Fact-checked by Grok 2 weeks ago

Biosynthesis

Biosynthesis is the production of complex organic compounds within living organisms from simpler precursors through a series of enzyme-catalyzed reactions that form part of . This process contrasts with , which breaks down molecules to release , and instead requires input from high-energy molecules such as ATP and reducing agents like NADPH to drive the synthesis of essential cellular components. In cells, biosynthesis encompasses multiple interconnected pathways that build the fundamental building blocks of life, including , , , , and proteins. For instance, carbohydrate biosynthesis via converts non-carbohydrate precursors like pyruvate or into glucose, a that consumes four ATP molecules, two GTP molecules, and two NADH molecules per glucose unit produced. Lipid biosynthesis, occurring primarily in the , assembles fatty acids from units, with each two-carbon addition requiring one ATP and two NADPH, culminating in molecules like palmitate for membrane formation. Amino acid and nucleotide biosynthesis further illustrate the complexity of these pathways, drawing from central metabolic intermediates such as glucose and for , or ribose-5-phosphate for purines and pyrimidines. Proteins are then polymerized from on ribosomes using mRNA templates, consuming one ATP and two GTP molecules per incorporated, while nucleic acids are assembled from triphosphates, releasing as a . These pathways are tightly regulated to ensure efficient and are conserved across prokaryotes and eukaryotes, underscoring their fundamental role in cellular growth, repair, and adaptation to environmental conditions.

Core Concepts

Definition and Scope

Biosynthesis refers to the anabolic processes by which living organisms construct complex biomolecules from simpler precursor molecules through a series of enzyme-catalyzed reactions. These pathways enable the endogenous production of essential cellular components, including proteins, nucleic acids, , and carbohydrates, which are fundamental to sustaining life. As part of , biosynthesis supports cellular by increasing , maintenance by repairing or replacing damaged structures, and evolutionary by allowing organisms to generate diverse molecules in response to environmental pressures. The scope of biosynthesis encompasses both , which builds molecules from basic building blocks like or sugars, and salvage pathways that recycle intermediates from the of existing compounds to conserve and resources. This anabolic focus distinguishes biosynthesis from , the degradative processes that break down complex molecules to extract in the form of ATP. A pivotal historical milestone in recognizing biosynthesis as a chemical process occurred in 1828, when synthesized from , disproving the vitalist doctrine that organic compounds could only arise from living matter and laying the groundwork for modern biochemistry. Beyond its core metabolic role, biosynthesis is indispensable for organismal , as it allows cells to produce specialized metabolites that enhance survival, such as stress-response proteins or secondary compounds in . In biotechnology, biosynthetic pathways form the basis of , where engineered organisms are designed to produce high-value compounds like pharmaceuticals or biofuels, optimizing natural processes for industrial applications. Key examples of biosynthetic classes include , formed by assembling carbon skeletons with nitrogen; , critical for genetic material; and , which construct cellular membranes—each illustrating the diversity and precision of these pathways without delving into specific mechanisms.

Thermodynamic and Kinetic Principles

Biosynthetic reactions are inherently endergonic, characterized by a positive change in (ΔG > 0), rendering them thermodynamically unfavorable under standard conditions. To render these processes feasible, cells couple them to exergonic reactions, such as the of high-energy phosphate bonds, which provide the necessary input to shift the overall ΔG to a negative value. This thermodynamic coupling ensures that the net change favors the formation of complex biomolecules from simpler precursors, maintaining cellular . The standard free energy change (ΔG°') for the hydrolysis of (ATP) to (ADP) and inorganic phosphate (Pi) is -30.5 kJ/mol under physiological conditions (pH 7, 25°C). This reaction is represented as: \text{ATP} + \text{H}_2\text{O} \rightarrow \text{ADP} + \text{P}_\text{i}, \quad \Delta G^{\circ\prime} = -30.5 \, \text{kJ/mol} In vivo, the actual ΔG is even more exergonic (approximately -50 to -60 kJ/mol) due to non-standard concentrations of ATP, ADP, and Pi, amplifying its role in driving endergonic biosynthesis. Similarly, (GTP) hydrolysis provides comparable energy in specific pathways, underscoring the reliance on nucleoside triphosphates for thermodynamic feasibility. The (K_eq) for a quantifies its thermodynamic favorability and is related to the standard free energy change by the equation: K_\text{eq} = e^{-\Delta G^{\circ\prime} / RT} where R is the (8.314 J/mol·K) and T is the absolute temperature in . For , this yields a large K_eq (on the order of 10^5), heavily favoring product formation. In cellular environments, the mass action (Γ), defined as the actual of product to reactant concentrations, often deviates significantly from K_eq in biosynthetic pathways, maintaining reactions far from to prevent reversal and ensure directional flux toward synthesis. Kinetically, biosynthetic reactions face high activation energy barriers (E_a) that impede spontaneous progression, even when thermodynamically coupled. Coupling to exergonic processes not only provides thermodynamic drive but also facilitates kinetic acceleration by stabilizing transition states, thereby lowering effective E_a and increasing reaction rates without altering the equilibrium position. This interplay ensures that biosynthesis proceeds at biologically relevant timescales despite inherent kinetic hurdles. Redox balance, mediated by the NAD+/NADH couple, further integrates and in biosynthesis, with the NAD+/NADH ratio (typically around 500–1000 in ) maintaining reducing power for reductive steps while coupling to exergonic oxidations in catabolic pathways. For instance, NADH generated from fuels reductive biosyntheses, preserving overall cellular essential for sustained endergonic flux.

Enzymatic Mechanisms and Cofactors

Enzymes play a central role in biosynthesis by accelerating the formation of complex molecules from simpler precursors through precise catalytic mechanisms. Key classes include transferases, which catalyze the transfer of functional groups such as methyl or phosphoryl groups between substrates, as seen in radical enzymes that initiate radical-based cascades in cofactor biosynthesis. Ligases and synthetases, subclasses of ligases that often couple reactions to , join molecules via formation of new bonds, such as amide linkages in synthetases (NRPS) during production. These types ensure specificity and efficiency in anabolic pathways. Multistep complexes, like synthases (PKSs), integrate multiple domains—including ketosynthase, acyltransferase, and ketoreductase—into a single megasynthase to perform iterative Claisen condensations, enabling the assembly of diverse natural products such as erythromycin. Biosynthetic enzymes employ mechanisms such as induced fit, where substrate binding triggers conformational changes in the enzyme's to align catalytic residues and stabilize states, enhancing specificity and rate. Acid-base is another prevalent mechanism, involving proton donation or abstraction by side chains to facilitate bond breaking and formation; for instance, in cyclases, aspartate or glutamate residues act as general bases to deprotonate substrates during generation for cyclization reactions. A representative example is fructose-1,6-bisphosphate aldolase in sugar metabolism, where class I aldolases use a residue for formation and /glutamate pairs for acid-base assisted , enabling reversible aldol condensations between and glyceraldehyde-3-phosphate to form fructose-1,6-bisphosphate. Cofactors are essential non-proteinaceous molecules that assist enzymes in biosynthesis, often derived from s. (), synthesized from pantothenate (vitamin B5), , and ATP, serves as a carrier for acyl groups in linkages, facilitating acyl transfer reactions critical for and synthesis; a simplified representation is the step: \text{R-COOH} + \text{CoA} \rightarrow \text{R-CO-SCoA} + \text{H}_2\text{O} though it typically involves ATP-dependent . Universal cofactors like ATP provide energy for bond formation in ligases and synthetases, driving endergonic steps through hydrolysis to and inorganic phosphate. NADPH acts as the primary reducing agent in anabolic processes, donating a hydride ion and proton to reduce substrates such as in elongation or synthesis, with its oxidation depicted as: \text{NADPH} \rightarrow \text{NADP}^+ + \text{H}^+ + 2\text{e}^- maintaining a high NADPH/NADP⁺ ratio via pathways like the to favor reductive biosynthesis.

Precursor Molecules and Sources

Carbon and Energy Precursors

In biosynthesis, carbon precursors provide the structural building blocks for macromolecules, while energy carriers supply the necessary driving force for anabolic reactions. The primary carbon sources vary by organism type, with glucose serving as the main entry point in heterotrophs and carbon dioxide (CO₂) as the inorganic source in autotrophs. Acetyl-coenzyme A (acetyl-CoA) acts as a pivotal intermediate, channeling carbon from catabolic pathways into diverse biosynthetic routes, including fatty acid and polyketide synthesis. Energy is predominantly harnessed through adenosine triphosphate (ATP), generated via oxidative phosphorylation, which couples electron transport to proton gradient-driven ATP synthesis in mitochondria or prokaryotic membranes. Complementing ATP, nicotinamide adenine dinucleotide phosphate (NADPH) provides reducing equivalents essential for reductive biosyntheses, such as lipid and nucleotide assembly, and is primarily produced in the oxidative phase of the pentose phosphate pathway. In heterotrophic organisms like and many microbes, organic compounds such as glucose from dietary or environmental sources are the dominant carbon precursors. Glucose is phosphorylated and metabolized through , yielding pyruvate and key intermediates that feed into the tricarboxylic acid () cycle or directly into biosynthesis. For instance, pyruvate is decarboxylated to form , which integrates carbon flux from carbohydrates, fats via β-oxidation, and some , positioning it as a central metabolic nexus into biosynthetic pathways. itself supplies versatile precursors; glyceraldehyde-3-phosphate (G3P), an early intermediate, is diverted toward glycerolipid production by serving as the backbone for in biogenesis. This reliance on pre-formed organic carbon underscores the trophic dependence of heterotrophs on external inputs. Autotrophs, including and photosynthetic , contrast sharply by assimilating inorganic CO₂ as their carbon source through the Calvin-Benson-Bassham (CBB) cycle, also known as the reductive . In this cycle, CO₂ is fixed onto ribulose-1,5-bisphosphate by the enzyme , generating 3-phosphoglycerate, which is then reduced to G3P using ATP and NADPH derived from . The CBB cycle not only produces sugars for immediate energy but also exports G3P and other triose phosphates as precursors for , , and , enabling autotrophic self-sufficiency. , as primary producers, thus convert atmospheric CO₂ into , forming the base of food webs that sustain heterotrophic animals, which cannot perform this fixation and must acquire carbon via consumption of plant-derived organics. The integration of energy carriers ensures biosynthetic efficiency across both lifestyles. Oxidative phosphorylation yields up to 30-32 ATP per glucose molecule in aerobes, far exceeding in , and powers endergonic steps like in the CBB or in elongation. Meanwhile, the generates approximately 60% of cellular NADPH in mammals by oxidizing glucose-6-phosphate, bypassing ATP production to prioritize reductive power for defenses against and for constructing complex molecules like and deoxyribonucleotides. These precursors and carriers converge at metabolic branch points, allowing cells to balance and based on nutritional status and environmental cues.

Nitrogen and Sulfur Sources

Nitrogen assimilation in organisms begins with the acquisition of inorganic , primarily as (NH₄⁺) derived from biological in prokaryotes such as and . This process is catalyzed by the , a complex that reduces atmospheric dinitrogen (N₂) to (NH₃) through the : \ce{N2 + 8H+ + 8e- -> 2NH3 + H2} This energy-intensive requires 16 molecules of ATP per N₂ fixed and occurs exclusively in diazotrophic microorganisms, providing bioavailable for ecosystems. Once formed, NH₄⁺ is toxic at high concentrations and must be rapidly assimilated into organic compounds, primarily via the /glutamate synthase (GS-GOGAT) cycle, which operates in , , and other organisms to incorporate into . In this cycle, (GS) first condenses glutamate with NH₄⁺ and ATP to form , while glutamate synthase (GOGAT) then transfers the from to α-ketoglutarate, yielding two molecules of glutamate. Glutamate serves as the primary nitrogen donor in biosynthetic pathways, facilitating the transfer of amino groups to various carbon skeletons through transamination reactions. A key step in this process is the GOGAT-mediated reaction: \ce{glutamine + \alpha-ketoglutarate + NADPH -> 2 glutamate + NADP+} This generates glutamate, which can donate its α-amino group to keto acids like oxaloacetate or pyruvate, producing aspartate or alanine, respectively, and regenerating α-ketoglutarate for entry into the tricarboxylic acid cycle. The GS-GOGAT cycle thus links nitrogen assimilation to central metabolism, ensuring efficient recycling of nitrogen donors while preventing ammonium accumulation. Sulfur for biosynthesis is primarily sourced from (SO₄²⁻) in the , which is reduced to (H₂S) through the assimilatory pathway in , fungi, and . This multi-step process involves activation of sulfate to adenosine 5'-phosphosulfate (APS), followed by sequential reductions using or NADPH-dependent enzymes to yield and ultimately sulfide, requiring energy input from ATP. The resulting H₂S is then incorporated into organic molecules, notably during biosynthesis, where it reacts with O-acetylserine (derived from serine) in a β-replacement reaction catalyzed by to form L-cysteine and . This pathway provides the sulfur backbone for essential biomolecules like and coenzymes. In humans, who lack the capacity for de novo nitrogen fixation or synthesis of certain amino acids, nitrogen requirements are met entirely through dietary sources, particularly proteins rich in essential amino acids that cannot be synthesized endogenously. This reliance underscores the importance of balanced nutrition to supply non-synthesizable nitrogen, with deficiencies leading to impaired protein synthesis and metabolic disorders.

Mineral Elements in Biosynthesis

Mineral elements, including metals and inorganic ions, play critical roles as cofactors and structural components in enzymes that drive biosynthetic processes across organisms. These elements facilitate key reactions such as , transformations, and substrate activation, enabling the assembly of complex biomolecules like proteins, nucleic acids, and pigments. Without adequate incorporation, biosynthetic pathways are disrupted, highlighting their indispensable function in cellular . Magnesium ions (Mg²⁺) are among the most ubiquitous mineral cofactors in biosynthesis, particularly in ATP-dependent reactions. Mg²⁺ stabilizes the ATP-Mg complex, which is essential for binding to active sites in kinases and synthetases, facilitating phosphoryl transfer during the synthesis of , , and . For instance, in nucleotide biosynthesis, Mg²⁺ coordinates with ATP to donate phosphate groups via (PRPP) synthetase, where ribose-5-phosphate reacts with ATP to form PRPP, the activated precursor for and assembly. Similarly, iron (Fe²⁺ or Fe³⁺) is incorporated into during the final step of heme biosynthesis by ferrochelatase, which inserts Fe²⁺ into to yield , a vital in hemoproteins like and . Trace elements such as (Zn²⁺) and (Mo) serve specialized roles in biosynthetic enzymes, often occupying active sites to enable . Zn²⁺ is the central metal in , a zinc metalloenzyme that accelerates the reversible hydration of CO₂ to (HCO₃⁻), supporting CO₂ fixation in photosynthetic organisms and providing carbon sources for biosynthesis. In , a key biosynthetic entry point for ammonia production, is integral to the iron-molybdenum cofactor (FeMo-co) of , where it coordinates with iron-sulfur clusters to catalyze the reduction of N₂ to NH₃. Beyond , these minerals stabilize enzyme active sites; for example, divalent cations like Mg²⁺ and Zn²⁺ often polarize substrates or maintain conformational integrity in biosynthetic synthases. Phosphate ions (Pᵢ), derived primarily from ATP hydrolysis, are incorporated into biomolecules during biosynthesis, underscoring their role as both cofactor and building block. In nucleotide synthesis, Pᵢ from ATP energizes the formation of high-energy bonds in PRPP and subsequent steps, ensuring the structural integrity of DNA and RNA precursors. Deficiencies in mineral elements can severely impair these pathways; magnesium shortage, for example, limits chlorophyll biosynthesis by halting Mg²⁺ insertion into protoporphyrin IX by magnesium chelatase, resulting in reduced photosynthetic pigment production and diminished carbon fixation efficiency in plants. Such disruptions emphasize the tight regulation required for mineral homeostasis in biosynthetic networks.

Biosynthesis of Amino Acids

Classification and Basic Structure

Amino acids are organic molecules that serve as the fundamental building blocks of proteins, sharing a common consisting of a central α-carbon atom bonded to an (-NH₂), a (-COOH), a , and a variable side chain (R group), represented by the general formula H₂N-CH(R)-COOH. This zwitterionic structure at physiological pH allows to participate in formation and diverse biochemical roles. The 20 standard proteinogenic are distinguished solely by the nature of their R groups, which determine physicochemical properties such as hydrophobicity, charge, and reactivity. These R groups enable classification into four main categories: nonpolar (e.g., , ), polar uncharged (e.g., serine, ), acidic (e.g., , ), and basic (e.g., , ). Except for , whose R group is a and thus lacks a chiral , all standard are chiral at the α-carbon and exist predominantly as L-enantiomers in living organisms, a that is essential for the specificity of enzymatic reactions and . This L-configuration arises from evolutionary selection and ribosomal synthesis mechanisms that incorporate only L-isomers into polypeptides. In contrast, D-amino acids, which are mirror images of L-forms, are rare in eukaryotic proteins but play structural roles in prokaryotes, particularly as components of cross-links in bacterial cell walls, conferring resistance to host proteases. Amino acids in humans are further classified as essential or non-essential based on nutritional requirements, with essential amino acids unable to be synthesized de novo due to the absence of specific biosynthetic enzymes. The nine amino acids—histidine, , , , , , , , and —must be acquired through dietary sources to support protein synthesis and metabolic functions; for example, and are critical branched-chain and basic amino acids, respectively. The 11 non-essential amino acids, such as and glutamate, can be produced endogenously from metabolic intermediates. The carbon frameworks for amino acid biosynthesis originate from glycolytic and tricarboxylic acid (TCA) cycle intermediates, linking central carbon metabolism to nitrogen assimilation. Glycolysis provides precursors like phosphoenolpyruvate and 3-phosphoglycerate for the synthesis of serine, glycine, and alanine, while TCA cycle compounds such as α-ketoglutarate (for glutamate and proline) and oxaloacetate (for aspartate and asparagine) supply skeletons for other families, with nitrogen typically incorporated from ammonia or glutamine.

Glutamate and Serine Families

The glutamate family of amino acids, comprising , , , and , originates from the tricarboxylic acid () cycle intermediate α-ketoglutarate (α-KG). In most organisms, serves as the primary entry point for incorporation into this family, synthesized through the reversible reaction catalyzed by (GDH), which facilitates the of α-KG with and NADPH: \alpha\text{-ketoglutarate} + \text{NH}_4^+ + \text{NADPH} \rightleftharpoons \text{L-glutamate} + \text{NADP}^+ + \text{H}_2\text{O}. This operates in mitochondria and plays a key role in assimilation and homeostasis, particularly in and where GDH activity is prominent under high conditions. In mammals, including humans, is also formed via reactions using 5'-phosphate () as a cofactor, transferring an amino group from other to α-KG, though the GDH pathway contributes to production during catabolic states. From glutamate, is derived by the addition of a second molecule via (GS), forming γ-glutamyl amide: glutamate + NH₄⁺ + ATP → + + Pi + H⁺. This step is crucial for nitrogen transport and storage, as serves as a non-toxic carrier of . biosynthesis branches from glutamate through a series of reductions: glutamate is first converted to glutamate-5-semialdehyde by glutamate-5-kinase and γ-glutamyl phosphate reductase, followed by spontaneous cyclization and reduction to . synthesis in this family proceeds via the conversion of glutamate to through acetylation and transamination steps, then incorporation into the (in animals) or analogous pathways (in and ) to form and ultimately . Transaminases, dependent on , are pivotal throughout these derivations, enabling amino group transfers with high specificity. In and , these pathways enable complete from inorganic precursors, whereas in humans, while the core enzymatic machinery exists for glutamate, , , and production, the latter is conditionally due to limited flux under physiological stress, relying partly on dietary input. The serine family includes serine and , derived from the glycolytic intermediate 3-phosphoglycerate (3-PG) in a phosphorylated pathway conserved across eukaryotes and prokaryotes. Serine biosynthesis begins with the oxidation of 3-PG to 3-phosphohydroxypyruvate by 3-phosphoglycerate dehydrogenase (PGDH), followed by to 3-phosphoserine (catalyzed by phosphoserine aminotransferase, PSAT, using ) and dephosphorylation to L-serine by phosphoserine phosphatase (PSP). This pathway is rate-limited by PGDH, which is feedback-inhibited by serine levels, ensuring balanced production for one-carbon metabolism and synthesis. is then generated from serine by (SHMT), which transfers a to tetrahydrofolate (THF): serine + THF → + 5,10-methylene-THF. The methylene-THF product links this family to folate-dependent reactions, supporting and biosynthesis. , another non-essential , is derived from serine via the transsulfuration pathway, where serine combines with to form cystathionine, which is then cleaved to , incorporating sulfur from metabolism. In humans, this pathway operates primarily in the of liver and kidney cells, rendering serine, , and non-essential fully synthesizable , unlike in auxotrophic mutants of or where environmental supplements are required. -dependent enzymes, such as PSAT and SHMT, underscore the reliance on for efficient nitrogen handling in both families.

Aspartate Family and Other Non-Essential Amino Acids

The biosynthesis of aspartate begins with the transamination of oxaloacetate, a key intermediate in the tricarboxylic acid cycle, using glutamate as the amino donor, catalyzed by aspartate aminotransferase (also known as aspartate transaminase). This reaction effectively incorporates nitrogen derived from ammonia assimilation into the carbon skeleton, yielding L-aspartate and α-ketoglutarate. In most organisms, including mammals, plants, and bacteria, this mitochondrial or cytosolic enzyme facilitates the reversible transfer, ensuring aspartate availability for protein synthesis and other metabolic roles. From aspartate, the pathway branches to produce amino acids including (non-essential in humans) and, in , , and other organisms capable of , , , and (essential in humans), primarily through a series of , , and steps. is synthesized directly from aspartate by asparagine synthetase, which transfers an amide group from to aspartate, producing glutamate as a ; this ATP-dependent reaction is crucial in storage and transport, particularly in . Methionine biosynthesis in and proceeds via the conversion of aspartate to β-aspartyl-phosphate, then to aspartate semialdehyde, and subsequently to homoserine, which is activated and methylated to form , involving enzymes like homoserine kinase and . is derived from homoserine through and by threonine synthase, while synthesis in and follows the diaminopimelate pathway, branching from aspartate semialdehyde to produce α,ε-diaminopimelate, which is decarboxylated to L-. These pathways are interconnected, sharing early intermediates to optimize resource allocation in response to cellular needs. Alanine, another non-essential amino acid, is synthesized via a straightforward reaction where pyruvate serves as the carbon precursor and glutamate donates the amino group, catalyzed by (also called glutamate-pyruvate transaminase). This reversible process links to metabolism, allowing alanine to act as a nitrogen carrier between tissues, such as from muscle to liver during . The reaction predominates in the and is essential for maintaining balance without requiring complex branching pathways. The aromatic , , and —are biosynthesized via the , a seven-step route absent in animals but present in , fungi, and ; and are essential in humans, while is non-essential and derived from by . The pathway commences with the condensation of phosphoenolpyruvate (PEP) from and erythrose-4-phosphate (E4P) from the , forming 3-deoxy-D-arabino-heptulosonate-7-phosphate (DAHP), catalyzed by DAHP : \text{PEP} + \text{E4P} \xrightarrow{\text{DAHP synthase}} \text{DAHP} + \text{P}_i Subsequent steps involve isomerization, oxidation, and cyclization to chorismate, the branch point intermediate; from chorismate, phenylalanine and tyrosine arise via prephenate, while tryptophan branches through anthranilate. This pathway not only supplies amino acids for protein synthesis but also precursors for secondary metabolites like flavonoids and lignin in plants. Regulation of these biosynthetic routes primarily occurs through feedback inhibition at key enzymatic steps to prevent overaccumulation of end products. In the aspartate family, aspartokinase—the first committed enzyme—is allosterically inhibited by and in a concerted manner in many , while homoserine dehydrogenase is sensitive to alone; plants exhibit isozyme-specific regulation to balance the branched pathways. For the , DAHP synthase is feedback-inhibited by , , and , often through multiple isozymes allowing fine-tuned control based on levels. These mechanisms ensure efficient carbon and nitrogen utilization, adapting to environmental and nutritional cues.

Essential Amino Acids in Humans

Humans require nine amino acids that cannot be synthesized due to the evolutionary loss of specific biosynthetic pathways, necessitating their acquisition through diet. These essential amino acids include , , , , , , , , and . In contrast to non-essential amino acids, which humans can produce from common metabolic intermediates like glucose or other , the essential ones depend on microbial or synthesis for their production. The branched-chain essential amino acids—valine, leucine, and isoleucine—are derived in bacteria and plants such that valine and leucine come from pyruvate through a shared pathway involving acetolactate synthase and subsequent isomeroreductases, reductoisomerases, and dehydratases, while isoleucine derives from threonine and pyruvate. Phenylalanine and tryptophan originate from the shikimate pathway, starting with phosphoenolpyruvate and erythrose-4-phosphate, leading to chorismate as a key intermediate in microorganisms and higher plants. Methionine, lysine, and threonine stem from the aspartate family pathway, beginning with aspartate and involving enzymes like homoserine dehydrogenase and cystathionine gamma-synthase in prokaryotes. Histidine is synthesized from phosphoribosyl pyrophosphate (PRPP) and ATP via a complex pathway with 10 enzymatic steps in bacteria. Humans lack the complete enzymatic machinery for these pathways, such as the acetolactate synthase required for synthesis from pyruvate, due to loss during metazoan evolution. For instance, while humans possess transaminases for , they cannot perform the initial condensation steps or the full reductive sequences needed for production. In microbial systems, leucine biosynthesis exemplifies this: α-ketoisocaproate, an intermediate formed from α-isopropylmalate via β-isopropylmalate dehydrogenase, is transaminated by a aminotransferase to yield . Nutritionally, these dependencies mean humans must consume adequate amounts of all essential amino acids to support protein synthesis and other metabolic functions, with requirements varying by age and physiological state—adults need approximately 19 mg/kg/day for , for example. Diets based on single plant sources, such as grains or , often provide incomplete proteins lacking one or more essential amino acids, requiring complementary foods like to achieve balance. This dietary necessity underscores the reliance on diverse food sources or microbial/plant-derived supplements to meet needs.

Biosynthesis of Nucleotides

Purine Biosynthesis Pathway

The de novo biosynthesis pathway assembles the purine ring stepwise onto a ribose-5-phosphate backbone, starting from 5-phosphoribosyl-1-pyrophosphate (PRPP) and culminating in inosine monophosphate () after 10 enzymatic reactions. This anabolic process is highly conserved across eukaryotes and prokaryotes, requiring input from , one-carbon units, and CO₂ to construct the bicyclic purine structure essential for synthesis and cellular signaling. The pathway operates primarily in the , with some enzymes forming multi-enzyme complexes to channel intermediates efficiently. The atoms contributing to the ring originate from specific metabolic : provides carbons 4, 5 and 7; aspartate donates 1; supplies nitrogens 3 and 9; 10-formyltetrahydrofolate (fTHF) contributes carbons 2 and 8; and CO₂ furnishes carbon 6. PRPP serves as the activated ribose donor, with its ribose-5-phosphate derived from the . The overall process consumes six ATP equivalents per IMP produced, highlighting its energy-intensive nature. The pathway's first committed step, catalyzed by glutamine-PRPP amidotransferase (PPAT, also called amidophosphoribosyltransferase), irreversibly activates PRPP by replacing its pyrophosphate group with ammonia from glutamine, forming 5-phosphoribosylamine (PRA). This rate-limiting reaction is allosterically inhibited by end products AMP and GMP to prevent overproduction. The equation for this step is: \text{PRPP} + \text{glutamine} + \text{H}_2\text{O} \rightarrow \text{PRA} + \text{glutamate} + \text{PP}_\text{i} Subsequent steps progressively build the imidazole ring (steps 2–5) and ring (steps 6–10) on the PRA scaffold, incorporating the remaining atoms via amidotransferases, carboxylases, and transformylases. Key enzymes include (GAR) synthase for addition, phosphoribosylaminoimidazole-succinocarboxamide synthase (PAICS) for aspartate incorporation and ring closure, and aminoimidazolecarboxamide ribonucleotide (AICAR) (ATIC) for the final and cyclization to IMP. The 10 steps are summarized below:
  1. PPAT: PRPP + → PRA + glutamate + (N9 addition).
  2. GAR synthase (GART): PRA + + ATP → glycinamide (GAR) + + Pi (C4, C5, N7 addition).
  3. GAR transformylase (GART): GAR + 10-formyl-THF → formylglycinamide ribonucleotide (FGAR) + THF (C8 addition).
  4. FGAR amidotransferase (PFAS): FGAR + + ATP + H₂O → formylglycinamidine ribonucleotide (FGAM) + glutamate + + Pi (N3 addition).
  5. AIR synthase (PFAS): FGAM + ATP → 5-aminoimidazole (AIR) + + Pi (imidazole ring closure).
  6. AIR carboxylase (PAICS): AIR + CO₂ → 4-carboxyaminoimidazole (CAIR) (C6 addition).
  7. SAICAR synthase (PAICS): CAIR + aspartate + ATP → 5-aminoimidazole-4-(N-succinylcarboxamide) (SAICAR) + + Pi (N1 addition).
  8. Adenylosuccinate lyase (): SAICAR → 5-aminoimidazole-4-carboxamide (AICAR) + fumarate.
  9. AICAR transformylase (ATIC): AICAR + 10-formyl-THF → 5-formaminoimidazole-4-carboxamide (FAICAR) + THF (C2 addition).
  10. IMP cyclohydrolase (ATIC): FAICAR → inosine monophosphate () (pyrimidine ring closure).
These reactions are tightly coupled, with bifunctional enzymes like GART (steps 2–3), (step 4–5), and PAICS (steps 6–7) and ATIC (steps 9–10) enhancing efficiency in humans. serves as the branch point for and . Conversion to () occurs in two steps: adenylosuccinate synthetase (ADSS) forms adenylosuccinate from , aspartate, and GTP, followed by cleavage by to and fumarate. For (), is oxidized to xanthosine monophosphate (XMP) by IMP dehydrogenase (IMPDH) using NAD⁺ and H₂O, then aminated by GMP synthetase (GMPS) with and ATP to yield GMP. These branches balance and production through cross-regulation, such as GTP stimulating AMP and ATP promoting GMP formation.

Pyrimidine Biosynthesis Pathway

The de novo pyrimidine biosynthesis pathway assembles the pyrimidine ring through a series of six enzymatic reactions, starting from simple precursors and culminating in the production of uridine monophosphate (UMP), the central precursor for all pyrimidine nucleotides. This linear process contrasts with purine biosynthesis by forming and closing the six-membered pyrimidine ring early, before attachment to the ribose-5-phosphate moiety derived from phosphoribosyl pyrophosphate (PRPP). The pathway is essential for providing building blocks for RNA, DNA, and coenzymes, and it is highly conserved across eukaryotes, with the first three steps catalyzed by the multifunctional CAD enzyme complex in mammals. The pathway initiates with the synthesis of from , (derived from CO₂), and two ATP molecules, catalyzed by the glutaminase and synthetase domains of carbamoyl phosphate synthetase II (CPSII), a component of the CAD complex. This is followed by the condensation of with aspartate to yield carbamoyl aspartate and inorganic phosphate (Pᵢ), mediated by the aspartate transcarbamoylase (ATCase) domain of CAD: \text{Carbamoyl phosphate} + \text{aspartate} \rightarrow \text{carbamoyl aspartate} + \text{P}_\text{i} Cyclization then occurs via the dihydroorotase (DHOase) domain of CAD, converting carbamoyl aspartate to L-dihydroorotate with the release of water.30130-2) Oxidation of L-dihydroorotate to orotate is carried out by dihydroorotate dehydrogenase (DHODH), a mitochondrial enzyme that uses quinone as an electron acceptor. The final two steps involve the phosphoribosyltransferase and decarboxylase activities of the bifunctional UMP synthase (UMPS): orotate reacts with PRPP to form orotidine 5'-monophosphate (OMP), which is then decarboxylated to UMP, releasing CO₂. From UMP, the pathway branches to produce other pyrimidines: UMP is phosphorylated to uridine triphosphate (UTP), which is aminated by CTP synthetase to form cytidine triphosphate (CTP), the precursor to CMP; alternatively, UMP is reduced to deoxyuridine monophosphate (dUMP), which is methylated by thymidylate synthase to yield thymidine monophosphate (TMP). All atoms of the pyrimidine ring originate from aspartate (providing carbons 4, 5, and 6, as well as nitrogen 1), CO₂ (carbon 2), and glutamine (nitrogen 3), highlighting the pathway's reliance on central metabolic intermediates for ring construction. This early ring closure distinguishes the process from purine synthesis, where the imidazole ring forms first on PRPP before the pyrimidine portion is added.

Nucleotide Salvage and Interconversion

Nucleotide salvage pathways enable cells to recycle purine and pyrimidine bases or nucleosides derived from the degradation of nucleic acids, thereby conserving energy and resources compared to de novo synthesis. These pathways are particularly vital in tissues with high rates of cell turnover, such as bone marrow and intestinal epithelium, where rapid nucleotide replenishment is essential for DNA and RNA production. By reutilizing pre-existing components, salvage mechanisms reduce the demand on ribose-5-phosphate and other precursors, making them an efficient alternative to the more ATP-intensive de novo routes. In salvage, the enzyme (HGPRT), also known as HPRT, plays a central role by catalyzing the conversion of free bases to using 5-phosphoribosyl-1-pyrophosphate (PRPP) as the ribose-phosphate donor. Specifically, HGPRT facilitates the reaction where hypoxanthine reacts with PRPP to form inosine monophosphate (IMP) and (PPi):
\text{hypoxanthine} + \text{PRPP} \rightarrow \text{IMP} + \text{PPi}
Similarly, is converted to (GMP) via the same enzyme. can also be salvaged by adenine phosphoribosyltransferase (APRT) to form (AMP). Defects in HGPRT activity, as seen in Lesch-Nyhan syndrome—an X-linked disorder caused by mutations in the HPRT1 —lead to impaired purine recycling, resulting in elevated levels, , neurological dysfunction, and self-mutilative behavior.
For pyrimidines, salvage primarily involves kinases that phosphorylate nucleosides to monophosphates. (TK), particularly the cytosolic isoform TK1, initiates the recycling of by phosphorylating it to deoxythymidine monophosphate (dTMP), which can then enter the pool for . This step is crucial in proliferating cells, where TK1 expression is cell cycle-regulated and peaks during the . and nucleosides are salvaged by uridine-cytidine kinase (UCK), converting them to (UMP) and (CMP), respectively. Nucleotide interconversions allow for the balanced production of different purine and pyrimidine forms from shared precursors. In purines, IMP serves as a branch point: it is aminated to AMP via adenylosuccinate synthetase and lyase, or converted to xanthosine monophosphate (XMP) by IMP dehydrogenase followed by amination to GMP; conversely, GMP can be deaminated back to IMP by GMP reductase, enabling interconversion between adenine and guanine nucleotides. For pyrimidines, UMP is sequentially phosphorylated to uridine diphosphate (UDP) by UMP kinase and then to uridine triphosphate (UTP) by nucleoside diphosphate kinase; UTP is subsequently aminated by CTP synthetase to form cytidine triphosphate (CTP). These conversions maintain nucleotide pools for RNA and DNA requirements. The importance of salvage and interconversion pathways lies in their , as they bypass the multi-step that consumes significant ATP and reducing equivalents, which is advantageous in rapidly dividing cells or under nutrient-limited conditions. In high-turnover tissues, these pathways can supply up to 90% of required , minimizing metabolic burden. Disruption, such as HGPRT deficiency in Lesch-Nyhan syndrome, not only causes purine overproduction and gout-like symptoms but also impairs brain development due to nucleotide imbalances. Regulation of these pathways is primarily governed by substrate availability, including PRPP levels for phosphoribosyltransferases and concentrations from degradation. For instance, elevated PRPP enhances HGPRT activity, while end-product inhibition by and GMP modulates interconversion enzymes like IMP dehydrogenase. In proliferating cells, increased uptake and expression further amplify salvage flux in response to demand.

Biosynthesis of Lipids

Fatty Acid Synthesis

Fatty acid synthesis, also known as de novo lipogenesis, is a that constructs saturated fatty acids from precursors, primarily in the of eukaryotic cells such as those in liver, , and lactating mammary glands. This process is essential for producing fatty acids used in formation, , and signaling molecules. , the starting substrate, is mainly derived from the oxidation of carbohydrates via the mitochondrial citrate shuttle, where citrate is exported to the cytosol and cleaved by ATP-citrate lyase to generate . The pathway is highly regulated to match cellular energy status and nutritional availability. The committed and rate-limiting step is the of to , catalyzed by the biotin-dependent enzyme (ACC). The reaction proceeds as follows: + HCO₃⁻ + ATP → + + Pᵢ. This step was first characterized by Wakil and colleagues in 1962, who identified ACC as the key regulatory enzyme in fatty acid biosynthesis. not only serves as the two-carbon donor for chain elongation but also inhibits , preventing simultaneous fatty acid synthesis and β-oxidation. Subsequent elongation occurs via the fatty acid synthase (FAS) complex, which iteratively adds two-carbon units from to a growing acyl chain attached to an (ACP). In eukaryotes, FAS operates as a type I system—a large, homodimeric multifunctional polypeptide (approximately 270 kDa per subunit) that houses all catalytic domains in a single protein, facilitating efficient substrate channeling; this contrasts with the type II FAS in and plant plastids, where discrete monofunctional enzymes perform each step. The process begins with the transfer of an acetyl group from to the ACP, followed by seven cycles of four reactions each: (1) condensation with malonyl-ACP by β-ketoacyl-ACP synthase to form β-ketoacyl-ACP and release CO₂; (2) of the β-keto group by β-ketoacyl-ACP reductase using NADPH; (3) to form trans-Δ²-enoyl-ACP; and (4) of the double bond by enoyl-ACP reductase using another NADPH molecule. These cycles, elucidated through pioneering work by Feodor Lynen in the 1950s and 1960s, build the chain to palmitate (C16:0), the primary product, after which a thioesterase domain hydrolyzes the palmitoyl-ACP to release free . The overall stoichiometry for palmitate synthesis requires eight molecules (one initial plus seven via ), seven ATP for , and 14 NADPH for reductions: 8 + 7 ATP + 14 NADPH + 14 H⁺ → palmitate + 7 + 7 Pᵢ + 14 NADP⁺ + 8 + 6 H₂O. Regulation occurs primarily at , which is allosterically activated by citrate (indicating ample carbon supply from the cycle) and inhibited by long-chain products, ensuring synthesis aligns with energy abundance.

Phospholipid and Sphingolipid Assembly

Phospholipid assembly in eukaryotic cells primarily occurs in the (), where phosphatidate serves as the central intermediate for synthesis. Phosphatidate is generated by sequential of glycerol-3-phosphate with two fatty molecules, catalyzed by glycerol-3-phosphate acyltransferase (GPAT) and acylglycerophosphate acyltransferase (AGPAT), predominantly in the and mitochondrial outer . This step incorporates fatty acids derived from prior synthesis, establishing the hydrophobic tails essential for integration. Dephosphorylation of phosphatidate by (lipins) yields diacylglycerol (DAG), which acts as the acceptor for polar head groups to form major such as (PC), (PE), (PS), and (PI). The pathway represents a key route for PC and biosynthesis, named after Eugene who elucidated its steps in the 1950s. In this pathway, free choline is phosphorylated by choline kinase (CHOK) to , primarily in the and . then reacts with CTP, catalyzed by CTP: cytidylyltransferase (), the rate-limiting , to produce CDP-choline. Finally, CDP-choline donates its group to DAG via choline/ phosphotransferase (CEPT) in the or choline phosphotransferase (CPT) in the Golgi, yielding PC. A parallel ethanolamine branch using ethanolamine kinase, ECT, and EPT forms . This pathway accounts for the majority of PC synthesis in mammalian cells, ensuring . For PS and PI, assembly involves base-exchange reactions or direct CDP-activated intermediates. PS is synthesized in the ER by phosphatidylserine synthase (PSS1 and PSS2), which exchanges the head group of PC or PE with serine, while PI forms from CDP-diacylglycerol (CDP-DAG, derived from phosphatidate and CTP via CDP-DAG synthase) and inositol via phosphatidylinositol synthase (PIS). These reactions maintain the diversity of head groups critical for membrane curvature and protein recruitment. Phospholipids contribute to membrane fluidity by modulating lipid packing and serve as precursors for signaling molecules like diacylglycerol and inositol phosphates. Sphingolipid assembly begins with biosynthesis in the , producing as the core scaffold for complex sphingolipids. The pathway initiates with the condensation of serine and palmitoyl-CoA by serine palmitoyltransferase (SPT), the rate-limiting enzyme, to form 3-ketodihydrosphingosine, which is reduced to sphinganine (dihydrosphingosine) by 3-ketosphinganine reductase. Sphinganine is then N-acylated by ceramide synthases (CerS1-6, with isoform specificity for fatty acid chain length) to dihydroceramide, followed by desaturation to . This backbone is transported to the Golgi for further modification. In the Golgi, is converted to by (SMS1/2), which transfers from PC to , or to glycosphingolipids via glucosylceramide (GCS) adding glucose, followed by further in the trans-Golgi network. , abundant in plasma membranes, influences raft formation and interactions, while glycosphingolipids contribute to cell recognition. regulate , vesicular trafficking, and signaling cascades, with acting as a pro-apoptotic mediator and promoting cell survival. Dysregulation of these pathways is implicated in diseases like cancer and neurodegeneration.

Sterol Biosynthesis Including Cholesterol

Sterol biosynthesis, a critical branch of the , produces and related essential for eukaryotic fluidity and precursor roles in various biomolecules. This pathway begins in the with the condensation of three molecules of to form 3-hydroxy-3-methylglutaryl-CoA (), catalyzed by acetoacetyl-CoA and HMG-CoA synthase, followed by the rate-limiting reduction of to mevalonate by , which consumes two NADPH molecules. Mevalonate is then sequentially phosphorylated and decarboxylated to yield isopentenyl (), the five-carbon building block, through the actions of mevalonate , phosphomevalonate , and mevalonate diphosphate decarboxylase. isomerizes to dimethylallyl (DMAPP), which condenses stepwise to form geranyl (GPP) and (FPP) via prenyltransferases. Two FPP molecules condense to presqualene pyrophosphate and are then reduced to by squalene synthase, marking the transition to the (ER) where subsequent cyclization occurs. is epoxidized to squalene-2,3-oxide by squalene epoxidase, and the oxide undergoes cyclization to via oxidosqualene cyclase ( synthase), establishing the tetracyclic core. is then transformed into through approximately 19 additional enzymatic steps, including demethylations at C4 and C14, isomerizations, and reductions, primarily catalyzed by enzymes like 14α-demethylase and Δ24-reductase, resulting in a total of about 30 steps from . In eukaryotes, the early mevalonate stages occur in the , while formation and downstream conversions localize to the ER membranes. Plants diverge by using cycloartenol as the initial cyclized intermediate instead of , leading to phytosterols, whereas fungi produce as their primary membrane through analogous but distinct modifications. The overall biosynthesis of one cholesterol molecule from acetyl-CoA can be simplified as 18 + 36 ATP + 16 NADPH + 16 H⁺ + 11 O₂ → + 16 NADP⁺ + 36 + 36 Pᵢ + 18 + 6 CO₂ + 11 H₂O, highlighting the high and reducing power demands. Regulation primarily occurs at the step, governed by sterol regulatory element-binding proteins (SREBPs), which, upon low levels, translocate to the nucleus to upregulate genes for and other pathway enzymes, ensuring feedback control based on cellular needs. This mechanism, elucidated through foundational studies on homeostasis, underscores the pathway's integration with broader .

Biosynthesis of Carbohydrates

Monosaccharide Production

Monosaccharide production encompasses several biosynthetic pathways that generate simple sugars from central metabolic intermediates, primarily in heterotrophic organisms via and the (PPP), as well as de novo routes for modified sugars like amino sugars and , and in autotrophs through the . These processes ensure the supply of glucose, , and other monosaccharides essential for energy storage, nucleotide synthesis, and glycoconjugate formation. In animals and many microorganisms, synthesizes glucose from non-carbohydrate precursors such as pyruvate, , and TCA cycle intermediates, bypassing the irreversible steps of through specialized enzymes. Gluconeogenesis proceeds as a reversal of from pyruvate to glucose-6-phosphate, primarily in the liver and kidneys, with key regulatory enzymes including (PEPCK), which converts oxaloacetate to phosphoenolpyruvate, and fructose-1,6-bisphosphatase (FBPase), which hydrolyzes fructose-1,6-bisphosphate to fructose-6-phosphate. Pyruvate is first carboxylated to oxaloacetate by in the mitochondria, followed by and via PEPCK to form phosphoenolpyruvate; subsequent steps mirror until FBPase and glucose-6-phosphatase complete the pathway to free glucose. This process is tightly regulated to prevent futile cycling with , with PEPCK and FBPase activated under fasting conditions by hormones like . In and , similar mechanisms operate but integrate with photosynthetic or fermentative metabolism. The provides ribose-5-phosphate for biosynthesis and generates NADPH for reductive reactions, branching from at glucose-6-phosphate. In the oxidative phase, glucose-6-phosphate is dehydrogenated to 6-phosphogluconolactone by , followed by , further oxidation, and to yield ribulose-5-phosphate, producing two NADPH per glucose-6-phosphate and releasing one CO₂. For balanced production, six glucose-6-phosphate molecules undergo the oxidative phase to form five ribulose-5-phosphate equivalents, with the overall reaction: $6 \text{ Glucose-6-P} + 12 \text{ NADP}^+ + 6 \text{ H}_2\text{O} \rightarrow 5 \text{ Ribose-5-P} + 6 \text{ CO}_2 + 12 \text{ NADPH} + 12 \text{ H}^+ The non-oxidative phase interconverts pentoses to glycolytic intermediates via transketolase and transaldolase, allowing flexible flux toward ribose-5-phosphate when demand is high, such as during cell proliferation. This pathway is irreversible in its oxidative steps and is a major source of ribose in non-photosynthetic tissues. De novo synthesis of amino sugars begins with the conversion of fructose-6-phosphate, a glycolytic intermediate, to glucosamine-6-phosphate by glutamine:fructose-6-phosphate amidotransferase (GFAT), the rate-limiting of the hexosamine biosynthetic pathway. This reaction incorporates an amino group from , producing glucosamine-6-phosphate and glutamate, which serves as the precursor for UDP-N-acetylglucosamine used in . GFAT activity is feedback-inhibited by UDP-GlcNAc and regulated by nutrient availability, linking carbohydrate flux to protein and lipid modification. In mammals, this pathway accounts for about 2-5% of glucose metabolism in most tissues but increases under hyperglycemic conditions. Mannose production occurs via the conversion of fructose-6-phosphate to mannose-6-phosphate by phosphomannose isomerase, followed by activation to GDP-mannose through mannose-6-phosphate isomerization to mannose-1-phosphate and guanylylation by GDP-mannose pyrophosphorylase. GDP-mannose acts as the donor for mannose incorporation into glycoproteins and cell wall polysaccharides, particularly in eukaryotes where it is essential for N-glycosylation and ascorbic acid biosynthesis in plants. This pathway ensures mannose availability from common hexose pools, with defects in phosphomannose isomerase causing congenital disorders of glycosylation. In autotrophic organisms like and , the fixes CO₂ into monosaccharides within chloroplasts, starting with ribulose-1,5-bisphosphate (RuBP) carboxylation by to form two molecules of 3-phosphoglycerate (3-PGA). The reductive phase reduces 3-PGA to glyceraldehyde-3-phosphate using ATP and NADPH from the light reactions, while the regenerative phase employs aldolase to condense and glyceraldehyde-3-phosphate into fructose-1,6-bisphosphate, and to transfer carbon units for RuBP regeneration, ultimately yielding net glucose equivalents. For every three CO₂ fixed, one glyceraldehyde-3-phosphate is exported for or synthesis, highlighting the cycle's role in converting inorganic carbon to organic monosaccharides. This process is light-dependent and accounts for nearly all global .

Polysaccharide Synthesis Pathways

Polysaccharide synthesis involves the enzymatic of activated units, primarily derived from sugars, to form linear or branched chains that serve as structural components or molecules in cells. These pathways typically proceed via glycosyltransferases that catalyze the formation of glycosidic bonds, adding units to the non-reducing end of growing chains in a processive manner. In structural polysaccharides like and , synthesis occurs at the plasma membrane, extruding polymers extracellularly, while storage forms such as and are assembled in the or plastids. Cellulose biosynthesis, a key pathway for plant cell wall formation, utilizes UDP-glucose as the activated substrate, which is polymerized into β-1,4-linked glucan chains by cellulose synthase enzymes (CesA). In plants, cellulose synthases are organized into rosette-shaped terminal complexes in the plasma membrane, each complex containing multiple CesA isoforms (e.g., AtCesA1, AtCesA3, AtCesA6) that coordinately synthesize 36 to 90 parallel glucan chains, forming microfibrils 3.5–10 nm in diameter. The mechanism involves processive addition of glucose units to the non-reducing end through an inverting glycosyltransferase reaction, with the chains elongating at rates of approximately 4–6 nm per second before crystallization into rigid microfibrils. In bacteria, such as Acetobacter xylinum, cellulose synthesis is similarly processive but mediated by linear terminal complexes producing ribbons of varying chain numbers, highlighting evolutionary adaptations in machinery organization. Chitin synthesis, essential for fungal cell walls and exoskeletons, employs UDP-N-acetylglucosamine (UDP-GlcNAc) as the , which is transferred by chitin synthases (CHS) to form β-1,4-linked polymers. Chitin synthases, belonging to family 2, facilitate processive at the non-reducing end via an S_N2 inverting , where a swinging loop motif (e.g., VLPGA) acts as a gate to direct sequential addition and translocation of the growing chain through a transmembrane . This pathway is conserved across eukaryotes, with CHS enzymes often zymogenic and activated by or partial , ensuring controlled deposition of microfibrils that provide tensile strength. Starch and amylopectin biosynthesis in plants occurs in plastids and involves ADP-glucose as the primary substrate, synthesized by ADP-glucose pyrophosphorylase (AGPase) from glucose-1-phosphate and ATP. Linear α-1,4-glucan chains are extended by soluble starch synthases (e.g., SSI, SSII, SSIII), which add glucose units processively to the non-reducing end of primer chains, while branching enzymes (e.g., BEI, BEII) introduce α-1,6 linkages through glucanotransferase activity, creating the branched architecture of amylopectin with branches every 20–30 residues. Initiation relies on malto-oligosaccharide primers or self-glucosylating protein scaffolds analogous to glycogenin, though plant-specific mechanisms may involve de novo priming by starch synthases using maltose. The pathway is regulated allosterically, with AGPase activated by glucose-6-phosphate to enhance flux toward storage under high carbon availability, and inhibited by inorganic phosphate to prevent synthesis during energy scarcity.

Glycogen and Starch Formation

Glycogen synthesis in animals begins with the initiation step catalyzed by glycogenin, a self-glucosylating protein that attaches glucose residues from UDP-glucose to a tyrosine residue on itself, forming a primer of approximately 8-12 glucose units linked by α-1,4-glycosidic bonds. This primer serves as the foundation for further elongation. Once the primer is established, glycogen synthase extends the chains by adding glucose units from UDP-glucose to the non-reducing ends, forming linear α-1,4-linked glucose polymers according to the reaction:
\text{UDP-glucose} + \text{glycogen}_{(n)} \rightarrow \text{glycogen}_{(n+1)} + \text{UDP}
catalyzed by glycogen synthase. Branching is introduced by glycogen branching enzyme, which transfers a segment of 6-7 glucose residues to create an α-1,6-glycosidic branch point approximately every 8-12 residues along the chain, enhancing the polymer's compactness and solubility.
In , starch biosynthesis follows a parallel but distinct pathway to form the storage , composed of (linear α-1,4-linked glucose) and (branched with α-1,6 links). is primarily synthesized by granule-bound (GBSS), which incorporates glucose from ADP-glucose into linear chains within granules. synthesis involves multiple soluble (SSI, SSII, SSIII, SSIV) that extend α-1,4 chains, similar to , while branching enzymes (SBE) create α-1,6 branches every 20-30 residues, resulting in a less frequent branching pattern than in . Unlike , initiation does not rely on a dedicated self-glucosylating protein like ; instead, it often begins on pre-existing primers or through facilitated by associating with nascent granules. Key differences between glycogen and starch arise from their structural and functional adaptations. 's higher degree of branching confers greater water solubility, allowing it to exist as a cytoplasmic hydrosol in animal cells for rapid mobilization, whereas 's lower branching and crystalline structure renders it insoluble, suitable for long-term storage in plastids. In animals, glycogen synthesis is tightly regulated by hormones; insulin promotes it by dephosphorylating and activating via signaling cascades involving 1. This hormonal control ensures glycogen accumulation in response to nutrient availability, contrasting with starch synthesis in , which is primarily influenced by and osmotic signals rather than animal-like endocrine .

Biosynthesis of Nucleic Acids

DNA Synthesis and Replication

DNA synthesis and replication involves the accurate polymerization of deoxyribonucleoside triphosphates (dNTPs) into a double-stranded DNA molecule, ensuring the faithful duplication of genetic information prior to cell division. This process is semiconservative, meaning each newly synthesized DNA molecule consists of one parental strand and one newly synthesized strand, as experimentally demonstrated by density gradient centrifugation experiments using isotopically labeled DNA in Escherichia coli. The dNTP precursors—dATP, dTTP, dGTP, and dCTP—are generated de novo primarily through the action of ribonucleotide reductase (RNR), which reduces ribonucleoside diphosphates (e.g., CDP + NADPH → dCDP) in a tightly regulated manner to maintain balanced nucleotide pools for replication. These dNTPs are then phosphorylated to their triphosphate forms and serve as substrates for DNA polymerases, with cellular dNTP concentrations typically in the micromolar range to support rapid synthesis rates exceeding 500 nucleotides per second in prokaryotes. In prokaryotes, such as E. coli, the core replicative enzyme is DNA polymerase III (Pol III) holoenzyme, a multisubunit complex discovered in the early , which catalyzes the addition of dNTPs to the 3' hydroxyl end of a growing DNA strand exclusively in the 5' → 3' direction. Replication initiates at the (oriC), where unwinds , creating a replication ; primase (DnaG protein) then synthesizes short RNA primers (10–12 ) complementary to the single-stranded DNA template to provide the necessary 3' OH group for Pol III extension. The leading strand is synthesized continuously toward the , while the lagging strand is produced discontinuously as short (1,000–2,000 each), each requiring a new RNA primer; after extension by Pol III, the RNA primers are removed by DNA polymerase I, gaps are filled with DNA, and DNA ligase seals the nicks by forming phosphodiester bonds between adjacent 3' OH and 5' phosphate groups, as first identified in enzymatic studies from the late 1960s. In eukaryotes, DNA replication is more complex due to larger genomes and multiple origins, but the fundamental mechanism parallels prokaryotes with DNA polymerases δ (Pol δ) and ε (Pol ε) serving as the primary replicative enzymes, identified through genetic and biochemical analyses in the and . Pol ε primarily synthesizes the leading strand, while Pol δ handles both strands on the lagging template, both operating in the 5' → 3' direction with high processivity conferred by PCNA sliding clamps. , part of the Pol α-primase complex, generates RNA-DNA hybrid primers (∼10 RNA + ∼20 DNA nucleotides) for , followed by discontinuous Okazaki fragment (100–200 nucleotides in mammals); ligase I then joins the fragments after primer removal and gap filling, ensuring chromosomal integrity. These polymerases associate with the CMG helicase complex (Cdc45-Mcm2-7-GINS) to coordinate fork progression. The fidelity of DNA synthesis is maintained through intrinsic proofreading mechanisms in replicative polymerases, which include 3' → 5' exonuclease activity that excises mismatched immediately after incorporation, reducing the base substitution error rate from approximately 10^{-5} to 10^{-7} per polymerized. In prokaryotic Pol III and eukaryotic Pol δ/ε, this is coupled to polymerization, with the domain actively selecting against non-Watson-Crick base pairs. Overall replication accuracy reaches about one error per 10^9 incorporated, underscoring the precision of these biosynthetic processes essential for genomic stability.

RNA Transcription and Processing

In prokaryotes, RNA transcription is carried out by a single multisubunit that synthesizes all types of using triphosphates (NTPs: ATP, CTP, GTP, UTP) as precursors, incorporating them in the 5' to 3' direction complementary to the DNA template. occurs at promoter sequences recognized by the polymerase holoenzyme, which includes a (σ) factor that confers promoter specificity; upon binding, the σ factor is released after promoter clearance, allowing at rates up to 50–90 per second. Transcription is often coupled directly to , as there is no membrane separating the processes, and termination happens at specific rho-dependent or rho-independent sites, releasing the without extensive processing. Prokaryotic mRNAs typically require little to no modification, though some tRNAs and rRNAs undergo modifications and for maturation. In eukaryotes, RNA biosynthesis involves the transcription of DNA templates into various RNA species by three distinct nuclear RNA polymerases, each specialized for specific RNA classes. (Pol I) transcribes the majority of (rRNA) precursors, (Pol II) synthesizes precursor (pre-mRNA) and certain small nuclear RNAs (snRNAs), and (Pol III) produces (tRNA), 5S rRNA, and additional snRNAs and small RNAs. These polymerases utilize DNA as a template to direct the synthesis of single-stranded RNA complementary to one DNA strand, ensuring fidelity through base-pairing rules. The process draws directly from cellular pools of nucleoside triphosphates (NTPs)—ATP, CTP, GTP, and UTP—as precursors, which are incorporated sequentially without prior reduction to deoxyribonucleotides, unlike DNA biosynthesis. The transcription mechanism begins with promoter recognition and initiation. For Pol II, which handles most protein-coding genes, general transcription factors such as TFIID bind to promoter elements like the TATA box, recruiting Pol II and forming the pre-initiation complex. Elongation proceeds as Pol II unwinds the DNA helix and adds nucleotides at rates of approximately 20–50 nucleotides per second in vivo, influenced by factors like pausing and elongation regulators. Pol I and Pol III employ analogous but distinct initiation complexes at their respective promoters, with Pol I achieving higher processivity for rRNA production and Pol III facilitating frequent reinitiation for abundant small RNAs. Termination occurs via polymerase release, often coupled to processing signals, ensuring efficient transcript release. Newly synthesized RNA transcripts require extensive processing to achieve maturity and functionality. For pre-mRNA transcribed by Pol II, co-transcriptional modifications include 5' capping with 7-methylguanosine shortly after (within the first 20–30 ), which stabilizes the transcript against exonucleases and facilitates export. Introns are excised through splicing mediated by the , a dynamic complex of small ribonucleoproteins (snRNPs) including U1, , U4, U5, and U6, which recognize splice sites and catalyze two reactions to join exons. The 3' end is formed by cleavage at a polyadenylation signal followed by addition of a poly-A (typically 200–250 adenines) by , enhancing stability and translation efficiency. rRNA processing, primarily from Pol I transcripts, involves multiple cleavages of the large pre-rRNA precursor to yield the mature 18S, 5.8S, and 28S rRNAs, coordinated in the . Extensive modifications, including over 100 sites of 2'-O- and pseudouridylation, are guided by small nucleolar ribonucleoproteins (snoRNPs) containing C/D box snoRNAs for and H/ACA box snoRNAs for pseudouridylation, which stabilize rRNA folding and assembly. tRNA maturation from Pol III transcripts entails endonucleolytic trimming of 5' and 3' extensions; the 5' leader is removed by RNase P, a ribozyme-containing complex, while 3' maturation involves endonuclease cleavage (e.g., by ELAC2 in humans) and addition of the tail by nucleotidyltransferase. tRNAs also undergo numerous base modifications, such as queuosine and wybutosine, to ensure accurate codon recognition and structural integrity. These processing steps integrate with transcription to produce functional RNAs essential for cellular biosynthesis.

Integration with Nucleotide Pools

Nucleotide pools serve as dynamic reservoirs that integrate with the demands of production, ensuring balanced availability of deoxyribonucleoside triphosphates (dNTPs) and ribonucleoside triphosphates (NTPs) for and transcription. During the , particularly in S-phase, dNTP pools expand to support replication fork progression, with levels increasing across all dNTPs. This expansion arises from cell cycle-regulated synthesis, where (RNR) activity peaks to convert ribonucleotides to dNTPs, while catabolic enzymes like SAMHD1 degrade excess dNTPs in to avoid overaccumulation. dNTP mechanisms maintain pool by allosterically inhibiting RNR when dNTP levels are high, preventing excessive diversion of precursors away from synthesis. Regulation of nucleotide pools aligns with cellular proliferation states, with the playing a central role in upregulating biosynthesis pathways during active growth. transcriptionally activates genes encoding enzymes for and synthesis, coordinating nucleotide production with increased demands for in proliferating cells. In contrast, quiescent or differentiated cells rely predominantly on salvage pathways, recycling nucleosides and bases from degraded nucleic acids to sustain low-level maintenance without . This shift minimizes energy expenditure, as salvage enzymes like efficiently replenish pools under nutrient-limited conditions. Imbalances in dNTP pools disrupt genomic integrity, with depletion during replication triggering fork stalling and increased rates. Low dNTP levels impair processivity, leading to error-prone extension and double-strand breaks, as observed in replication stress models where deficiency elevates frequencies. by ATP and dATP fine-tunes RNR to counteract such imbalances; ATP binds the activity site to activate the enzyme and promote balanced dNTP production, while dATP binds the specificity site for inhibition, suppressing overproduction and restoring equilibrium. This dual control ensures dNTP levels match replication needs without excess that could induce misincorporation. Mitochondrial nucleic acid synthesis draws directly from cytosolic nucleotide pools, linking organellar function to cytoplasmic . dNTPs and NTPs are imported into mitochondria via nucleotide transporters, where imbalances in cytosolic pools propagate to alter mtDNA replication and mtRNA transcription fidelity. For instance, cytosolic dNTP depletion reduces mitochondrial import, stalling mtDNA synthesis and accumulating ribonucleotides in mtDNA, which compromises organelle stability. Salvage mechanisms within mitochondria further recycle imported precursors, but reliance on cytosolic supply underscores the integrated nature of these pools across compartments.

Protein Biosynthesis

mRNA Preparation and Genetic Code

In eukaryotic cells, mRNA preparation begins with post-transcriptional modifications of the primary transcript (pre-mRNA) to ensure its stability, proper localization, and functionality for . The 5' ping process adds a 7-methylguanosine to the 5' end shortly after transcription initiation, which protects the mRNA from 5' exonucleases, facilitates splicing and export, and aids in ribosome recruitment during . at the 3' end involves cleavage of the pre-mRNA followed by addition of a poly(A) tail, typically 200-250 long, which enhances mRNA stability by preventing 3' degradation and interacts with poly(A)-binding proteins to promote nuclear export and cytoplasmic longevity; the 3' (UTR) further modulates stability through regulatory elements that influence decay rates. These modifications, coupled with splicing to remove introns, mature the pre-mRNA into export-competent mRNP complexes. Mature mRNA is then exported from the to the through complexes (NPCs), large protein assemblies that selectively transport molecules larger than 40 . Export relies on the complex, which links transcription, processing, and export by recruiting the NXF1-NXT1 heterodimer (also known as TAP-p15) to the mRNP; the 5' cap and poly(A) tail serve as identity marks that prevent retention of unprocessed transcripts and ensure directionality through the NPC via Ran-GTP hydrolysis. Once in the , the mRNA serves as a template for protein , interpreted via the , a triplet system where 64 possible codons (from four bases: A, U, C, G) specify 20 standard plus three stop signals (UAA, UAG, UGA). The code exhibits degeneracy, meaning multiple codons encode the same —for instance, UUU and UUC both specify —primarily in the third position, reducing the impact of mutations and allowing economy with fewer tRNAs. This degeneracy is explained by the wobble hypothesis, proposed by , which posits flexible base-pairing at the third codon position (wobble position) between codon and anticodon on tRNA; for example, (I) in the anticodon can pair with U, C, or A, while G can pair with U or C, enabling one tRNA to recognize multiple codons. Before participating in translation, tRNAs must be charged with their cognate through aminoacylation, catalyzed by specific aminoacyl-tRNA synthetases (ARSs), a family of 20 enzymes (one per ) that ensure fidelity. The reaction occurs in two steps: first, the (AA) reacts with ATP to form aminoacyl-adenylate (AA-AMP) and (), \text{AA} + \text{ATP} \rightleftharpoons \text{AA-AMP} + \text{PP}_\text{i} followed by transfer of the aminoacyl group to the 3'-CCA end of tRNA, releasing AMP: \text{AA-AMP} + \text{tRNA} \rightleftharpoons \text{AA-tRNA} + \text{AMP}. ARSs achieve accuracy through proofreading mechanisms, hydrolyzing mischarged tRNAs at rates up to 1 in 10,000 errors. The genetic code is nearly universal across all life forms, from bacteria to humans, reflecting a common evolutionary origin, but minor exceptions exist, particularly in organelles. In mammalian mitochondria, for example, AUA codes for methionine (Met) instead of isoleucine (Ile), and UGA serves as a tryptophan (Trp) codon rather than a stop, adaptations possibly arising from the reduced tRNA set (22 tRNAs) in mitochondrial genomes to optimize translation efficiency. These variations highlight context-dependent decoding while preserving the core triplet structure.

Translation Initiation and Elongation

Translation initiation begins with the assembly of the ribosomal initiation complex, marking the start of protein synthesis on the mRNA template. In prokaryotes, the small 30S ribosomal subunit first binds to the Shine-Dalgarno sequence upstream of the start codon AUG on the mRNA, facilitated by initiation factors IF1, IF2, and IF3. The initiator methionyl-tRNA (Met-tRNAi) then pairs with the AUG codon in the P-site, with IF2-GTP promoting this binding. Hydrolysis of GTP by IF2 triggers the release of initiation factors and the joining of the large 50S subunit to form the 70S initiation complex, ready for elongation. In eukaryotes, the process is more complex, involving the 40S small subunit scanning from the 5' cap of the mRNA via eIF4 factors to locate the AUG start codon, guided by the Kozak consensus sequence. The ternary complex of eIF2-GTP-Met-tRNAi binds to the 40S subunit, and subsequent GTP hydrolysis by eIF5B enables the 60S large subunit to join, forming the 80S initiation complex.80268-8) During , the sequentially adds to the growing polypeptide chain following the . , delivered to the A-site by EF-Tu (or eEF1A in eukaryotes) in complex with GTP, undergoes codon-anticodon matching; correct pairing accelerates GTP , releasing EF-Tu and allowing accommodation into the A-site. The center of the then catalyzes formation between the peptidyl-tRNA in the P-site and the in the A-site, a reaction that requires no direct energy input as it relies on the activated bond of the . Translocation follows, where the advances one codon along the mRNA, moving the deacylated tRNA to the E-site, peptidyl-tRNA to the , and opening the A-site for the next cycle; this step is powered by EF-G (or ) binding with GTP, whose drives the conformational change in the . The process ensures through kinetic , where initial selection and steps discriminate against near-cognate tRNAs via differential rates of GTP and , enhancing accuracy beyond simple base-pairing selectivity. The cycle in proceeds at a rate of 15-20 per second under optimal conditions, reflecting efficient coordination of tRNA delivery and ribosomal movement. This speed is modulated by factors like tRNA availability and mRNA sequence but maintains high throughput for rapid . Energetically, consumes two GTP molecules per incorporated: one for EF-Tu-mediated delivery and one for EF-G-driven translocation, underscoring the high metabolic cost of protein synthesis.

Translation Termination and Protein Folding

Translation termination occurs when the ribosome encounters one of three stop codons—UAA, UAG, or UGA—in the mRNA, which do not code for amino acids but signal the end of protein synthesis. In prokaryotes, release factor 1 (RF1) recognizes UAA and UAG, while release factor 2 (RF2) recognizes UAA and UGA; both RF1 and RF2 mimic the structure of tRNA to bind the ribosomal A site and catalyze the hydrolysis of the ester bond linking the completed polypeptide to the peptidyl-tRNA in the P site, thereby releasing the nascent chain. Release factor 3 (RF3), a GTPase, then facilitates the dissociation of RF1 or RF2 from the ribosome to recycle them for subsequent termination events, ensuring efficient ribosomal availability. In eukaryotes, the process is analogous but mediated by eukaryotic release factor 1 (eRF1), which recognizes all three stop codons, and eRF3, a GTPase that promotes eRF1 recycling after hydrolysis.80845-4) Following release, the ribosomal subunits dissociate, completing the termination phase that finalizes polypeptide synthesis. Upon release, the nascent polypeptide chain undergoes folding to achieve its functional three-dimensional structure, a process governed by Anfinsen's principle that the native conformation is thermodynamically favored under physiological conditions.80928-9) Folding often proceeds spontaneously through hydrophobic collapse, where nonpolar residues cluster inward to minimize exposure to the aqueous environment, stabilized by interactions such as hydrogen bonds, van der Waals forces, and ionic bridges. In oxidizing environments like the eukaryotic , disulfide bonds form between residues to further lock the structure, catalyzed by enzymes such as (PDI).80928-9) However, many proteins require assistance from molecular chaperones to avoid kinetic traps and misfolding; heat shock protein 70 () binds hydrophobic regions of unfolded chains in an ATP-dependent manner to prevent aggregation and promote correct folding, while chaperonins like (in prokaryotes) or TRiC (in eukaryotes) provide an enclosed cavity for iterative folding cycles. Folding frequently begins co-translationally as the polypeptide emerges from the , influenced by targeting signals that direct the chain to specific cellular compartments. Signal peptides, short N-terminal sequences (typically 15-30 ), mediate co-translational translocation to the via the (SRP), where folding occurs in the assisted by ER chaperones like BiP (an homolog). For mitochondrial proteins, analogous presequences guide import, often coupling with co-translational folding near the outer membrane to ensure proper insertion into the organelle. This vectorial from the ribosome imposes a folding , with domains folding sequentially to maintain solubility and functionality. Errors in folding can lead to misfolded proteins that aggregate into toxic structures, as seen in prion diseases where the cellular prion protein (PrP^C) misfolds into a beta-sheet-rich isoform (PrP^Sc) that templates further aggregation, propagating neuropathology.80232-9) Such aggregates, including , overwhelm cellular and contribute to diseases like Alzheimer's and Parkinson's, highlighting the critical role of chaperones in preventing these outcomes.

Regulation of Biosynthetic Processes

Allosteric and Feedback Mechanisms

Allosteric and are critical intracellular regulatory strategies in biosynthesis that maintain metabolic by modulating activity in response to the concentrations of pathway intermediates and end products. inhibition occurs when the end product of a biosynthetic pathway binds to and inhibits an early in the same pathway, preventing overaccumulation of the product. A classic example is the inhibition of deaminase, the first committed in the aspartate family pathway, by , which ensures balanced production of , , , and . Allosteric regulation involves the binding of effectors at sites distinct from the , inducing conformational changes in the that alter its activity. This can lead to either or inhibition and often exhibits , where of one effector molecule facilitates of subsequent ones. is quantitatively described by the Hill equation, which models the fractional saturation Y of sites as a function of concentration [L], K_d, and Hill coefficient n (indicating cooperativity degree): Y = \frac{[L]^n}{K_d + [L]^n} This equation highlights how allosteric enzymes can respond sharply to small changes in effector levels, providing sensitive control over biosynthetic flux. In biosynthesis, () exerts feedback inhibition on phosphoribosyl pyrophosphate amidotransferase, the first committed step, by binding allosterically to reduce enzyme affinity for its substrate and prevent excess purine nucleotide accumulation. Similarly, in biosynthesis, levels indirectly regulate the pathway through reciprocal control, as elevated inhibits , suppressing oxidation and favoring net synthesis, while long-chain acyl-CoA products like palmitoyl-CoA allosterically inhibit to curtail further production. Feed-forward activation complements these inhibitory mechanisms by stimulating pathway enzymes when substrates are abundant. For instance, citrate, an early intermediate derived from , allosterically activates in by promoting enzyme polymerization and increasing catalytic efficiency, thereby accelerating malonyl-CoA formation when energy is plentiful. These molecular controls, observed across , , and pathways, ensure efficient resource allocation without relying on transcriptional changes.

Hormonal and Genetic Regulation

Hormonal regulation plays a crucial role in modulating biosynthetic pathways by integrating extracellular signals with changes. In mammals, insulin promotes the biosynthesis of glucose-derived metabolites by activating sterol regulatory element-binding protein 1c (SREBP1c), which transcriptionally upregulates genes involved in and while repressing gluconeogenic enzymes such as (PEPCK) and glucose-6-phosphatase (G6Pase). This dual action ensures efficient nutrient utilization during fed states, with SREBP1c mediating insulin's effects through the liver X receptor (LXR) pathway to enhance from glycolytic intermediates. Conversely, , released during , represses glycolytic gene expression, notably by suppressing (PK) transcription and accelerating its mRNA degradation, thereby diverting carbon flux toward and maintaining blood glucose levels. Genetic regulation of biosynthesis occurs through precise transcriptional controls that respond to cellular needs, often via operons in prokaryotes and enhancers in eukaryotes. In , the tryptophan () operon exemplifies , a where high levels promote the formation of a terminator hairpin in the leader RNA, halting transcription of genes encoding anthranilate synthase and subsequent enzymes in the biosynthetic pathway; this fine-tunes production without requiring a protein. In eukaryotes, hypoxia-inducible factor 1 (HIF-1) acts as a transcriptional enhancer complex that binds to hypoxia response elements (HREs) in promoters of glycolytic genes like enolase 1 (ENO1) and (LDHA), upregulating their expression under low-oxygen conditions to sustain energy production via . Epigenetic modifications further govern access to biosynthetic genes by altering structure. acetylation, catalyzed by acetyltransferases such as p300/CBP, neutralizes positive charges on residues (e.g., H3K9ac, H3K27ac), loosening packing and facilitating binding to promoters of biosynthetic loci, including those for secondary metabolites like in . This modification is dynamically balanced by deacetylases and links metabolic states—such as availability—to gene activation, ensuring biosynthetic pathways align with environmental cues. Steroid hormones provide a notable example of self-regulation in biosynthesis, as they derive from and feedback to control precursor production. In the , glucocorticoids like inhibit their own synthesis by repressing (CRH) and (ACTH) expression, which in turn reduces mobilization and the activity of steroidogenic enzymes such as (CYP11A1). This loop prevents overproduction and maintains hormonal , integrating with broader biosynthetic regulation via SREBP2.

Compartmentalization in Cells

Cellular compartmentalization organizes biosynthetic reactions within membrane-bound organelles in eukaryotic cells, allowing for specialized microenvironments that optimize enzyme activity, substrate availability, and protection from incompatible intermediates. This spatial segregation enhances metabolic efficiency and coordination, as opposed to prokaryotic cells, which lack such organelles and perform most biosynthetic processes in the cytoplasm or at the plasma membrane. In prokaryotes like bacteria, the nucleoid region loosely organizes DNA for transcription, but lipid and protein synthesis occurs without discrete compartments, relying on diffusion and membrane association for localization. In the , proceeds via the complex, utilizing and NADPH to produce palmitate, while glycogen biosynthesis assembles glucose units into storage granules through . The () serves as the primary site for phospholipid synthesis, where enzymes like choline phosphotransferase generate and from diacylglycerol precursors, and the rough facilitates co-translational insertion and folding of and secretory proteins by ribosomes. host key steps in biosynthesis, including the final ferrochelatase-catalyzed insertion of iron into , and initiate production through the () in the inner . initiate biosynthesis, a class of phospholipids, via dihydroxyacetone phosphate acyltransferase and alkyl-dihydroxyacetone phosphate synthase, which establish the linkage before completion in the . Membrane domains, such as lipid rafts enriched in cholesterol and sphingolipids, facilitate the localized assembly of biosynthetic signaling complexes at the plasma membrane, though primary lipid synthesis occurs elsewhere. Biosynthetic products are distributed via vesicular transport, where COPII-coated vesicles bud from the ER to deliver newly synthesized lipids and proteins to the Golgi apparatus for further modification and sorting, ensuring targeted delivery to cellular destinations. Mitochondrial DNA replication, a biosynthetic process, occurs within the organelle matrix, supporting organelle maintenance.

Disorders of Biosynthesis

Genetic Defects in Pathways

Genetic defects in biosynthetic pathways, also known as , are inherited disorders that impair the function of enzymes or transporters involved in the synthesis of essential biomolecules such as , , and . These defects typically result from mutations in genes encoding pathway components, leading to the accumulation of toxic precursors, deficiency of downstream products, or both, which can cause severe physiological disruptions. Most such disorders follow an autosomal recessive inheritance pattern, requiring biallelic mutations for clinical manifestation, though rare autosomal dominant or X-linked forms exist. The molecular mechanisms underlying these defects often involve loss-of-function mutations that abolish or diminish enzymatic activity. Common mutation types include missense variants that alter the enzyme's , impair binding, or destabilize , as well as nonsense mutations that introduce premature stop codons, resulting in truncated, nonfunctional proteins. For instance, in phenylketonuria (PKU), mutations in the PAH gene encoding lead to reduced conversion of to , causing hyperphenylalaninemia and toxic buildup of and its metabolites, which can result in if untreated. Similarly, in adenosine deaminase (ADA) deficiency, a cause of (SCID), mutations in the ADA gene disrupt the purine salvage pathway, leading to accumulation of and its toxic derivatives that selectively impair development and function. Another example is hereditary orotic aciduria, stemming from biallelic mutations in the UMPS gene, which encodes uridine-5'-monophosphate synthase; this bifunctional enzyme catalyzes the final steps of pyrimidine biosynthesis, and its deficiency causes orotic acid accumulation and pyrimidine shortfall, manifesting as and developmental delays. Diagnosis of these defects relies on detecting biochemical imbalances, often through programs that employ to identify elevated substrate levels in blood spots. For PKU, screening measures concentrations above 2 mg/dL, enabling early detection in over 99% of cases in screened populations. In ADA-SCID, deoxyadenosine metabolites are assayed, while is confirmed by urinary excretion exceeding 100 mmol/mol . via sequencing of the affected gene confirms the diagnosis and identifies specific variants, such as the common R408W in PAH for PKU, which accounts for up to 30% of alleles in European populations. These pathways, including hydroxylation, recycling, and synthesis, are briefly referenced here as they underpin the defects discussed.

Nutritional Deficiencies and Impacts

Nutritional deficiencies in essential vitamins and minerals disrupt key biosynthetic pathways in humans, leading to impaired production of critical biomolecules such as , , and . These deficiencies arise from inadequate dietary intake of precursors that serve as cofactors or substrates in enzymatic reactions, compromising cellular metabolism and tissue integrity. For instance, vitamins like B6 (in the form of pyridoxal 5'-phosphate, or ) are vital for reactions in biosynthesis, where acts as a coenzyme for aminotransferases that facilitate the interconversion of . Deficiency in impairs this process and hinders the conversion of to , resulting in pellagra-like symptoms including , , and neurological disturbances. Minerals such as iron are indispensable for heme biosynthesis, where iron is incorporated into protoporphyrin IX by ferrochelatase to form the heme prosthetic group essential for hemoglobin and cytochromes. Iron deficiency limits this final step, reducing heme production and causing microcytic hypochromic anemia, as the body cannot adequately synthesize functional red blood cells. Similarly, vitamin C (ascorbic acid) is a cofactor for prolyl and lysyl hydroxylases in collagen biosynthesis, enabling the hydroxylation of proline and lysine residues that stabilize the collagen triple helix. In scurvy, vitamin C deficiency halts this hydroxylation, leading to unstable collagen, weakened connective tissues, bleeding gums, and poor wound healing. Thiamine (vitamin B1) deficiency, as seen in beriberi, disrupts the pyruvate dehydrogenase complex (PDH), preventing the conversion of pyruvate to acetyl-CoA and impairing the entry of carbohydrates into the citric acid cycle and fatty acid biosynthesis, which manifests as cardiovascular and neurological symptoms. Protein-energy malnutrition, particularly , exemplifies the consequences of insufficient intake of essential , which humans cannot synthesize and must obtain from diet (as detailed in the section on Biosynthesis of Amino Acids). Low availability of these , such as sulfur-containing ones like and , limits hepatic protein synthesis, resulting in and subsequent osmotic imbalance that causes peripheral and facial . This impaired biosynthesis of plasma proteins like reduces , exacerbating fluid retention despite adequate caloric intake from carbohydrates. Globally, deficiencies contribute to widespread stunting in children, affecting linear growth and through disrupted biosynthesis of hormones, enzymes, and structural proteins. In 2024, approximately 150 million children under five were stunted due to chronic deficiencies in iron, , , and others, leading to irreversible impairments in height, immune function, and metabolic pathways. These gaps hinder overall biosynthetic capacity, perpetuating cycles of undernutrition in vulnerable populations.

Therapeutic Interventions

Therapeutic interventions for biosynthetic disorders aim to restore deficient enzymatic activities, modulate overactive pathways, or compensate for metabolic imbalances through targeted pharmacological, dietary, or genetic approaches. These strategies address disruptions in key biosynthetic processes, such as , , , and , often linked to genetic defects or nutritional shortfalls. Enzyme replacement therapy (ERT) provides exogenous enzymes to compensate for deficiencies in lysosomal storage disorders, exemplified by , where mutations in the GBA gene impair activity, leading to accumulation. Intravenous administration of recombinant , such as imiglucerase or velaglucerase alfa, targets macrophages via mannose-6-phosphate receptors, reducing , improving hematological parameters, and enhancing in type 1 Gaucher patients. Long-term ERT has demonstrated sustained clinical benefits, with home-based infusions further improving patient adherence and resource utilization. Nutritional supplementation corrects deficiencies in cofactor-dependent biosynthetic pathways, such as for synthesis in . , or vitamin B9, is essential for thymidylate and synthesis via one-carbon ; its deficiency impairs in erythroid precursors, causing . Oral folic acid supplementation (1-5 mg daily) rapidly reverses hematological abnormalities by replenishing tetrahydrofolate pools, though concurrent assessment is critical to avoid masking . Pharmacological inhibition of rate-limiting enzymes modulates lipid biosynthesis in . Statins, such as and simvastatin, competitively inhibit , the key enzyme in the for synthesis, reducing hepatic low-density lipoprotein production and increasing receptor-mediated clearance. This lowers serum levels by up to 50% and reduces cardiovascular risk, with prolonged inhibition enhancing efficacy over shorter-acting agents. Gene therapy addresses monogenic biosynthetic defects by correcting underlying mutations, as in adenosine deaminase (ADA) deficiency causing severe combined immunodeficiency (SCID). Autologous hematopoietic stem cell gene therapy using retroviral or lentiviral vectors to insert functional ADA genes has shown immune recovery in up to 95% of ADA-SCID patients, with durable T-cell function persisting beyond two years post-infusion. Emerging preclinical approaches, such as CRISPR/Cas9 editing to correct specific mutations like Q3X, aim to restore enzyme function without viral integration risks but are not yet in clinical use. Dietary interventions restrict substrate accumulation in amino acid biosynthetic disorders like (PKU), caused by deficiency. A lifelong low-phenylalanine diet, typically limiting intake to 20-40 mg/kg/day via phenylalanine-free formula and controlled natural protein sources, prevents neurotoxic buildup and supports normal when initiated neonatally. Sapropterin dihydrochloride may adjunctively enhance residual activity in responsive patients, allowing dietary liberalization. For overproduction in , inhibitors like block the final steps of biosynthesis. and its metabolite oxypurinol competitively inhibit , reducing hypoxanthine and conversion to , thereby lowering serum urate levels by 60-70% and preventing formation. This therapy is particularly effective in underexcretors, with dose adjustments based on renal function to minimize risks.

References

  1. [1]
    The Biosynthesis of Cell Constituents - NCBI - NIH
    One major biosynthetic pathway, the synthesis of carbohydrates from CO2 and H2O during the dark reactions of photosynthesis, was discussed in the preceding ...
  2. [2]
  3. [3]
    5.12B: Biosynthesis and Energy - Biology LibreTexts
    Nov 23, 2024 · Biosynthesis and Energy. Biosynthesis in living organisms is a process in which substrates are converted to more complex products.
  4. [4]
    22.4: Biosynthesis and Degradation of Nucleotides
    Sep 7, 2025 · Mammals have two pathways for purine synthesis: a de novo pathway and a salvage pathway, which allows for the recycling of nucleotide bases. The ...
  5. [5]
    Nucleotide Synthesis via Salvage Pathway - Wiley Online Library
    Dec 9, 2014 · The pathways of de novo synthesis are the same in animals and microorganisms. Salvage pathways are considerably more energy-efficient than de ...
  6. [6]
    Vitalism and synthesis of urea. From Friedrich Wöhler to Hans A. Krebs
    In 1828, Friedrich Wöhler, a German physician and chemist by training, published a paper that describes the formation of urea, known since 1773 to be a ...
  7. [7]
    Beyond energy and growth: the role of metabolism in developmental ...
    Oct 26, 2023 · During proliferation, metabolism is geared toward biosynthesis to double the amount of lipids, nucleotides and proteins required to produce a ...
  8. [8]
    Synthetic biology strategies for microbial biosynthesis of plant ...
    May 13, 2019 · In this review, we describe recent developments in metabolic engineering at the level of host, pathway, and enzyme, and discuss how the field is approaching ...
  9. [9]
    Metabolic Engineering and Synthetic Biology - NIH
    Mar 4, 2019 · Metabolic engineering seeks for the optimization of cellular processes, endemic to a specific organism, to produce a compound of interest from a substrate.
  10. [10]
    ATP cycle and reaction coupling | Energy (article) - Khan Academy
    While ATP hydrolysis is exergonic (releases energy), coupling it to an unfavorable reaction through an unstable intermediate requires energy to form that ...
  11. [11]
    The Essence of ATP Coupling - PMC - NIH
    The traditional explanation of ATP coupling is based on the raising of the equilibrium constants of the biochemical reactions.
  12. [12]
    Pathway Thermodynamics Highlights Kinetic Obstacles in Central ...
    Feb 20, 2014 · It is given by ΔrG′ = ΔrG′o+RT·ln(Q), where Q is the reaction quotient (also known as the mass action ratio). Because of its more intuitive ...
  13. [13]
    Enzymes: principles and biotechnological applications - PMC
    This chapter covers the basic principles of enzymology, such as classification, structure, kinetics and inhibition, and also provides an overview of industrial ...
  14. [14]
    NAD(H) and NADP(H) Redox Couples and Cellular Energy ... - NIH
    The NAD+/NADH redox couple is known as a regulator of cellular energy metabolism, that is, of glycolysis and mitochondrial oxidative phosphorylation. By ...
  15. [15]
    Enzymatic Cascade Reactions in Biosynthesis - PMC
    This Review highlights four types of enzymatic cascade strategies, mediated by nucleophilic, electrophilic, pericyclic, and radical-based reactions.
  16. [16]
    STRUCTURAL ENZYMOLOGY OF POLYKETIDE SYNTHASES - PMC
    This chapter describes structural and associated enzymological studies of polyketide synthases, including isolated single domains and multidomain fragments.
  17. [17]
    Role of Induced Fit in Enzyme Specificity: A Molecular Forward ... - NIH
    The slow release of a correct substrate commits it to the forward reaction so that specificity is determined solely by the rate of substrate binding, including ...
  18. [18]
    General base-general acid catalysis by terpenoid cyclases - Nature
    Apr 13, 2016 · Terpenoid cyclases catalyze the most complex reactions in biology, in that more than half of the substrate carbon atoms often undergo changes in bonding.
  19. [19]
    Research - Boston University
    Aldolase is involved in glycolysis, gluconeogenesis, and fructose metabolism. The catalytic mechanism of a class I aldolase utilizes a Schiff-base intermediate.
  20. [20]
    Harnessing enzyme cofactors and plant metabolism: an essential ...
    Mar 2, 2023 · Coenzyme A. CoA is an essential metabolic cofactor that is synthesized from vitamin B5 (pantothenate), ATP, and cysteine.Coenzyme A · Vitamin Coenzymes... · Thiamin (vitamin B1)
  21. [21]
    NADPH—The Forgotten Reducing Equivalent - PMC - NIH
    NADPH is the major reducing equivalent driving de novo synthesis of fatty acids, cholesterol, amino acids, and nucleotides.
  22. [22]
    Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage
    N2 fixation occurs in three different ways: (i) through geochemical processes such as lightning, (ii) biologically through the action of the enzyme, nitrogenase ...
  23. [23]
    The Role of Glutamine Synthetase (GS) and Glutamate Synthase ...
    The glutamine synthetase (GS)/glutamate synthase (GOGAT) cycle represents a crucial metabolic step of N assimilation, regulating crop yield.
  24. [24]
    Glutamine Synthetase-Glutamate Synthase Pathway and Glutamate ...
    The first step of [15N]ammonium entry is into the Gln-amide group by GS. Then GOGAT transfers the 15N amide group of Gln to the 2-oxoglutarate position yielding ...
  25. [25]
    Recent Advances in Metabolic Pathways of Sulfate Reduction ... - NIH
    Mar 12, 2020 · In this metabolic pathway, sulfate is used as an electron acceptor and is reduced to hydrogen sulfide, which is subsequently released into the ...
  26. [26]
    The Role of Compartment-Specific Cysteine Synthesis for Sulfur ...
    The amino acid precursor used for cysteine synthesis, OAS, is provided by serine acetyltransferase (SAT). SAT and OAS-TL can interact to form the cysteine ...
  27. [27]
    An Evolutionary Perspective on Amino Acids - Nature
    In humans, reduced nitrogen enters the physiological system in dietary sources containing amino acids. All organisms contain the enzymes glutamate dehydrogenase ...
  28. [28]
    The Role of Minerals in the Optimal Functioning of the Immune System
    This review reports on the minerals that are essential for the immune system's function and inflammation regulation.
  29. [29]
    Magnesium: Biochemistry, Nutrition, Detection, and Social Impact of ...
    Many of the glycolytic enzymes are sensitive to magnesium, whose principal function is to facilitate the transfer of high energy phosphate. Thus, hexokinase, ...
  30. [30]
    Nucleotide Metabolism - PMC - NIH
    The first reaction is catalyzed by ribose 5-phosphate pyrophosphokinase to generate PRPP from ribose 5-phosphate and the two high-energy phosphates of ATP.
  31. [31]
    Review Biosynthesis of heme in mammals - ScienceDirect.com
    Ferrochelatase inserts iron into protoporphyrin IX to form heme which is incorporated into hemoglobin and cytochromes, the dominant hemoproteins in mammals.
  32. [32]
    Plant Carbonic Anhydrases: Structures, Locations, Evolution, and ...
    Sep 16, 2016 · Carbonic anhydrases (CAs) are zinc metalloenzymes that catalyze the interconversion of CO2 and HCO3− and are ubiquitous in nature.
  33. [33]
    Biosynthesis of the Iron-Molybdenum Cofactor of Nitrogenase - PMC
    Here, we briefly review the recent progress in this research area, highlighting our work on the molybdenum nitrogenase from Azotobacter vinelandii that ...
  34. [34]
    Magnesium Deficiency Triggers SGR–Mediated Chlorophyll ...
    In conclusion, the experiments herein demonstrate that Mg deficiency initiates OsSGR-mediated chlorophyll degradation in mid-aged leaves of rice. Upregulation ...Missing: impact | Show results with:impact
  35. [35]
    Amino acid structure and classifications (article) - Khan Academy
    This article reviews the 20 proteinogenic amino acids, their structures, symbols, and abbreviations, and related information important for the MCAT.
  36. [36]
    Emerging knowledge of regulatory roles of d-amino acids in bacteria
    For example, l-amino acids are the predominant building blocks of proteins; d-amino acids cannot be incorporated into proteins via ribosomal synthesis.Missing: paper | Show results with:paper
  37. [37]
    New Insights Into the Mechanisms and Biological Roles of D-Amino ...
    The presence of D-amino acids in the peptide moieties of the PG of bacteria makes the cell wall invulnerable to most proteases designed to cleave between L- ...Abstract · Introduction · Role of D-Amino Acids in... · Bacteria-Host Interactions...
  38. [38]
    Biochemistry, Essential Amino Acids - StatPearls - NCBI Bookshelf
    Apr 30, 2024 · Each amino acid has an α-carboxyl group, a primary α-amino group, and a side chain called the R group (see Image. Amino Acid Generic Structure).
  39. [39]
    Synthesis and Degradation of Amino Acids - Basicmedical Key
    Aug 7, 2022 · Four amino acids (serine, glycine, cysteine, and alanine) are produced from glucose through components of the glycolytic pathway. TCA cycle ...A. Serine · B. Glycine · C. Cysteine
  40. [40]
    Aspartic Acid in Health and Disease - PMC - NIH
    Most L-Asp is synthesized by mitochondrial aspartate aminotransferase from oxaloacetate and glutamate acquired by glutamine deamidation, particularly in the ...
  41. [41]
    Aspartate-Derived Amino Acid Biosynthesis in Arabidopsis thaliana
    Jun 10, 2009 · The aspartate-derived amino acid pathway in plants leads to the biosynthesis of lysine, methionine, threonine, and isoleucine. These four amino ...
  42. [42]
    The aspartate-family pathway of plants - PubMed Central - NIH
    The Asp-family pathway of plants has been extensively studied in respect to the regulation of the biosynthesis of its amino acids Lys, Thr, Ile, Met and Gly.
  43. [43]
    T cell activation depends on extracellular alanine - PubMed Central
    Sep 17, 2020 · Alanine can be synthesized from glucose through pyruvate transamination by the enzyme alanine aminotransferase, also known as glutamate-pyruvate ...
  44. [44]
    The shikimate pathway: gateway to metabolic diversity - PMC
    The shikimate pathway is the metabolic process responsible for the biosynthesis of the aromatic amino acids phenylalanine, tyrosine, and tryptophan.
  45. [45]
    The shikimate pathway and aromatic amino Acid biosynthesis in plants
    This review highlights recent molecular identification of enzymes of the pathway and summarizes the pathway organization and the transcriptional/ ...
  46. [46]
    Mechanism of the feedback-inhibition resistance in aspartate kinase ...
    Dec 22, 2020 · In microorganisms and plants, aspartate kinase (AK) is the initial committed enzyme of the biosynthesis of the aspartate acid family amino acids
  47. [47]
    Insight into de-regulation of amino acid feedback inhibition
    Aug 23, 2023 · Aspartate family includes lysine, threonine and methionine that regulate their own synthesis by end product feedback inhibition. Lysine and ...
  48. [48]
    Synthesis of Essential Amino Acids - News-Medical
    Aug 23, 2018 · The essential amino acids are not synthesized in humans but are synthesized in either plants or bacteria. The 9 essential amino acids are ...
  49. [49]
    Bacterial Branched-Chain Amino Acid Biosynthesis - NIH
    The first committed step in the biosynthesis of l-isoleucine is catalyzed by the ilvA-encoded threonine dehydratase/deaminase (EC 4.3.1.19, TD). This enzyme ...
  50. [50]
    Essential Amino Acid - an overview | ScienceDirect Topics
    These 'essential' amino acids are produced only in microorganisms and plants, with their biosynthetic pathways lost early in animal evolution.
  51. [51]
    Retention and Loss of Amino Acid Biosynthetic Pathways Based on ...
    In addition to losing all of the genes in pathways for the human-essential amino acids, they have lost genes required for serine and arginine biosynthesis.
  52. [52]
    Resurrecting essential amino acid biosynthesis in mammalian cells
    Sep 27, 2022 · Branching from core metabolism, the nine EAA biosynthesis pathways missing from metazoans involve over 40 genes (Figure 1B, Supplementary files ...
  53. [53]
    Purine and Pyrimidine Metabolism
    Apr 12, 1997 · The final atom of the purine ring, carbon 2, is supplied by 10-Formyl tetrahydrofolate. Ring closure produces the purine nucleotide, IMP. Note ...Missing: origins | Show results with:origins
  54. [54]
    De novo and Salvage Purine Synthesis Pathways Across Tissues ...
    The de novo purine synthesis pathway is a metabolically costly process (6 ATP molecules per molecule of purine synthesized) that involves 10 catalytic steps to ...
  55. [55]
  56. [56]
  57. [57]
  58. [58]
  59. [59]
    Pentose Phosphate Pathway (PPP), Purine and Pyrimidine ...
    There are two primary enzymes involved in the salvage pathway: adenine phosphoribosyltransferase (APRT) and xanthine-guanine phosphoribosyltransferase (HGPRT) ( ...
  60. [60]
    Inborn Errors of Purine Salvage and Catabolism - PMC
    Hypoxanthine-Guanine Phosphoribosyltransferase. HPRT catalyzes the salvage of the purine bases, hypoxanthine and guanine, converting them into their ...
  61. [61]
    Lesch-Nyhan Syndrome - StatPearls - NCBI Bookshelf - NIH
    Apr 24, 2023 · Lesch Nyhan syndrome is an inborn disorder caused by a deficiency of hypoxanthine-guanine phosphoribosyltransferase (HPRT) enzyme, an enzyme of purine salvage ...Continuing Education Activity · Pathophysiology · History and Physical · Evaluation
  62. [62]
    Lesch Nyhan Syndrome - Symptoms, Causes, Treatment | NORD
    Lesch-Nyhan syndrome is a rare inborn error of purine metabolism characterized by the absence or deficiency of the activity of the enzyme hypoxanthine-guanine ...
  63. [63]
    Role of Pyrimidine Salvage Pathway in the Maintenance ... - PubMed
    Thymidine kinase (TK) is the first enzyme in the salvage pathway to recycle thymidine nucleosides as it phosphorylates thymidine to yield thymidine ...
  64. [64]
    Pyrimidine Salvage Enzymes Are Essential for De Novo ... - NIH
    Nov 7, 2016 · Thymidine kinase (TK) catalyzes the formation of dTMP and dUMP and is one of several salvage enzymes that appear redundant to the de novo ...
  65. [65]
    Nucleotides: Biosynthesis and Catabolism
    IMP represents a branch point for purine biosynthesis, because it can be converted into either AMP or GMP through two distinct reaction pathways. The pathway ...
  66. [66]
    Nucleotide Salvage - an overview | ScienceDirect Topics
    Nucleotide salvage pathways recover bases and nucleosides, from RNA and DNA degradation or from exogenous sources, to convert them back to nucleotides.
  67. [67]
    Hypoxanthine-guanine phosophoribosyltransferase (HPRT) deficiency
    Deficiency of hypoxanthine-guanine phosphoribosyltransferase (HPRT) activity is an inborn error of purine metabolism associated with uric acid overproduction.
  68. [68]
    Regulation of mammalian nucleotide metabolism and biosynthesis
    Substrate availability in eukaryotic cells can be an important means of controlling flux through pathways, either by physical compartmentation (important in ...Genes And Enzymes Involved... · Nucleic Acid Synthesis... · Stable Isotope Tracing Of...
  69. [69]
    Membrane phospholipid synthesis and endoplasmic reticulum function
    This review is focused on membrane phospholipid biogenesis in mammalian cells with a particular emphasis on the role played by the endoplasmic reticulum (ER).
  70. [70]
    The Kennedy pathway—De novo synthesis of ... - IUBMB Journal
    May 6, 2010 · This is the last step in a biosynthetic pathway known as the Kennedy pathway, so named after Eugene Kennedy who elucidated it over 50years ago.INTRODUCTION · CHOLINE AND... · THE CHOLINE AND... · The CTP...
  71. [71]
    The Kennedy pathway--De novo synthesis of ... - PubMed
    This is the last step in a biosynthetic pathway known as the Kennedy pathway, so named after Eugene Kennedy who elucidated it over 50 years ago.<|control11|><|separator|>
  72. [72]
  73. [73]
    Sphingolipid Metabolic Pathway: An Overview of Major Roles ...
    This review discusses the role of the sphingolipid metabolic pathway in diabetes, Alzheimer's disease, and hepatocellular carcinoma2. Sphingolipid Metabolism... · 2.2. Ceramide Homeostasis · 3. Sphingolipid Metabolism...Missing: seminal | Show results with:seminal
  74. [74]
    Biosynthesis of Cholesterol and Other Sterols | Chemical Reviews
    Sep 8, 2011 · The choice of pathway is determined by the stage at which the double bond at C24 in the sterol side chain is reduced. If C24 double bond ...Introduction · Recent Advances in Sterol... · Sterol Enzyme Action · Biography<|control11|><|separator|>
  75. [75]
    Biochemistry, Cholesterol - StatPearls - NCBI Bookshelf - NIH
    ... biosynthesis involves condensing 2 molecules of acetyl-CoA to form acetoacetyl-CoA. Next, a cytosolic enzyme HMG-CoA synthase adds a third molecule of acetyl- ...
  76. [76]
    Unraveling the Regulation of Hepatic Gluconeogenesis - PMC
    ... pyruvate kinase (PK) (7), which are overcome by four unique gluconeogenic enzymes including PC, PEPCK, FBPase, and G6Pase (Figure 1). Given their importance ...
  77. [77]
  78. [78]
    [PDF] Metabolism Lecture 5 — PENTOSE PHOSPHATE PATHWAY
    Oxidative Phase: Glucose-6-P → Ribose-5-P. Glucose 6-phosphate dehydrogenase. First enzymatic step in oxidative phase, converting NADP. + to NADPH. Glucose 6 ...
  79. [79]
    13.4: Pentose Phosphate Pathway of Glucose Oxidation
    Aug 23, 2025 · When ribose 5-phosphate is in high demand, its cellular concentration will decrease, pulling the ribose 5-phosphate isomerase reaction forward ...
  80. [80]
    Presence of glutamine:fructose-6-phosphate amidotransferase for ...
    Our results show the presence of glutamine:fructose-6-phosphate amidotransferase for de novo glucosamine synthesis in endothelial cells.
  81. [81]
    Structural analysis of human glutamine:fructose‐6‐phosphate ...
    Dec 6, 2008 · The first step in the de novo biosynthesis of hexosamines, the formation of glucosamine-6-phosphate (GlcN6P) from fructose-6-phosphate ...
  82. [82]
    MetaCyc GDP-mannose biosynthesis - BioCyc
    GDP-α-D-mannose is the critical metabolite of the mannose activation pathway for glycoconjugate synthesis in eukaryotes, as it is utilized directly or ...
  83. [83]
    Genetic evidence for the role of GDP-mannose in plant ascorbic acid ...
    This enzyme provides GDP-mannose, which is used for cell wall carbohydrate biosynthesis and protein glycosylation as well as for AsA biosynthesis. In addition ...
  84. [84]
    Distinct plastid fructose bisphosphate aldolases function in ...
    The fixation of atmospheric CO2 into organic compounds by plants occurs via the Calvin–Benson cycle (Calvin, 1962). The cycle is organized into three phases: a ...
  85. [85]
    The importance of the photosynthetic Gibbs effect in the elucidation ...
    Observations by Calvin that the pool sizes of ribulose 1,5-bisphosphate (RuBP) and PGA changed in a reciprocal fashion when algae were subjected to changes in ...
  86. [86]
    Cellulose Biosynthesis: Current Views and Evolving Concepts - PMC
    Aims To outline the current state of knowledge and discuss the evolution of various viewpoints put forth to explain the mechanism of cellulose biosynthesis.
  87. [87]
  88. [88]
  89. [89]
    Structural basis for directional chitin biosynthesis - PMC
    Sep 21, 2022 · The key reactions of chitin biosynthesis are catalysed by chitin synthase, a membrane-integrated glycosyltransferase that transfers GlcNAc from ...
  90. [90]
  91. [91]
    Formation of starch in plant cells - PMC - PubMed Central - NIH
    The synthesis of ADPglucose by AGPase is often regarded as the “committed step” of starch synthesis. There is appreciable evidence that the step is regulated ...
  92. [92]
  93. [93]
  94. [94]
    Catalytic activities of glycogenin additional to autocatalytic self ...
    Jun 23, 1995 · Glycogenin is the autocatalytic, self-glucosylating protein that initiates glycogen synthesis in muscle and other tissues.
  95. [95]
    Biochemistry, Glycogen - StatPearls - NCBI Bookshelf - NIH
    Glycogen branching is essential because it allows for increased water solubility and several sites to break it down; this allows for easy and quick glycogen ...
  96. [96]
    Structural basis of glycogen branching enzyme deficiency and ...
    Glycogen branching enzyme 1 (GBE1) plays an essential role in glycogen biosynthesis by generating α-1,6-glucosidic branches from α-1,4-linked glucose chains.
  97. [97]
    Competition between Granule Bound Starch Synthase and ... - NIH
    Dec 23, 2019 · Starch branching enzymes (SBE) and granule-bound starch synthase (GBSS) are two important enzymes for starch biosynthesis.Fitting Parameters · Amylopectin Clds · Amylose Clds
  98. [98]
    Starch biosynthesis and crop bioengineering - ScienceDirect.com
    For amylopectin biosynthesis, a number of starch synthase isoforms are needed as they transfer glucosyl moieties to form different chain lengths, required for ...Review Article · 2. Starch Biosynthesis In... · 3. Plant Genetic Engineering...
  99. [99]
    Glycogen and its metabolism: some new developments and old ...
    Glycogen is a branched polymer of glucose that acts as a store of energy in times of nutritional sufficiency for utilization in times of need.
  100. [100]
    Glycogen metabolism and structure: A review - ScienceDirect.com
    Dec 15, 2024 · Insulin, glucagon and epinephrine are key hormonal regulators, modulating the activity of glycogen synthase (GS) and glycogen phosphorylase ...
  101. [101]
    ATR kinase supports normal proliferation in the early S phase by ...
    Jun 19, 2023 · Mec1 is activated at the onset of normal S phase by low-dNTP pools impeding DNA replication. Mol. Cell 78, 396–410 (2020). Article CAS ...
  102. [102]
    The deoxynucleotide triphosphohydrolase SAMHD1 is a major ...
    Synthesis by the reductase peaks during S-phase, and catabolism by SAMHD1 is maximal during G1 phase when large dNTP pools would prevent cells from preparing ...
  103. [103]
    Regulation of cancer cell metabolism: oncogenic MYC in the driver's ...
    Jul 10, 2020 · MYC enhances nucleotide synthesis by inducing a series of genes involved in this process, further to prepare the cells ready for cell cycle ...Myc Regulation Of Glucose... · Myc Regulation Of Amino Acid... · Targeting Myc-Driven...
  104. [104]
    Regulation of nucleotide metabolism in cancers and immune disorders
    Cancer cells rewire key metabolic pathways to increase substrate availability for nucleotide synthesis and escape DNA damage induced by chemotherapeutic agents.
  105. [105]
    Nucleoside salvage pathway kinases regulate hematopoiesis by ...
    Nov 12, 2012 · It is through the kinase action of TK1 that exogenously added thymidine is trapped in cells, thus enabling its conversion to dTTP in the cytosol ...
  106. [106]
    An essential role for dNTP homeostasis following CDK-induced ...
    Mar 25, 2019 · Subsequent dNTP depletion leads to inefficient DNA replication, DNA damage and to genome instability. Cells respond to this replication ...
  107. [107]
    How ATP and dATP Act as Molecular Switches to Regulate ...
    Aug 20, 2024 · Class Ia ribonucleotide reductases (RNRs) are allosterically regulated by ATP and dATP to maintain the appropriate deoxyribonucleotide levels inside the cell ...
  108. [108]
    Convergent allostery in ribonucleotide reductase - Nature
    Jun 14, 2019 · Allosteric regulation of RNRs plays a key role in maintaining the appropriate balance of intracellular nucleotides required for DNA replication ...
  109. [109]
    NME6: ribonucleotide salvage sustains mitochondrial transcription
    Aug 7, 2023 · The building blocks for RNA and DNA are made in the cytosol, meaning mitochondria depend on the import and salvage of ribonucleoside ...
  110. [110]
    Ribonucleotides incorporated by the yeast mitochondrial DNA ...
    Nov 6, 2017 · We show that imbalances in the total cellular dNTP pool are transmitted into the mitochondria and that they alter the frequency of ribonucleotides.
  111. [111]
    Mitochondrial Deoxyribonucleotides, Pool Sizes, Synthesis, and ...
    It is generally believed that special dNTP pools in mitochondria not in equilibrium with cytosolic dNTP pools serve the synthesis of mt DNA. Seminal reports ...
  112. [112]
    mRNA capping: biological functions and applications
    ... mRNA splicing, polyadenylation and nuclear export. It also acts as the anchor for the recruitment of initiation factors that initiate protein synthesis and ...Abstract · INTRODUCTION · CONCLUDING REMARKS
  113. [113]
    Polyadenylation and nuclear export of mRNAs - PubMed Central - NIH
    In eukaryotes, the separation of translation from transcription by the nuclear envelope enables mRNA modifications such as capping, splicing, and
  114. [114]
    Mechanisms of nuclear mRNA export: a structural perspective - PMC
    Sep 12, 2019 · This review discusses ongoing research, mainly from a structural perspective, to provide an overview of the key steps in nuclear mRNA export.Missing: paper | Show results with:paper
  115. [115]
    Codon—anticodon pairing: The wobble hypothesis - ScienceDirect
    This hypothesis is explored systematically, and it is shown that such a wobble could explain the general nature of the degeneracy of the genetic code.
  116. [116]
    Aminoacyl-tRNA synthetases - PMC - PubMed Central
    The first stage of tRNA binding is relatively fast and unspecific, driven mainly by the electrostatic interactions between positively charged residues of the ...
  117. [117]
    Mitochondrial and plastid genome architecture: Reoccurring themes ...
    Mar 26, 2015 · Overall, mitochondria that retain the universal code are relatively rare exceptions. In contrast, no noncanonical genetic codes have been ...Sign Up For Pnas Alerts · Noncoding Nucleotides · Convergent Organelle Genome...
  118. [118]
    How EF-Tu can contribute to efficient proofreading of aa-tRNA by the ...
    Oct 31, 2016 · Proofreading contributes to the total selectivity of translation according to I × Pf, where I is the selectivity of initial selection. To ...
  119. [119]
    Maintenance of translational elongation rate underlies the survival of ...
    May 27, 2019 · When E. coli cells grow in rich nutrient conditions, the ribosome translates proteins at a high elongation rate (16–17 amino acids per sec, aa/s) ...
  120. [120]
    The Energy Costs of Protein Metabolism: Lean and Mean on Uncle ...
    ... GTP per peptide bond is required for initiation, two GTP per bond for elongation, and one GTP per peptide for termination. However, recent evidence suggests ...Bookshelf · The Role Of Protein And... · Protein Metabolism As An...
  121. [121]
    Structural aspects of translation termination on the ribosome - PMC
    In eubacteria, translation termination is mediated by release factor RF1 in response to a UAG or UAA stop codon, and by RF2 in response to a UGA or UAA codon.
  122. [122]
    The structural basis for release-factor activation during translation ...
    Jun 12, 2019 · In termination of translation, the complete protein is released from the ribosome by a class-1 release factor (RF) recognizing one of the ...
  123. [123]
    Termination of translation: interplay of mRNA, rRNAs and release ...
    From these experiments, it could be concluded that the only function of RF3 in translational termination is to remove RF1 or RF2 from ribosomes in their post‐ ...
  124. [124]
    Bacterial Hsp70 resolves misfolded states and accelerates ... - Nature
    Jan 17, 2020 · Together these findings demonstrate that the Hsp70 chaperone system actively promotes client protein folding in an ATP-dependent manner, by ...
  125. [125]
    Co-translational Mechanisms of Protein Biogenesis and Complex ...
    Apr 24, 2024 · Here we review the constraints and mechanisms governing eukaryotic co- and post-translational protein folding and assembly.
  126. [126]
    Principles of cotranslational mitochondrial protein import
    Oct 2, 2025 · Nearly all mitochondrial proteins are translated on cytosolic ribosomes. How these proteins are subsequently delivered to mitochondria ...
  127. [127]
    Navigating the landscape of protein folding and proteostasis - Nature
    Oct 23, 2025 · Co-chaperones such as HOP and BAG1 promote protein folding by facilitating the transfer of certain substrates from HSP70 to HSP90. In contrast, ...
  128. [128]
    Evidence for a Negative-Feedback Mechanism in the Biosynthesis ...
    Evidence for a Negative-Feedback Mechanism in the Biosynthesis of Isoleucine. H. Edwin Umbarger, Science, 1956. This Week in Science. Science. Generating a ...
  129. [129]
    SREBP1c-CRY1 signalling represses hepatic glucose production by ...
    Jul 14, 2016 · These data suggest that insulin-activated SREBP1c downregulates gluconeogenesis through CRY1-mediated FOXO1 degradation and that ...
  130. [130]
    Pancreatic regulation of glucose homeostasis - Nature
    Mar 11, 2016 · As glucagon's opponent, insulin stimulates glycolysis via enhanced expression of the hepatic glucokinase gene, a key enzyme that converts ...
  131. [131]
    Biology of HIF-1α | Cell Death & Differentiation - Nature
    Feb 15, 2008 · The transcriptional complex hypoxia-inducible factor has emerged as a key regulator of the molecular hypoxic response, mediating a wide range of physiological ...
  132. [132]
    Regulation of chromatin by histone modifications | Cell Research
    Feb 15, 2011 · Here, we describe the known histone modifications, define where they are found genomically and discuss some of their functional consequences.
  133. [133]
    The Compartmentalization of Cells - Molecular Biology ... - NCBI - NIH
    In this introductory section we present a brief overview of the compartments of the cell and the relationships between them. In doing so, we organize the ...
  134. [134]
    Principles and functions of metabolic compartmentalization - PMC
    Oct 20, 2022 · Within cells, membrane-bound organelles subdivide the cytoplasm into chemically and physically unique reaction compartments. These compartments ...
  135. [135]
    Prokaryotic and eukaryotic cells (article) - Khan Academy
    Overall, prokaryotic cells are less complex than eukaryotic cells. Prokaryotic cells are considered to be the simplest and most ancient forms of life. The ...Missing: biosynthesis | Show results with:biosynthesis
  136. [136]
    Synthesis of Fatty Acids - The Medical Biochemistry Page
    Oct 7, 2025 · All of the reactions of fatty acid synthesis are carried out by the multiple enzymatic activities of FAS. Like fat oxidation, fat synthesis ...Malic Enzymes · Reactions of Fatty Acid Synthesis · Regulation of Fatty Acid...<|control11|><|separator|>
  137. [137]
    The Major Sites of Cellular Phospholipid Synthesis and Molecular ...
    Oct 13, 2017 · Phosphatidylcholine and phosphatidylethanolamine are the two main phospholipids in eukaryotic cells comprising ∼50 and 25% of phospholipid mass, respectively.
  138. [138]
    From Synthesis to Utilization: The Ins and Outs of Mitochondrial Heme
    In this review, we summarize current knowledge about heme biosynthesis and trafficking within and outside of the mitochondria.2. Heme Biosynthesis · 3. From Heme B To Hemes C, O... · 3.3. Heme A Pathway
  139. [139]
    Steroid hormone synthesis in mitochondria - ScienceDirect.com
    Oct 15, 2013 · Mitochondria are essential sites for steroid hormone biosynthesis. Mitochondria in the steroidogenic cells of the adrenal, gonad, placenta and brain contain ...
  140. [140]
    regulation of plasmalogen biosynthesis and its ... - PubMed
    Plasmalogen synthesis is initiated in peroxisomes and completed in the endoplasmic reticulum. The absence of plasmalogens in several organs of peroxisome ...
  141. [141]
    Peroxisome-driven ether-linked phospholipids biosynthesis is ...
    Mar 17, 2021 · Ether phospholipids are initially synthesized in peroxisomes and processed in the endoplasmic reticulum (ER) [1,2,3]. Plasmalogens are the most ...
  142. [142]
    Vesicular and non-vesicular lipid export from the ER to the secretory ...
    In this review we will discuss the mechanism and roles of vesicular and non-vesicular transport of lipids from the ER to other organelles of the secretory ...
  143. [143]
    11.6: Vesicular Transport - Biology LibreTexts
    Dec 30, 2022 · The major types of coat proteins used in vesicle formation are COPII, COPI, and clathrin. COPII coat proteins form the vesicles that move from ...
  144. [144]
    Inborn Errors of Metabolism - StatPearls - NCBI Bookshelf
    Inborn errors of metabolism are a heterogeneous group of disorders that may be inherited or may occur as the result of spontaneous mutations.Continuing Education Activity · Epidemiology · History and Physical · Evaluation
  145. [145]
    Inborn Errors of Metabolism: Background, Pathophysiology, Etiology
    Apr 3, 2025 · Most of the disorders are inherited as autosomal recessive, but some are autosomal dominant or X-linked.Background · Pathophysiology · Etiology · Epidemiology
  146. [146]
    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 ...
  147. [147]
    Phenylketonuria: An Inborn Error of Phenylalanine Metabolism - PMC
    Phenylketonuria (PKU) is an autosomal recessive inborn error of phenylalanine (Phe) metabolism resulting from deficiency of phenylalanine hydroxylase (PAH).
  148. [148]
    Adenosine deaminase deficiency: a review
    Apr 24, 2018 · Adenosine deaminase (ADA) deficiency leads to an accumulation of toxic purine degradation by-products, most potently affecting lymphocytes.
  149. [149]
    Hereditary orotic aciduria (HOA): A novel uridine-5-monophosphate ...
    Jan 9, 2021 · Hereditary orotic aciduria (HOA) is a very rare inborn error of pyrimidine metabolism. It results from a defect of the uridine-5-monophosphate synthase (UMPS) ...
  150. [150]
    About Newborn Screening - CDC
    Dec 17, 2024 · The conditions screened for include spinal muscular atrophy, cystic fibrosis, sickle cell disease and other hemoglobinopathies, endocrine ...
  151. [151]
    Vitamin B6 Deficiency - StatPearls - NCBI Bookshelf
    It plays a role in gluconeogenesis and glycogenolysis. Vitamin B6 is critical in transamination and decarboxylation, the initial steps of porphyrin synthesis.Continuing Education Activity · Introduction · Evaluation · Treatment / ManagementMissing: biosynthesis | Show results with:biosynthesis
  152. [152]
    A peculiarly characterised case of isoniazid-induced pellagra- 2 Ds ...
    She complained of a generalised rash, sloughing off, and darkening of the skin on the face, neck, forearms, and dorsum of both feet. A physical examination ...
  153. [153]
    Iron, Heme Synthesis and Erythropoietic Porphyrias - NIH
    Nov 23, 2021 · In patients with functional heme biosynthesis pathway, a chronic depletion in iron stores leads to microcytic hypochromic anemia with a moderate ...
  154. [154]
    Vitamin C Deficiency - StatPearls - NCBI Bookshelf
    Nov 12, 2023 · Scurvy arises due to inadequate vitamin C intake, which is critical in synthesizing collagen. Collagen is a vital structural protein essential ...Missing: biosynthesis | Show results with:biosynthesis
  155. [155]
    Hiding in Plain Sight: Modern Thiamine Deficiency - PMC
    Without thiamine, PDC activity diminishes and both oxidative capacity and lipid biosynthesis falter. Pyruvate, not converted to acetyl CoA, is shunted through ...
  156. [156]
    Dietary intake of sulfur amino acids and risk of kwashiorkor ...
    May 8, 2021 · Light-colored unkinked hair, facial edema, and bilateral pitting edema in the feet were recorded as potential signs of kwashiorkor (32).
  157. [157]
    Oedema in kwashiorkor is caused by hypoalbuminaemia - PMC - NIH
    It has been argued that the oedema of kwashiorkor is not caused by hypoalbuminaemia because the oedema disappears with dietary treatment before the plasma ...Missing: essential | Show results with:essential
  158. [158]
    Fact sheets - Malnutrition
    Mar 1, 2024 · Globally in 2022, 149 million children under 5 were estimated to be stunted (too short for age), 45 million were estimated to be wasted (too ...
  159. [159]
    Micronutrients - World Health Organization (WHO)
    Micronutrient deficiencies can cause visible and dangerous health conditions, but they can also lead to less clinically notable reductions in energy level, ...
  160. [160]
    Macrophage-Targeted Glucocerebrosidase for Gaucher's Disease
    May 23, 1991 · We conclude that intravenous infusion of macro-phage-targeted glucocerebrosidase is a safe and highly beneficial treatment for selected patients ...
  161. [161]
    How I treat Gaucher disease | Blood | American Society of Hematology
    Aug 11, 2011 · The purpose of this review is to present a cohesive approach to treating patients with Gaucher disease (GD).
  162. [162]
    Home Enzyme Replacement Therapy in Gaucher Disease: A Review
    Home ERT has consistently been shown to be safe, to improve patients' quality of life, to reduce the utilization of hospital resources, and to pose no ...
  163. [163]
    Folic Acid - StatPearls - NCBI Bookshelf - NIH
    Aug 8, 2023 · Folic acid (vitamin B9) is a water-soluble vitamin used to manage and treat megaloblastic anemia. Folic acid has FDA approval for treating megaloblastic and ...
  164. [164]
    Megaloblastic Anemia - StatPearls - NCBI Bookshelf - NIH
    Apr 3, 2023 · ... DNA synthesis abnormalities. Both vitamin B12 and folate deficiencies may cause defective DNA synthesis. Subsequently, the nucleus and ...
  165. [165]
    HMG-CoA Reductase Inhibitors - StatPearls - NCBI Bookshelf
    HMG-CoA reductase inhibitors (statins) are lipid-lowering medications used in the primary and secondary prevention of coronary heart disease.
  166. [166]
    Prolonged inhibition of cholesterol synthesis explains the ... - PubMed
    These findings suggest that the greater efficacy of atorvastatin compared with simvastatin is due to more prolonged inhibition of HMG-CoA reductase.
  167. [167]
    140. Point Mutation Correction for ADA SCID - Cell Press
    We have designed a CRISPR/Cas9 strategy that can target the nonsense Q3X (ADA c7C>T) point mutation endemic to ADA SCID patients of Somali origin residing in ...
  168. [168]
    Gene Therapy for Immunodeficiency Due to Adenosine Deaminase ...
    Gene therapy, combined with reduced-intensity conditioning, is a safe and effective treatment for SCID in patients with ADA deficiency.
  169. [169]
    PKU dietary handbook to accompany PKU guidelines
    Aug 4, 2020 · A strict, lifelong low phenylalanine diet is the principle treatment in PKU. It may be the only treatment or used in combination with drug ...
  170. [170]
    What are common treatments for phenylketonuria (PKU)? | NICHD
    Jan 16, 2024 · These treatments include large neutral amino acid supplementation, which may help prevent phenylalanine from entering the brain, and enzyme replacement therapy.
  171. [171]
    Allopurinol - StatPearls - NCBI Bookshelf - NIH
    Allopurinol and oxypurinol both inhibit xanthine oxidase, an enzyme in the purine catabolism pathway that converts hypoxanthine to xanthine to uric acid.
  172. [172]
    Allopurinol (Xanthine Oxidase Inhibitor) in the Treatment of ...
    Allopurinol was given to 24 patients with gout resistant or intolerant to uricosuric drugs; in each case a substantial fall in serum uric acid level resulted.