Fact-checked by Grok 2 weeks ago

Citric acid cycle

The citric acid cycle, also known as the Krebs cycle or tricarboxylic acid (TCA) cycle, is a central series of enzymatic reactions in aerobic organisms that oxidizes acetyl-coenzyme A () derived from the breakdown of carbohydrates, fats, and proteins to produce , while generating high-energy electron carriers for ATP synthesis. This cycle serves as the final common pathway for the catabolism of these macronutrients, linking their initial metabolism to the for efficient energy production. The cycle operates primarily in the of eukaryotic cells, with one key enzyme, , embedded in the as part of complex II of the . It begins with the condensation of (a two-carbon unit) with oxaloacetate (a four-carbon molecule) to form citrate, catalyzed by , followed by a series of seven additional transformations that regenerate oxaloacetate and release two molecules of . Per turn of the cycle, one yields three molecules of NADH, one FADH₂, and one GTP (or ATP via ), providing reducing equivalents that drive to produce up to 10 additional ATP molecules per . Discovered by Hans Adolf Krebs in 1937 through studies on pigeon muscle tissue, the cycle was initially elucidated as a mechanism for oxidizing pyruvate-derived intermediates, earning Krebs the in Physiology or Medicine in 1953. Beyond energy generation, the citric acid cycle functions as a metabolic hub, supplying precursors for biosynthetic pathways such as the production of (e.g., aspartate from oxaloacetate, glutamate from α-ketoglutarate), , and , while its intermediates like citrate also regulate . Regulation of the cycle occurs at three irreversible steps—, , and α-ketoglutarate dehydrogenase—primarily through allosteric inhibition by high levels of NADH, ATP, and , ensuring coordination with cellular energy demands and the availability of NAD⁺ and . In prokaryotes, the cycle occurs in the or associated membranes, and variations exist in conditions or certain pathogens, but its core role in aerobic remains conserved across life forms.

Introduction and Overview

Definition and Role in Metabolism

The citric acid cycle, also known as the tricarboxylic acid (TCA) cycle, is an eight-step aerobic metabolic pathway that oxidizes acetyl-coenzyme A (acetyl-CoA) to carbon dioxide (CO₂) while generating reducing equivalents—three molecules of nicotinamide adenine dinucleotide (NADH) and one molecule of flavin adenine dinucleotide (FADH₂)—along with one molecule of guanosine triphosphate (GTP) or adenosine triphosphate (ATP) per cycle. This process occurs in the mitochondrial matrix of eukaryotic cells and in the cytoplasm of prokaryotes, requiring oxygen indirectly through its linkage to oxidative phosphorylation. The cycle integrates the breakdown products from various catabolic routes, with acetyl-CoA serving as the universal entry point derived from pyruvate produced by glycolysis in carbohydrate metabolism, beta-oxidation of fatty acids, or degradation of certain amino acids. As the final common oxidative pathway for carbohydrates, fats, and proteins, the citric acid cycle funnels the carbon skeletons of these nutrients into a centralized hub that connects to the (), where NADH and FADH₂ donate electrons to drive ATP synthesis via . This linkage enables the cycle to account for the majority of energy extraction from nutrient oxidation, yielding approximately 10 ATP molecules per through the combined action of the cycle and subsequent processes. Beyond energy production, the citric acid cycle holds profound significance in cellular metabolism by providing intermediates that act as precursors for anabolic pathways, such as the synthesis of amino acids (e.g., from α-ketoglutarate and oxaloacetate), porphyrins, and fatty acids, thereby supporting biosynthesis under varying physiological demands. Intermediates like citrate also function in cellular signaling, serving as regulators of enzymes in glycolysis and fatty acid synthesis to coordinate metabolic flux. The cycle's circular architecture underscores its efficiency: condenses with the four-carbon oxaloacetate to initiate the pathway as , followed by sequential dehydrogenations, decarboxylations, and rearrangements that release two CO₂ molecules and regenerate oxaloacetate, allowing the cycle to turn repeatedly without depleting its catalytic intermediates.

Historical Discovery

The isolation of from lemon juice was first achieved in 1784 by Swedish chemist , who crystallized it as , marking the initial recognition of this key in biological systems. In the early 1930s, the elucidation of the Embden-Meyerhof-Parnas pathway—commonly known as —by Gustav Embden, Otto Meyerhof, and Jakub Karol Parnas provided critical context for subsequent research, as it identified pyruvate as the primary product of glucose breakdown available for aerobic oxidation in tissues. Building on these foundations, Hungarian biochemist demonstrated in minced muscle preparations that dicarboxylic acids such as fumarate and succinate catalytically enhanced respiration, earning him the 1937 in Physiology or Medicine for revealing their role in biological oxidation processes. In 1937, while at the , Hans Adolf Krebs and his graduate student William Arthur proposed a cyclic pathway integrating these observations, dubbing it the "citric acid cycle" based on experiments with minced pigeon breast muscle. Using manometric techniques to measure oxygen consumption and production, they found that adding citrate or related tricarboxylic acids to the preparations dramatically accelerated pyruvate oxidation, indicating a regenerative rather than a linear degradation. This work resolved ongoing debates in the field, where earlier models favored straight-chain oxidations of pyruvate without regeneration of catalysts, by showing how two-carbon units from pyruvate condense with oxaloacetate to form citrate, which then undergoes sequential transformations back to oxaloacetate. Further validation came in 1941 through isotope-tracer studies by Earl A. Evans Jr. and Leonidas Slotin, who used 13C-carboxyl-labeled pyruvate in pigeon liver minces and observed the label's specific incorporation into the carboxyl groups of α-ketoglutarate, confirming the cyclic intermediacy and ruling out alternative linear pathways. Krebs later renamed the pathway the "tricarboxylic acid cycle" to emphasize its key intermediates, though it is also commonly called the Krebs cycle in his honor. For this discovery, Krebs shared the 1953 in or with Fritz Albert Lipmann, recognizing the cycle's central role in metabolic integration.

Core Mechanism of the Cycle

Reaction Steps

The citric acid cycle consists of eight sequential enzymatic reactions that occur primarily in the of eukaryotic cells, with one exception embedded in the . These steps oxidize the from to two molecules of CO₂, generating reduced electron carriers NADH and FADH₂ that feed into the for ATP production. Each reaction involves specific enzymes, cofactors, and intermediates, ensuring the cycle's efficiency in energy extraction. Step 1: Citrate formation
The cycle begins with the irreversible condensation of acetyl-CoA and oxaloacetate to form citrate, catalyzed by citrate synthase in the mitochondrial matrix. The balanced equation is:
\ce{Acetyl-CoA + oxaloacetate + H2O -> citrate + CoA-SH}
This reaction proceeds via a citryl-CoA intermediate, driven by a highly negative standard free energy change (ΔG°' ≈ -32 kJ/mol), making it effectively irreversible under physiological conditions. No additional cofactors are required beyond the substrates themselves.
Step 2: Isomerization to isocitrate
Citrate is then to isocitrate through a two-part and rehydration process involving the cis-aconitate, catalyzed by aconitase in the . The overall reaction is:
\ce{Citrate <=> isocitrate}
Aconitase utilizes a [4Fe-4S] iron-sulfur cluster as a cofactor to facilitate the dehydration/rehydration, with ΔG°' ≈ +6.3 kJ/mol, rendering it reversible but pulled forward by subsequent steps. This rearrangement positions the hydroxyl group for oxidation in the next reaction.
Step 3: Oxidative decarboxylation to α-ketoglutarate
Isocitrate undergoes oxidative to form α-ketoglutarate, catalyzed by in the . The reaction is:
\ce{Isocitrate + NAD+ -> α-ketoglutarate + CO2 + NADH + H+}
This irreversible step (ΔG°' ≈ -8.4 kJ/mol) requires NAD⁺ as a cofactor and Mg²⁺ for activity, involving first the oxidation to oxalosuccinate (a β-keto acid ) followed by decarboxylation. It represents the first CO₂ release and NADH generation in the cycle.
Step 4: Oxidative decarboxylation to succinyl-CoA
The α-ketoglutarate dehydrogenase complex, located in the mitochondrial matrix, catalyzes the oxidative decarboxylation of α-ketoglutarate to succinyl-CoA, analogous to the pyruvate dehydrogenase complex. The balanced equation is:
\ce{α-ketoglutarate + NAD+ + CoA-SH -> succinyl-CoA + CO2 + NADH + H+}
This irreversible reaction (ΔG°' ≈ -30 kJ/mol) employs multiple cofactors including thiamine pyrophosphate (for decarboxylation), lipoic acid (for acyl transfer), CoA, FAD, and NAD⁺, occurring via a multi-enzyme complex that ensures efficient substrate channeling. It releases the second CO₂ and produces another NADH.
Step 5: Substrate-level phosphorylation to succinate
Succinyl-CoA is converted to succinate with the concomitant synthesis of GTP from GDP and inorganic phosphate, catalyzed by succinyl-CoA synthetase (also known as succinate thiokinase) in the mitochondrial matrix. The reaction is:
\ce{Succinyl-CoA + GDP + P_i -> succinate + GTP + CoA-SH}
This reversible step (ΔG°' ≈ -3.3 kJ/mol) involves substrate-level phosphorylation, where the high-energy thioester bond of succinyl-CoA drives GTP formation via a phosphohistidine intermediate on the enzyme; no additional cofactors are needed. GTP can be converted to ATP via nucleoside diphosphate kinase.
Step 6: Oxidation to fumarate
Succinate is oxidized to fumarate by succinate dehydrogenase, a flavoprotein embedded in the inner mitochondrial membrane as complex II of the electron transport chain. The reaction is:
\ce{Succinate + FAD -> fumarate + FADH2}
This reversible step (ΔG°' ≈ 0 kJ/mol) uses FAD as a tightly bound cofactor to abstract electrons, forming a trans double bond; the FADH₂ directly reduces ubiquinone in the membrane, linking the cycle to oxidative phosphorylation.
Step 7: Hydration to malate
Fumarase catalyzes the reversible hydration of fumarate to form L-malate in the . The reaction is:
\ce{Fumarate + H2O <=> L-malate}
With ΔG°' ≈ -3.8 kJ/mol, this stereospecific trans-addition of water across the requires no cofactors and proceeds via a intermediate stabilized by the . It introduces asymmetry to the for the final oxidation.
Step 8: Oxidation to oxaloacetate
The cycle closes with the reversible oxidation of L-malate to oxaloacetate, catalyzed by in the . The reaction is:
\ce{L-malate + NAD+ <=> oxaloacetate + NADH + H+}
This endergonic step (ΔG°' ≈ +30 kJ/mol) relies on NAD⁺ as a cofactor and is thermodynamically unfavorable but driven forward by the highly exergonic reaction that consumes oxaloacetate; it generates the final NADH of the cycle.

Stoichiometric Products

The net reaction for one complete turn of the citric acid cycle, starting from the condensation of acetyl-CoA with oxaloacetate, is given by: \text{Acetyl-CoA} + 3\, \text{NAD}^+ + \text{FAD} + \text{GDP} + \text{P}_\text{i} + 2\, \text{H}_2\text{O} \rightarrow 2\, \text{CO}_2 + 3\, \text{NADH} + \text{FADH}_2 + \text{GTP} + 3\, \text{H}^+ + \text{CoA-SH} This equation summarizes the overall stoichiometry, where the two-carbon acetyl group is oxidized, producing energy-rich molecules and releasing carbon dioxide. The primary stoichiometric products per cycle include three molecules of NADH, one molecule of FADH₂, one molecule of GTP, and two molecules of CO₂, along with the regeneration of free (CoA-SH). The GTP is generated via at the succinyl-CoA synthetase step and can be readily converted to ATP through the action of , providing a direct equivalent. However, the cycle itself yields no net ATP beyond this single GTP; the bulk of the energetic output is captured in the reducing equivalents NADH and FADH₂, which donate electrons to the for . In terms of carbon fate, the two carbon atoms from the acetyl moiety of are completely oxidized and released as the two CO₂ molecules during the decarboxylation reactions at and α-ketoglutarate dehydrogenase; the four-carbon skeleton of oxaloacetate is fully regenerated at the end of the cycle, with no net consumption or loss of its carbons. The NADH and FADH₂ produced serve as key electron donors: each NADH is estimated to generate approximately 2.5 ATP molecules, while each FADH₂ yields about 1.5 ATP through proton pumping and in the respiratory chain, underscoring the cycle's role in coupling carbon oxidation to respiratory energy production.

Thermodynamic Efficiency

The citric acid cycle exhibits high thermodynamic efficiency in converting the of into usable forms, primarily through the production of high-energy carriers. The overall standard change (ΔG°') for the reactions of the (sum of individual steps) is approximately -44 kJ/mol, rendering the process exergonic and effectively irreversible under physiological conditions. This substantial negative ΔG°' ensures unidirectional flux through the , preventing significant back-reactions despite some individual steps having near-equilibrium . In the context of complete glucose oxidation, which proceeds through two turns of the (yielding two acetyl-CoA molecules), the citric acid cycle contributes roughly 20 ATP equivalents out of a total yield of 30-32 ATP per glucose molecule. The overall reaction for glucose oxidation is: \mathrm{C_6H_{12}O_6 + 6\, O_2 \rightarrow 6\, CO_2 + 6\, H_2O} with ΔG°' ≈ -2870 kJ/, wherein the cycle accounts for about 50% of the total energy release by oxidizing the carbon skeleton to CO₂ while generating NADH, FADH₂, and GTP. Using the physiological of ATP hydrolysis (≈ -50 kJ/), the energetic of this process reaches 60-70% of the theoretical maximum, far superior to the 100% dissipation in non-biological of glucose. Much of the cycle's energy is conserved in the reduction potentials of NADH (E°' ≈ -0.32 V) and FADH₂ (E°' ≈ -0.22 V), which fuel by driving proton translocation across the to establish a proton motive force ( ≈ 200 mV). This coupling minimizes loss compared to uncoupled oxidation, though approximately 30-40% of the is inevitably released as to maintain the second law of thermodynamics. The cycle directly captures a small portion of this as GTP (equivalent to ATP) via at synthetase. Efficiency can vary due to mitochondrial factors, such as the strength of the proton motive force, which optimizes activity (requiring ≈ 3-4 H⁺ per ATP), and the presence of uncoupling proteins (e.g., in ) that dissipate the gradient as heat, reducing ATP yield by up to 50% in thermogenic tissues while preventing excessive production.

Regulation Mechanisms

Enzymatic Control Points

The citric acid cycle is primarily regulated at its three irreversible steps, which serve as key control points to modulate metabolic flux in response to cellular energy demands. These steps are catalyzed by (step 1), (step 3), and α-ketoglutarate dehydrogenase (step 4), as regulation at these committed, exergonic reactions allows efficient prevention of intermediate accumulation and wasteful cycling. The rationale for targeting these points lies in their thermodynamic favorability (with large negative ΔG°' values, such as -32 kJ/mol for ), making reversal unlikely and thus ideal for flux control without reversing the pathway. Citrate synthase, the entry point enzyme, condenses and oxaloacetate to form citrate and follows Michaelis-Menten kinetics with low values for its substrates (e.g., ~1-5 μM for oxaloacetate in mammalian mitochondria), ensuring efficient response to substrate availability. It is allosterically inhibited by high-energy signals including ATP, NADH, and , which bind to reduce enzyme activity when cellular energy is abundant, while acts as an activator to promote flux under energy-deficient conditions. Isocitrate dehydrogenase, a rate-limiting , oxidatively decarboxylates isocitrate to α-ketoglutarate and exhibits sigmoidal kinetics modulated by allosteric effectors, with a low for isocitrate (~20-50 μM in the activated state) that heightens sensitivity to substrate levels and regulatory inputs. It is activated by and Ca²⁺, which lower the Km for isocitrate and enhance Vmax to accelerate the cycle during energy need or signaling events like , while inhibited by ATP and NADH to slow activity when energy carriers are plentiful. α-Ketoglutarate dehydrogenase, a multi-enzyme complex analogous to , decarboxylates α-ketoglutarate to and operates under Michaelis-Menten kinetics with regulation focused on product inhibition. It is inhibited by , NADH, and ATP, which bind allosterically to decrease activity and prevent overproduction of reducing equivalents, while Ca²⁺ activation reduces the Km for α-ketoglutarate to fine-tune flux in response to calcium signals. Entry into the cycle is further controlled upstream by , which converts pyruvate to and undergoes covalent modification via . phosphorylates and inactivates the enzyme complex in response to high NADH/NAD⁺ and / ratios, while pyruvate dehydrogenase dephosphorylates and activates it under conditions of low energy charge, thereby linking glycolytic flux to citric acid cycle demand.
EnzymeKey Substrates (Km examples)ActivatorsInhibitors
Citrate synthase (~10-50 μM), oxaloacetate (~1-5 μM)ATP, NADH, , citrate
Isocitrate dehydrogenaseIsocitrate (~20-50 μM activated), Ca²⁺ATP, NADH
α-Ketoglutarate dehydrogenaseα-Ketoglutarate (~100-200 μM)Ca²⁺, NADH, ATP
Pyruvate dehydrogenase (upstream)Pyruvate (~50-100 μM)Dephosphorylation (by PDP)Phosphorylation (by PDK), NADH,
This table summarizes the primary regulatory features, emphasizing how kinetic properties and modifiers align cycle activity with bioenergetic status.

Allosteric and Substrate Regulation

The citric acid cycle is subject to by end products and indicators to fine-tune according to cellular demands. NADH acts as a by competitively binding to dehydrogenases, such as those in the cycle, thereby reducing their activity when capacity is saturated and preventing over-reduction that could lead to accumulation. Similarly, the energy charge, reflected in the ATP/ADP ratio, modulates activity; elevated ATP levels inhibit this enzyme, slowing cycle entry during energy surplus to avoid unnecessary oxidation. Substrate availability exerts direct control over cycle initiation and progression, ensuring coordination with upstream metabolic states. The concentration of oxaloacetate, replenished via , limits citrate formation by , as low levels prevent efficient condensation with and thereby restrict overall flux. The / ratio further influences entry, with high inhibiting activity while promotes its activation, balancing substrate supply to match oxidative needs. Hormonal signals integrate the cycle with systemic , particularly during or . Glucagon and epinephrine elevate cAMP levels, activating , which promotes dephosphorylation and activation of upstream enzymes like , thereby increasing entry and enhancing cycle flux to support and energy production. Mitochondrial compartmentalization imposes regulatory constraints by separating cycle intermediates from cytosolic processes. Export of citrate from mitochondria to the cytosol via the citrate-malate shuttle supports but depletes mitochondrial citrate pools, indirectly inhibiting cycle progression by reducing substrate for subsequent enzymes. The NAD+/NADH serves as a critical sensor, linking cycle activity to oxygen availability and electron transport efficiency. A high NADH/NAD+ , indicative of or impaired , allosterically inhibits key dehydrogenases, downregulating flux to prevent metabolic imbalance, while a favorable promotes oxidation. For instance, activity is particularly sensitive to this , adjusting the cycle's pace in response to respiratory conditions.

Variations Across Organisms and Conditions

Organism-Specific Modifications

In prokaryotes, the citric acid cycle operates in the , lacking the compartmentalization seen in eukaryotic cells, which allows for flexible integration with other cytosolic metabolic pathways. Many prokaryotes exhibit variations, such as the , which enables the net conversion of derived from two-carbon units (like or fatty acids) into four-carbon intermediates for , bypassing the decarboxylation steps of and α-ketoglutarate dehydrogenase through the enzymes isocitrate lyase and malate synthase. This modification is particularly prominent in capable of growing on C2 substrates, enhancing their adaptability to nutrient-limited environments. In eukaryotes, the cycle is localized to the , where it couples closely with the for efficient ATP production via . However, certain anaerobic prokaryotes employ a reverse tricarboxylic acid () cycle for autotrophic CO₂ fixation, running the pathway in the reductive direction to synthesize organic compounds from inorganic carbon. For instance, green sulfur bacteria like Chlorobium limicola utilize this reductive TCA cycle, where enzymes such as α-ketoglutarate:ferredoxin and operate reversibly to assimilate CO₂ into citrate and subsequent intermediates. Anaerobic adaptations further diversify the cycle's function across organisms. In fermentative like species, the TCA cycle is incomplete, lacking key enzymes such as α-ketoglutarate dehydrogenase, , and , which limits it to a partial oxidative branch that supports and production rather than full energy generation. Similarly, in the parasitic , the cycle is incomplete and operates primarily in the reductive direction under conditions in the host intestine, generating succinate and propionate as end products via a branched pathway involving fumarate reductase, without completing the full oxidative loop. In , the glyoxylate shunt—a specialized variant of the —plays a critical role during seed germination, particularly in oil-rich seeds like those of sunflower or castor bean. This shunt allows the conversion of stored lipids into carbohydrates for , bypassing the CO₂-releasing steps of the standard cycle to conserve carbon atoms as succinate, which is then used to synthesize glucose for early growth. The enzymes isocitrate lyase and malate synthase are highly expressed in glyoxysomes during this phase, ensuring efficient mobilization of triacylglycerols. Recent investigations into extremophiles have revealed adaptations in the citric acid cycle that enhance survival in harsh environments. In hyperthermophiles such as Archaeoglobus fulgidus, enzymes like exhibit exceptional , maintaining activity above 80°C due to reinforced ion-pair networks and hydrophobic cores that prevent denaturation, allowing the cycle to function in high-temperature niches like hydrothermal vents. These modifications underscore the cycle's evolutionary plasticity while conserving core catalytic steps across domains of life.

Pathological Alterations in Disease

In cancer, the citric acid cycle undergoes significant reprogramming, exemplified by the Warburg effect, where tumor cells preferentially utilize aerobic for energy production, leading to reduced flux through the TCA cycle despite adequate oxygen availability. This metabolic shift supports rapid proliferation by diverting carbon from to biosynthetic pathways. Mutations in (SDH), the enzyme catalyzing the sixth step of the cycle, disrupt this process and are associated with hereditary paragangliomas and pheochromocytomas, resulting in succinate accumulation that inhibits α-ketoglutarate-dependent dioxygenases and promotes pseudohypoxic signaling. Similarly, mutations in 1 and 2 (IDH1/2), which normally convert isocitrate to α-ketoglutarate in the cycle, produce the oncometabolite 2-hydroxyglutarate (2-HG), a competitive inhibitor of α-ketoglutarate-dependent enzymes, thereby blocking and and sustaining oncogenic . Mitochondrial diseases often stem from defects in (PDH) or TCA cycle enzymes, impairing the cycle's ability to generate reducing equivalents for the . For instance, in , mutations affecting complex I of the respiratory chain hinder NADH oxidation, leading to energy deficits and that indirectly suppress TCA cycle activity through feedback inhibition. These disruptions manifest as progressive neurological deterioration due to inadequate ATP production in high-energy tissues like the . In neurodegenerative disorders, cycle alterations contribute to cellular dysfunction. In , pathological α-synuclein aggregates inhibit complex I activity, reducing NADH utilization and causing a backup of TCA intermediates, which exacerbates and dopaminergic neuron loss. features reduced citrate levels in affected brain regions, linked to impaired aconitase activity and overall TCA flux decline, correlating with amyloid-β accumulation and . Diabetes involves impaired TCA cycle regulation, where insulin resistance diminishes glucose oxidation and pyruvate entry into the cycle, leading to altered flux and accumulation of upstream metabolites that promote lipotoxicity and β-cell exhaustion. This dysregulation contributes to hyperglycemia and long-term complications by uncoupling mitochondrial energy production from insulin signaling. Recent research from 2020 to 2025 highlights glutamine-dependent anaplerosis as a key adaptation in tumors, where glutaminolysis replenishes TCA intermediates to sustain proliferation despite glycolytic dominance, as evidenced in glutamine-addicted cancer models. TCA cycle intermediates serve as biomarkers for disease states; for example, elevated succinate levels signal inflammation by stabilizing hypoxia-inducible factor 1α and promoting immune cell activation in conditions like pulmonary fibrosis. Therapeutically, IDH inhibitors such as vorasidenib and ivosidenib target mutant IDH1/2 in gliomas, reducing 2-HG production, restoring TCA cycle function, and inducing differentiation of tumor cells, with clinical trials showing prolonged progression-free survival in low-grade gliomas.

Integration with Broader Metabolism

Anaplerotic and Cataplerotic Pathways

The citric acid cycle requires continuous replenishment of its intermediates to sustain , as these compounds are frequently diverted for biosynthetic purposes. Anaplerotic pathways provide this replenishment by introducing new carbon skeletons into the cycle, primarily through the conversion of precursors like pyruvate and into key intermediates such as oxaloacetate and α-ketoglutarate. These reactions are essential in tissues with high metabolic demands, such as liver and , where the cycle supports not only energy production but also and homeostasis. A primary anaplerotic reaction is catalyzed by pyruvate carboxylase, a biotin-dependent mitochondrial enzyme that carboxylates pyruvate to form oxaloacetate. The reaction proceeds as follows: \text{Pyruvate} + \text{CO}_2 + \text{ATP} \rightarrow \text{oxaloacetate} + \text{ADP} + \text{P}_\text{i} This pathway is activated by acetyl-CoA and inhibited by α-ketoglutarate, ensuring coordination with cycle activity. Originally identified in avian liver, pyruvate carboxylase plays a crucial role in mammals by linking glycolysis-derived pyruvate to the cycle, particularly during fasting when gluconeogenesis is active.96061-8/fulltext) Other anaplerotic routes include the action of phosphoenolpyruvate carboxykinase (PEPCK) in certain organisms and conditions, where phosphoenolpyruvate (PEP) is carboxylated to oxaloacetate, and glutamate dehydrogenase (GDH), which oxidatively deaminates glutamate to α-ketoglutarate. GDH catalyzes: \text{Glutamate} + \text{NAD(P)}^+ + \text{H}_2\text{O} \rightarrow \alpha\text{-ketoglutarate} + \text{NH}_4^+ + \text{NAD(P)H} + \text{H}^+ This enzyme is allosterically regulated, with high NADH levels inhibiting it to prevent excess reducing equivalents. In mammals, GDH contributes significantly to anaplerosis from glutamine-derived glutamate, especially in cancer cells and kidney. PEPCK-mediated anaplerosis is more prominent in prokaryotes and yeast but can support cycle filling in mammalian models under specific metabolic stresses.70110-9/fulltext) Cataplerotic pathways counteract anaplerosis by exporting intermediates to meet biosynthetic needs, preventing cycle stagnation. Citrate is transported out of mitochondria via the citrate-malate shuttle and cleaved by ATP-citrate lyase to and oxaloacetate, fueling in lipogenic tissues like liver and adipose. Malate exits via the malate-aspartate shuttle for cytosolic use in , where it is converted to oxaloacetate and then to phosphoenolpyruvate by PEPCK. Additionally, α-ketoglutarate is withdrawn through conversion to glutamate by transaminases, supporting and biosynthesis. These exports are balanced to maintain intermediate pools without net accumulation or depletion.70110-9/fulltext) The balance between anaplerosis and cataplerosis is critical to prevent intermediate depletion during periods of high biosynthetic demand, such as or , ensuring sustained flux for ATP production. This is regulated by the cellular state; for instance, elevated NADH inhibits GDH and other anaplerotic enzymes, while low (high AMP/ATP) promotes replenishment. Dysregulation can impair metabolism, as seen in metabolic disorders where insufficient anaplerosis limits .70110-9/fulltext) Anaplerotic and cataplerotic pathways are tightly linked to metabolism through reactions. For example, aspartate aminotransferase interconverts aspartate and α-ketoglutarate to oxaloacetate and glutamate, while links to pyruvate, indirectly supporting oxaloacetate formation via . These reactions allow to serve as both anaplerotic substrates and cataplerotic products, integrating nitrogen and carbon fluxes with the cycle.70110-9/fulltext)

Inputs from Glycolysis and Other Routes

The citric acid cycle receives its primary carbon input as , which is generated from the catabolism of carbohydrates, lipids, and proteins through various upstream pathways. From , the end product pyruvate is transported into the mitochondria, where it undergoes oxidative catalyzed by the (PDC). This multienzyme complex facilitates the irreversible reaction: pyruvate + CoA + NAD⁺ → + CO₂ + NADH, linking directly to the cycle by providing the two-carbon acetyl unit for condensation with oxaloacetate to form citrate. In conditions of metabolism or high glycolytic flux, such as in during intense exercise, accumulates and is released into the bloodstream. Circulating can be taken up by other tissues, particularly the liver and heart, where (LDH) reversibly converts it back to pyruvate: + NAD⁺ ⇌ pyruvate + NADH. The regenerated pyruvate then enters the mitochondria to form via PDC, allowing to serve as an indirect fuel source for the citric acid cycle in these extrahepatic tissues. Lipid catabolism contributes acetyl-CoA through β-oxidation of fatty acids in the . Long-chain fatty acids are activated to fatty in the , transported across the via the carnitine shuttle, and sequentially shortened by two carbons per cycle of β-oxidation, yielding units that directly enter the citric acid cycle. Each round of β-oxidation also produces NADH and FADH₂, enhancing the cycle's reducing power. For even-chain fatty acids, all carbons are converted to ; however, odd-chain fatty acids leave a three-carbon propionyl-CoA remnant, which is carboxylated to D-methylmalonyl-CoA, isomerized to L-methylmalonyl-CoA, and rearranged by (vitamin B₁₂-dependent) to , an intermediate that feeds into the cycle. Protein breakdown supplies and other precursors via . Ketogenic , such as and , are degraded to or acetoacetyl-CoA, entering the without net production of glucose precursors. Glucogenic , like and aspartate, yield pyruvate or intermediates: is transaminated to pyruvate, which proceeds to via PDC, while aspartate is converted to oxaloacetate through , replenishing the cycle's catalytic pool. and are both ketogenic and glucogenic, producing both and . These pathways ensure carbons integrate into the based on their carbon skeletons./25:_Protein_and_Amino_Acid_Metabolism/25.05:Amino_Acid_Catabolism-_The_Carbon_Atoms) During fasting or prolonged exercise, produced in the liver from excess provide an alternative input to extrahepatic tissues. Acetoacetate and β-hydroxybutyrate are released into circulation; in peripheral tissues like muscle and , β-hydroxybutyrate is oxidized to acetoacetate by β-hydroxybutyrate dehydrogenase, and acetoacetate is activated to acetoacetyl-CoA by succinyl-CoA:acetoacetate CoA transferase (using from the cycle). Acetoacetyl-CoA is then cleaved by to two molecules of , which enter the citric acid cycle to sustain energy production. Quantitatively, complete oxidation of one glucose molecule via yields two pyruvates, each producing one , resulting in two turns of the citric acid cycle and the release of two CO₂ from PDC. This underscores the cycle's role in fully oxidizing the six carbons of glucose, with the remaining four entering as two acetyl groups.

Biosynthetic Diversions from Intermediates

The tricarboxylic acid () cycle, also known as the citric acid cycle, not only generates but also supplies key intermediates for biosynthetic pathways essential for cellular and . These intermediates are diverted from the cycle through specific enzymatic reactions, allowing the synthesis of , , , and other biomolecules. Such diversions are particularly prominent in metabolically active tissues where biosynthetic demands are high, ensuring a balance between catabolic and anabolic processes. α-Ketoglutarate, an intermediate in the TCA cycle, serves as a central precursor for the of several via reactions. It is converted to glutamate by or through with or aspartate, catalyzed by glutamate-oxaloacetate or glutamate-pyruvate . Glutamate then acts as a nitrogen donor for the of via , which adds in an ATP-dependent manner; is crucial for and transport. Additionally, glutamate is used to produce through pyrroline-5-carboxylate synthetase and Δ1-pyrroline-5-carboxylate reductase, and it contributes to via the intermediates and . These pathways highlight α-ketoglutarate's role in and homeostasis. Oxaloacetate, another TCA cycle intermediate, is primarily diverted for aspartate family and synthesis. Through with glutamate, catalyzed by aspartate aminotransferase, oxaloacetate forms aspartate, which is then amidated to by asparagine synthetase using as the nitrogen source. Aspartate also serves as a precursor for , , , and in plants and , though in mammals, it links to these via additional pathways. Furthermore, aspartate condenses with —generated from , CO2, and ATP by —to form carbamoyl aspartate, the first step in leading to UMP and other . These diversions underscore oxaloacetate's importance in nitrogen-containing production. Succinyl-CoA, formed from α-ketoglutarate in the cycle, is a critical precursor for . It condenses with in a pyridoxal phosphate-dependent reaction catalyzed by δ-aminolevulinic acid synthase (ALAS), the rate-limiting enzyme of the heme pathway, to produce δ-aminolevulinic acid (), CO2, and . Two molecules of ALA then condense to form porphobilinogen, which progresses through several steps to yield , ultimately incorporating iron to form . This diversion is vital for , , and production, with ALAS activity regulated by heme feedback inhibition. Citrate, the first TCA cycle intermediate, is exported from mitochondria via the citrate carrier to the , where it is cleaved by ATP-citrate lyase into and oxaloacetate, consuming ATP and . The resulting serves as the primary substrate for and biosynthesis, fueling in tissues like liver and adipose. Cytosolic oxaloacetate is reduced to malate by , regenerating NAD+ and allowing malate to re-enter mitochondria or contribute to other pathways. This citrate-malate shuttle links mitochondrial to cytosolic lipid synthesis, particularly under nutrient-rich conditions. Fumarate and malate, late TCA cycle intermediates, participate in the and . In the , argininosuccinate lyase cleaves argininosuccinate to and fumarate; the fumarate is hydrated to malate by cytosolic and oxidized to oxaloacetate by , integrating urea synthesis with TCA flux. For , malate exits mitochondria via the malate-aspartate shuttle, where it is converted to oxaloacetate in the , then decarboxylated by to phosphoenolpyruvate, initiating glucose production primarily in and . These roles connect and synthesis. In proliferating cells, such as hepatocytes in the liver during regeneration or tumor cells, a significant portion of intermediates—up to half in some estimates—is diverted toward to support rapid growth, with often serving as a major carbon and source to replenish the cycle in cancer contexts. This metabolic flexibility ensures sustained flux despite high anabolic demands.

Evolutionary Perspectives

Origins in Early Life

The prebiotic roots of the citric acid cycle trace to geochemical processes in environments, particularly alkaline hydrothermal vents, where reverse tricarboxylic acid (rTCA) cycle reactions could have facilitated the abiotic of key intermediates. In these settings, (CO₂) and (H₂S) from geochemical cycles served as precursors, enabling non-enzymatic formation of molecules like pyruvate and through mineral-catalyzed reactions involving iron-sulfur clusters. The pathway, observed in modern acetogens, is proposed as a analog, linking CO₂ reduction to production under , vent-like conditions with pH gradients driving proton motive force. Autotrophic origins of the cycle emerged in primitive microbes, where the reductive TCA cycle enabled carbon fixation by assimilating CO₂ into organic compounds. Green sulfur bacteria such as those in the Chlorobiaceae family utilize this pathway for autotrophic growth, employing enzymes like ATP-citrate lyase to reverse key oxidative steps, suggesting an ancient adaptation in . Genomic analyses indicate that the (LUCA) possessed core TCA enzymes, including subunits of oxoglutarate oxidoreductase and , supporting the cycle's presence in early cellular life as an , CO₂-fixing . The citric acid cycle likely emerged between 3.5 and 4 billion years ago, coinciding with the advent of microbial life in oceans, as inferred from phylogenetic reconstructions of LUCA's . preserved in approximately 3.5 billion-year-old provides evidence of early in prokaryotes, with the —relying on from intermediates—proposed as a key route for isoprenoid production in ancient , though the oldest confirmed isoprenoid biomarkers date to around 1.6 billion years ago. Hypotheses emphasize geochemical gradients in hydrothermal vents as drivers of citrate formation, with and disparities promoting protometabolic cycles akin to oxidative of glyoxylate. Recent simulations in the , including a study, have recreated these conditions, demonstrating abiotic production of all intermediates from simple precursors like CO₂ and under irradiation (simulated via electron bombardment of interstellar ices), mimicking prebiotic vent chemistry.

Comparative Evolution in Eukaryotes and Prokaryotes

The citric acid cycle exhibits considerable diversity in prokaryotes, reflecting adaptations to varied metabolic niches. In many bacteria, particularly methylotrophs, branched pathways such as the ethylmalonyl-CoA pathway replace or supplement the standard cycle to assimilate one-carbon compounds, converting acetyl-CoA to glyoxylate without relying on the glyoxylate shunt, as observed in Methylobacterium extorquens AM1 during growth on acetate. Obligate anaerobes often operate incomplete versions of the cycle, such as the reductive branch in methanogens like Methanococcus maripaludis, which supports carbon fixation and energy generation under oxygen-free conditions without full oxidation. These variations highlight the cycle's plasticity in prokaryotes, enabling survival in environments ranging from aerobic soils to anaerobic sediments, and are frequently shaped by horizontal gene transfer events that introduce or modify key enzymes across bacterial lineages. In eukaryotes, the cycle was acquired through endosymbiosis of an α-proteobacterium approximately 1.5–2 billion years ago, which provided the ancestral mitochondrial machinery for oxidative metabolism. Following this event, extensive gene transfer from the endosymbiont to the host nucleus occurred, resulting in most TCA cycle enzymes—such as citrate synthase, isocitrate dehydrogenase, and α-ketoglutarate dehydrogenase—being nuclear-encoded and targeted back to mitochondria via import signals. This relocation streamlined eukaryotic gene regulation while preserving the cycle's core function. Compartmentalization within the mitochondrial matrix further enhanced efficiency by concentrating intermediates and enzymes, minimizing diffusion losses and integrating the cycle tightly with the electron transport chain for ATP production. Evolutionary adaptations in eukaryotes include lineage-specific modifications, such as the loss of the glyoxylate shunt in , which prevents net carbon assimilation from and commits the cycle primarily to generation, whereas and fungi retain this shunt for from during or . Phylogenetic analyses reveal high sequence conservation of enzymes across domains, with showing up to 50–70% identity between prokaryotic and eukaryotic orthologs, underscoring the cycle's ancient origins while allowing domain-specific divergences. Recent metagenomic studies from 2020–2025 have uncovered novel variants in uncultured microbes, revealing branched or incomplete cycles in diverse environments like deep-sea sediments and host-associated communities, which expand our understanding of prokaryotic diversity beyond cultured representatives. These findings also illuminate the cycle's role in , where the endosymbiont's efficient TCA-driven energy production met the high ATP demands of the emerging , facilitating complex cellular processes and integration.

References

  1. [1]
    Biochemistry, Citric Acid Cycle - StatPearls - NCBI Bookshelf - NIH
    The citric acid cycle serves as the mitochondrial hub for the final steps in carbon skeleton oxidative catabolism for carbohydrates, amino acids, and fatty ...
  2. [2]
    Physiology, Krebs Cycle - StatPearls - NCBI Bookshelf
    Nov 23, 2022 · The tricarboxylic acid (TCA) cycle, also known as the Krebs or citric acid cycle, is an important cell's metabolic hub.Introduction · Issues of Concern · Cellular Level · Function
  3. [3]
  4. [4]
  5. [5]
    [PDF] Tricarboxylic acid cycle
    The TCA cycle, also known as the citric acid cycle or the Krebs cycle, is a cyclic series of enzymatically catalyzed reactions, carried out by a multienzyme ...
  6. [6]
    Citric Acid - an overview | ScienceDirect Topics
    Citric acid was first isolated in 1784 by Scheele, who precipitated it as calcium citrate by adding calcium hydroxide to lemon juice. Prior to 1920, citric acid ...
  7. [7]
    Glycolysis - PMC - NIH
    Glycolysis was the first metabolic pathway elucidated and is also referred to as the Embden–Meyerhof–Parnas pathway (see Box 1). The word “glycolysis” is ...
  8. [8]
    Metabolism of ketonic acids in animal tissues | Biochemical Journal
    Get Permissions. Citation. Hans Adolf Krebs, William Arthur Johnson; Metabolism of ketonic acids in animal tissues. Biochem J 1 April 1937; 31 (4): 645–660.
  9. [9]
    [PDF] Hans A. Krebs - Nobel Lecture
    It was from such considerations that the term "citric acid cycle" was proposed in 1937. The evidence in support of the cycle mentioned so far comes under two.
  10. [10]
    Hans Krebs – Facts - NobelPrize.org
    ... Hans Krebs was able to present a complete picture of an important part of metabolism—the citric acid cycle. In this process, which is cyclical and has ...
  11. [11]
  12. [12]
    Citric Acid Cycle | Pathway - PubChem - NIH
    May 31, 2019 · The citric acid cycle was discovered in 1937 by Hans Adolf Krebs while he worked at the University of Sheffield in England (PMID: 16746382). ...
  13. [13]
    How Cells Obtain Energy from Food - NCBI - NIH
    Simple overview of the citric acid cycle. The reaction of acetyl CoA with oxaloacetate starts the cycle by producing citrate (citric acid). In each turn of the ...
  14. [14]
    Biochemistry, Electron Transport Chain - StatPearls - NCBI Bookshelf
    Sep 4, 2023 · The acetyl CoA is then used in the citric acid cycle, which is a chain of chemical reactions that produce CO2, NADH, flavin adenine dinucleotide ...
  15. [15]
    Thermodynamic Constraints on the Citric Acid Cycle and Related ...
    We quantify the thermodynamic viability and energetics of various reactions of interest to metabolism at pressures and temperatures relevant to ocean worlds.
  16. [16]
    [PDF] The Thermodynamics of the Krebs Cycle and Related Compounds
    Oct 15, 2009 · The Krebs cycle or citric acid cycle is the major path- way in aerobic organisms for the oxidation of sugars, fatty acids and some amino acids.
  17. [17]
    Metabolic Energy - The Cell - NCBI Bookshelf - NIH
    The standard free-energy change (ΔG°) of a reaction therefore determines its chemical equilibrium and predicts in which direction the reaction will proceed ...
  18. [18]
    The Overall Efficiency of Oxidative Phosphorylation – BIOC*2580
    ATP yield from complete oxidation of glucose. Thermodynamic efficiency of Glucose oxidation. Glucose: C6H12O6 + 6 O2 → 6 CO2 + 6 H2O. ΔG˚ = -2,840 kJ/mol. 32 ...
  19. [19]
    Towards an evolutionary theory of the origin of life based on kinetics ...
    A much higher potential (ca 150 kJ mol–1 [42,43]) than the free energy ... citric acid cycle [66] and the efficiency of the formation of a variety of ...
  20. [20]
    Regulation and function of the mammalian tricarboxylic acid cycle
    The TCA cycle plays a central role in theories for the chemical origins of life because it supplies acetyl-CoA, pyruvate, OAA, succinate, and alpha- ...
  21. [21]
    Citric Acid Cycle | Pathway - PubChem - NIH
    May 29, 2019 · The citric acid cycle is regulated in a number of ways but the primary mechanism is by product inhibition. For instance, NADH inhibits pyruvate ...
  22. [22]
    Citric Acid Cycle Regulation - News-Medical
    Because citrate synthase is inhibited by the final product of the citric acid cycle as ATP, ADP (adenosine diphosphate) works as an allosteric activator of the ...
  23. [23]
    Isocitrate Dehydrogenase (NAD) - an overview | ScienceDirect Topics
    In mammals, Ca2+ activates NAD-isocitrate dehydrogenase by reducing the Km for the substrate threo-ds-isocitrate [94,104,105].Abbreviations · Iii Subcellular Calcium... · A Steady-State Distribution<|control11|><|separator|>
  24. [24]
    Physiology, Krebs Cycle - PubMed
    Nov 23, 2022 · Regulation of the TCA cycle occurs at 3 distinct points, including the following enzymes: citrate synthase, isocitrate dehydrogenase, and alpha ...
  25. [25]
    Pyruvate dehydrogenase kinases (PDKs): an overview toward ...
    Under normal conditions, pyruvate dehydrogenase phosphatase inhibits PDK and activates PDC, which then catalyzes pyruvate in the tricarboxylic acid cycle to ...Pdks And Cancer · Table 1. Pyruvate... · Pdks Inhibitors
  26. [26]
    Studies on the activation of rat liver pyruvate dehydrogenase and 2 ...
    Studies on the activation of rat liver pyruvate dehydrogenase and 2-oxoglutarate dehydrogenase by adrenaline and glucagon. Role of increases in ...Missing: dephosphorylation | Show results with:dephosphorylation
  27. [27]
    Mitochondrial TCA cycle metabolites control physiology and disease
    Jan 3, 2020 · Another intrinsic regulator is succinyl-CoA, which inhibits both citrate synthase and α-KG dehydrogenase to slow the cycle down. Likewise ...
  28. [28]
    The Citric Acid Cycle and Fatty Acid Biosynthesis - NCBI - NIH
    The oxidative decarboxylation of pyruvate is an important reaction in archaea, bacteria, and eukaryotes alike, generating acetyl-CoA necessary for CAC reactions ...Missing: net | Show results with:net
  29. [29]
    Revisiting the glyoxylate cycle: alternate pathways for microbial ...
    Jun 15, 2006 · The glyoxylate cycle, identified by Kornberg et al. in 1957, provides a simple and efficient strategy for converting acetyl-CoA into anapleurotic and ...
  30. [30]
    glyoxylate cycle | Pathway - PubChem - NIH
    The glyoxylate cycle, found in fungi, plants, and bacteria, is essential for growth on two-carbon compounds and provides precursors for biosynthesis. It ...
  31. [31]
    Oxidation of Pyruvate and the Citric Acid Cycle – Biology
    Steps three and four are both oxidation and decarboxylation steps, which release electrons that reduce NAD+ to NADH and release carboxyl groups that form CO2 ...
  32. [32]
    Both forward and reverse TCA cycles operate in green sulfur bacteria
    Nov 12, 2010 · The anoxygenic green sulfur bacteria (GSBs) assimilate CO(2) autotrophically through the reductive (reverse) tricarboxylic acid (RTCA) cycle ...
  33. [33]
    Genome-scale metabolic modeling of the human milk ...
    Feb 16, 2024 · The production of lactate and ethanol is useful for NADH regeneration (11), and the TCA cycle is incomplete in most Bifidobacterium (12). In ...
  34. [34]
    Biochemistry and Evolution of Anaerobic Energy Metabolism in ...
    The Krebs cycle is incomplete and is likely used in the reductive direction (177). ... Electron-transfer flavoprotein from anaerobic Ascaris suum mitochondria and ...
  35. [35]
    Re-examining the role of the glyoxylate cycle in oilseeds - PubMed
    After germination, this reserve is mobilized in order to support growth during early seedling development. The glyoxylate cycle is instrumental in this ...
  36. [36]
    Postgerminative growth and lipid catabolism in oilseeds lacking the ...
    The glyoxylate cycle is regarded as essential for postgerminative growth and seedling establishment in oilseed plants. We have identified two allelic ...
  37. [37]
    Structure determination, thermal stability and catalytic mechanism of ...
    Sep 2, 2024 · IDH is an enzyme in the citric-acid cycle and is widely distributed in the three domains of life: Archaea, Bacteria and Eukarya. It ...
  38. [38]
    Ins and Outs of the TCA Cycle: The Central Role of Anaplerosis
    Oct 11, 2021 · Liver glutamate dehydrogenase controls whole-body energy partitioning through amino acid–derived gluconeogenesis and ammonia homeostasis ...
  39. [39]
    The Glutamate Dehydrogenase Pathway and Its Roles in Cell and ...
    Glutamate dehydrogenase (GDH) is a hexameric enzyme that catalyzes the reversible conversion of glutamate to α-ketoglutarate and ammonia while reducing NAD(P)+ ...
  40. [40]
    Phosphoenolpyruvate Carboxykinase as the Sole Anaplerotic ...
    This study investigates whether and under which conditions PEPCK can replace the anaplerotic function of pyruvate carboxylase in S. cerevisiae.
  41. [41]
    Pyruvate Dehydrogenase and Pyruvate Dehydrogenase ... - NIH
    Pyruvate dehydrogenase (PDH) catalyzes the conversion of pyruvate to acetyl-coenzyme A, which enters into the Krebs cycle, providing adenosine triphosphate (ATP) ...
  42. [42]
    Lactate: the ugly duckling of energy metabolism - PMC - NIH
    Jul 20, 2020 · This process requires lactate dehydrogenase (LDH) and monocarboxylic transporter (MCT) activity. ... Once inside cells, pyruvate and lactate are ...
  43. [43]
    Lactate is always the end product of glycolysis - Frontiers
    Feb 26, 2015 · Transport across the mitochondrial inner membrane with subsequent conversion to Acetyl-CoA via the pyruvate dehydrogenase (PDH) reaction ...
  44. [44]
    Biochemistry, Fatty Acid Oxidation - StatPearls - NCBI Bookshelf - NIH
    Mitochondrial beta-oxidation can be used to supply acetyl coenzyme A (CoA) to 2 separate pathways, depending on which tissue oxidation occurs. In skeletal ...
  45. [45]
    Propionyl-CoA catabolism - Reactome Pathway Database
    The three reactions of this pathway convert propionyl-CoA to succinyl-CoA, an intermediate of the citric acid cycle.
  46. [46]
    Amino Acid Catabolism - The Medical Biochemistry Page
    Glucogenic amino acids are those that give rise to a net production of pyruvate or TCA cycle intermediates, such as 2-oxoglutarate (α-ketoglutarate) or ...<|separator|>
  47. [47]
    Ketones and the Heart: Metabolic Principles and Therapeutic ...
    Mar 30, 2023 · The ketone bodies beta-hydroxybutyrate and acetoacetate are hepatically produced metabolites catabolized in extrahepatic organs.
  48. [48]
    Ketone bodies: from enemy to friend and guardian angel
    Dec 9, 2021 · Secreted βOHB and acetoacetate are taken up by extrahepatic cells and converted back to acetyl-CoA. The latter can be entered into the TCA ...
  49. [49]
    Alpha-Ketoglutarate: Physiological Functions and Applications - PMC
    Jan 1, 2016 · It is a nitrogen scavenger and a source of glutamate and glutamine that stimulates protein synthesis and inhibits protein degradation in muscles ...
  50. [50]
    Supporting aspartate biosynthesis is an essential function of ... - NIH
    For instance, pyruvate carboxylation produces oxaloacetate that can be transaminated to form aspartate. Alternatively, reductive AKG carboxylation can generate ...
  51. [51]
    5-AMINOLEVULINATE SYNTHASE: CATALYSIS OF THE FIRST ...
    The reaction cycle involves condensation of glycine with succinyl-coenzyme A to yield 5-aminolevulinate, carbon dioxide, and CoA. Mutations in the human ...
  52. [52]
    A Role for the Krebs Cycle Intermediate Citrate in Metabolic ...
    Feb 5, 2018 · Citrate Provides a Bridge Between Carbohydrate and Fatty Acid Metabolism. Citrate is produced in the Krebs cycle (also known as the citric acid ...Missing: paper | Show results with:paper
  53. [53]
    Amino Acid Metabolism - PMC - PubMed Central - NIH
    The other molecule that is recycled here is aspartate. The fumarate generated in the urea cycle in the cytosol is converted by cytosolic fumarase to malate, ...Metabolic Pathways That... · The Urea Cycle Is Necessary... · Figure 5
  54. [54]
    Cancer Cell Metabolism: One Hallmark, Many Faces - PMC
    Proliferating cells are able to sustain the TCA cycle by replenishing these depleted intermediates through a process called anaplerosis (25). It is well ...
  55. [55]
    Glutamine Metabolism in Cancer - NCBI - NIH
    May 21, 2021 · Cancer cells use precursors derived from the TCA cycle intermediates to synthesize proteins, lipids, and nucleic acids. In order to maintain ...
  56. [56]
    Origin of the Reductive Tricarboxylic Acid (rTCA) Cycle-Type CO2 ...
    Our view supports the theory of an autotrophic origin of life from primordial carbon assimilation within a sulfide-rich hydrothermal vent.
  57. [57]
    On the origin of biochemistry at an alkaline hydrothermal vent
    A model for the origin of biochemistry at an alkaline hydrothermal vent has been developed that focuses on the acetyl-CoA (Wood–Ljungdahl) pathway of CO2 ...Missing: TCA | Show results with:TCA
  58. [58]
    The reductive tricarboxylic acid cycle of carbon dioxide assimilation
    Aug 1, 1997 · One of the key enzymes, ATP-citrate lyase, was purified to apparent homogeneity from the moderately thermophilic green sulfur bacterium Chlorobium tepidum.Missing: primitive Chlorobiaceae LUCA citric
  59. [59]
    The nature of the last universal common ancestor and its impact on ...
    Jul 12, 2024 · LUCA was probably capable of gluconeogenesis/glycolysis in that we find support for most subunits of enzymes involved in these pathways ( ...
  60. [60]
  61. [61]
    The Ethylmalonyl-CoA Pathway Is Used in Place of the Glyoxylate ...
    Significance: Tight coordination must exist for operation of the citric acid cycle in conjunction with the ethylmalonyl-CoA pathway.Missing: diversity | Show results with:diversity
  62. [62]
    MetaCyc incomplete reductive TCA cycle - BioCyc
    Enzymes of an incomplete reductive TCA cycle shown here have been experimentally demonstrated in the autotrophic methanogen Methanococcus maripaludis [Shieh87].
  63. [63]
    Role of horizontal gene transfers and microbial ecology in the ...
    Apr 12, 2023 · In this study, we examined the evolution of key enzymes in the rTCA, which are rare in extant organisms, occurring in a few groups of Bacteria and Archaea.
  64. [64]
    (PDF) The origin and early evolution of mitochondria - ResearchGate
    Aug 7, 2025 · These organelles, descended from ancient α-proteobacteria through endosymbiosis approximately 1.5-2 billion years ago (Gray et al., 2001) ...Abstract And Figures · References (48) · Recommended Publications
  65. [65]
    Origin and Evolution of the Mitochondrial Proteome - PubMed Central
    The three identifiable ancestral sources to the proteome of mitochondria are proteins descended from the ancestral α-proteobacteria symbiont.Missing: citric | Show results with:citric
  66. [66]
    Principles and functions of metabolic compartmentalization - PMC
    Oct 20, 2022 · Compartmentalization promotes metabolic efficiency by enhancing the physical proximity of components in functionally related pathways and ...
  67. [67]
    Evolution of glyoxylate cycle enzymes in Metazoa - PubMed Central
    Oct 23, 2006 · The glyoxylate cycle is thought to be present in bacteria, protists, plants, fungi, and nematodes, but not in other Metazoa.Missing: retention | Show results with:retention
  68. [68]
    Evolution of the enzymes of the citric acid cycle and the glyoxylate ...
    Feb 14, 2002 · Here, we investigate the individual evolutionary histories of all of the enzymes of the tricarboxylic acid cycle and the glyoxylate cycle using protein maximum ...
  69. [69]
    Opportunities and challenges of using metagenomic data to bring ...
    May 12, 2022 · In this paper, we discuss a range of successful methodologies and applications that have underpinned recent metagenome-guided isolation and cultivation of ...Introduction · The Uncultured Majority · Culture Medium Optimization
  70. [70]
    TCA cycle signalling and the evolution of eukaryotes - PubMed Central
    Oct 30, 2020 · We propose an evolutionary concept in which mitochondrial TCA cycle signalling was also a crucial player during eukaryogenesis enabling the dynamic control of ...