Metabolism

Source: Wikipedia, the free encyclopedia.

Simplified view of the cellular metabolism
Structure of adenosine triphosphate (ATP), a central intermediate in energy metabolism

Metabolism (

structures, and respond to their environments. The word metabolism can also refer to the sum of all chemical reactions that occur in living organisms, including digestion
and the transportation of substances into and between different cells, in which case the above described set of reactions within the cells is called intermediary (or intermediate) metabolism.

Metabolic reactions may be categorized as

anabolic – the building up (synthesis
) of compounds (such as proteins, carbohydrates, lipids, and nucleic acids). Usually, catabolism releases energy, and anabolism consumes energy.

The chemical reactions of metabolism are organized into metabolic pathways, in which one chemical is transformed through a series of steps into another chemical, each step being facilitated by a specific enzyme. Enzymes are crucial to metabolism because they allow organisms to drive desirable reactions that require energy and will not occur by themselves, by coupling them to spontaneous reactions that release energy. Enzymes act as catalysts – they allow a reaction to proceed more rapidly – and they also allow the regulation of the rate of a metabolic reaction, for example in response to changes in the cell's environment or to signals from other cells.

The metabolic system of a particular organism determines which substances it will find nutritious and which poisonous. For example, some prokaryotes use hydrogen sulfide as a nutrient, yet this gas is poisonous to animals.[1] The basal metabolic rate of an organism is the measure of the amount of energy consumed by all of these chemical reactions.

A striking feature of metabolism is the similarity of the basic metabolic pathways among vastly different species.

type II diabetes, metabolic syndrome, and cancer, normal metabolism is disrupted.[6] The metabolism of cancer cells is also different from the metabolism of normal cells, and these differences can be used to find targets for therapeutic intervention in cancer.[7]

Key biochemicals

triacylglycerol
lipid
This is a diagram depicting a large set of human metabolic pathways.[image reference needed]

Most of the structures that make up animals, plants and microbes are made from four basic classes of

macromolecules of life.[8]

Type of molecule Name of monomer forms Name of polymer forms Examples of polymer forms
Amino acids Amino acids Proteins (made of polypeptides) Fibrous proteins and globular proteins
Carbohydrates Monosaccharides Polysaccharides Starch, glycogen and cellulose
Nucleic acids Nucleotides Polynucleotides DNA and RNA

Amino acids and proteins

Proteins are made of

tricarboxylic acid cycle),[11] especially when a primary source of energy, such as glucose, is scarce, or when cells undergo metabolic stress.[12]

Lipids

Lipids are the most diverse group of biochemicals. Their main structural uses are as part of

organic solvents such as ethanol, benzene or chloroform.[13] The fats are a large group of compounds that contain fatty acids and glycerol; a glycerol molecule attached to three fatty acids by ester linkages is called a triacylglyceride.[14] Several variations on this basic structure exist, including backbones such as sphingosine in sphingomyelin, and hydrophilic groups such as phosphate as in phospholipids. Steroids such as sterol are another major class of lipids.[15]

Carbohydrates

The straight chain form consists of four C H O H groups linked in a row, capped at the ends by an aldehyde group C O H and a methanol group C H 2 O H. To form the ring, the aldehyde group combines with the O H group of the next-to-last carbon at the other end, just before the methanol group.
Glucose can exist in both a straight-chain and ring form.

Carbohydrates are

hydroxyl groups attached, that can exist as straight chains or rings. Carbohydrates are the most abundant biological molecules, and fill numerous roles, such as the storage and transport of energy (starch, glycogen) and structural components (cellulose in plants, chitin in animals).[10] The basic carbohydrate units are called monosaccharides and include galactose, fructose, and most importantly glucose. Monosaccharides can be linked together to form polysaccharides in almost limitless ways.[16]

Nucleotides

The two nucleic acids, DNA and

heterocyclic rings containing nitrogen, classified as purines or pyrimidines. Nucleotides also act as coenzymes in metabolic-group-transfer reactions.[18]

Coenzymes

acetyl group
is bonded to the sulfur atom at the extreme left.

Metabolism involves a vast array of chemical reactions, but most fall under a few basic types of reactions that involve the transfer of

substrate for a set of enzymes that produce it, and a set of enzymes that consume it. These coenzymes are therefore continuously made, consumed and then recycled.[20]

One central coenzyme is adenosine triphosphate (ATP), the energy currency of cells. This nucleotide is used to transfer chemical energy between different chemical reactions. There is only a small amount of ATP in cells, but as it is continuously regenerated, the human body can use about its own weight in ATP per day.[20] ATP acts as a bridge between catabolism and anabolism. Catabolism breaks down molecules, and anabolism puts them together. Catabolic reactions generate ATP, and anabolic reactions consume it. It also serves as a carrier of phosphate groups in phosphorylation reactions.[21]

A

reductases in the cell that need to transfer hydrogen atoms to their substrates.[23] Nicotinamide adenine dinucleotide exists in two related forms in the cell, NADH and NADPH. The NAD+/NADH form is more important in catabolic reactions, while NADP+/NADPH is used in anabolic reactions.[24]

The structure of iron-containing hemoglobin. The protein subunits are in red and blue, and the iron-containing heme groups in green. From PDB: 1GZX​.

Mineral and cofactors

Inorganic elements play critical roles in metabolism; some are abundant (e.g. sodium and potassium) while others function at minute concentrations. About 99% of a human's body weight is made up of the elements carbon, nitrogen, calcium, sodium, chlorine, potassium, hydrogen, phosphorus, oxygen and sulfur. Organic compounds (proteins, lipids and carbohydrates) contain the majority of the carbon and nitrogen; most of the oxygen and hydrogen is present as water.[25]

The abundant inorganic elements act as

ion gradients across cell membranes maintains osmotic pressure and pH.[26] Ions are also critical for nerve and muscle function, as action potentials in these tissues are produced by the exchange of electrolytes between the extracellular fluid and the cell's fluid, the cytosol.[27] Electrolytes enter and leave cells through proteins in the cell membrane called ion channels. For example, muscle contraction depends upon the movement of calcium, sodium and potassium through ion channels in the cell membrane and T-tubules.[28]

Transition metals are usually present as trace elements in organisms, with zinc and iron being most abundant of those.[29] Metal cofactors are bound tightly to specific sites in proteins; although enzyme cofactors can be modified during catalysis, they always return to their original state by the end of the reaction catalyzed. Metal micronutrients are taken up into organisms by specific transporters and bind to storage proteins such as ferritin or metallothionein when not in use.[30][31]

Catabolism

Catabolism is the set of metabolic processes that break down large molecules. These include breaking down and oxidizing food molecules. The purpose of the catabolic reactions is to provide the energy and components needed by anabolic reactions which build molecules.

organic molecules that are broken down to simpler molecules, such as carbon dioxide and water. Photosynthetic organisms, such as plants and cyanobacteria, use similar electron-transfer reactions to store energy absorbed from sunlight.[34]

Classification of organisms based on their metabolism [35]
Energy source sunlight photo-   -troph
molecules chemo-
Hydrogen or electron donor organic compound   organo-  
inorganic compound litho-
Carbon source organic compound   hetero-
inorganic compound auto-

The most common set of catabolic reactions in animals can be separated into three main stages. In the first stage, large organic molecules, such as proteins, polysaccharides or lipids, are digested into their smaller components outside cells. Next, these smaller molecules are taken up by cells and converted to smaller molecules, usually acetyl coenzyme A (acetyl-CoA), which releases some energy. Finally, the acetyl group on acetyl-CoA is oxidized to water and carbon dioxide in the citric acid cycle and electron transport chain, releasing more energy while reducing the coenzyme nicotinamide adenine dinucleotide (NAD+) into NADH.[32]

Digestion

Macromolecules cannot be directly processed by cells. Macromolecules must be broken into smaller units before they can be used in cell metabolism. Different classes of enzymes are used to digest these polymers. These

monosaccharides.[36]

Microbes simply secrete digestive enzymes into their surroundings,[37][38] while animals only secrete these enzymes from specialized cells in their guts, including the stomach and pancreas, and in salivary glands.[39] The amino acids or sugars released by these extracellular enzymes are then pumped into cells by active transport proteins.[40][41]

A simplified outline of the catabolism of proteins, carbohydrates and fats[image reference needed]

Energy from organic compounds

Carbohydrate catabolism is the breakdown of carbohydrates into smaller units. Carbohydrates are usually taken into cells after they have been digested into

NADPH and produces pentose sugars such as ribose, the sugar component of nucleic acids.[citation needed
]

Carbon Catabolism pathway map for free energy including carbohydrate and lipid sources of energy

Fats are catabolized by hydrolysis to free fatty acids and glycerol. The glycerol enters glycolysis and the fatty acids are broken down by beta oxidation to release acetyl-CoA, which then is fed into the citric acid cycle. Fatty acids release more energy upon oxidation than carbohydrates. Steroids are also broken down by some bacteria in a process similar to beta oxidation, and this breakdown process involves the release of significant amounts of acetyl-CoA, propionyl-CoA, and pyruvate, which can all be used by the cell for energy. M. tuberculosis can also grow on the lipid cholesterol as a sole source of carbon, and genes involved in the cholesterol-use pathway(s) have been validated as important during various stages of the infection lifecycle of M. tuberculosis.[45]

glutamate.[47] The glucogenic amino acids can also be converted into glucose, through gluconeogenesis (discussed below).[48]

Energy transformations

Oxidative phosphorylation

In oxidative phosphorylation, the electrons removed from organic molecules in areas such as the citric acid cycle are transferred to oxygen and the energy released is used to make ATP. This is done in eukaryotes by a series of proteins in the membranes of mitochondria called the electron transport chain. In prokaryotes, these proteins are found in the cell's inner membrane.[49] These proteins use the energy from reduced molecules like NADH to pump protons across a membrane.[50]

Mechanism of ATP synthase. ATP is shown in red, ADP and phosphate in pink and the rotating stalk subunit in black.

Pumping protons out of the mitochondria creates a proton concentration difference across the membrane and generates an electrochemical gradient.[51] This force drives protons back into the mitochondrion through the base of an enzyme called ATP synthase. The flow of protons makes the stalk subunit rotate, causing the active site of the synthase domain to change shape and phosphorylate adenosine diphosphate – turning it into ATP.[20]

Energy from inorganic compounds

soil fertility.[56][57]

Energy from light

The energy in sunlight is captured by plants, cyanobacteria, purple bacteria, green sulfur bacteria and some protists. This process is often coupled to the conversion of carbon dioxide into organic compounds, as part of photosynthesis, which is discussed below. The energy capture and carbon fixation systems can, however, operate separately in prokaryotes, as purple bacteria and green sulfur bacteria can use sunlight as a source of energy, while switching between carbon fixation and the fermentation of organic compounds.[58][59]

In many organisms, the capture of solar energy is similar in principle to oxidative phosphorylation, as it involves the storage of energy as a proton concentration gradient. This proton motive force then drives ATP synthesis.[60] The electrons needed to drive this electron transport chain come from light-gathering proteins called photosynthetic reaction centres. Reaction centers are classified into two types depending on the nature of photosynthetic pigment present, with most photosynthetic bacteria only having one type, while plants and cyanobacteria have two.[61]

In plants, algae, and cyanobacteria, photosystem II uses light energy to remove electrons from water, releasing oxygen as a waste product. The electrons then flow to the cytochrome b6f complex, which uses their energy to pump protons across the thylakoid membrane in the chloroplast.[34] These protons move back through the membrane as they drive the ATP synthase, as before. The electrons then flow through photosystem I and can then be used to reduce the coenzyme NADP+.[62] This coenzyme can enter the Calvin cycle, which is discussed below, or be recycled for further ATP generation.[citation needed]

Anabolism

Anabolism is the set of constructive metabolic processes where the energy released by catabolism is used to synthesize complex molecules. In general, the complex molecules that make up cellular structures are constructed step-by-step from smaller and simpler precursors. Anabolism involves three basic stages. First, the production of precursors such as amino acids, monosaccharides, isoprenoids and nucleotides, secondly, their activation into reactive forms using energy from ATP, and thirdly, the assembly of these precursors into complex molecules such as proteins, polysaccharides, lipids and nucleic acids.[63]

Anabolism in organisms can be different according to the source of constructed molecules in their cells. Autotrophs such as plants can construct the complex organic molecules in their cells such as polysaccharides and proteins from simple molecules like carbon dioxide and water. Heterotrophs, on the other hand, require a source of more complex substances, such as monosaccharides and amino acids, to produce these complex molecules. Organisms can be further classified by ultimate source of their energy: photoautotrophs and photoheterotrophs obtain energy from light, whereas chemoautotrophs and chemoheterotrophs obtain energy from oxidation reactions.[63]

Carbon fixation

Plant cells (bounded by purple walls) filled with chloroplasts (green), which are the site of photosynthesis

Photosynthesis is the synthesis of carbohydrates from sunlight and

glycerate 3-phosphate, which can then be converted into glucose. This carbon-fixation reaction is carried out by the enzyme RuBisCO as part of the Calvin – Benson cycle.[64] Three types of photosynthesis occur in plants, C3 carbon fixation, C4 carbon fixation and CAM photosynthesis. These differ by the route that carbon dioxide takes to the Calvin cycle, with C3 plants fixing CO2 directly, while C4 and CAM photosynthesis incorporate the CO2 into other compounds first, as adaptations to deal with intense sunlight and dry conditions.[65]

In photosynthetic prokaryotes the mechanisms of carbon fixation are more diverse. Here, carbon dioxide can be fixed by the Calvin – Benson cycle, a reversed citric acid cycle,[66] or the carboxylation of acetyl-CoA.[67][68] Prokaryotic chemoautotrophs also fix CO2 through the Calvin–Benson cycle, but use energy from inorganic compounds to drive the reaction.[69]

Carbohydrates and glycans

In carbohydrate anabolism, simple organic acids can be converted into

glucose-6-phosphate through a series of intermediates, many of which are shared with glycolysis.[43] However, this pathway is not simply glycolysis run in reverse, as several steps are catalyzed by non-glycolytic enzymes. This is important as it allows the formation and breakdown of glucose to be regulated separately, and prevents both pathways from running simultaneously in a futile cycle.[70][71]

Although fat is a common way of storing energy, in

oxaloacetate, where it can be used for the production of glucose.[72][74] Other than fat, glucose is stored in most tissues, as an energy resource available within the tissue through glycogenesis which was usually being used to maintained glucose level in blood.[75]

Polysaccharides and

hydroxyl groups on the ring of the substrate can be acceptors, the polysaccharides produced can have straight or branched structures.[76] The polysaccharides produced can have structural or metabolic functions themselves, or be transferred to lipids and proteins by enzymes called oligosaccharyltransferases.[77][78]

Fatty acids, isoprenoids and sterol

shown. Some intermediates are omitted for clarity.

Fatty acids are made by fatty acid synthases that polymerize and then reduce acetyl-CoA units. The acyl chains in the fatty acids are extended by a cycle of reactions that add the acyl group, reduce it to an alcohol, dehydrate it to an alkene group and then reduce it again to an alkane group. The enzymes of fatty acid biosynthesis are divided into two groups: in animals and fungi, all these fatty acid synthase reactions are carried out by a single multifunctional type I protein,[79] while in plant plastids and bacteria separate type II enzymes perform each step in the pathway.[80][81]

sterol biosynthesis. Here, the isoprene units are joined to make squalene and then folded up and formed into a set of rings to make lanosterol.[86] Lanosterol can then be converted into other sterols such as cholesterol and ergosterol.[86][87]

Proteins

Organisms vary in their ability to synthesize the 20 common amino acids. Most bacteria and plants can synthesize all twenty, but mammals can only synthesize eleven nonessential amino acids, so nine

glutamate and glutamine. Nonessensial amino acid synthesis depends on the formation of the appropriate alpha-keto acid, which is then transaminated to form an amino acid.[89]

Amino acids are made into proteins by being joined in a chain of

primary structure. Just as the letters of the alphabet can be combined to form an almost endless variety of words, amino acids can be linked in varying sequences to form a huge variety of proteins. Proteins are made from amino acids that have been activated by attachment to a transfer RNA molecule through an ester bond. This aminoacyl-tRNA precursor is produced in an ATP-dependent reaction carried out by an aminoacyl tRNA synthetase.[90] This aminoacyl-tRNA is then a substrate for the ribosome, which joins the amino acid onto the elongating protein chain, using the sequence information in a messenger RNA.[91]

Nucleotide synthesis and salvage

Nucleotides are made from amino acids, carbon dioxide and

orotate, which is formed from glutamine and aspartate.[95]

Xenobiotics and redox metabolism

All organisms are constantly exposed to compounds that they cannot use as foods and that would be harmful if they accumulated in cells, as they have no metabolic function. These potentially damaging compounds are called

A related problem for

disulfide bonds during protein folding produce reactive oxygen species such as hydrogen peroxide.[103] These damaging oxidants are removed by antioxidant metabolites such as glutathione and enzymes such as catalases and peroxidases.[104][105]

Thermodynamics of living organisms

Living organisms must obey the laws of thermodynamics, which describe the transfer of heat and work. The second law of thermodynamics states that in any isolated system, the amount of entropy (disorder) cannot decrease. Although living organisms' amazing complexity appears to contradict this law, life is possible as all organisms are open systems that exchange matter and energy with their surroundings. Living systems are not in equilibrium, but instead are dissipative systems that maintain their state of high complexity by causing a larger increase in the entropy of their environments.[106] The metabolism of a cell achieves this by coupling the spontaneous processes of catabolism to the non-spontaneous processes of anabolism. In thermodynamic terms, metabolism maintains order by creating disorder.[107]

Regulation and control

As the environments of most organisms are constantly changing, the reactions of metabolism must be finely regulated to maintain a constant set of conditions within cells, a condition called homeostasis.[108][109] Metabolic regulation also allows organisms to respond to signals and interact actively with their environments.[110] Two closely linked concepts are important for understanding how metabolic pathways are controlled. Firstly, the regulation of an enzyme in a pathway is how its activity is increased and decreased in response to signals. Secondly, the control exerted by this enzyme is the effect that these changes in its activity have on the overall rate of the pathway (the flux through the pathway).[111] For example, an enzyme may show large changes in activity (i.e. it is highly regulated) but if these changes have little effect on the flux of a metabolic pathway, then this enzyme is not involved in the control of the pathway.[112]

plasma membrane and influx of glucose (3), glycogen synthesis (4), glycolysis (5) and fatty acid synthesis (6).[image reference needed
]

There are multiple levels of metabolic regulation. In intrinsic regulation, the metabolic pathway self-regulates to respond to changes in the levels of substrates or products; for example, a decrease in the amount of product can increase the flux through the pathway to compensate.[111] This type of regulation often involves allosteric regulation of the activities of multiple enzymes in the pathway.[113] Extrinsic control involves a cell in a multicellular organism changing its metabolism in response to signals from other cells. These signals are usually in the form of water-soluble messengers such as hormones and growth factors and are detected by specific receptors on the cell surface.[114] These signals are then transmitted inside the cell by second messenger systems that often involved the phosphorylation of proteins.[115]

A very well understood example of extrinsic control is the regulation of glucose metabolism by the hormone

blood glucose levels. Binding of the hormone to insulin receptors on cells then activates a cascade of protein kinases that cause the cells to take up glucose and convert it into storage molecules such as fatty acids and glycogen.[117] The metabolism of glycogen is controlled by activity of phosphorylase, the enzyme that breaks down glycogen, and glycogen synthase, the enzyme that makes it. These enzymes are regulated in a reciprocal fashion, with phosphorylation inhibiting glycogen synthase, but activating phosphorylase. Insulin causes glycogen synthesis by activating protein phosphatases and producing a decrease in the phosphorylation of these enzymes.[118]

Evolution

Evolutionary tree showing the common ancestry of organisms from all three domains of life. Bacteria are colored blue, eukaryotes red, and archaea green. Relative positions of some of the phyla included are shown around the tree.

The central pathways of metabolism described above, such as glycolysis and the citric acid cycle, are present in all

RNA world.[122]

Many models have been proposed to describe the mechanisms by which novel metabolic pathways evolve. These include the sequential addition of novel enzymes to a short ancestral pathway, the duplication and then divergence of entire pathways as well as the recruitment of pre-existing enzymes and their assembly into a novel reaction pathway.[123] The relative importance of these mechanisms is unclear, but genomic studies have shown that enzymes in a pathway are likely to have a shared ancestry, suggesting that many pathways have evolved in a step-by-step fashion with novel functions created from pre-existing steps in the pathway.[124] An alternative model comes from studies that trace the evolution of proteins' structures in metabolic networks, this has suggested that enzymes are pervasively recruited, borrowing enzymes to perform similar functions in different metabolic pathways (evident in the MANET database)[125] These recruitment processes result in an evolutionary enzymatic mosaic.[126] A third possibility is that some parts of metabolism might exist as "modules" that can be reused in different pathways and perform similar functions on different molecules.[127]

As well as the evolution of new metabolic pathways, evolution can also cause the loss of metabolic functions. For example, in some

parasites metabolic processes that are not essential for survival are lost and preformed amino acids, nucleotides and carbohydrates may instead be scavenged from the host.[128] Similar reduced metabolic capabilities are seen in endosymbiotic organisms.[129]

Investigation and manipulation

Metabolic network of the Arabidopsis thaliana citric acid cycle. Enzymes and metabolites are shown as red squares and the interactions between them as black lines.

Classically, metabolism is studied by a reductionist approach that focuses on a single metabolic pathway. Particularly valuable is the use of radioactive tracers at the whole-organism, tissue and cellular levels, which define the paths from precursors to final products by identifying radioactively labelled intermediates and products.[130] The enzymes that catalyze these chemical reactions can then be purified and their kinetics and responses to inhibitors investigated. A parallel approach is to identify the small molecules in a cell or tissue; the complete set of these molecules is called the metabolome. Overall, these studies give a good view of the structure and function of simple metabolic pathways, but are inadequate when applied to more complex systems such as the metabolism of a complete cell.[131]

An idea of the complexity of the metabolic networks in cells that contain thousands of different enzymes is given by the figure showing the interactions between just 43 proteins and 40 metabolites to the right: the sequences of genomes provide lists containing anything up to 26.500 genes.[132] However, it is now possible to use this genomic data to reconstruct complete networks of biochemical reactions and produce more holistic mathematical models that may explain and predict their behavior.[133] These models are especially powerful when used to integrate the pathway and metabolite data obtained through classical methods with data on gene expression from proteomic and DNA microarray studies.[134] Using these techniques, a model of human metabolism has now been produced, which will guide future drug discovery and biochemical research.[135] These models are now used in network analysis, to classify human diseases into groups that share common proteins or metabolites.[136][137]

Bacterial metabolic networks are a striking example of bow-tie[138][139][140] organization, an architecture able to input a wide range of nutrients and produce a large variety of products and complex macromolecules using a relatively few intermediate common currencies.[141]

A major technological application of this information is metabolic engineering. Here, organisms such as yeast, plants or bacteria are genetically modified to make them more useful in biotechnology and aid the production of drugs such as antibiotics or industrial chemicals such as 1,3-propanediol and shikimic acid.[142][143][144] These genetic modifications usually aim to reduce the amount of energy used to produce the product, increase yields and reduce the production of wastes.[145]

History

The term metabolism is derived from the Ancient Greek word μεταβολή – "Metabole" for "a change" which derived from μεταβάλλ –"Metaballein" means "To change"[146]

Aristotle's metabolism as an open flow model

Greek philosophy

The Parts of Animals sets out enough details of his views on metabolism for an open flow model to be made. He believed that at each stage of the process, materials from food were transformed, with heat being released as the classical element of fire, and residual materials being excreted as urine, bile, or faeces.[147]

Al-Risalah al-Kamiliyyah fil Siera al-Nabawiyyah (The Treatise of Kamil on the Prophet's Biography) which included the following phrase "Both the body and its parts are in a continuous state of dissolution and nourishment, so they are inevitably undergoing permanent change."[148]

Application of the scientific method and Modern metabolic theories

The history of the scientific study of metabolism spans several centuries and has moved from examining whole animals in early studies, to examining individual metabolic reactions in modern biochemistry. The first controlled

sleep, working, sex, fasting, drinking, and excreting. He found that most of the food he took in was lost through what he called "insensible perspiration
".

Santorio Santorio in his steelyard balance, from Ars de statica medicina, first published 1614

In these early studies, the mechanisms of these metabolic processes had not been identified and a

Friedrich Wöhler in 1828 of a paper on the chemical synthesis of urea,[152]
and is notable for being the first organic compound prepared from wholly inorganic precursors. This proved that the organic compounds and chemical reactions found in cells were no different in principle than any other part of chemistry.

It was the discovery of

radioisotopic labelling, electron microscopy and molecular dynamics simulations. These techniques have allowed the discovery and detailed analysis of the many molecules and metabolic pathways in cells.[citation needed
]

See also

References

  1. ^
    PMID 9328649. {{cite book}}: |journal= ignored (help
    )
  2. .
  3. ^ .
  4. ^ .
  5. ^ .
  6. .
  7. .
  8. ^ Cooper GM (2000). "The Molecular Composition of Cells". The Cell: A Molecular Approach (2nd ed.). Archived from the original on 27 August 2020. Retrieved 25 June 2020.
  9. S2CID 4550126
    .
  10. ^ .
  11. .
  12. .
  13. .
  14. ^ "Lipid nomenclature Lip-1 & Lip-2". qmul.ac.uk. Archived from the original on 6 June 2020. Retrieved 6 June 2020.
  15. OCLC 913469736
    .
  16. .
  17. .
  18. ^ .
  19. .
  20. ^ .
  21. .
  22. ^ Berg JM, Tymoczko JL, Stryer L (2002). "Vitamins Are Often Precursors to Coenzymes". Biochemistry. 5th Edition. Archived from the original on 15 December 2020. Retrieved 9 June 2020.
  23. PMID 17295611
    .
  24. .
  25. .
  26. ^ "Electrolyte Balance". Anatomy and Physiology. OpenStax. Archived from the original on 2 June 2020. Retrieved 23 June 2020.
  27. ^ Lodish H, Berk A, Zipursky SL, Matsudaira P, Baltimore D, Darnell J (2000). "The Action Potential and Conduction of Electric Impulses". Molecular Cell Biology (4th ed.). Archived from the original on 30 May 2020. Retrieved 23 June 2020 – via NCBI.
  28. S2CID 37462321
    .
  29. .
  30. from the original on 25 June 2020. Retrieved 24 June 2020.
  31. .
  32. ^ a b Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002). "How Cells Obtain Energy from Food". Molecular Biology of the Cell (4th ed.). Archived from the original on 5 July 2021. Retrieved 25 June 2020 – via NCBI.
  33. from the original on 25 June 2020. Retrieved 25 June 2020.
  34. ^ .
  35. .
  36. .
  37. .
  38. .
  39. .
  40. .
  41. .
  42. .
  43. ^ .
  44. .
  45. .
  46. .
  47. .
  48. .
  49. .
  50. (PDF) from the original on 22 January 2020. Retrieved 11 November 2019.
  51. .
  52. .
  53. from the original on 2 May 2019. Retrieved 6 October 2019.
  54. .
  55. .
  56. .
  57. .
  58. .
  59. .
  60. ^ Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002). "Energy Conversion: Mitochondria and Chloroplasts". Molecular Biology of the Cell (4th ed.). Archived from the original on 15 December 2020. Retrieved 3 July 2020.
  61. S2CID 21596537
    .
  62. .
  63. ^ a b Mandal A (26 November 2009). "What is Anabolism?". News-Medical.net. Archived from the original on 5 July 2020. Retrieved 4 July 2020.
  64. PMID 6351728
    .
  65. .
  66. .
  67. .
  68. .
  69. .
  70. .
  71. .
  72. ^ .
  73. .
  74. ^ .
  75. from the original on 31 October 2020. Retrieved 28 August 2020.
  76. PMID 28876856. Archived from the original on 24 February 2022. Retrieved 8 July 2020.{{cite book}}: CS1 maint: DOI inactive as of January 2024 (link
    )
  77. .
  78. .
  79. .
  80. .
  81. .
  82. S2CID 27523830. Archived from the original
    (PDF) on 15 April 2007.
  83. ^ .
  84. .
  85. .
  86. ^ .
  87. .
  88. .
  89. .
  90. PMID 11375928. Archived from the original
    on 1 May 2011.
  91. .
  92. ^ .
  93. .
  94. (PDF) from the original on 24 October 2020. Retrieved 18 September 2019.
  95. .
  96. .
  97. .
  98. .
  99. .
  100. .
  101. (PDF) from the original on 11 November 2019. Retrieved 11 November 2019.
  102. .
  103. .
  104. .
  105. .
  106. .
  107. from the original on 4 August 2020. Retrieved 22 September 2019.
  108. .
  109. from the original on 29 March 2007. Retrieved 12 March 2007.
  110. .
  111. ^ .
  112. .
  113. .
  114. .
  115. .
  116. .
  117. .
  118. (PDF) from the original on 19 June 2007. Retrieved 25 March 2007.
  119. .
  120. PMID 9889982. {{cite book}}: |journal= ignored (help
    )
  121. .
  122. .
  123. .
  124. .
  125. .
  126. .
  127. .
  128. .
  129. .
  130. .
  131. .
  132. .
  133. .
  134. .
  135. .
  136. .
  137. .
  138. .
  139. .
  140. .
  141. ^ "Macromolecules: Nutrients, Metabolism, and Digestive Processes | Virtual High School - KeepNotes". keepnotes.com. Retrieved 29 December 2023.
  142. PMID 12749845
    .
  143. .
  144. .
  145. .
  146. ^ "metabolism | Origin and meaning of metabolism by Online Etymology Dictionary". www.etymonline.com. Archived from the original on 21 September 2017. Retrieved 23 July 2020.
  147. .
  148. ^ Al-Roubi AS (1982). Ibn Al-Nafis as a philosopher. Symposium on Ibn al-Nafis, Second International Conference on Islamic Medicine. Kuwait: Islamic Medical Organization.
  149. S2CID 32900603
    .
  150. ^ Williams HA (1904). Modern Development of the Chemical and Biological Sciences. A History of Science: in Five Volumes. Vol. IV. New York: Harper and Brothers. pp. 184–185. Retrieved 26 March 2007.
  151. PMID 8595136
    .
  152. .
  153. ^ Eduard Buchner's 1907 Nobel lecture Archived 8 July 2017 at the Wayback Machine at http://nobelprize.org Archived 5 April 2006 at the Wayback Machine Accessed 20 March 2007
  154. S2CID 28092593
    .
  155. .
  156. .

Further reading

Introductory

Advanced

External links

General information

Human metabolism

Databases

Metabolic pathways