initiative arose from the rapid growth in enzyme discoveries during the mid-20th century, necessitating a systematic approach to organize the expanding body of knowledge. The first report of the Enzyme Commission was published in 1961, laying the foundation for the hierarchical classification that has since become the global standard. Enzymes are grouped into seven main classes under the EC System, each defined by the fundamental type of chemical transformation they facilitate. Oxidoreductases (EC 1) catalyze oxidation-reduction reactions involving , such as dehydrogenases that manage processes in . Transferases (EC 2) facilitate the transfer of functional groups like amino or phosphate groups from one to another. Hydrolases (EC 3) promote reactions, breaking bonds by adding , as seen in proteases and lipases. Lyases (EC 4) add or eliminate groups to form double bonds without or oxidation, including enzymes like decarboxylases. Isomerases (EC 5) enable intramolecular rearrangements, converting a into one of its isomers. Ligases (EC 6), also known as synthetases, join two molecules using from . Translocases (EC 7), added in 2018, catalyze the movement of ions or molecules across membranes or their relocation within membranes. The classification employs a four-digit numerical code in the format EC a.b.c.d, reflecting a hierarchical structure that refines enzyme specificity step by step. The first digit (a) indicates the main (1 through 7), the second (b) denotes the subclass based on the type of reaction or substrate, the third (c) specifies the sub-subclass by further reaction details or bond involvement, and the fourth (d) identifies the of the specific enzyme within that group. For instance, EC designates , an that acts on primary or secondary alcohols using NAD+ as an acceptor. This numbering ensures unique identification and allows for systematic expansion as new enzymes are characterized. The System is dynamically maintained through the ExplorEnz database, the official IUBMB repository for enzyme nomenclature and classification, which undergoes regular updates to incorporate newly discovered or reclassified enzymes. As of the 2025.11 release, the database includes over 6,900 entries, with proposed changes subjected to a four-week public review period before integration. This ongoing curation by the Nomenclature Committee of the IUBMB ensures the system's relevance and accuracy in reflecting advances in enzymology.

Nomenclature Conventions

Enzymes are assigned both trivial and systematic names to facilitate clear and unambiguous communication in scientific literature. Trivial names, also known as recommended or accepted names, are concise and descriptive terms that often reflect the enzyme's function, source, or historical discovery, such as "trypsin" for a serine protease that hydrolyzes peptide bonds in proteins. These names are preferred for general use due to their brevity and familiarity but must be linked to the specific reaction catalyzed to avoid confusion. Systematic names, in contrast, provide a more precise description of the biochemical reaction by specifying the substrates involved and the type of transformation, typically in the format "substrate:product [reaction type]" or similar. For example, the systematic name for is "protein + H₂O = protein + ," emphasizing its activity on peptide bonds. This naming convention ensures that the name directly mirrors the catalyzed , using accepted trivial names for complex substrates where possible to maintain usability. The International Union of Biochemistry and (IUBMB), through its Nomenclature Committee (NC-IUBMB), established these conventions in the 1978 recommendations, which have been periodically updated to incorporate new enzymatic activities and refine terminology. The guidelines mandate that enzyme names reflect the overall reaction catalyzed, based on experimental evidence of physiological substrates, and prohibit names derived solely from sequence similarity or non-enzymatic properties. Updates, such as those in the 1992 edition and subsequent supplements through 2025, address evolving knowledge while preserving the core principles. For multi-substrate reactions, nomenclature prioritizes the primary physiological substrates, listing additional ones in parentheses if they significantly contribute to the enzyme's function, to avoid overly complex names. For instance, in acetyl-CoA C-acetyltransferase (EC 2.3.3.9), the name specifies "acetyl-CoA:acetyl-CoA C-acetyltransferase" but includes qualifiers like "(thioester-hydrolysing, carboxymethyl-forming)" for clarity on mechanism. Ambiguities are resolved by creating separate entries for enzymes with distinct specificities or mechanisms, even if reactions appear similar, ensuring each name corresponds to a unique catalytic profile. Recommended names are finalized after review, while provisional or working names may be used during initial characterization. Enzyme nomenclature integrates with gene naming systems by associating EC numbers with gene symbols in databases, allowing sequence-based annotation of function; for example, the ECOD database uses structural homology to link protein domains to EC classifications, aiding in gene-enzyme correspondence. This complementary approach supports bioinformatics tools in predicting enzymatic roles from genomic data without altering the core reaction-based naming rules.

Structure

Chemical Composition

Enzymes are predominantly composed of proteins, which are linear polymers formed from 20 standard amino acids linked by peptide bonds. These amino acids include glycine, alanine, valine, leucine, isoleucine, phenylalanine, tyrosine, tryptophan, serine, threonine, cysteine, methionine, asparagine, glutamine, aspartic acid, glutamic acid, lysine, arginine, histidine, and proline, each contributing unique side chains that determine the enzyme's functional properties. The primary structure of these polypeptide chains, dictated by the genetic code, serves as the foundation for higher-order organization, ultimately influencing the enzyme's catalytic efficiency. Although most enzymes are proteins, a small subset known as ribozymes are composed of molecules capable of catalytic activity. Notable examples include self-splicing introns in thermophila, where the RNA folds to form an that cleaves and rejoins phosphodiester bonds without protein assistance. Ribozymes demonstrate that RNA can mimic protein enzyme functions, supporting theories on the hypothesis in early evolution. Post-translational modifications significantly alter the chemical composition of enzymes, enhancing their stability, localization, or activity. Common modifications include , where groups are covalently attached to , serine, or residues, and , which adds groups to serine, , or side chains, often regulating enzymatic function through charge alterations. These modifications can introduce functional groups that fine-tune specificity or enable interactions with cellular components. Enzyme molecular weights vary widely but typically range from 10 to 100 kDa for monomeric forms, reflecting the number of (roughly 100-900 residues). Larger enzymes, such as DNA polymerase III, can exceed 300 kDa due to multi-subunit assemblies, allowing for complex functions like . Specific play critical roles in ; for instance, in serine proteases like , acts as a general base to facilitate nucleophilic attack by the serine hydroxyl group. This composition directly impacts the enzyme's ability to fold into functional conformations.

Three-Dimensional Architecture

The three-dimensional architecture of enzymes, as specialized proteins, is organized hierarchically across primary, secondary, , and quaternary levels, with each level contributing to the stability and functional specificity essential for catalysis. The primary structure consists of the linear of covalently linked by bonds, which serves as the foundational blueprint determining all subsequent folding patterns, as established by Christian Anfinsen's thermodynamic hypothesis through experiments on A folding. This sequence encodes the information needed for the protein to achieve its native conformation under physiological conditions, ensuring the enzyme's structural integrity and reactivity. At the secondary structure level, segments of the polypeptide chain fold into local conformations such as α-helices, β-sheets, and unstructured loops, primarily stabilized by hydrogen bonds between backbone atoms. These elements form the building blocks of the enzyme's scaffold, providing rigidity and flexibility that facilitate positioning and dynamic movements during . The tertiary structure represents the global three-dimensional arrangement of these secondary elements into a compact fold, driven by non-covalent interactions including hydrophobic effects that bury nonpolar residues in the core, electrostatic forces, van der Waals interactions, and covalent bridges in some cases. This folding creates a stable globular domain that shields reactive groups and orients functional regions precisely, enabling efficient enzymatic activity. For enzymes requiring cooperative or regulated function, a quaternary structure assembles multiple polypeptide subunits into a functional complex, often through similar intermolecular interactions as in tertiary folding. A representative example is lactate dehydrogenase, a tetrameric enzyme with four identical subunits that enhance allosteric regulation and catalytic efficiency. Common structural motifs recur across enzyme families, underscoring evolutionary conservation for functional versatility; the TIM barrel, characterized by eight alternating α-helices and β-strands forming a cylindrical core, is prevalent in metabolic enzymes like triosephosphate isomerase, providing a robust framework for diverse reactions. Similarly, the Rossmann fold, featuring parallel β-sheets flanked by α-helices, dominates in nucleotide-binding dehydrogenases such as alcohol dehydrogenase, optimizing cofactor interactions. These folds and overall architectures are systematically classified in databases like SCOP (Structural Classification of Proteins) and CATH (Class, Architecture, Topology, Homologous superfamily), which organize enzyme domains by structural similarity to reveal evolutionary relationships and design principles.

Active Site Characteristics

The active site of an enzyme is a specialized region, typically a pocket or cleft on the protein's surface, formed by the precise arrangement of specific residues that enable substrate recognition and . These residues, often distant in the primary sequence, converge in the three-dimensional structure to create a microenvironment optimized for chemical reactions, with the site's dictating the enzyme's specificity and efficiency. For instance, in serine proteases like , the active site features a composed of serine, , and aspartate residues, where the nucleophilic serine hydroxyl group is positioned for attack on the , facilitated by hydrogen bonding from histidine and stabilization by aspartate. Binding pockets within the are tailored to accommodate based on their chemical properties; hydrophobic pockets, lined with non-polar such as or , interact with non-polar substrates through van der Waals forces, while charged or polar pockets incorporate residues like or glutamate to form electrostatic interactions or bonds with polar or ionic substrates. Specificity is further determined by the active site's , , and electrostatic profile, which ensure selective substrate binding and exclude non-cognate molecules—for example, the oxyanion hole in proteases, formed by backbone groups, stabilizes negatively charged states through bonding, enhancing catalytic precision. Visualization of active site characteristics has been advanced through techniques such as , (NMR) spectroscopy, and cryo-electron microscopy (cryo-EM), which reveal atomic-level details of the site's architecture. A classic example is hen egg-white , where in the disclosed a long, narrow cleft in the that accommodates the , with key residues like Glu35 and Asp52 positioned for . These methods, often complemented by computational modeling, continue to elucidate how subtle variations in active site features underpin enzymatic diversity across biological systems.

Mechanism of Catalysis

Substrate Binding

Substrate binding is the initial step in , where the substrate molecule is recognized and attached to the enzyme's through specific interactions that ensure high selectivity. The classical model for this process, proposed by in 1894, describes the enzyme and as possessing complementary geometric shapes, analogous to a , allowing only the correct substrate to fit precisely into the rigid of the enzyme. This lock-and-key hypothesis emphasized the structural specificity that prevents non-substrate molecules from binding effectively, thereby explaining the observed selectivity of enzymatic reactions. Subsequent observations revealed limitations in the rigid lock-and-key model, leading to the induced fit model introduced by Daniel E. Koshland in 1958. In this framework, the enzyme undergoes a conformational change upon initial contact, adjusting its to achieve a tighter, more complementary fit that not only enhances specificity but also positions catalytic residues optimally for the reaction. This dynamic adjustment is particularly evident in enzymes like , where binding induces a large-scale closure of the cleft, excluding water and stabilizing the . The energy of substrate binding arises primarily from non-covalent interactions between the substrate and amino acid residues in the active site, including hydrogen bonds, ionic bonds, and van der Waals forces, which collectively contribute to the enzyme's specificity by discriminating against incorrect substrates. These interactions lower the free energy of the enzyme-substrate complex relative to the unbound state, with hydrogen bonds often providing directional precision and van der Waals forces enabling close-range attractions. The strength of substrate binding is quantitatively assessed by the dissociation constant K_d, defined as K_d = \frac{[E][S]}{[ES]}, where [E] is the concentration of free enzyme, [S] is the concentration of free , and [ES] is the concentration of the enzyme-substrate complex at . A lower K_d value indicates higher affinity, reflecting tighter interactions that are crucial for efficient in physiological conditions.

Catalytic Process

Enzymes facilitate the transformation of substrates into products by stabilizing the of the reaction, thereby lowering the barrier through mechanisms such as electrostatic interactions, acid-base , and the formation of covalent intermediates. Electrostatic involves charged residues in the that stabilize polar transition states via hydrogen bonding or ionic interactions, while acid-base employs proton donors and acceptors, often from side chains like or aspartate, to facilitate proton transfer during the reaction. Covalent occurs when a nucleophilic group on the enzyme, such as a serine residue, forms a transient with the , creating a that lowers the energy of subsequent steps. In hydrolases, such as serine proteases, catalysis often proceeds via nucleophilic attack by the enzyme's serine hydroxyl group on the substrate's carbonyl carbon, forming a tetrahedral that is stabilized by the oxyanion hole in the . For oxidoreductases, exemplified by , the involves hydride transfer from the to the NAD+ coenzyme, facilitated by coordination that polarizes the and positions it for efficient transfer. These catalytic strategies enable enzymes to accelerate rates by factors up to 10^{20}-fold compared to uncatalyzed reactions in . The , denoted as k_{cat}, represents the maximum number of molecules converted to product per enzyme per second under saturating conditions, reflecting the enzyme's catalytic once the substrate is bound.

Conformational Dynamics

Enzymes exhibit intrinsic conformational dynamics that involve flexible movements such as hinge bending and loop closing, which are essential for their . These motions allow the protein to between open and closed states, facilitating the of and the stabilization of states during . For instance, in , binding of glucose induces a large hinge-bending motion between its two domains, closing the cleft by approximately 8 Å to enclose the substrate and exclude water, thereby enhancing specificity and efficiency. Molecular dynamics (MD) simulations and are key techniques for probing these dynamics at atomic and temporal resolutions. MD simulations model the time evolution of enzyme structures, revealing how drive hinge and loop motions on picosecond to microsecond timescales, as demonstrated in studies of where hinge unfolding correlates with substrate access. , such as single-molecule (smFRET), captures real-time conformational changes, for example, in , where active-site fluctuations occur on scales and synchronize with catalytic turnover. These dynamics play a critical role in enabling substrate entry and exit while stabilizing the transition state to lower activation barriers. In the catalytic cycle, loop closing motions position catalytic residues optimally, reducing entropy loss and promoting efficient proton transfer or nucleophilic attack. For HIV-1 protease, conformational fluctuations in the flap regions—opening and closing on millisecond timescales—allow substrate binding and have informed inhibitor design by targeting semi-open states to trap the enzyme in non-productive conformations, improving antiviral efficacy. Such intrinsic flexibility underpins the induced fit model, where dynamics adapt the enzyme to substrates without requiring external regulators.

Cofactors and Coenzymes

Inorganic Cofactors

Inorganic cofactors, primarily divalent and ions such as Zn²⁺, Mg²⁺, Fe²⁺/Fe³⁺, and Cu²⁺, serve as essential non-organic auxiliaries in , enabling reactions without being consumed themselves. These ions are present in more than one-third of known proteins, where they coordinate with specific residues to facilitate diverse biochemical processes. Representative examples illustrate their catalytic roles. In II, Zn²⁺ acts as a acid by coordinating a , lowering its from approximately 10 to 7 and generating a nucleophilic ion at physiological to hydrate CO₂ into . Similarly, Mg²⁺ in kinases like coordinates the phosphate groups of ATP and , optimizing the nucleophilic attack angle to nearly 180° and enhancing phosphoryl transfer efficiency by over 40,000-fold. In , Fe²⁺/Fe³⁺ undergoes cycling within the group, transferring electrons in mitochondrial with a standard around +260 mV that supports efficient energy capture. The primary functions of these cofactors include , where -inert ions like Zn²⁺ polarize substrates to activate nucleophiles; mediation, as seen with ions that alternate oxidation states to shuttle electrons; and structural stabilization, such as Zn²⁺ reinforcing protein folds in zinc-finger motifs to maintain integrity. Metal ions typically bind through coordinate covalent interactions with side chains, often forming tetrahedral geometries with imidazole nitrogens—for example, Zn²⁺ in is ligated by His94, His96, and His119 alongside a . These bindings integrate seamlessly with the apoenzyme structure to form the active holoenzyme. Disruptions in metal can severely impair enzyme . In , mutations in the ATP7B gene cause hepatic accumulation, reducing activity in Cu-dependent enzymes such as (a ferroxidase), (essential for ATP synthesis), and (an ), leading to , liver damage, and neurological deficits.

Organic Coenzymes

Organic coenzymes are non-protein organic compounds, typically derived from , that serve as transient carriers of chemical groups during enzymatic reactions. Unlike prosthetic groups tightly bound to enzymes, these coenzymes loosely associate with the and as second substrates, accepting or donating moieties such as ions, electrons, acyl groups, or amino groups to facilitate . Their vitamin origins ensure dietary dependence, as most organisms cannot synthesize these precursors de novo. A key example is (NAD⁺/NADH), biosynthesized from (vitamin B₃). In dehydrogenases, NAD⁺ acts as an oxidant by accepting a ion (H⁻) from the at the C4 position of its ring, forming NADH and enabling oxidation reactions such as the conversion of to pyruvate. NADH subsequently donates the hydride to other acceptors, regenerating NAD⁺ for reuse in a cyclic manner, distinct from the consumption of primary substrates. Flavin adenine dinucleotide (FAD), derived from (vitamin B₂), participates in electron transport within oxidoreductases. FAD accepts two electrons and two protons to form FADH₂ through semiquinone intermediates, supporting reactions like succinate oxidation to fumarate; it is then reoxidized by downstream electron acceptors such as ubiquinone, allowing continuous cycling without net consumption. Coenzyme A (CoA), synthesized from (vitamin B₅), functions as an carrier in enzymes involved in . The terminal (-SH) group of CoA forms high-energy linkages with acyl moieties, as in during fatty acid β-oxidation or citrate synthesis; upon transfer of the acyl group to an acceptor, free CoA is released and recycled for subsequent activations. Pyridoxal 5'-phosphate (), derived from (vitamin B₆), enables amino group transfer in transaminases and other amino acid-modifying enzymes. forms a covalent with the amino acid via its group, facilitating proton abstraction and group exchange—such as converting an amino acid to its corresponding α-keto acid—before reacting with a second to regenerate the free coenzyme. These coenzymes are regenerated through coupled metabolic pathways, ensuring their availability for multiple turnover events in enzymatic catalysis.

Thermodynamics

Energy Barriers

Enzymes accelerate chemical reactions by lowering the activation energy barrier, E_a, which is the energy required to reach the transition state from the reactants. According to the Arrhenius equation, the rate constant k of a reaction is given by k = A e^{-E_a / RT}, where A is the pre-exponential factor, R is the gas constant, and T is the temperature in Kelvin; a reduction in E_a exponentially increases the reaction rate. In enzymatic catalysis, this lowering of E_a enables reactions to proceed at biologically relevant rates and temperatures, often by orders of magnitude faster than the uncatalyzed counterparts. Transition state theory provides the thermodynamic framework for understanding this effect, positing that enzymes stabilize the transition state (TS) more tightly than the ground-state substrates or products, thereby reducing the of , \Delta G^\ddagger. This preferential binding to the TS, first proposed by , decreases \Delta G^\ddagger and thus E_a, as the enzyme's is complementary to the TS geometry and charge distribution. For instance, in chorismate mutase, electrostatic interactions at the stabilize the TS of the , contributing up to 5.9 kcal/mol to the \Delta G^\ddagger reduction through specific residues like Arg90. The reduction in often involves both enthalpic and contributions, with enzymes minimizing the penalty associated with organizing reactants for reaction. In solution, reactions incur a high cost due to the loss of translational and rotational freedom upon forming the reactive complex, as well as reorganization . Enzymes counteract this through pre-organization of the , which is already configured to complement the TS, thereby reducing the need for large conformational adjustments and associated loss during . This pre-organization effectively lowers the overall \Delta G^\ddagger by favoring enthalpic stabilization while mitigating barriers, as seen in cases where the enzyme environment substitutes for in a low-dielectric . A striking example is orotidine 5'-monophosphate decarboxylase (OMPDC), which catalyzes the of orotidine 5'-monophosphate to uridine 5'-monophosphate with a rate acceleration of approximately $10^{17}-fold compared to the uncatalyzed reaction. This enormous enhancement arises primarily from electrostatic stabilization of the vinyl anion-like , where active site residues such as Lys72 position charges to delocalize the negative charge developing during . While additional interactions, like a from Ser127 to the substrate's O4, contribute modestly (about $10^2-fold), the dominant effect is the enthalpic stabilization of the TS through electrostatic pre-organization, underscoring how enzymes exploit such mechanisms to surmount energy barriers efficiently.

Equilibrium and Reversibility

Enzymes catalyze chemical reactions by lowering the barrier, thereby accelerating the rate at which the system approaches , but they do not alter the position of or the standard change (ΔG°) of the reaction. The change for a reaction is given by ΔG = ΔH - TΔS, where ΔH is the change, T is the absolute , and ΔS is the change; enzymes influence neither the thermodynamic parameters ΔH nor ΔS, nor the resulting ΔG°, which determines the spontaneity and of the reaction. Instead, enzymes facilitate the reaction along a pathway that stabilizes the , allowing the system to reach the same ratio of reactants to products more rapidly than in the uncatalyzed case. Most enzymatic reactions are reversible, meaning the enzyme can catalyze the reaction in either direction depending on and product concentrations, as long as the reaction does not violate thermodynamic constraints. For instance, isomerases such as triose phosphate isomerase in interconvert and in a readily reversible manner, with an close to unity that reflects minimal difference between substrates. In contrast, some reactions appear effectively irreversible under physiological conditions due to a highly negative ΔG, such as the of ATP to ADP and inorganic catalyzed by ATPases, where ΔG° is approximately -30.5 kJ/mol, driving the reaction overwhelmingly toward product formation and preventing significant reversal. The relationship between enzymatic kinetic parameters and is encapsulated in the Haldane equation, derived from steady-state kinetics, which connects the (K_eq = [P]/[S] at ) to the maximum velocities (k_cat) and Michaelis constants (K_m) for the forward and reverse reactions: K_eq = (k_cat^f / K_m^f) / (k_cat^r / K_m^r), where superscripts f and r denote forward and reverse directions. This equation demonstrates that while enzymes enhance rates through favorable kinetic constants, these parameters must align with the underlying to maintain the position dictated by ΔG°. In metabolic pathways, enzymes often participate in coupled reactions to drive thermodynamically unfavorable (endergonic) steps by linking them to highly exergonic reactions, such as , through shared intermediates that make the overall process spontaneous. For example, the endergonic synthesis of from glutamate and (ΔG° ≈ +14 kJ/mol) is coupled to via , resulting in a net ΔG° of about -16 kJ/mol and rendering the coupled reaction effectively irreversible in the forward direction. This coupling ensures that non-spontaneous transformations proceed efficiently within cellular constraints, without enzymes altering the intrinsic ΔG° of individual steps.

Kinetics

Rate Laws

The rate of an enzyme-catalyzed reaction is defined as the v = \frac{d[P]}{dt}, where [P] is the product concentration, and this rate depends on the total enzyme concentration [E], concentration [S], and environmental conditions such as and . This general law provides the foundation for analyzing how enzymes accelerate reactions by stabilizing states, though rates are ultimately constrained by thermodynamic principles governing energies. Enzyme reaction orders vary with substrate concentration, reflecting the saturation behavior of the enzyme active site. At saturating [S], where all enzyme molecules are bound to substrate, the reaction exhibits zero-order kinetics: the rate becomes independent of further increases in [S] and is determined solely by the enzyme's catalytic capacity, often expressed as v = k_{\text{cat}} [E]. In contrast, at low [S] where substrate binding is limiting, the reaction follows first-order kinetics, with v proportional to [S] as the probability of enzyme-substrate encounters increases linearly. These orders highlight the transition from substrate-limited to enzyme-limited regimes in typical enzyme kinetics profiles. For enzymes acting on multiple substrates, rate laws adopt more complex forms based on the and . Sequential require all substrates to bind to the enzyme before any product is released; in ordered sequential , substrates bind in a specific sequence, while random sequential allows any order, leading to rate equations that include terms for , , and higher complexes. Ping-pong , also known as double-displacement, involve the enzyme first binding one to form and release a product, modifying the enzyme (e.g., via covalent ), before the second binds to the altered form; this results in parallel line patterns in double-reciprocal plots and rate laws featuring reciprocal terms without cross-interaction products. These distinctions, formalized by Cleland's nomenclature, enable mechanistic diagnosis through initial velocity studies varying concentrations. A key approximation in deriving enzyme rate laws is the steady-state assumption, proposed by Briggs and Haldane, which holds that during the initial phase of the reaction—before significant product accumulation or substrate depletion—the concentration of the enzyme-substrate complex [ES] remains nearly constant, such that \frac{d[ES]}{dt} \approx 0. This condition arises because the rates of ES formation and breakdown equilibrate rapidly compared to overall product formation, simplifying the differential equations for [ES] and allowing focus on initial velocities under controlled conditions. The assumption is valid when [E] << [S] and is widely applied in kinetic analyses to predict rates without solving full time-dependent systems.

Michaelis-Menten Model

The Michaelis-Menten model describes the kinetics of enzyme-catalyzed reactions involving a single substrate, providing a foundational framework for understanding how reaction velocity depends on substrate concentration. Originally proposed by and in 1913 based on equilibrium assumptions for the invertase reaction, the model was refined in 1925 by and using a steady-state approximation, which is the standard form used today. The model assumes the reaction proceeds via the formation of an enzyme-substrate (ES) complex: E + S \underset{k_{-1}}{\stackrel{k_1}{\rightleftharpoons}} ES \stackrel{k_{\text{cat}}}{\rightarrow} E + P where E is the , S the , P the , k_1 the association rate constant, k_{-1} the dissociation rate constant, and k_{\text{cat}} (also called k_2) the catalytic rate constant. The initial reaction velocity v is given by the : v = \frac{V_{\max} [S]}{K_m + [S]} Here, V_{\max} is the maximum velocity achieved when the enzyme is saturated with substrate, defined as V_{\max} = k_{\text{cat}} [E]_{\text{total}}, where [E]_{\text{total}} is the total enzyme concentration; K_m is the , representing the substrate concentration at which v = V_{\max}/2, and given by K_m = (k_{-1} + k_{\text{cat}})/k_1. This hyperbolic relationship indicates that velocity increases with substrate concentration but approaches V_{\max} asymptotically. The steady-state derivation begins by applying the quasi-steady-state approximation to the ES complex, assuming d[ES]/dt \approx 0 after an initial transient phase. The rate of ES formation equals its rate of depletion: k_1 [E] [S] = (k_{-1} + k_{\text{cat}}) [ES] Since [E] = [E]_{\text{total}} - [ES], solving for [ES] yields: [ES] = \frac{[E]_{\text{total}} [S]}{K_m + [S]} The velocity v = k_{\text{cat}} [ES] then substitutes to give the . This approach relaxes the rapid equilibrium assumption of the original model, making it applicable to a broader range of enzymes where the catalytic step is not necessarily slow compared to dissociation. To estimate K_m and V_{\max} experimentally, the Lineweaver-Burk double-reciprocal plot linearizes the equation: \frac{1}{v} = \frac{K_m}{V_{\max}} \cdot \frac{1}{[S]} + \frac{1}{V_{\max}} Plotting $1/v versus $1/[S] produces a straight line with slope K_m / V_{\max}, y-intercept $1/V_{\max}, and x-intercept -1/K_m. Introduced by and in 1934, this transformation facilitates parameter determination from initial rate data but can amplify errors at low substrate concentrations. The model relies on key assumptions, including measurement of initial rates where product accumulation is negligible (avoiding product inhibition or reverse reactions) and the absence of complicating factors like multiple substrates or enzyme instability. It applies well to non-allosteric enzymes but has limitations for allosteric enzymes, where cooperative substrate binding leads to sigmoidal rather than hyperbolic kinetics, as described in the .

Inhibition

Reversible Inhibition

Reversible inhibition occurs when an inhibitor binds non-covalently to an enzyme, forming a reversible complex that can dissociate, thereby modulating enzyme activity without permanent alteration. This type of inhibition is characterized by equilibrium binding and can be analyzed using modifications of the , where the inhibition constant K_i quantifies the inhibitor's affinity for the enzyme, with lower values indicating stronger binding. In competitive inhibition, the inhibitor binds exclusively to the free enzyme at the active site, competing directly with the substrate and preventing substrate binding. This increases the apparent Michaelis constant K_m while leaving the maximum velocity V_{max} unchanged, as higher substrate concentrations can outcompete the inhibitor. The modified velocity equation is: v = \frac{V_{max} [S]}{K_m (1 + \frac{[I]}{K_i}) + [S]} A representative example is the action of statins, such as lovastatin, which competitively inhibit by mimicking the substrate and binding to its active site, thereby reducing cholesterol biosynthesis. Non-competitive inhibition involves the inhibitor binding to a site distinct from the active site on either the free enzyme or the enzyme-substrate complex with equal affinity, thereby reducing the enzyme's catalytic efficiency without affecting substrate binding. This decreases the apparent V_{max} but leaves K_m unchanged, as the inhibitor does not interfere with substrate affinity. The velocity equation becomes: v = \frac{V_{max} [S]}{(K_m + [S]) (1 + \frac{[I]}{K_i})} Heavy metals like mercury or lead exemplify non-competitive inhibitors, binding to sulfhydryl groups on enzymes such as pyruvate kinase and impairing function regardless of substrate presence. Uncompetitive inhibition is distinguished by the inhibitor binding solely to the enzyme-substrate complex, stabilizing it and preventing product formation. This results in a decrease in both apparent K_m and V_{max}, with the reduction in K_m arising from the inhibitor's enhancement of substrate affinity in the complex. The kinetic equation is: v = \frac{V_{max} [S]}{K_m + [S] (1 + \frac{[I]}{K_i})} Lithium serves as an uncompetitive inhibitor of inositol monophosphatase, binding to the enzyme-substrate complex and inhibiting dephosphorylation of inositol phosphates, a mechanism implicated in its therapeutic effects on bipolar disorder. Mixed inhibition encompasses cases where the inhibitor binds to both the free enzyme and the enzyme-substrate complex, but with differing affinities, combining elements of competitive and non-competitive inhibition. It alters both K_m and V_{max}, with the degree of change depending on the relative dissociation constants K_i (for free enzyme) and K_i' (for the complex). The general velocity equation is: v = \frac{V_{max} [S]}{K_m (1 + \frac{[I]}{K_i}) + [S] (1 + \frac{[I]}{K_i'})} The inhibition constant K_i is formally defined as the dissociation constant for the enzyme-inhibitor complex, K_i = \frac{[E][I]}{[EI]}, providing a measure of binding strength that is central to comparing inhibitor potencies across these reversible mechanisms.

Irreversible Inhibition

Irreversible inhibition occurs when an inhibitor forms a covalent bond with the enzyme, permanently inactivating it and preventing substrate binding or catalysis. This contrasts with reversible inhibition, where the inhibitor can dissociate from the enzyme. The process typically involves a two-step mechanism: an initial non-covalent binding step followed by irreversible covalent modification. A common mechanism is the nucleophilic attack by an enzyme residue, such as a serine hydroxyl group, on an electrophilic center in the inhibitor, leading to covalent adduct formation. For instance, acts as an irreversible inhibitor of bacterial by mimicking the D-Ala-D-Ala substrate; the β-lactam ring opens upon binding, allowing the serine nucleophile to acylate the inhibitor, blocking peptidoglycan cross-linking in cell walls. The kinetics of irreversible inhibition are characterized by time-dependent loss of enzyme activity, often modeled as pseudo-first-order inactivation. The observed rate constant k_{\text{obs}} follows the equation: k_{\text{obs}} = \frac{k_{\text{inact}} [I]}{K_I + [I]} where k_{\text{inact}} is the maximum inactivation rate constant, [I] is the inhibitor concentration, and K_I is the dissociation constant for the initial enzyme-inhibitor complex. This hyperbolic relationship allows determination of potency through k_{\text{inact}}/K_I, a second-order rate constant reflecting overall efficiency. Representative examples include aspirin, which irreversibly acetylates Ser530 in the of cyclooxygenase-1 (COX-1), inhibiting synthesis and platelet aggregation. Organophosphates, such as those in pesticides, phosphorylate the active-site serine in , preventing hydrolysis and causing cholinergic toxicity. Suicide inhibitors, also known as mechanism-based inhibitors, are prodrugs activated by the target enzyme's catalytic machinery to generate a reactive species that covalently modifies the enzyme. A key example is 5-fluorouracil (5-FU), metabolized to 5-fluoro-2'-deoxyuridine-5'-monophosphate (FdUMP), which forms a stable ternary complex with and 5,10-methylenetetrahydrofolate, irreversibly inhibiting the enzyme and disrupting in cancer cells.

Regulation

Allosteric Effects

Allosteric effects in enzymes arise from the binding of regulatory molecules, termed effectors, to dedicated sites remote from the catalytic active site, inducing conformational changes that alter the enzyme's affinity for its substrate or its catalytic rate. This non-competitive modulation allows precise control of metabolic pathways without directly competing at the active site. Unlike simple inhibition or activation at the active site, allostery enables integrated responses to cellular signals, often in oligomeric enzymes where subunit interactions propagate the effect across the protein structure. Allosteric sites are structurally distinct from the , typically located at subunit interfaces or on non-catalytic domains, enabling specific recognition of effectors. A well-characterized example is aspartate transcarbamoylase (ATCase), the committed enzyme in pyrimidine biosynthesis, which consists of catalytic and regulatory subunits. (CTP), the pathway's end product, binds to allosteric sites on the regulatory subunits, stabilizing a low-affinity tense state and inhibiting activity by enhancing the sigmoidal response to aspartate, the substrate; this reduces enzyme velocity at physiological concentrations. In contrast, (ATP), signaling abundance, binds to the same sites, competing with CTP to favor a high-affinity relaxed state and activate the enzyme. These heterotropic effects fine-tune nucleotide balance without covalent modification. Theoretical models elucidate how allosteric binding translates to functional changes. The concerted Monod-Wyman-Changeux (MWC) model describes the enzyme as existing in equilibrium between a tense (T) state of low substrate affinity and a relaxed (R) state of high affinity; all subunits transition simultaneously upon effector binding, shifting the T-R equilibrium without hybrid intermediates. This symmetry-conserving mechanism explains both homotropic substrate and heterotropic regulation, where inhibitors stabilize the T state and activators favor the R state. Alternatively, the sequential Koshland-Némethy-Filmer (KNF) model proposes an induced-fit process: binding to one subunit triggers a localized conformational change that sequentially alters adjacent subunits' affinities, allowing asymmetric intermediates and greater flexibility in cooperativity patterns. These models, while idealized, capture the essence of allosteric propagation in multisubunit enzymes. Cooperativity, a hallmark of allosteric enzymes, manifests as interdependent sites, yielding sigmoidal kinetics that amplify responses to concentration changes. The Hill equation quantifies this: \theta = \frac{[S]^{n_H}}{K_{0.5}^{n_H} + [S]^{n_H}} where \theta represents fractional , [S] is concentration, K_{0.5} is the concentration for half-maximal , and n_H (the Hill coefficient) indicates degree: n_H > 1 for positive (enhanced after initial ), n_H < 1 for negative, and n_H = 1 for non-cooperative (Michaelis-Menten) . In ATCase, for instance, n_H \approx 1.5-2 reflects moderate positive homotropic for aspartate. Allostery distinguishes homotropic effects, where the itself acts as effector to drive its own , from heterotropic effects, where non- molecules like CTP or ATP modulate independent of . These interactions underpin sensitive regulatory switches in cellular .

Covalent Modifications

Covalent modifications represent a key mechanism for the post-translational regulation of enzyme activity, involving the addition or removal of chemical groups that can reversibly activate or deactivate enzymes through enzymatic control. These modifications allow cells to rapidly respond to signals by altering enzyme function without synthesizing new proteins. is one of the most prevalent covalent modifications, where kinases transfer a group from ATP to , , or residues on the enzyme, often inactivating it, while phosphatases remove the phosphate to restore activity. For instance, , which catalyzes synthesis, is inactivated by multi-site on serine residues by kinases such as cAMP-dependent and glycogen synthase kinase-3, increasing its Km for UDP-glucose and reducing catalytic efficiency; dephosphorylation by protein phosphatase-1 reactivates it. This reversible cycle exemplifies ultrasensitive , enabling switch-like responses to hormonal cues like or insulin. Other covalent modifications include , where acetyl groups are added to residues by histone acetyltransferases (e.g., p300) and removed by deacetylases (e.g., HDACs or sirtuins), modulating enzyme activity in metabolic pathways; , involving methyltransferases adding methyl groups to or , which can alter substrate binding or stability; and ubiquitination, where is attached via E1, E2, and E3 enzymes to mark enzymes for proteasomal degradation, thereby controlling protein levels and indirectly regulating activity. A notable example of irreversible covalent modification is activation, such as the proteolytic cleavage of to active by in the , which removes an N-terminal and forms stabilizing salt bridges, enabling the enzyme's catalytic function in protein . In signal transduction cascades, these modifications often operate reversibly; for example, in insulin signaling, phosphorylation of downstream enzymes like by insulin-stimulated kinases promotes , while fine-tunes the response to maintain .

Environmental Factors

Enzyme activity is profoundly influenced by environmental , which modulates the ionization states of catalytic residues such as , aspartate, and glutamate in the . Deviations from the optimal pH can protonate or deprotonate these residues, disrupting binding or . For instance, , a in the gastric environment, exhibits maximal activity at pH 1.5–2, where acidic conditions facilitate its function in protein . In contrast, , involved in , achieves peak activity at pH 9–10, reflecting adaptation to alkaline cellular or extracellular compartments. The pH dependence typically manifests as a bell-shaped activity , with the optimum corresponding to the average of key ionizable groups; activity declines sharply outside this range due to altered electrostatic interactions. Temperature exerts a dual effect on enzymes, initially enhancing reaction rates through increased molecular collisions and , as described by the , which relates the rate constant k to T via k = A e^{-E_a / RT}, where A is the , E_a is the , and R is the . Most mammalian enzymes operate optimally near 37°C, but exceeding this threshold leads to thermal denaturation, where hydrophobic interactions and hydrogen bonds weaken, causing irreversible unfolding and loss of native structure. Thermostable enzymes, such as Taq from , maintain activity up to 72–80°C, enabling applications like due to their resistance to denaturation. Ionic strength and specific salts impact enzyme function by modulating electrostatic forces within the protein and between the enzyme and . Higher can screen charges, stabilizing folded states or alleviating repulsion in active sites, though excessive levels may disrupt salt bridges. Divalent cations like Ca²⁺ often serve as essential cofactors, binding to specific sites to rigidify structures or participate in ; for example, A₂ requires Ca²⁺ for interfacial activation and . Monovalent ions such as Na⁺ or K⁺ can activate certain enzymes at low concentrations by facilitating conformational changes, but inhibitory effects emerge at higher levels. Denaturation represents a critical environmental , particularly from , resulting in irreversible unfolding that exposes hydrophobic cores and promotes aggregation. The melting (T_m), defined as the midpoint where 50% of the protein population is unfolded, quantifies thermal stability and is measured via techniques like . For many enzymes, T_m values range from 40–60°C, but engineered or thermophilic variants exceed 80°C, correlating with enhanced hydrogen bonding and hydrophobic packing. Once denatured, recovery of activity is rare without chaperones, underscoring the importance of physiological .

Biological Functions

Metabolic Pathways

Enzymes are integral to metabolic pathways, orchestrating the transformation of substrates in interconnected catabolic and anabolic networks that support cellular energy production and . In the glycolytic pathway, which converts glucose to pyruvate under conditions, ten distinct enzymes catalyze sequential reactions, with initiating the process by phosphorylating glucose to glucose-6-phosphate. This pathway exemplifies how enzymes enable efficient breakdown of carbohydrates, yielding ATP and NADH for cellular use. The tricarboxylic acid (TCA) cycle, a amphibolic pathway in aerobic , relies on eight enzymes to oxidize derived from or fatty acid breakdown, generating reducing equivalents for the . Citrate synthase, the first enzyme, catalyzes the condensation of with oxaloacetate to produce citrate, linking upstream catabolism to downstream energy yield. In autotrophic organisms, photosynthetic pathways such as the Calvin-Benson-Bassham cycle depend on enzymes like ribulose-1,5-bisphosphate carboxylase/oxygenase (), which fixes atmospheric CO2 into 3-phosphoglycerate, facilitating carbon assimilation and the synthesis of sugars. Metabolic flux through these pathways is tightly controlled by rate-limiting enzymes that dictate overall throughput based on substrate availability and energy demands. For example, in catalyzes the irreversible of fructose-6-phosphate to fructose-1,6-bisphosphate, serving as a primary regulatory point to prevent unnecessary glucose consumption when energy is abundant. Such control ensures balanced integration of catabolic and anabolic processes across cellular compartments. Spatial compartmentalization enhances pathway efficiency and specificity, with glycolytic enzymes localized in the to rapidly process cytoplasmic glucose, while TCA cycle enzymes reside in the , coupling oxidation to . This segregation prevents interference between pathways and optimizes metabolite gradients. Multi-enzyme complexes, known as metabolons, further streamline reactions; the , bridging and the cycle, assembles multiple subunits to channel pyruvate-derived directly to , minimizing diffusion losses and intermediate exposure. Regulation mechanisms briefly coordinate these pathways to synchronize flux with cellular needs.

Cellular Control Mechanisms

Cells maintain metabolic through precise control of enzyme availability and positioning, ensuring that catalytic activities align with cellular demands. One primary mechanism involves of enzyme , where environmental signals modulate the synthesis of specific enzymes. For instance, in , the exemplifies inducible expression: in the presence of , the protein dissociates from the operator region, allowing to transcribe genes encoding β-galactosidase, lactose permease, and thiogalactoside transacetylase, thereby enabling lactose metabolism only when glucose is scarce. Enzyme protein levels are further regulated post-transcriptionally by balancing synthesis and degradation rates. Protein synthesis rates are influenced by translational efficiency and mRNA stability, while degradation primarily occurs via the ubiquitin-proteasome pathway, where enzymes targeted for turnover are polyubiquitinated and degraded by the 26S proteasome complex. This selective degradation prevents accumulation of unnecessary or damaged enzymes, maintaining optimal concentrations. Additionally, molecular chaperones, such as and GroEL/GroES, assist in proper folding of newly synthesized enzymes, preventing aggregation and ensuring functional maturation; misfolded enzymes may be directed to degradation pathways if refolding fails. Subcellular localization restricts enzyme activity to specific compartments, enhancing efficiency and preventing off-target effects. Lysosomal hydrolases, like acid phosphatases and cathepsins, are trafficked to lysosomes via mannose-6-phosphate receptors in the Golgi apparatus, where they function optimally in the acidic lumen to degrade macromolecules. proteins further organize enzymes into multi-enzyme complexes or signaling hubs, localizing them to precise cellular locales; for example, A-kinase anchoring proteins (AKAPs) tether and phosphatases to maintain localized signaling, indirectly influencing enzyme regulation. Feedback loops provide dynamic control by integrating enzyme activity with pathway outputs, particularly through product inhibition. In metabolic pathways, end products often bind to upstream enzymes, allosterically inhibiting their activity to prevent overproduction; a classic case is the inhibition of aspartate transcarbamoylase by CTP in , which halts the pathway when levels are sufficient. These mechanisms collectively ensure that enzyme quantities and localizations adapt to maintain balanced across cellular contexts.

Pathological Roles

Enzyme deficiencies arising from genetic mutations can lead to severe pathological conditions by disrupting critical metabolic processes. (PKU), an autosomal recessive disorder, results from mutations in the PAH gene, causing a deficiency in (PAH), the enzyme responsible for converting to ; this leads to toxic accumulation of , resulting in , seizures, and behavioral issues if untreated. Lysosomal storage diseases, such as , exemplify another category of enzyme deficiencies, where mutations in the GBA1 gene impair activity, causing accumulation of in lysosomes; this manifests as , , , and in type 1 , with neurological involvement in types 2 and 3. In cancer, enzyme overactivity often drives uncontrolled and immortality. , a ribonucleoprotein enzyme that maintains length, is reactivated and overexpressed in approximately 90% of human cancers, enabling limitless replicative potential and tumor progression by preventing shortening-induced . Similarly, oncogenic kinases like BCR-ABL, a fusion resulting from the translocation, constitutively activate signaling pathways in chronic (CML), promoting leukemic cell survival, proliferation, and resistance to . Therapeutic strategies targeting pathological enzyme activities have revolutionized treatment for enzyme-related diseases. Enzyme inhibitors, such as , a selective , bind to the BCR-ABL kinase domain and block its ATP-binding site, inducing remission in over 90% of CML patients by halting aberrant signaling. For deficiencies, enzyme replacement therapy (ERT) delivers recombinant enzymes intravenously; in , imiglucerase (recombinant ) reduces substrate accumulation, alleviating visceral and skeletal symptoms. Advanced approaches include antibody-linked enzymes in antibody-directed enzyme prodrug therapy (ADEPT), where tumor-targeted antibodies conjugate enzymes like carboxypeptidase to activate non-toxic s selectively at cancer sites, minimizing systemic toxicity. Enzymes also serve as biomarkers for diagnosing pathological conditions through their abnormal levels in bodily fluids. Elevated kinase-MB (CK-MB), a cardiac-specific isoenzyme, in indicates , rising within 3-6 hours of injury due to cardiomyocyte and peaking at 16-30 hours, aiding rapid diagnosis when combined with troponins.

Evolution

Origins and Ancestry

The evolutionary origins of enzymes trace back to prebiotic chemistry on , where simple catalytic molecules likely facilitated the emergence of life before the dominance of protein-based enzymes. In the proposed hypothesis, molecules served as both genetic material and catalysts, known as , which performed essential reactions without protein assistance. These ribozymes are considered precursors to modern enzymes, enabling self-replication and basic metabolism in a pre-protein era. A prominent example is the center of the , which functions as a ribozyme to catalyze formation during protein synthesis, suggesting that RNA-based catalysis predated the protein world. Genomic and fossil evidence places the timeline of enzyme origins around 4 billion years ago, coinciding with the formation of the first cellular forms shortly after Earth's oceans stabilized. This period aligns with the appearance of the (LUCA), a hypothetical progenitor from which all extant descends, possessing a core set of enzymes essential for basic cellular functions. Notably, , which generates ATP via proton gradients across membranes, is conserved across bacterial and archaeal domains and is inferred to have been present in LUCA, indicating its ancient role in . Throughout evolutionary history, has significantly influenced enzyme distribution, allowing rapid dissemination of catalytic capabilities across microbial lineages. For instance, genes encoding enzymes that confer , such as beta-lactamases, have spread via plasmids and other mobile elements, accelerating in response to environmental pressures. This mechanism highlights how enzyme evolution extended beyond vertical inheritance, shaping microbial diversity from early prokaryotic communities. Modern enzyme structures often retain ancient protein folds, echoing these catalytic motifs.

Adaptive Diversification

Enzymes achieve adaptive diversification through evolutionary mechanisms that enable the emergence of new functions, primarily via , , , and laboratory-directed , allowing organisms to respond to changing environmental pressures and metabolic demands. Gene duplication events provide a key for this process by creating paralogous copies that initially retain the original function but face relaxed selective constraints, permitting mutations to accumulate without immediate fitness costs. This can lead to neofunctionalization, where one copy evolves a novel catalytic activity while the other maintains the ancestral role, thereby expanding the enzyme repertoire without disrupting existing pathways. In the alpha-amylase family, gene duplications have driven neofunctionalization, resulting in diverse starch-degrading enzymes adapted to specific ecological niches, such as the evolution of beta-amylases in angiosperms that exhibit sub- or neo-functionalization through extensive duplication events across eight distinct clades. These duplications allow for , for instance, in hydrolyzing different glycosidic bonds under varying conditions like or in and microbial lineages. Similarly, in ancestral enzymes—characterized by low substrate specificity—serves as an evolutionary starting point, enabling weak side activities to be refined into high-efficiency new functions under selective pressure, as seen in the transition from generalist hydrolases to specialized lipases or esterases. This promiscuous foundation facilitates innovation by providing latent catalytic potential that can be co-opted for novel metabolisms. Moonlighting enzymes further exemplify adaptive versatility, where a single protein performs multiple, often unrelated functions depending on cellular context, such as localization or binding partners, without requiring sequence divergence. Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), primarily involved in glycolysis, also acts in apoptosis by translocating to the nucleus to bind DNA and promote cell death signaling, a role that enhances cellular control mechanisms in response to stress. This multifunctionality arises evolutionarily from structural features that allow conformational flexibility, enabling the enzyme to switch roles without gene duplication. In laboratory settings, directed evolution mimics these natural processes by iteratively applying random mutagenesis and selection to engineer enzymes with desired properties, as demonstrated by Frances Arnold's development of cytochrome P450 variants capable of stereoselective cyclopropanation, achieving up to 99% enantioselectivity for non-natural reactions like carbene transfer to alkenes. These engineered P450s, derived from promiscuous ancestors, illustrate how targeted selection can rapidly diversify enzyme functions for biotechnological applications.

Applications

Industrial Processes

Enzymes play a pivotal role in by enabling efficient, sustainable manufacturing through biocatalysis, which often reduces energy consumption and minimizes environmental impact compared to traditional chemical methods. In sectors like food production, biofuels, and pharmaceuticals, enzymes facilitate large-scale transformations of raw materials into valuable products, enhancing yield and product quality while allowing for milder reaction conditions. In the , amylases are widely employed to hydrolyze into simpler sugars, improving processing efficiency. In , α-amylases break down in to produce dextrins, resulting in better volume, texture, and by acting as antistaling agents. Similarly, in , these enzymes convert barley into fermentable sugars during , boosting yield and enabling the of lighter beers with reduced calories through glucoamylase action on dextrins. , primarily from microbial sources, is essential for cheese , where it hydrolyzes κ-casein in to promote and formation, accelerating and enhancing development. Lipases find application in formulations to degrade lipid-based stains, improving performance in processes, though they also contribute to enhancement in dairy products like cheese by liberating free fatty acids. For biofuel production, are critical in breaking down into fermentable sugars for synthesis. These enzyme cocktails, including endoglucanases, exoglucanases, and β-glucosidases, hydrolyze pretreated plant materials like , with companies such as (formed by the 2024 merger of and ) developing optimized formulations that have contributed to reducing production costs, with minimum fuel selling prices averaging $2.65 per (range $0.90–$6.00/) as of recent analyses through improved yields and stability. In pharmaceutical manufacturing, penicillin acylase catalyzes the hydrolysis of penicillin G to produce 6-aminopenicillanic acid (6-APA), a key intermediate for semisynthetic β-lactam antibiotics like amoxicillin. This enzymatic process offers high specificity and efficiency, replacing harsher chemical methods and enabling scalable synthesis of antibiotics with minimal byproducts. To enhance economic viability, enzyme immobilization techniques are routinely applied in these industries, allowing repeated use and recovery. Adsorption involves reversible binding of enzymes to solid supports like ion-exchange resins, offering simplicity and low cost but risking desorption under operational stresses. Entrapment, by contrast, confines enzymes within polymer matrices such as alginate beads or gels, providing robust protection and high loading but potentially introducing diffusion limitations that reduce reaction rates. Key challenges include maintaining enzymatic stability against thermal denaturation, pH shifts, and mechanical shear during prolonged reuse, which can limit operational cycles and overall productivity. Engineered enzymes, optimized for such immobilization, further improve process robustness in industrial settings.

Biomedical Uses

Enzymes play a pivotal role in biomedical diagnostics through techniques that leverage their catalytic properties for sensitive detection of biomolecules. In , (HRP) is commonly conjugated to antibodies to amplify signals via chromogenic or fluorescent substrates, enabling the quantification of antigens or antibodies at picomolar levels in clinical samples such as or . This method has become a cornerstone for diagnosing infectious diseases, autoimmune disorders, and cancers, with HRP's high turnover rate allowing for rapid readout in under an hour. In therapeutics, enzymes are administered to replace deficient activities or target pathological processes. For , alginate lyase has been investigated as an enzyme replacement therapy to degrade alginate biofilms produced by in lung mucus, potentially improving and reducing severity in preclinical models. Similarly, thrombolytic enzymes like activate plasminogen to , dissolving clots in acute or , achieving reperfusion in approximately 65% of cases when administered promptly. These applications draw from pathological roles where enzyme dysregulation contributes to disease, informing targeted interventions. Emerging biotechnologies harness engineered enzymes for precise genetic and metabolic modifications. The CRISPR-Cas9 system utilizes the Cas9 nuclease enzyme, developed in 2012, to create targeted double-strand breaks in DNA guided by RNA, facilitating gene editing for treating genetic disorders like sickle cell disease and certain cancers in clinical trials. Directed evolution techniques iteratively mutate and select cytochrome P450 enzymes—key drug-metabolizing proteins—to enhance their specificity and stability, enabling personalized medicine applications such as detoxifying xenobiotics or optimizing prodrug activation in vivo. Despite these advances, enzyme therapeutics face challenges including and delivery barriers. Foreign enzymes can elicit responses, reducing efficacy and causing , as observed in 1.6% to 4.4% of patients receiving from major clinical trials. To mitigate this, —covalent attachment of —extends plasma from minutes to days and masks immunogenic epitopes, as demonstrated in approved therapies like pegademase for . In the 2020s, progress with mRNA-encoded enzymes, such as transient expression of via lipid nanoparticles, circumvents immunogenicity by leveraging the patient's cellular machinery, showing promise in phase I trials for liver-directed editing without persistent protein exposure, with updated data confirming safety as of 2025.

References

  1. [1]
    Enzyme - National Human Genome Research Institute
    An enzyme is a biological catalyst that is usually a protein but could be RNA. The point of a catalyst is to increase the speed with which a reaction happens.
  2. [2]
    RNA catalysis—is that it? - PMC - PubMed Central - NIH
    Ribozymes are RNA molecules that accelerate chemical reactions, enzymes that happen to be made of RNA rather than protein.
  3. [3]
    Biochemistry, Proteins Enzymes - StatPearls - NCBI Bookshelf - NIH
    Enzymes are proteins that act upon substrate molecules and decrease the activation energy necessary for a chemical reaction to occur by stabilizing the ...
  4. [4]
    Enzymes | Idaho State University
    Enzymes are biological catalysts, compounds that speed up chemical reactions without being used up or altered in the reaction.
  5. [5]
    Enzyme nomenclature and classification: the state of the art - PubMed
    There were originally six recognized groups of enzymes: Oxidoreductases (EC 1), Transferases (EC 2), Hydrolases (EC 3), Lyases (EC 4), Isomerases (EC 5) and ...
  6. [6]
    [PDF] Enzyme Nomenclature
    Enzymes are classified into six different groups according to the reaction being catalyzed. The nomenclature was determined by the Enzyme Commission in 1961 ...
  7. [7]
    Looking Back: A Short History of the Discovery of Enzymes and How ...
    In this review, the origin of modern biocatalysis is summarized (Figure 1), starting with the discovery of enzymes, and exploring the principles of ...
  8. [8]
    A Broader View: Microbial Enzymes and Their Relevance in ... - NIH
    In 1833, French chemist Anselme Payen discovered the first enzyme, diastase [4]. In 1835, the hydrolysis of starch by diastase was acknowledged as a catalytic ...
  9. [9]
    Enzymes: principles and biotechnological applications - PMC
    Enzymes are biological catalysts (also known as biocatalysts) that speed up biochemical reactions in living organisms. They can also be extracted from cells and ...
  10. [10]
    The history of biocatalysis - Nobel Prize
    1877, Wilhelm Kuhne, Germany introduces the term “enzyme” and distinguishes between enzymes and bacteria. ; 1893, Wilhelm Ostwald, Latvia classifies enzymes as ...Missing: etymology | Show results with:etymology
  11. [11]
    History of Enzyme Chemistry - Cornish‐Bowden - Wiley Online Library
    May 16, 2011 · Justus von Liebig, for example, did not accept that yeast was a living organism, ridiculing those who did, and arguing that fermentation ...
  12. [12]
    A Brief History of Fermentation, East and West - SoyInfo Center
    ... Justus von Liebig, J.J. Berzelius, and Friedrich Woehler. This view seemed ... From that time on the term "enzyme" came to be applied to all ferments.
  13. [13]
    Eduard Buchner – Facts - NobelPrize.org
    In 1897 Eduard Buchner discovered that yeast extract with no living yeast fungi can form alcohol from a sugar solution. The conclusion was that biochemical ...Missing: organized soluble
  14. [14]
    Fermented Cereal-based Products: Nutritional Aspects, Possible ...
    The etymology of the term fermentation comes from the Latin verb “fevere”, which means “to boil”. Fermentation is one of the oldest known food processing ...
  15. [15]
    Enzyme Power | Current History | University of California Press
    Nov 1, 2023 · It was gradually regularized by the coining of the term enzyme in the 1860s and the decisive demonstration in 1897 that enzymes, though the ...
  16. [16]
  17. [17]
  18. [18]
    Translation of the 1913 Michaelis-Menten Paper - PMC - NIH
    Michaelis and Menten tested the postulate that the rate of an enzyme-catalyzed reaction could be described by a constant term (c) multiplied by the ...
  19. [19]
    Press release: The 1989 Nobel Prize in Chemistry - NobelPrize.org
    This year's Nobel Prize in chemistry has been awarded to Sidney Altman, USA and Thomas Cech, USA for their discovery that RNA (ribonucleic acid) in living ...
  20. [20]
    Frances H. Arnold – Facts – 2018 - NobelPrize.org
    In 1993, Arnold conducted the first directed evolution of enzymes, which are proteins that catalyze chemical reactions. The uses of her results include more ...
  21. [21]
    [PDF] Current IUBMB recommendations on enzyme nomenclature and ...
    Mar 27, 2014 · To overcome the risk of imminent chaos, the IUB set up the Enzyme Commission in 19566 which presented its Report in 1961 (IUB, 1961), in ...
  22. [22]
    [PDF] A Brief Guide to Enzyme Nomenclature and Classification - IUBMB
    Enzymes are classified by EC numbers, based on the substrates transformed and products formed, and divided into 6 major classes and 7 total.
  23. [23]
    1.1.1.1 alcohol dehydrogenase - Expasy - ENZYME
    ENZYME entry: EC 1.1.1.1 · Acts on primary or secondary alcohols or hemi-acetals with very broad specificity; however the enzyme oxidizes methanol much more ...
  24. [24]
    About ExplorEnz — The Enzyme Database
    There are currently over 8300 entries, and proposed additions and amendments to the Enzyme Nomenclature list are made available on this site for public review ...
  25. [25]
    Enzyme Classification - IUBMB Nomenclature
    To produce usable systematic names, accepted trivial names of substrates forming part of the enzyme names should be used. Where no accepted and convenient ...
  26. [26]
    Enzyme Nomenclature
    Oct 20, 2025 · This page contains general information on enzyme nomenclature. It includes links to individual documents, and the number of these will increase as more ...Missing: establishment 1956
  27. [27]
    ECOD: An Evolutionary Classification of Protein Domains - PMC - NIH
    Dec 4, 2014 · Here we introduce the ECOD (Evolutionary Classification Of protein Domains) database. Our goal is threefold: (1) to construct a comprehensive ...
  28. [28]
    Enzyme nomenclature and classification: the state of the art
    Nov 12, 2021 · Systematic names can be quite long and complicated. An example is the systematic name for the enzyme riboflavin synthase (EC 2.5.1.9), 6,7 ...
  29. [29]
    The Anfinsen Dogma: Intriguing Details Sixty-Five Years Later - NIH
    Jul 14, 2022 · The pioneering experiments of Anfinsen on the oxidative folding of RNase have been revisited discovering some details, which update the statement of his dogma.
  30. [30]
    The Shape and Structure of Proteins - Molecular Biology of the Cell
    The full three-dimensional organization of a polypeptide chain is sometimes referred to as the protein's tertiary structure, and if a particular protein ...
  31. [31]
    Biochemistry, Tertiary Protein Structure - StatPearls - NCBI Bookshelf
    This structure encompasses four levels—primary, secondary, tertiary, and quaternary, each contributing to the overall shape and function of the protein. Primary ...
  32. [32]
    The TIM-barrel fold: a versatile framework for efficient enzymes
    Mar 16, 2001 · Recent studies on triosephosphate isomerase (TIM)-barrel enzymes highlight the remarkable versatility of the TIM-barrel scaffold.
  33. [33]
    An Ancient Fingerprint Indicates the Common Ancestry of Rossmann ...
    Mar 3, 2016 · The Rossmann fold is one of the most ancient and functionally diverse protein folds, and most Rossmann enzymes utilize nucleoside-based ...
  34. [34]
    The history of the CATH structural classification of protein domains
    This article presents a historical review of the protein structure classification database CATH. Together with the SCOP database, CATH remains comprehensive ...
  35. [35]
    SCOP: a Structural Classification of Proteins database - PMC - NIH
    The Structural Classification of Proteins (SCOP) database provides a detailed and comprehensive description of the relationships of known protein structures.
  36. [36]
    Molecular Docking: From Lock and Key to Combination Lock - NIH
    In 1894, Emil Fischer suggested that the specificity of an enzyme ... lock and key' model. Koshland's suggestion was that active sites of enzymes ...
  37. [37]
    Application of a Theory of Enzyme Specificity to Protein Synthesis ...
    derived from the induced-fit theory of specificity. For example, Figure 2 shows a possible sequence of reactions in the formation of a peptide bondin the ...
  38. [38]
    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.
  39. [39]
    Key difference between transition state stabilization and ground ...
    2.2. Enzyme–substrate noncovalent interactions that modulate ΔG‡s. For an enzyme to catalyze the reaction of a substrate, the substrate enters the enzyme ...<|separator|>
  40. [40]
    Role of non-covalent enzyme-substrate interactions in the reaction ...
    The data show that hydrogen-bonding interactions between the 1-OH and 4-OH and charged groups on the enzyme stabilize the ground-state ternary complex of the ...
  41. [41]
    Titration ELISA as a Method to Determine the Dissociation Constant ...
    Feb 15, 2018 · The dissociation constant describes the interaction between two partners in the binding equilibrium and is a measure of their affinity.
  42. [42]
    Simple methods to determine the dissociation constant, Kd - PMC
    Sep 16, 2024 · Kd is the dissociation constant, indicating the concentration of enzyme at which 50% of the DNA is bound.
  43. [43]
    The Central Role of Enzymes as Biological Catalysts - The Cell - NCBI
    Moreover, the transition state is stabilized by its tight binding to the enzyme, thereby lowering the required energy of activation.
  44. [44]
    Chapter 7: Catalytic Mechanisms of Enzymes - Chemistry
    Electrostatic catalysis occurs when the enzyme active site stabilizes the transition state of the reaction by forming electrostatic interactions with the ...
  45. [45]
    Design of activated serine-containing catalytic triads with atomic ...
    Serine hydrolases utilize a conserved nucleophilic serine to hydrolyze ester, amide or thio-ester bonds in proteins and small molecules and constitute one of ...
  46. [46]
    Another Look at the Mechanisms of Hydride Transfer Enzymes ... - NIH
    YADH catalyzes the oxidation of alcohols to aldehydes using the coenzyme nicotinamide adenine dinucleotide (NAD).
  47. [47]
    How an enzyme surmounts the activation energy barrier - PNAS
    Enzymes are powerful catalysts, some accelerating their target reactions by factors that can exceed 1020 (4). Chorismate mutases accelerate their target ...
  48. [48]
    Deep learning-based kcat prediction enables improved enzyme ...
    Jun 16, 2022 · The enzyme turnover number (kcat), which defines the maximum chemical conversion rate of a reaction, is a critical parameter for understanding ...
  49. [49]
    Molecular mechanism of glycolytic flux control intrinsic to ... - PNAS
    Dec 10, 2021 · The phosphotransfer reaction proceeds by adopting a fully closed conformation through the hinge-bending motion demonstrated by small-angle X-ray ...
  50. [50]
    Yeast hexokinase in solution exhibits a large conformational change ...
    Yeast hexokinase in solution exhibits a large conformational change upon binding glucose or glucose 6-phosphate. Click to copy article linkArticle link copied!
  51. [51]
    Illuminating the mechanistic roles of enzyme conformational dynamics
    Many enzymes mold their structures to enclose substrates in their active sites such that conformational remodeling may be required during each catalytic ...
  52. [52]
    On the Roles of Substrate Binding and Hinge Unfolding in ... - NIH
    We characterized the conformational change of adenylate kinase (AK) between open and closed forms by conducting five all-atom molecular-dynamics simulations.Missing: hexokinase | Show results with:hexokinase
  53. [53]
    Probing conformational dynamics of an enzymatic active site ... - PNAS
    Dec 8, 2016 · Here we report our study on the conformational dynamics of horseradish peroxidase using single-molecule multiparameter photon time-stamping spectroscopy with ...Sign Up For Pnas Alerts · Results · Enzymatic Reaction Essay
  54. [54]
    Resolving the complex role of enzyme conformational dynamics in ...
    We show that Kramers' rate theory can be used to present a clear rationale of how ECD affects the reaction dynamics and catalytic rates.
  55. [55]
    Protein conformational dynamics in the mechanism of HIV-1 ... - PNAS
    Dec 8, 2011 · We show that conformational isomerization in the flaps is correlated with structural reorganization of residues in the active site.
  56. [56]
    Role of Conformational Fluctuations in the Enzymatic Reaction of ...
    The emergence of compensatory drug-resistant mutations in HIV-1 protease challenges the common view of the reaction mechanism of this enzyme.
  57. [57]
    The Fe and Zn cofactor dilemma - PMC - PubMed Central
    1. Introduction. Metal ions are critical and ubiquitous inorganic cofactors in biological systems, occurring in more than one-third of known proteins.
  58. [58]
    Elucidating the role of metal ions in carbonic anhydrase catalysis
    Sep 11, 2020 · The catalytic zinc ion in CA II serves as a Lewis acid; its primary role is to lower the pKa of the Zn-bound water from 10 to 7, allowing the ...
  59. [59]
    Magnesium induced structural reorganization in the active site ... - NIH
    Aug 9, 2024 · Magnesium optimizes the angle defining the nucleophilic attack of substrates in phosphoryl transfer catalyzed by adenylate kinase.
  60. [60]
    Redox properties of cytochrome c - PubMed
    The redox properties of cytochromes (cyt) c, a ubiquitous class of heme-containing electron transport proteins, have been extensively investigated over the ...
  61. [61]
    Wilson Disease - StatPearls - NCBI Bookshelf - NIH
    Apr 3, 2025 · In Wilson disease, dysfunctional ATP7B leads to copper accumulation in the liver, and ceruloplasmin is secreted in a form that lacks copper and ...
  62. [62]
    [PDF] Chapter 23: Coenzymes
    • many enzymes are inactive without cofactors. • cofactors are 1) metal ions or 2) coenzymes. • coenzymes - organic molecules, derived from vitamins ... examples ...
  63. [63]
    [PDF] Mechanisms of actions of coenzymes
    A portion of ATP molecule which is transferred to the vitamin is the group which binds the coenzyme for the enzyme active centers. Vitamins soluble in water are ...
  64. [64]
    Enzymology below 200 K: The kinetics and thermodynamics ... - PNAS
    ... activation energy was calculated by using the Arrhenius equation: ln k = (−Eact/RT) + c, [2]. where k is the rate constant, Eact is the activation energy, R ...
  65. [65]
    Molecular Architecture and Biological Reactions - ACS Publications
    Answers to many basic problems of biology—nature of growth, mechanism of duplication of viruses and genes, action of enzymes, mechanism, of physiological ...
  66. [66]
    Electrostatic transition state stabilization rather than reactant ... - PNAS
    For more than half a century, transition state theory has provided a useful framework for understanding the origins of enzyme catalysis.
  67. [67]
  68. [68]
    Investigating the role of a backbone to substrate hydrogen bond in ...
    Oct 1, 2014 · Orotidine 5′-monophosphate decarboxylase has attracted intense enzymological interest, because it achieves a very large rate enhancement ...
  69. [69]
    Gibbs Free Energy of Activation - an overview | ScienceDirect Topics
    2 Chemical and Biochemical Catalysis: The Enzymes. Catalysts accelerate the reactions by decreasing the Gibbs energy of activation without affecting equilibrium ...
  70. [70]
    A Note on the Kinetics of Enzyme Action - PubMed
    A Note on the Kinetics of Enzyme Action. Biochem J. 1925;19(2):338-9. doi: 10.1042/bj0190338. Authors. G E Briggs , J B Haldane. Affiliation.Missing: equilibrium | Show results with:equilibrium
  71. [71]
    (PDF) Review of the 1925 Briggs-Haldane Enzyme Kinetics Paper ...
    Jul 28, 2022 · This preprint is a brief note reviewing the contents of a 1925 paper by Briggs and Haldane that introduced the concept of steady-state conditions in enzyme ...
  72. [72]
    Basics of Metabolic Reactions - PMC - NIH
    Endergonic reactions can also be coupled to exergonic reactions involving NADH, FADH2, and NADPH conversion into NAD+, FAD, and NADP+, respectively.
  73. [73]
    The kinetics of enzyme-catalyzed reactions with two or more ...
    A nomenclature is proposed to facilitate discussion of possible mechanisms for enzyme-catalyzed reactions with more than one substrate or product.
  74. [74]
    A Note on the Kinetics of Enzyme Action - PMC - NIH
    A Note on the Kinetics of Enzyme Action · View on publisher site · PDF (147.8 KB) · Cite · Collections · Permalink. PERMALINK. Copy ...
  75. [75]
    The Determination of Enzyme Dissociation Constants
    Hans Lineweaver · Dean Burk. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1934, 56, 3, 658–666. Click to ...
  76. [76]
    A guide to the Michaelis–Menten equation: steady state and beyond
    Jul 16, 2021 · Subsequently, G.E. Briggs and J.B.S. Haldane derived the equation with steady-state assumptions and this is the version that we popularly ...
  77. [77]
    ON THE NATURE OF ALLOSTERIC TRANSITIONS: A PLAUSIBLE ...
    ON THE NATURE OF ALLOSTERIC TRANSITIONS: A PLAUSIBLE MODEL. J Mol Biol. 1965 May:12:88-118. doi: 10.1016/s0022-2836(65)80285-6.
  78. [78]
    Mechanism of Action Assays for Enzymes - NCBI
    May 1, 2012 · Types of Inhibition. There are three main types of inhibition (competitive, noncompetitive, and uncompetitive) that are most commonly used to ...
  79. [79]
    Discovery of a potent HMG-CoA reductase degrader that eliminates ...
    Dec 3, 2018 · Statins act by competitively binding to the catalytic domain of 3-hydroxy-3-methylglutaryl coenzyme A reductase (HMG-CoA reductase, HMGCR) and ...
  80. [80]
    Non-Competitive Inhibition - an overview | ScienceDirect Topics
    The inhibitory effects of heavy metals, and of cyanide on cytochrome oxidase and of arsenate on glyceraldehyde phosphate dehydrogenase, are examples of non- ...
  81. [81]
    an example of uncompetitive inhibition and its pharmacological ...
    The first indications that lithium may inhibit inositol monophosphatase in oivo came from AIlison, Sherman and col- leagues (see Ref. 13 for a review). Plasma ...
  82. [82]
    Assessment of Enzyme Inhibition: A Review with Examples from the ...
    This review describes key principles and approaches for the reliable determination of enzyme activities and inhibition as well as some of the methods that are ...
  83. [83]
    Strategies for discovering and derisking covalent, irreversible ...
    This article presents several covalent inhibitors, including examples of successful drugs, as well as highly selective, irreversible inhibitors of emerging ...
  84. [84]
    Demystifying Functional Parameters for Irreversible Enzyme Inhibitors
    Irreversible enzyme inhibitors typically operate through a two-step mechanism that involves initial reversible binding of the inhibitor to form E·I, governed k ...
  85. [85]
    Penicillin - StatPearls - NCBI Bookshelf - NIH
    The catalyst for this reaction is penicillin-binding proteins, such as the enzyme DD-transpeptidase. Penicillin's 4-membered β-lactam ring can bind to DD ...
  86. [86]
    The mechanism of action of penicillin. Penicillin acylates ... - PubMed
    May 10, 1980 · Penicillin kills susceptible bacteria by specifically inhibiting the transpeptidase that catalyzes the final step in cell wall biosynthesis, the cross-linking ...
  87. [87]
    Rapid Determination of the Specificity Constant of Irreversible ...
    Nov 16, 2015 · A simple approach to measuring kinact/KI was developed that makes use of an irreversible probe for competitive assays run to completion against test compounds.
  88. [88]
    [PDF] Enzyme Catalysis: inhibition
    Sep 10, 2015 · Example of an irreversible inhibitor: Page 12. Inhibitor Types: Mechanism-based irreversible inhibitors. 10 September 2015. PHRM 836 - Devlin ...
  89. [89]
    Residual cyclooxygenase activity of aspirin-acetylated COX-2 forms ...
    Aspirin inhibits the cyclooxygenase (COX) enzymes via a unique mechanism (1). Aspirin covalently modifies the enzyme by transfer of its acetyl group to Ser530 ...
  90. [90]
    Organophosphate Toxicity - StatPearls - NCBI Bookshelf - NIH
    Nov 12, 2023 · These insecticides inactivate AChE by phosphorylating the serine hydroxyl group on the enzyme. As AChE is essential in acetylcholine degradation ...
  91. [91]
    An outlook on suicide enzyme inhibition and drug design - PMC
    Oct 9, 2021 · Mechanism of suicide enzyme inhibition​​ The precise mechanism of inactivation process will differ from one enzyme to another, but the catalytic ...
  92. [92]
    Mechanisms of action of FdUMP[10]: metabolite activation and ...
    FdUMP[10] is a multimer of FdUMP, a suicide inhibitor of thymidylate synthase (TS), and was designed to bypass resistance to 5-fluorouracil (5FU).
  93. [93]
    From feedback inhibition to allostery: the enduring ... - FEBS Press
    Aug 16, 2013 · ATCase is the first enzyme of the pyrimidine biosynthetic pathway, and is inhibited by CTP and UTP, the nucleotide end-products of the pathway.<|separator|>
  94. [94]
    None
    ### Summary of the Monod-Wyman-Changeux Model for Allosteric Transitions
  95. [95]
    Comparison of Experimental Binding Data and Theoretical Models ...
    Comparison of Experimental Binding Data and Theoretical Models in Proteins Containing Subunits*. Click to copy article linkArticle link copied! D. E. Koshland, ...
  96. [96]
    Cooperative Binding - PMC - NIH
    The Hill Equation​​ This means that cooperativity is assumed to be fixed, i.e., it does not change with saturation. It also means that binding sites always ...
  97. [97]
    Protein posttranslational modifications in health and diseases
    Phosphorylation is a reversible PTM, and the reverse process of phosphorylation is called dephosphorylation catalyzed by phosphatases (Figure 2A).
  98. [98]
    Glycogen Synthase - an overview | ScienceDirect Topics
    Glycogen synthase exists in at least two forms: a phosphorylated form, arising from covalent modification of serine residues by ATP; and a dephosphorylated ...
  99. [99]
    Quantification of the glycogen cascade system: the ultrasensitive ...
    Regulation of glycogen synthase and phosphorylase by reversible covalent modification cycles exemplifies signal transduction by enzyme cascades. Although ...
  100. [100]
    Mechanistic insights into the regulation of metabolic enzymes by ...
    Examples are emerging where acetylation plays a similar role in directly regulating the amount of metabolic enzymes through targeting the substrate to ...
  101. [101]
    Biochemistry, Ubiquitination - StatPearls - NCBI Bookshelf - NIH
    Ubiquitination is a 3-step process involving 3 enzymes: ubiquitin-activating enzyme (E1), ubiquitin-conjugating enzyme (E2), and ubiquitin-protein ligase (E3).
  102. [102]
    Zymogen Activation Confers Thermodynamic Stability on a Key ...
    Trypsinogen is normally activated in the duodenum upon cleavage by enteropeptidase, but under pathological conditions can be susceptible to autoactivation in ...
  103. [103]
    Regulation of insulin secretion by the post-translational modifications
    Aug 4, 2023 · Phosphorylation mediated signaling cascades play an important role in maintaining glucose homeostasis (Roder et al., 2016). The insulin ...
  104. [104]
    Physiology, Pepsin - StatPearls - NCBI Bookshelf - NIH
    May 1, 2023 · A low pH (1.5 to 2) activates pepsin. Acetylcholine, gastrin, and histamine stimulate the proton pump in parietal cells to release hydrogen ions ...Introduction · Issues of Concern · Function · Pathophysiology
  105. [105]
    Effect of pH, temperature and media on acid and alkaline ... - NIH
    The pH optimum for alkaline phosphatase activity was 9.0 for all of the isolates tested. The pH optimum for acid phosphatase activity was 5.8 for 67% of the ...
  106. [106]
    The combined effects of reactant kinetics and enzyme stability ... - NIH
    Apr 23, 2017 · The Arrhenius equation was originally applied to describe the temperature dependence of chemical reaction rates in controlled settings, but it ...
  107. [107]
    Deoxyribonucleic acid polymerase from the extreme thermophile ...
    A stable deoxyribonucleic acid (DNA) polymerase (EC 2.7.7.7) with a temperature optimum of 80 degrees C has been purified from the extreme thermophile Thermus
  108. [108]
    The Effect of Ionic Strength and Specific Anions on Substrate ... - NIH
    With all salts there is an increase in the activity at low salt concentration (due to the activation by Na+) followed by a decline in rate of hydrolysis with ...
  109. [109]
    Interfacial Catalysis: The Mechanism of Phospholipase A2 - PMC - NIH
    Calcium ion is essential both for the binding of substrate and for catalysis. Figure 4 shows the coordination geometry of the calcium-ion cofactor in the two ...<|separator|>
  110. [110]
    Thermal unfolding methods in drug discovery - PMC - PubMed Central
    May 16, 2023 · The melting temperature (Tm) (dotted line) can be defined as the midpoint of the transition, or the temperature at which 50% of the protein ...
  111. [111]
    Thermally denatured state determines refolding in lipase - NIH
    Upon denaturation of the parent at 75°C, possibly the protein unfolds into an intermediate that was prone to aggregation. At 95°C accumulation of the residual ...
  112. [112]
    Glycolysis: A multifaceted metabolic pathway and signaling hub - PMC
    The pathway consists of 10 metabolic enzymes which work ... Enzymes of glycolysis are functionally associated with the mitochondrion in Arabidopsis cells.
  113. [113]
    Biochemistry, Glycolysis - StatPearls - NCBI Bookshelf - NIH
    As described before, many enzymes are involved in the glycolytic pathway by converting one intermediate to another. Control of these enzymes, such as ...
  114. [114]
    Physiology, Krebs Cycle - StatPearls - NCBI Bookshelf
    Nov 23, 2022 · Krebs Cycle). It comprises 8 enzymes within the mitochondrial ... number of enzymes produced in fasting and fed conditions. In the fed ...
  115. [115]
    Evolution and Functional Implications of the Tricarboxylic Acid Cycle ...
    The tricarboxylic acid (TCA) cycle, a crucial component of respiratory metabolism, is composed of a set of eight enzymes present in the mitochondrial matrix.
  116. [116]
    A faster Rubisco with potential to increase photosynthesis in crops
    Sep 17, 2014 · In photosynthetic organisms, d-ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) is the major enzyme assimilating atmospheric CO2 ...Missing: enzyme | Show results with:enzyme
  117. [117]
    Structural basis for allosteric regulation of human ... - PubMed - NIH
    Aug 25, 2024 · Phosphofructokinase-1 (PFK1) catalyzes the rate-limiting step of glycolysis, committing glucose to conversion into cellular energy.
  118. [118]
    Glycolysis Rate-Limiting Enzymes: Novel Potential Regulators of ...
    Nov 24, 2021 · Both key glycolysis rate-limiting enzymes, hexokinase-II (HK-II), phosphofructokinase-1 (PFK-1), and pyruvate kinase M2 (PKM2), as well as indirect rate- ...
  119. [119]
    Compartmentalization and metabolic regulation of glycolysis - PMC
    Glycolysis enzymes are typically diffuse, but under hypoxic stress, they form dynamic structures like G bodies, which are condensates, and may act as ...
  120. [120]
    Tracing Compartmentalized NADPH Metabolism in the Cytosol and ...
    Apr 22, 2014 · Here we develop an approach to resolve NADP(H)-dependent pathways present within both the cytosol and the mitochondria. By tracing hydrogen in ...
  121. [121]
    Protein-protein interactions and metabolite channelling in the plant ...
    May 16, 2017 · Protein complexes of sequential metabolic enzymes, often termed metabolons, may permit direct channelling of metabolites between the enzymes ...
  122. [122]
    The Roles of Molecular Chaperones in Regulating Cell Metabolism
    While chaperones have been shown to directly impact the folding of a given client, they can also serve as mediators to regulate metabolic enzyme degradation ...
  123. [123]
    Signalling scaffolds and local organization of cellular behaviour - PMC
    This Review examines the features of scaffold proteins and highlights examples of locally organized groups of signalling enzymes that drive essential ...
  124. [124]
    Phenylalanine Hydroxylase Deficiency - GeneReviews - NCBI - NIH
    Mar 13, 2025 · Maternal phenylketonuria (MPKU) syndrome occurs in offspring of mothers with inadequately treated PAH deficiency during pregnancy and results ...Summary · Clinical Characteristics · Management · Genetic Counseling
  125. [125]
    Gaucher Disease - StatPearls - NCBI Bookshelf
    Nov 12, 2023 · Gaucher disease results from mutations in the GBA1 gene, leading to deficient glucocerebrosidase activity within lysosomes.Continuing Education Activity · Etiology · Evaluation · Treatment / Management
  126. [126]
    Telomerase in Cancer: Function, Regulation, and Clinical Translation
    Feb 5, 2022 · To do so, malignant cells reactivate telomerase to extend their telomeres and achieve cellular immortality, which is a “Hallmark of Cancer”.
  127. [127]
    Molecular biology of bcr-abl1–positive chronic myeloid leukemia
    Decades of research zeroing in on the role of BCR-ABL1 kinase in the pathogenesis of CML have culminated in the development of highly efficacious therapeutics ...
  128. [128]
    Imatinib in Chronic Myeloid Leukemia: an Overview - PMC - NIH
    Imatinib directly inhibits the constitutive tyrosine kinase activity. Imatinib binds to BCR-ABL kinase domain by preventing the transfer of a phosphate group to ...
  129. [129]
    Enzyme Replacement in Gaucher Disease - PMC - PubMed Central
    What Does the Future Hold? The quality of life for patients with Gaucher disease has been greatly improved by the development of enzyme replacement therapy.
  130. [130]
    Antibody-directed enzyme prodrug therapy (ADEPT) for cancer
    The principle of antibody-directed enzyme prodrug therapy is to use an antibody directed at a tumor-associated antigen to vector an enzyme to tumor sites. The ...
  131. [131]
    Myocardial Infarction Serum Markers - StatPearls - NCBI Bookshelf
    CK-MB is released from skeletal muscle, and this can lead to falsely positive elevation. Per gram of myocardial tissue, more troponin is present than CK-MB.Introduction · Diagnostic Tests · Testing Procedures
  132. [132]
    The lost language of the RNA World | Science Signaling
    Jun 13, 2017 · The possibility of an RNA World is based on the notion that life on Earth passed through a primitive phase without proteins.
  133. [133]
    The RNA World as a Model System to Study the Origin of Life
    Oct 5, 2015 · We present a current view of the biochemistry of the origin of life, focusing on issues surrounding the emergence of an RNA World.<|separator|>
  134. [134]
    RNA, the first macromolecular catalyst: the ribosome is a ribozyme
    The larger ribosomal subunit contains the site of peptide-bond formation – the peptidyl transferase centre. The ribosome has three. tRNA-binding sites: the A- ...
  135. [135]
    The nature of the last universal common ancestor and its impact on ...
    Jul 12, 2024 · The last universal common ancestor (LUCA) is the node on the tree of life from which the fundamental prokaryotic domains (Archaea and Bacteria) diverge.
  136. [136]
    The very early evolution of biological complexity: Trends in Genetics
    Sep 25, 2024 · The dating of the LUCA to around 4.2 billion years ago agrees with this broader timeline. However, this dating also gives only 100–200 million ...
  137. [137]
    ATP synthase evolution on a cross-braced dated tree of life - Nature
    Nov 17, 2023 · The ATP synthase complex is thought to have originated prior to the Last Universal Common Ancestor (LUCA) and analyses of ATP synthase genes, ...
  138. [138]
    Functions predict horizontal gene transfer and the emergence of ...
    Oct 22, 2021 · Interphylum HGT events are of particular importance as they likely contribute to the recent emergence of antibiotic resistance in pathogenic ...
  139. [139]
    Horizontal antimicrobial resistance transfer drives epidemics of ...
    Apr 13, 2018 · Horizontal gene transfer has played a role in developing the global public health crisis of antimicrobial resistance (AMR).Results · Co-Circulating Msma... · Methods<|separator|>
  140. [140]
    Enzyme Promiscuity: A Mechanistic and Evolutionary Perspective
    We discuss the mechanistic enzymology of promiscuity--how enzymes, which generally exert exquisite specificity, catalyze other, and sometimes barely related, ...
  141. [141]
    The evolution of functional complexity within the β-amylase gene ...
    Feb 28, 2019 · We identified eight distinct clades in angiosperms, which results from extensive gene duplications and sub- or neo-functionalization. We ...
  142. [142]
    Enzyme promiscuity: engine of evolutionary innovation - PubMed - NIH
    Oct 31, 2014 · Catalytic promiscuity and substrate ambiguity are keys to evolvability, which in turn is pivotal to the successful acquisition of novel biological functions.
  143. [143]
    Nuclear moonlighting of cytosolic glyceraldehyde-3-phosphate ...
    Jul 10, 2020 · Enzyme moonlighting is a phenomenon whereby an enzyme can perform secondary non-enzymatic functions acquired through evolution, such as ...
  144. [144]
    Directed Evolution of a Cytochrome P450 Carbene Transferase for ...
    Jun 5, 2019 · We addressed this challenge by evolving a cytochrome P450 for highly efficient carbene transfer to indoles, pyrroles, and cyclic alkenes.
  145. [145]
    Review Industrial enzyme applications - ScienceDirect.com
    Aug 1, 2002 · The majority of currently used industrial enzymes are hydrolytic in action, being used for the degradation of various natural substances.
  146. [146]
    Applications of Microbial Enzymes in Food Industry - PMC - NIH
    The present review discusses the recent advancement in enzyme technology for food industries. A comprehensive list of enzymes used in food processing, the ...
  147. [147]
  148. [148]
  149. [149]
  150. [150]
    Enzyme Linked Immunosorbent Assay - StatPearls - NCBI Bookshelf
    ELISA is a technique using enzymes to detect and quantify immunologic reactions, where a component is bound to a solid phase.
  151. [151]
    Horseradish peroxidase-triggered direct in situ fluorescent ...
    Nov 21, 2021 · Usually, ELISA uses an enzyme, typically horseradish peroxidase (HRP) or alkaline phosphatase (ALP), to convert the corresponding substrate into ...
  152. [152]
    Enzymatic depolymerization of alginate by two novel thermostable ...
    Sep 20, 2022 · Their application for degradation of alginate polymers produced by Pseudomonas aeruginosa in cystic fibrosis patients has been studied (Mrsny et ...Missing: therapy | Show results with:therapy
  153. [153]
    Thrombolytic Therapy - StatPearls - NCBI Bookshelf - NIH
    Thrombolytic treatment, also known as fibrinolytic therapy, dissolves dangerous intravascular clots to prevent ischemic damage by improving blood flow.
  154. [154]
    Streptokinase for Treatment of Thrombotic Disorders: The End ... - NIH
    PAs convert the endogenous human plasminogen (HPG: the inactive proenzyme) to plasmin (HPM: the active enzyme), which cleaves the fibrin network of the blood ...
  155. [155]
    CRISPR/Cas9 therapeutics: progress and prospects - Nature
    Jan 16, 2023 · Promising applications for this technology include the treatment of cancers, cardiovascular diseases, sickle cell anemia, and neurodegenerative ...
  156. [156]
    Engineering cytochrome P450 enzyme systems for biomedical ... - NIH
    Directed evolution has been widely applied to engineer P450s, the structures of which are often unknown, for desired properties under artificial selective ...
  157. [157]
    Streptokinase—a clinically useful thrombolytic agent - ScienceDirect
    Native streptokinase is useful for cost-effective thrombolytic therapy in clinical practice, but its use is not risk free. Large quantities of streptokinase ...Research Review Paper · Thrombolysis Or Fibrinolysis · Streptokinase
  158. [158]
    PEGylated therapeutics in the clinic - PMC - PubMed Central
    Sep 22, 2023 · Specifically, PEGylation can improve the parent drug's solubility, extend its circulation time, and reduce its immunogenicity, with minimal ...
  159. [159]
    Application of CRISPR-Cas9 genome editing technology in various ...
    In biomedical research, CRISPR-Cas9 has facilitated the study of gene functions and disease mechanisms. It has enabled researchers to create targeted gene ...