Oxidative phosphorylation

Jump to navigation Jump to search
File:Mitochondrial electron transport chain—Etc4.svg
The electron transport chain in the mitochondrion is the site of oxidative phosphorylation in eukaryotes. The NADH and succinate generated in the citric acid cycle is oxidized, releasing energy to power the ATP synthase.

WikiDoc Resources for Oxidative phosphorylation

Articles

Most recent articles on Oxidative phosphorylation

Most cited articles on Oxidative phosphorylation

Review articles on Oxidative phosphorylation

Articles on Oxidative phosphorylation in N Eng J Med, Lancet, BMJ

Media

Powerpoint slides on Oxidative phosphorylation

Images of Oxidative phosphorylation

Photos of Oxidative phosphorylation

Podcasts & MP3s on Oxidative phosphorylation

Videos on Oxidative phosphorylation

Evidence Based Medicine

Cochrane Collaboration on Oxidative phosphorylation

Bandolier on Oxidative phosphorylation

TRIP on Oxidative phosphorylation

Clinical Trials

Ongoing Trials on Oxidative phosphorylation at Clinical Trials.gov

Trial results on Oxidative phosphorylation

Clinical Trials on Oxidative phosphorylation at Google

Guidelines / Policies / Govt

US National Guidelines Clearinghouse on Oxidative phosphorylation

NICE Guidance on Oxidative phosphorylation

NHS PRODIGY Guidance

FDA on Oxidative phosphorylation

CDC on Oxidative phosphorylation

Books

Books on Oxidative phosphorylation

News

Oxidative phosphorylation in the news

Be alerted to news on Oxidative phosphorylation

News trends on Oxidative phosphorylation

Commentary

Blogs on Oxidative phosphorylation

Definitions

Definitions of Oxidative phosphorylation

Patient Resources / Community

Patient resources on Oxidative phosphorylation

Discussion groups on Oxidative phosphorylation

Patient Handouts on Oxidative phosphorylation

Directions to Hospitals Treating Oxidative phosphorylation

Risk calculators and risk factors for Oxidative phosphorylation

Healthcare Provider Resources

Symptoms of Oxidative phosphorylation

Causes & Risk Factors for Oxidative phosphorylation

Diagnostic studies for Oxidative phosphorylation

Treatment of Oxidative phosphorylation

Continuing Medical Education (CME)

CME Programs on Oxidative phosphorylation

International

Oxidative phosphorylation en Espanol

Oxidative phosphorylation en Francais

Business

Oxidative phosphorylation in the Marketplace

Patents on Oxidative phosphorylation

Experimental / Informatics

List of terms related to Oxidative phosphorylation


Overview

Oxidative phosphorylation is a metabolic pathway that uses energy released by the oxidation of nutrients to produce adenosine triphosphate (ATP). Although the many forms of life on earth use a range of different nutrients, almost all carry out oxidative phosphorylation to produce ATP, the molecule that supplies energy to metabolism. This pathway is probably so pervasive because it is a highly efficient way of storing energy, compared to alternative fermentation processes such as glycolysis.

During oxidative phosphorylation, electrons are transferred from electron donors to electron acceptors such as oxygen, in a redox reaction. These redox reactions release energy, which is used to form ATP. In eukaryotes these redox reactions are carried out by a series of protein complexes within mitochondria, whereas in prokaryotes, these proteins are located in the cells' inner membranes. These linked sets of enzymes are called electron transport chains. In eukaryotes, five main protein complexes are involved, whereas in prokaryotes many different enzymes are present, using a variety of electron donors and acceptors.

The energy released as electrons flow through this electron transport chain is used to transport protons across the inner mitochondrial membrane, in a process called chemiosmosis. This generates potential energy in the form of a pH gradient and an electrical potential across this membrane. This store of energy is tapped by allowing protons to flow back across the membrane and down this gradient, through a large enzyme called ATP synthase. This enzyme uses this energy to generate ATP from adenosine diphosphate (ADP), in a phosphorylation reaction. Unusually, the ATP synthase works by the proton flow driving the rotation of part of the enzyme—it is a rotary mechanical motor.

Although oxidative phosphorylation is a vital part of metabolism, it produces reactive oxygen species such as superoxide and hydrogen peroxide that lead to propagation of free radicals, damaging cells and contributing to ageing and disease. The enzymes carrying out this metabolic pathway are also the target of many drugs and poisons that inhibit their activities.

Overview of energy transfer by chemiosmosis

Oxidative phosphorylation works by using energy-releasing chemical reactions to drive energy-requiring reactions: the two sets of reactions are said to be coupled. This means one cannot occur without the other. The flow of electrons through the electron transport chain, from electron donors such as NADH to electron acceptors such as oxygen, is an exergonic process – it releases energy; while the synthesis of ATP is an endergonic process that requires an input of energy.

Both the electron transport chain and the ATP synthase are embedded in a membrane, and energy is transferred from electron transport chain to the ATP synthase by movements of protons across this membrane, in a process called chemiosmosis.[1] In practice, this is like a simple electric circuit, with a current of protons being driven from the negative N-side of the membrane to the positive P-side by the proton-pumping enzymes of the electron transport chain. These enzymes are like a battery, as they perform work to drive current through the circuit. The movement of protons creates an electrochemical gradient across the membrane, which is often called the proton-motive force. This gradient has two components: a difference in proton concentration (a pH gradient) and a difference in electric potential, with the N-side having a negative charge. The energy is stored largely as the difference of electric potentials in mitochondria, but as a pH gradient in chloroplasts.[2]

ATP synthase releases this stored energy by completing the circuit and allowing protons to flow down the electrochemical gradient, back to the N-side of the membrane.[3] This enzyme is like an electric motor as it uses the proton-motive force to drive the rotation of part of its structure and couples this motion to the synthesis of ATP.

The amount of energy released by oxidative phosphorylation is high, compared with the amount produced by anaerobic fermentation. Glycolysis only produces 2 ATP molecules, but 26 ATPs are produced by the oxidative phosphorylation of the 10 NADH and 2 succinate molecules made by converting one molecule of glucose to carbon dioxide and water.[4] This ATP yield is the theoretical maximum value; in practice some protons leak across the membrane, lowering the yield of ATP.[5]

Electron and proton transfer molecules

File:Ubiquinone–ubiquinol conversion.svg
Reduction of coenzyme Q from its ubiquinone form (Q) to the reduced ubiquinol form (QH2).

The electron transport chain carries both protons and electrons, passing electrons from donors to acceptors, and transporting protons across a membrane. These processes use both soluble and protein-bound transfer molecules. In mitochondria, electrons are transferred within the intermembrane space by the water-soluble electron transfer protein cytochrome c.[6] This carries only electrons, and these are transferred by the reduction and oxidation of an iron atom that the protein holds within a heme group in its structure. Cytochrome c is also found in some bacteria, where it is located within the periplasmic space.[7]

Within the inner mitochondrial membrane, the lipid-soluble electron carrier coenzyme Q10 (Q) carries both electrons and protons by a redox cycle.[8] This small benzoquinone molecule is very hydrophobic, so it diffuses freely within the membrane. When Q accepts two electrons and two protons, it becomes reduced to the ubiquinol form (QH2); when QH2 releases two electrons and two protons, it becomes oxidized back to the ubiquinone (Q) form. As a result, if two enzymes are arranged so that Q is reduced on one side of the membrane and QH2 oxidized on the other, ubiquinone will couple these reactions and shuttle protons across the membrane.[9] Some bacterial electron transport chains use different quinones, such as menaquinone, in addition to ubiquinone.[10]

Within proteins, electrons are transferred between flavin cofactors [11][3], iron–sulfur clusters, and cytochromes. There are several types of iron–sulfur cluster. The simplest kind found in the electron transfer chain consists of two iron atoms joined by two atoms of inorganic sulfur; these are called [2Fe–2S] clusters. The second kind, called [4Fe–4S], contains a cube of four iron atoms and four sulfur atoms. Each iron atom in these clusters is coordinated by an additional amino acid, usually by the sulfur atom of cysteine. Metal ion cofactors undergo redox reactions without binding or releasing protons, so in the electron transport chain they serve solely to transport electrons through proteins. Electrons move through quite long distances in proteins by hopping along between chains of these cofactors.[12] This occurs by quantum tunnelling, which is rapid over distances of less than 14 Å.[13]

Eukaryotic electron transport chains

Many catabolic biochemical processes, such as glycolysis, the citric acid cycle and beta oxidation, produce the reduced coenzyme NADH. This coenzyme contains electrons that have a high transfer potential; in other words, they will release a large amount of energy upon oxidation. However, the cell does not release this energy all at once, as this would be an uncontrollable reaction. Instead, the electrons are removed from NADH and passed to oxygen through a series of enzymes that each release a small amount of the energy. This set of enzymes, consisting of complexes I through IV, is called the electron transport chain and is found in the inner membrane of the mitochondrion. Succinate is also oxidized by the electron transport chain, but feeds into the pathway at a different point.

In eukaryotes, the enzymes in this electron transport system use the energy released from the oxidation of NADH to pump protons across the inner membrane of the mitochondrion. This causes protons to build up in the intermembrane space, and generates an electrochemical gradient across the membrane. The energy stored in this potential is then used by ATP synthase to produce ATP. Oxidative phosphorylation in the eukaryotic mitochondrion is the best-understood example of this process. The mitochondrion is present in almost all eukaryotes, with the exception of anaerobic protozoa such as Trichomonas vaginalis that instead reduce protons to hydrogen in a remnant mitochondrion called a hydrogenosome.[14]

NADH-coenzyme Q oxidoreductase (complex I)

File:Complex I.svg
Complex I or NADH-Q oxidoreductase. The abbreviations are discussed in the text. In all diagrams of respiratory complexes, the matrix is at the bottom, with the intermembrane space above.

NADH-coenzyme Q oxidoreductase, also known as NADH dehydrogenase or complex I, is the first protein in the electron transport chain.[15] Complex I is a giant enzyme with the mammalian complex I having 46 subunits and a molecular mass of about 1,000 kilodaltons (kDa).[16] The structure is known in detail only from a bacterium [17]; in most organisms the complex resembles a boot with a large “ball” poking out from the membrane into the mitochondrion.[18][19] The genes that encode the individual proteins are contained in both the cell nucleus and the mitochondrial genome, as is the case for many enzymes present in the mitochondrion.

The reaction which is catalyzed by this enzyme is the two electron reduction by NADH of coenzyme Q10 or ubiquinone (represented as Q in the equation below), a lipid-soluble quinone that is found in the mitochondrion membrane:

  <math>\mbox{NADH + Q + 5H}^{+}_{matrix} \rightarrow \mbox{NAD}^+ + \mbox{QH}_2 + \mbox{4H}^+_{cytosol}</math>

The start of the reaction, and indeed of the entire electron chain, is the binding of a NADH molecule to complex I and the donation of two electrons. The electrons enter complex I via a prosthetic group attached to the complex, flavin mononucleotide (FMN). The addition of electrons to FMN converts it to its reduced form, FMNH2. The electrons are then transferred through a series of iron–sulfur clusters: the second kind of prosthetic group present in the complex.[20] There are both [2Fe–2S] and [4Fe–4S] iron–sulfur clusters in complex I.

As the electrons pass through this complex, four protons are pumped from the matrix into the intermembrane space. Exactly how this occurs is unclear, but it seems to involve conformational changes in complex I that cause the protein to bind protons on the N-side of the membrane and then move them onto the P-side of the membrane.[21] Finally, the electrons are transferred from the chain of iron–sulfur clusters to a ubiquinone molecule in the membrane.[15] Reduction of ubiquinone also contributes to the generation of a proton gradient, as two protons are taken up from the matrix as it is reduced to ubiquinol (QH2).

File:Complex II.svg
Complex II: Succinate-Q oxidoreductase.

Succinate-Q oxidoreductase (complex II)

Succinate-Q oxidoreductase, also known as complex II, is a second entry point to the electron transport chain.[22] It is unusual as it is the only enzyme that participates in both the citric acid cycle and the electron transport chain. Complex II consists of four protein subunits and contains a bound flavin adenine dinucleotide (FAD) cofactor, iron–sulfur clusters, and a heme group which does not participate in electron transfer to coenzyme Q, but is believed to be important in decreasing production of reactive oxygen species.[23][24] It oxidizes succinate to fumarate and reduces ubiquinone. As this reaction releases less energy than the oxidation of NADH, complex II does not transport protons across the membrane and does not contribute to the proton gradient.

  <math>\mbox{Succinate} + \mbox{Q} \rightarrow \mbox{Fumarate} + \mbox{QH}_2 \, </math>

In some eukaryotes, such as the parasitic worm Ascaris suum, an enzyme similar to complex II, fumarate reductase (menaquinol:fumarate oxidoreductase, or QFR) operates in reverse to oxidize ubiquinol and reduce fumarate. This allows the worm to survive in the anaerobic environment of the large intestine, carrying out anaerobic oxidative phosphorylation with fumarate as the electron acceptor.[25] Another unconventional function of complex II is seen in the malaria parasite Plasmodium falciparum. Here, the reversed action of complex II as an oxidase is important in regenerating ubiquinol, which the parasite uses in an unusual form of pyrimidine biosynthesis.[26]

Electron transfer flavoprotein-Q oxidoreductase

Electron transfer flavoprotein-ubiquinone oxidoreductase (ETF-Q oxidoreductase), also known as electron transferring-flavoprotein dehydrogenase, is a third entry point to the electron transport chain. It is an enzyme that accepts electrons from electron-transferring flavoprotein in the mitochondrial matrix, and uses these electrons to reduce ubiquinone.[27] This enzyme contains a flavin and a [4Fe–4S] cluster, but unlike the other respiratory complexes, it attaches to the surface of the membrane and does not cross the lipid bilayer.[28]

  <math>\mbox{ETF}_{red} + \mbox{Q} \rightarrow \mbox{ETF}_{ox} + \mbox{QH}_2 \, </math>

In mammals, this metabolic pathway is important in beta oxidation of fatty acids and catabolism of amino acids and choline, as it accepts electrons from multiple acetyl-CoA dehydrogenases.[29][30] In plants, ETF-Q oxidoreductase is also important in the metabolic responses that allow survival in extended periods of darkness.[31]

File:Complex III reaction(2).png
The two electron transfer steps in complex III: Q-cytochrome c oxidoreductase. After each step, Q (in the upper part of the figure) leaves the enzyme.

Q-cytochrome c oxidoreductase (complex III)

Q-cytochrome c oxidoreductase is also known as cytochrome c reductase, cytochrome bc1 complex, or simply complex III.[32][33] In mammals, this enzyme is a dimer, with each subunit complex containing 11 protein subunits, an [2Fe-2S] iron–sulfur cluster and three cytochromes; one cytochrome c1 and two b cytochromes.[34] A cytochrome is a kind of electron-transferring protein which contains at least one heme group. The iron atoms inside complex III’s heme groups alternate between a reduced ferrous (+2) and oxidized ferric (+3) state as the electrons are transferred through the protein.

The reaction catalyzed by complex III is the oxidation of one molecule of ubiquinol and the reduction of two molecules of cytochrome c, a heme protein loosely associated with the mitochondrion. Unlike coenzyme Q, which carries two electrons, cytochrome c only carries one electron.

  <math>\mbox{QH}_2 + \mbox{2Cyt c}_{ox} + \mbox{2H}^+_{matrix} \rightarrow \mbox{Q} + \mbox{2Cyt c}_{red} + \mbox{4H}^+_{cytosol} \, </math>

As only one of the electrons can be transferred from the QH2 donor to a cytochrome c acceptor at a time, the reaction mechanism of complex III is more elaborate than those of the other respiratory complexes, and occurs in two steps called the Q cycle.[35] In the first step, the enzyme binds three substrates. Firstly QH2, which is then oxidized, with one electron being passed to the second substrate, cytochrome c. The two protons released from QH2 pass into the intermembrane space. The third substrate is Q, which accepts the second electron from the QH2 and is reduced to Q.-, which is the ubisemiquinone free radical. The first two substrates are released, but this ubisemiquinone intermediate remains bound. In the second step, a second molecule of QH2 is bound and again passes its first electron to a cytochrome c acceptor. The second electron is passed to the bound ubisemiquinone, reducing it to QH2 as it gains two protons from the mitochondrial matrix. This QH2 is then released from the enzyme.[36]

As coenzyme Q is reduced to ubiquinol on the inner side of the membrane and oxidized to ubiquinone on the other, this causes the net transfer of protons across the membrane, adding to the proton gradient.[3] The rather complex two-step mechanism by which this occurs is important as it increases the efficiency of proton transfer. If instead of the Q cycle, one molecule of QH2 was used to directly reduce two molecules of cytochrome c, the efficiency would be halved, with only one proton transferred per cytochrome c reduced.[3]

File:Complex IV.svg
Complex IV: cytochrome c oxidase.

Cytochrome c oxidase (complex IV)

Cytochrome c oxidase, also known as complex IV, is the final protein complex in the electron transport chain.[37] The mammalian enzyme has an extremely complex structure and contains 13 subunits, two heme groups, as well as multiple metal ion cofactors –- in all three atoms of copper, one of magnesium and one of zinc.[38]

This enzyme mediates the final reaction in the electron transport chain and transfers electrons to oxygen, while pumping protons across the membrane.[39] The final electron acceptor oxygen, which is also called the terminal electron acceptor, is reduced to water in this step. Both the direct pumping of protons and the consumption of matrix protons in the reduction of oxygen contribute to the proton gradient. The reaction catalyzed is the oxidation of cytochrome c and the reduction of oxygen:

  <math> \mbox{4Cyt c}_{red} + \mbox{O}_{2} + \mbox{8H}^+_{matrix} \rightarrow \mbox{4Cyt c}_{ox} + \mbox{2H}_2\mbox{O} + \mbox{4H}^+_{cytosol} \, </math>

For a detailed discussion of the mechanism of reduction of oxygen, and a better figure illustrating the redox centers in the protein, see the linked article Cytochrome c oxidase.

Alternative reductases and oxidases

Many eukaryotic organisms have electron transport chains that differ from the well-studied mammalian enzymes described above. For example, in plants, alternative NADH oxidases exist that oxidize NADH in the cytosol, rather than the mitochondrial matrix, and pass these electrons to the ubiquinone pool.[40] These enzymes do not transport protons and therefore reduce ubiquinone without altering the electrochemical gradient across the inner membrane.[41]

Another example of a divergent electron transport chain is the alternative oxidase, which is found in plants, as well as some fungi, protists and possibly some animals.[42][43] This enzyme transfers electrons directly from ubiquinol to oxygen.[44]

The electron transport pathways produced by these alternative NADH and ubiquinone oxidases have lower ATP yields than the full pathway. The advantages produced by a shortened pathway are not entirely clear. However, the alternative oxidase is produced in response to stresses such as cold, reactive oxygen species and infection by pathogens, as well as other factors that inhibit the full electron transport chain.[45][46] Alternative pathways might therefore enhance an organisms' resistance to injury, by reducing oxidative stress.[47]

Organization of complexes

The original model for how the respiratory chain complexes are organized was that they diffuse freely and independently in the mitochondrial membrane.[16] However, recent data suggest that the complexes might form higher-order structures called supercomplexes or "respirasomes".[48] In this model, the various complexes exist as organized sets of interacting enzymes.[49] These associations might allow channeling of substrates between the various enzyme complexes, increasing the rate and efficiency of electron transfer.[50] Within such mammalian supercomplexes, some components would be present in higher amounts than others, with a ratio between complexes I/II/III/IV and the ATP synthase of approximately 1:1:3:7:4.[51] However, the debate over this supercomplex model is not completely resolved, as some data do not appear to fit with this model.[52][16]

Prokaryotic electron transport chains

In contrast to the general similarity in structure and function of the electron transport chains in eukaryotes, bacteria and archaea possess a large variety of electron-transfer enzymes. These use an equally wide set of chemicals as substrates.[53] In common with eukaryotes, prokaryotic electron transport uses the energy released from the oxidation of a substrate to pump ions across a membrane and generate an electrochemical gradient. In the bacteria, oxidative phosphorylation in Escherichia coli is understood in most detail, while archaeal systems are at present poorly-understood.[54]

The main difference between eukaryotic and prokaryotic oxidative phosphorylation is that bacteria and archaea use many different substances to donate or accept electrons. This allows prokaryotes to grow under a wide variety of environmental conditions.[55] In E. coli, for example, oxidative phosphorylation can be driven by a large number of pairs of reducing agents and oxidizing agents, which are listed below. The midpoint potential of a chemical measures how much energy is released when it is oxidized or reduced, with reducing agents having negative potentials and oxidizing agents positive potentials.

Respiratory enzymes and substrates in E. coli [56]
Respiratory enzyme Redox pair  Midpoint potential 

(Volts)

 Formate dehydrogenase Bicarbonate / Formate −0.43
 Hydrogenase Proton / Hydrogen −0.42
 NADH dehydrogenase NAD+ / NADH −0.32
 Glycerol-3-phosphate dehydrogenase DHAP / Gly-3-P −0.19
 Pyruvate oxidase  Acetate + Carbon dioxide / Pyruvate   ?
 Lactate dehydrogenase Pyruvate / Lactate −0.19
 D-amino acid dehydrogenase  2-oxoacid + ammonia / D-amino acid   ?
 Glucose oxidase Glucose / Gluconate −0.14
 Succinate dehydrogenase Succinate / Fumarate +0.03
 Ubiquinol oxidase Oxygen / Water +0.82
 Nitrate reductase Nitrate / Nitrite +0.42
 Nitrite reductase Nitrite / Ammonia +0.36
 Dimethyl sulfoxide reductase DMSO / DMS +0.16
 Trimethylamine N-oxide reductase TMAO / TMA +0.13
 Fumarate reductase Fumarate / Succinate +0.03

As shown above, E. coli can grow with reducing agents such as formate, hydrogen or lactate as electron donors, and nitrate, DMSO or oxygen as acceptors.[55] The larger the difference in midpoint potential between an oxidizing and reducing agent, the more energy is released when they react. Out of these compounds, the succinate/fumarate pair is unusual, as its midpoint potential is close to zero. Succinate can therefore be oxidized to fumarate if a strong oxidizing agent such as oxygen is available, or fumarate can be reduced to succinate using a strong reducing agent such as formate. These alternative reactions are catalyzed by succinate dehydrogenase and fumarate reductase, respectively.[57]

Some prokaryotes use redox pairs that have only a small difference in midpoint potential. For example, nitrifying bacteria such as Nitrobacter oxidize nitrite to nitrate, donating the electrons to oxygen. The small amount of energy released in this reaction is enough to pump protons and generate ATP, but not enough to produce NADH or NADPH directly for use in anabolism.[58] This problem is solved by using a nitrite oxidoreductase to produce enough proton-motive force to run part of the electron transport chain in reverse, causing complex I to generate NADH.[59][60]

Prokaryotes control their use of these electron donors and acceptors by varying which enzymes are produced, in response to environmental conditions.[61] This flexibility is possible because different oxidases and reductases use the same ubiquinone pool. This allows many combinations of enzymes to function together, linked by the common ubiquinol intermediate.[56] These respiratory chains therefore have a modular design, with easily interchangeable sets of enzyme systems.

In addition to this metabolic diversity, prokaryotes also possess a range of isozymes – different enzymes that catalyze the same reaction. For example, in E. coli there are two different types of ubiquinol oxidase using oxygen as an electron acceptor. Under highly aerobic conditions the cell uses an oxidase with a low affinity for oxygen that can transport two protons per electron. However, if levels of oxygen fall, they switch to an oxidase that only transfers one proton per electron, but has a high affinity for oxygen.[62]

ATP synthase

ATP synthase. The FO proton channel and stalk are shown in blue, the F1 synthase domain in red and the membrane in gray.

ATP synthase, also called complex V, is the final enzyme in the oxidative phosphorylation pathway. This enzyme is found in all forms of life and functions in the same way in both prokaryotes or eukaryotes.[63] The enzyme uses the energy stored in a proton gradient across a membrane to drive the synthesis of ATP from ADP and phosphate (Pi). Estimates of the number of protons required to synthesise one ATP have ranged from three to four,[64][65] with some suggesting cells can vary this ratio, to suit different conditions.[66]

  <math>\mbox{ADP} + \mbox{P}_{i} + \mbox{4H}^+_{cytosol} \rightleftharpoons \mbox{ATP} + \mbox{H}_{2}\mbox{O} + \mbox{4H}^+_{matrix} </math>

This phosphorylation reaction is an equilibrium, which can be shifted by altering the proton-motive force. In the absence of a proton-motive force, the ATP synthase reaction will run from right to left, hydrolyzing ATP and pumping protons out of the matrix across the membrane. However, when the proton-motive force is high, the reaction is forced to run in the opposite direction; it proceeds from left to right, allowing protons to flow down their concentration gradient and turning ADP into ATP.[63] Indeed, in the closely related vacuolar type H+-ATPases the same reaction is used to acidify cellular compartments, by pumping protons and hydrolysing ATP.[67]

ATP synthase is a massive protein complex with a mushroom-like shape. The mammalian enzyme complex contains 16 subunits and has a mass of approximately 600 kilodaltons.[68] The portion embedded within the membrane is called FO and contains a ring of c subunits and the proton channel. The stalk and the ball-shaped headpiece is called F1 and is the site of ATP synthesis. The ball-shaped complex at the end of the F1 portion contains six proteins of two different kinds (three α subunits and three β subunits), while the "stalk" consists of one protein: the γ subunit, with the tip of the stalk extending into the ball of α and β subunits.[69] Both the α and β subunits bind nucleotides, but only the β subunits catalyze the ATP synthesis reaction. Reaching along the side of the F1 portion and back into the membrane is a long rod-like subunit that anchors the α and β subunits into the base of the enzyme.

As protons cross the membrane through the channel in the base of ATP synthase this causes the FO proton-driven motor to rotate.[70] Rotation might be caused by changes in the ionization of amino acids in the ring of c subunits causing electrostatic interactions that propel the ring of c subunits past the proton channel.[71] This rotating ring in turn drives the rotation of the central axle (the γ subunit stalk) within the α and β subunits. The α and β subunits are prevented from rotating themselves by the side-arm, which acts as a stator. This movement of the tip of the γ subunit within the ball of α and β subunits provides the energy for the active sites in the β subunits to undergo a cycle of movements that produces and then releases ATP.[2]

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

This ATP synthesis reaction is called the binding change mechanism and involves the active site of a β subunit cycling between three states.[72] In the "open" state, ADP and phosphate enter the active site (shown in brown in the diagram). The protein then closes up around the molecules and binds them loosely – the "loose" state (shown in red). The enzyme then changes shape again and forces these molecules together, with the active site in the resulting "tight" state (shown in pink) binding the newly produced ATP molecule with very high affinity. Finally, the active site cycles back to the open state, releasing ATP and binding more ADP and phosphate, ready for the next cycle.

In some bacteria and archaea, ATP synthesis is driven by the movement of sodium ions through the cell membrane, rather than the movement of protons.[73][74] Archaea such as Methanococcus also contain the A1Ao synthase, a form of the enzyme that contains additional proteins with little similarity in sequence to other bacterial and eukaryotic ATP synthase subunits. It is possible that in some species the A1Ao form of the enzyme is a specialized sodium-driven ATP synthase,[75] but this might not be true in all cases.[74]

Reactive oxygen species

Molecular oxygen is an ideal terminal electron acceptor because it is a strong oxidizing agent. The reduction of oxygen does involve potentially harmful intermediates.[76] Although the transfer of four electrons and four protons reduces oxygen to water, which is harmless, transfer of one or two electrons produces superoxide anion and peroxide.

<math> \begin{matrix} \quad & {e^-} & \quad & {e^-} \\ {\mbox{O}_{2}} & \longrightarrow & \mbox{O}_2^{\underline{\bullet}} & \longrightarrow & \mbox{O}_2^{2-} \\ \quad & \quad & \mbox{Superoxide} & \quad & \mbox{Peroxide} \\ \quad & \quad \end{matrix}</math>

These reactive oxygen species and their reaction products, such as the hydroxyl radical, are very harmful to cells, as they oxidize proteins and cause mutations in DNA. This cellular damage might contribute to disease and is proposed as one cause of ageing.[77][78]

The cytochrome c oxidase complex is highly efficient at reducing oxygen to water and it releases very few partly reduced intermediates; however small amounts of superoxide anion and peroxide are produced by the electron transport chain.[79] Particularly important is the reduction of coenzyme Q in complex III, as a highly reactive ubisemiquinone free radical is formed as an intermediate in the Q cycle. This unstable species can lead to electron "leakage" when electrons transfer directly to oxygen, forming superoxide.[80]

To counteract these reactive oxygen species, cells contain numerous antioxidant systems, including antioxidant vitamins such as vitamin C and vitamin E, and antioxidant enzymes such as superoxide dismutase, catalase, and peroxidases.[76]which detoxify the reactive species, limiting damage to the cell.

Inhibitors

There are several well-known drugs and toxins that inhibit oxidative phosphorylation. Although any one of these toxins only inhibits one enzyme in the electron transport chain, inhibition of any step in this process will halt the rest of the process. For example, if oligomycin inhibits ATP synthase, protons cannot pass back into the mitochondrion.[81] As a result, the proton pumps are unable to operate, as the gradient becomes too strong for them to overcome. NADH is then no longer oxidized and the citric acid cycle ceases to operate because the concentration of NAD+ falls below the concentration that these enzymes can use.

Compounds Use Effect on oxidative phosphorylation
Cyanide
Carbon monoxide
Poisons Inhibit the electron transport chain by binding more strongly than oxygen to the FeCu center in cytochrome c oxidase, preventing the reduction of oxygen.[82]
Oligomycin Antibiotic Inhibits ATP synthase by blocking the flow of protons through the Fo subunit.[81]
CCCP
2,4-Dinitrophenol
Poisons Ionophores that disrupt the proton gradient by carrying protons across the membrane. This uncouples proton pumping from ATP synthesis.[83]
Rotenone Pesticide Prevents the transfer of electrons from complex I to ubiquinone by blocking to the ubiquinone-binding site.[84]

Not all inhibitors of oxidative phosphorylation are toxins. In brown adipose tissue regulated proton channels called uncoupling proteins can uncouple respiration from ATP synthesis.[85] This rapid respiration produces heat, and is particularly important as a way of maintaining body temperature during hibernation, although these proteins may also have a more general function in cells' responses to stress.[86]

History

The field of oxidative phosphorylation began with the report in 1906 by Arthur Harden of a vital role for phosphate in cellular fermentation, but initially only sugar phosphates were known to be involved.[87] However, in the early 1940s, the link between the oxidation of sugars and the generation of ATP was firmly established by Herman Kalckar,[88] confirming the central role of ATP in energy transfer that had been proposed by Fritz Albert Lipmann in 1941.[89] Later, in 1949, Morris Friedkin and Albert L. Lehninger proved that the coenzyme NADH linked metabolic pathways such as the citric acid cycle and the synthesis of ATP.[90]

For another twenty years the mechanism by which ATP was generated remained mysterious, with scientists searching for an elusive "high-energy intermediate" that would link oxidation and phosphorylation reactions.[91] This puzzle was solved by Peter D. Mitchell with the publication of the chemiosmotic theory in 1961.[92] At first this proposal was highly controversial, but it was slowly accepted and Mitchell was awarded a Nobel prize in 1978.[93][94] Subsequent research concentrated on purifying and characterizing the enzymes involved, with major contributions being made by David E. Green on the complexes of the electron-transport chain, as well as Efraim Racker on the ATP synthase.[95] A critical step towards solving the mechanism of the ATP synthase was provided by Paul D. Boyer, by his development in 1973 of the "binding change" mechanism, followed by his radical proposal of rotational catalysis in 1982.[96][72] More recent work has included structural studies on the enzymes involved in oxidative phosphorylation by John E. Walker, with Walker and Boyer being awarded a Nobel prize in 1997.[97]

References

  1. Mitchell P, Moyle J (1967). "Chemiosmotic hypothesis of oxidative phosphorylation". Nature. 213 (5072): 137&ndash, 9. PMID 4291593.
  2. 2.0 2.1 Dimroth P, Kaim G, Matthey U (2000). "Crucial role of the membrane potential for ATP synthesis by F(1)F(o) ATP synthases". J. Exp. Biol. 203 (Pt 1): 51&ndash, 9. PMID 10600673.
  3. 3.0 3.1 3.2 3.3 Schultz B, Chan S (2001). "Structures and proton-pumping strategies of mitochondrial respiratory enzymes". Annu Rev Biophys Biomol Struct. 30: 23&ndash, 65. PMID 11340051.
  4. Rich PR (2003). "The molecular machinery of Keilin's respiratory chain". Biochem. Soc. Trans. 31 (Pt 6): 1095&ndash, 105. PMID 14641005.
  5. Porter RK, Brand MD (1995). "Mitochondrial proton conductance and H+/O ratio are independent of electron transport rate in isolated hepatocytes". Biochem. J. 310 ( Pt 2): 379&ndash, 82. PMID 7654171.
  6. Mathews FS (1985). "The structure, function and evolution of cytochromes". Prog. Biophys. Mol. Biol. 45 (1): 1&ndash, 56. PMID 3881803.
  7. Wood PM (1983). "Why do c-type cytochromes exist?". FEBS Lett. 164 (2): 223&ndash, 6. PMID 6317447.
  8. Crane FL (2001). "Biochemical functions of coenzyme Q10". Journal of the American College of Nutrition. 20 (6): 591&ndash, 8. PMID 11771674.
  9. Mitchell P (1979). "Keilin's respiratory chain concept and its chemiosmotic consequences". Science. 206 (4423): 1148&ndash, 59. PMID 388618.
  10. Søballe B, Poole RK (1999). "Microbial ubiquinones: multiple roles in respiration, gene regulation and oxidative stress management" (PDF). Microbiology (Reading, Engl.). 145 ( Pt 8): 1817&ndash, 30. PMID 10463148.
  11. Johnson D, Dean D, Smith A, Johnson M (2005). "Structure, function, and formation of biological iron-sulfur clusters". Annu Rev Biochem. 74: 247&ndash, 81. PMID 15952888.
  12. Page CC, Moser CC, Chen X, Dutton PL (1999). "Natural engineering principles of electron tunnelling in biological oxidation-reduction". Nature. 402 (6757): 47&ndash, 52. PMID 10573417.
  13. Leys D, Scrutton NS (2004). "Electrical circuitry in biology: emerging principles from protein structure". Curr. Opin. Struct. Biol. 14 (6): 642&ndash, 7. PMID 15582386.
  14. Boxma B, de Graaf RM, van der Staay GW; et al. (2005). "An anaerobic mitochondrion that produces hydrogen". Nature. 434 (7029): 74&ndash, 9. PMID 15744302.
  15. 15.0 15.1 Hirst J (2005). "Energy transduction by respiratory complex I--an evaluation of current knowledge" (PDF). Biochem. Soc. Trans. 33 (Pt 3): 525&ndash, 9. PMID 15916556.
  16. 16.0 16.1 16.2 Lenaz G, Fato R, Genova M, Bergamini C, Bianchi C, Biondi A (2006). "Mitochondrial Complex I: structural and functional aspects". Biochim Biophys Acta. 1757 (9–10): 1406&ndash, 20. PMID 16828051.
  17. Sazanov L.A., Hinchliffe P. (2006) Structure of the hydrophilic domain of respiratory complex I from Thermus thermophilus. Science 311, 1430-1436
  18. Baranova EA, Holt PJ, Sazanov LA (2007). "Projection structure of the membrane domain of Escherichia coli respiratory complex I at 8 A resolution". J. Mol. Biol. 366 (1): 140&ndash, 54. PMID 17157874.
  19. Friedrich T, Böttcher B (2004). "The gross structure of the respiratory complex I: a Lego System". Biochim. Biophys. Acta. 1608 (1): 1&ndash, 9. PMID 14741580.
  20. Sazanov L.A., Hinchliffe P. (2006) Structure of the hydrophilic domain of respiratory complex I from Thermus thermophilus. Science 311, 1430-1436
  21. Brandt U, Kerscher S, Dröse S, Zwicker K, Zickermann V (2003). "Proton pumping by NADH:ubiquinone oxidoreductase. A redox driven conformational change mechanism?". FEBS Lett. 545 (1): 9&ndash, 17. PMID 12788486.
  22. Cecchini G (2003). "Function and structure of complex II of the respiratory chain". Annu Rev Biochem. 72: 77&ndash, 109. PMID 14527321.
  23. Yankovskaya V., Horsefield R., Tornroth S., Luna-Chavez C., Miyoshi H., Leger C., Byrne B., Cecchini G., Iwata S. (2003) Architecture of succinate dehydrogenase and reactive oxygen species generation. Science 299, 700-704
  24. Horsefield R, Iwata S, Byrne B (2004). "Complex II from a structural perspective". Curr. Protein Pept. Sci. 5 (2): 107&ndash, 18. PMID 15078221.
  25. Kita K, Hirawake H, Miyadera H, Amino H, Takeo S (2002). "Role of complex II in anaerobic respiration of the parasite mitochondria from Ascaris suum and Plasmodium falciparum". Biochim. Biophys. Acta. 1553 (1–2): 123&ndash, 39. PMID 11803022.
  26. Painter HJ, Morrisey JM, Mather MW, Vaidya AB (2007). "Specific role of mitochondrial electron transport in blood-stage Plasmodium falciparum". Nature. 446 (7131): 88&ndash, 91. PMID 17330044.
  27. Ramsay RR, Steenkamp DJ, Husain M (1987). "Reactions of electron-transfer flavoprotein and electron-transfer flavoprotein: ubiquinone oxidoreductase". Biochem. J. 241 (3): 883&ndash, 92. PMID 3593226.
  28. Zhang J, Frerman FE, Kim JJ (2006). "Structure of electron transfer flavoprotein-ubiquinone oxidoreductase and electron transfer to the mitochondrial ubiquinone pool". Proc. Natl. Acad. Sci. U.S.A. 103 (44): 16212&ndash, 7. doi:10.1073/pnas.0604567103. PMID 17050691.
  29. Ikeda Y, Dabrowski C, Tanaka K (1983). "Separation and properties of five distinct acyl-CoA dehydrogenases from rat liver mitochondria. Identification of a new 2-methyl branched chain acyl-CoA dehydrogenase". J. Biol. Chem. 258 (2): 1066&ndash, 76. PMID 6401712.
  30. Ruzicka FJ, Beinert H (1977). "A new iron-sulfur flavoprotein of the respiratory chain. A component of the fatty acid beta oxidation pathway" (PDF). J. Biol. Chem. 252 (23): 8440&ndash, 5. PMID 925004.
  31. Ishizaki K, Larson TR, Schauer N, Fernie AR, Graham IA, Leaver CJ (2005). "The critical role of Arabidopsis electron-transfer flavoprotein:ubiquinone oxidoreductase during dark-induced starvation". Plant Cell. 17 (9): 2587&ndash, 600. PMID 16055629.
  32. Berry E, Guergova-Kuras M, Huang L, Crofts A (2000). "Structure and function of cytochrome bc complexes". Annu Rev Biochem. 69: 1005&ndash, 75. PMID 10966481.
  33. Crofts AR (2004). "The cytochrome bc1 complex: function in the context of structure". Annu. Rev. Physiol. 66: 689&ndash, 733. PMID 14977419.
  34. Iwata S, Lee JW, Okada K; et al. (1998). "Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex". Science. 281 (5373): 64&ndash, 71. PMID 9651245.
  35. Trumpower BL (1990). "The protonmotive Q cycle. Energy transduction by coupling of proton translocation to electron transfer by the cytochrome bc1 complex" (PDF). J. Biol. Chem. 265 (20): 11409&ndash, 12. PMID 2164001.
  36. Hunte C, Palsdottir H, Trumpower BL (2003). "Protonmotive pathways and mechanisms in the cytochrome bc1 complex". FEBS Lett. 545 (1): 39&ndash, 46. PMID 12788490.
  37. Calhoun M, Thomas J, Gennis R (1994). "The cytochrome oxidase superfamily of redox-driven proton pumps". Trends Biochem Sci. 19 (8): 325&ndash, 30. PMID 7940677.
  38. Tsukihara T, Aoyama H, Yamashita E, Tomizaki T, Yamaguchi H, Shinzawa-Itoh K, Nakashima R, Yaono R, Yoshikawa S. (1996). "TThe whole structure of the 13-subunit oxidized cytochrome c oxidase at 2.8 A.". Science. 272 (5265): 1136&ndash, 44. PMID 8638158.
  39. Yoshikawa S, Muramoto K, Shinzawa-Itoh K; et al. (2006). "Proton pumping mechanism of bovine heart cytochrome c oxidase". Biochim. Biophys. Acta. 1757 (9–10): 1110&ndash, 6. PMID 16904626.
  40. Rasmusson AG, Soole KL, Elthon TE (2004). "Alternative NAD(P)H dehydrogenases of plant mitochondria". Annual review of plant biology. 55: 23&ndash, 39. PMID 15725055.
  41. Menz RI, Day DA (1996). "Purification and characterization of a 43-kDa rotenone-insensitive NADH dehydrogenase from plant mitochondria". J. Biol. Chem. 271 (38): 23117&ndash, 20. PMID 8798503.
  42. McDonald A, Vanlerberghe G (2004). "Branched mitochondrial electron transport in the Animalia: presence of alternative oxidase in several animal phyla". IUBMB Life. 56 (6): 333&ndash, 41. PMID 15370881.
  43. Sluse FE, Jarmuszkiewicz W (1998). "Alternative oxidase in the branched mitochondrial respiratory network: an overview on structure, function, regulation, and role". Braz. J. Med. Biol. Res. 31 (6): 733&ndash, 47. PMID 9698817.
  44. Moore AL, Siedow JN (1991). "The regulation and nature of the cyanide-resistant alternative oxidase of plant mitochondria". Biochim. Biophys. Acta. 1059 (2): 121&ndash, 40. PMID 1883834.
  45. Vanlerberghe GC, McIntosh L (1997). "Alternative oxidase: From Gene to Function". 48: 703&ndash, 34. PMID 15012279.
  46. Ito Y, Saisho D, Nakazono M, Tsutsumi N, Hirai A (1997). "Transcript levels of tandem-arranged alternative oxidase genes in rice are increased by low temperature". Gene. 203 (2): 121&ndash, 9. PMID 9426242.
  47. Maxwell DP, Wang Y, McIntosh L (1999). "The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells". Proc. Natl. Acad. Sci. U.S.A. 96 (14): 8271&ndash, 6. PMID 10393984.
  48. Heinemeyer J, Braun HP, Boekema EJ, Kouril R (2007). "A structural model of the cytochrome C reductase/oxidase supercomplex from yeast mitochondria". J. Biol. Chem. 282 (16): 12240&ndash, 8. PMID 17322303.
  49. Schägger H, Pfeiffer K (2000). "Supercomplexes in the respiratory chains of yeast and mammalian mitochondria". EMBO J. 19 (8): 1777&ndash, 83. PMID 10775262.
  50. Schägger H (2002). "Respiratory chain supercomplexes of mitochondria and bacteria". Biochim. Biophys. Acta. 1555 (1–3): 154&ndash, 9. PMID 12206908.
  51. Schägger H, Pfeiffer K (2001). "The ratio of oxidative phosphorylation complexes I-V in bovine heart mitochondria and the composition of respiratory chain supercomplexes". J. Biol. Chem. 276 (41): 37861&ndash, 7. PMID 11483615.
  52. Gupte S, Wu ES, Hoechli L; et al. (1984). "Relationship between lateral diffusion, collision frequency, and electron transfer of mitochondrial inner membrane oxidation-reduction components". Proc. Natl. Acad. Sci. U.S.A. 81 (9): 2606&ndash, 10. PMID 6326133.
  53. Nealson KH (1999). "Post-Viking microbiology: new approaches, new data, new insights". Origins of life and evolution of the biosphere : the journal of the International Society for the Study of the Origin of Life. 29 (1): 73&ndash, 93. PMID 11536899.
  54. Schäfer G, Engelhard M, Müller V (1999). "Bioenergetics of the Archaea". Microbiol. Mol. Biol. Rev. 63 (3): 570–620. PMID 10477309.
  55. 55.0 55.1 Ingledew WJ, Poole RK (1984). "The respiratory chains of Escherichia coli". Microbiol. Rev. 48 (3): 222&ndash, 71. PMID 6387427.
  56. 56.0 56.1 Unden G, Bongaerts J (1997). "Alternative respiratory pathways of Escherichia coli: energetics and transcriptional regulation in response to electron acceptors". Biochim. Biophys. Acta. 1320 (3): 217&ndash, 34. PMID 9230919.
  57. Cecchini G, Schröder I, Gunsalus RP, Maklashina E (2002). "Succinate dehydrogenase and fumarate reductase from Escherichia coli". Biochim. Biophys. Acta. 1553 (1–2): 140&ndash, 57. PMID 11803023.
  58. Freitag A, Bock E (1990). "Energy conservation in Nitrobacter". FEMS Microbiology Letters. 66 (1&ndash, 3): 157&ndash:62. doi:10.1111/j.1574-6968.1990.tb03989.x.
  59. Starkenburg SR, Chain PS, Sayavedra-Soto LA; et al. (2006). "Genome sequence of the chemolithoautotrophic nitrite-oxidizing bacterium Nitrobacter winogradskyi Nb-255". Appl. Environ. Microbiol. 72 (3): 2050&ndash, 63. PMID 16517654.
  60. Yamanaka T, Fukumori Y (1988). "The nitrite oxidizing system of Nitrobacter winogradskyi". FEMS Microbiol. Rev. 4 (4): 259&ndash, 70. PMID 2856189.
  61. Iuchi S, Lin EC (1993). "Adaptation of Escherichia coli to redox environments by gene expression". Mol. Microbiol. 9 (1): 9&ndash, 15. PMID 8412675.
  62. Calhoun MW, Oden KL, Gennis RB, de Mattos MJ, Neijssel OM (1993). "Energetic efficiency of Escherichia coli: effects of mutations in components of the aerobic respiratory chain" (PDF). J. Bacteriol. 175 (10): 3020&ndash, 5. PMID 8491720.
  63. 63.0 63.1 Boyer PD (1997). "The ATP synthase--a splendid molecular machine". Annu. Rev. Biochem. 66: 717&ndash, 49. PMID 9242922.
  64. Van Walraven HS, Strotmann H, Schwarz O, Rumberg B (1996). "The H+/ATP coupling ratio of the ATP synthase from thiol-modulated chloroplasts and two cyanobacterial strains is four". FEBS Lett. 379 (3): 309–13. PMID 8603713.
  65. Yoshida M, Muneyuki E, Hisabori T (2001). "ATP synthase--a marvellous rotary engine of the cell". Nat. Rev. Mol. Cell Biol. 2 (9): 669–77. PMID 11533724.
  66. Schemidt RA, Qu J, Williams JR, Brusilow WS (1998). "Effects of carbon source on expression of F0 genes and on the stoichiometry of the c subunit in the F1F0 ATPase of Escherichia coli". J. Bacteriol. 180 (12): 3205–8. PMID 9620972.
  67. Nelson N, Perzov N, Cohen A, Hagai K, Padler V, Nelson H (2000). "The cellular biology of proton-motive force generation by V-ATPases". J. Exp. Biol. 203 (Pt 1): 89&ndash, 95. PMID 10600677.
  68. Rubinstein JL, Walker JE, Henderson R (2003). "Structure of the mitochondrial ATP synthase by electron cryomicroscopy". EMBO J. 22 (23): 6182&ndash, 92. PMID 14633978.
  69. Leslie AG, Walker JE (2000). "Structural model of F1-ATPase and the implications for rotary catalysis". Philos. Trans. R. Soc. Lond., B, Biol. Sci. 355 (1396): 465&ndash, 71. PMID 10836500.
  70. Noji H, Yoshida M (2001). "The rotary machine in the cell, ATP synthase". J. Biol. Chem. 276 (3): 1665–8. PMID 11080505.
  71. Capaldi R, Aggeler R (2002). "Mechanism of the F(1)F(0)-type ATP synthase, a biological rotary motor". Trends Biochem Sci. 27 (3): 154&ndash, 60. PMID 11893513.
  72. 72.0 72.1 Gresser MJ, Myers JA, Boyer PD (1982). "Catalytic site cooperativity of beef heart mitochondrial F1 adenosine triphosphatase. Correlations of initial velocity, bound intermediate, and oxygen exchange measurements with an alternating three-site model". J. Biol. Chem. 257 (20): 12030&ndash, 8. PMID 6214554.
  73. Dimroth P (1994). "Bacterial sodium ion-coupled energetics". Antonie Van Leeuwenhoek. 65 (4): 381&ndash, 95. PMID 7832594.
  74. 74.0 74.1 Becher B, Müller V (1994). "Delta mu Na+ drives the synthesis of ATP via a delta mu Na(+)-translocating F1F0-ATP synthase in membrane vesicles of the archaeon Methanosarcina mazei Gö1". J. Bacteriol. 176 (9): 2543&ndash, 50. PMID 8169202.
  75. Müller V (2004). "An exceptional variability in the motor of archael A1A0 ATPases: from multimeric to monomeric rotors comprising 6–13 ion binding sites". J. Bioenerg. Biomembr. 36 (1): 115&ndash, 25. PMID 15168615.
  76. 76.0 76.1 Davies K (1995). "Oxidative stress: the paradox of aerobic life". Biochem Soc Symp. 61: 1&ndash, 31. PMID 8660387.
  77. Rattan SI (2006). "Theories of biological aging: genes, proteins, and free radicals". Free Radic. Res. 40 (12): 1230&ndash, 8. PMID 17090411.
  78. Valko M, Leibfritz D, Moncol J, Cronin MT, Mazur M, Telser J (2007). "Free radicals and antioxidants in normal physiological functions and human disease". Int. J. Biochem. Cell Biol. 39 (1): 44&ndash, 84. PMID 16978905.
  79. Raha S, Robinson B (2000). "Mitochondria, oxygen free radicals, disease and ageing". Trends Biochem Sci. 25 (10): 502&ndash, 8. PMID 11050436.
  80. Finkel T, Holbrook NJ (2000). "Oxidants, oxidative stress and the biology of ageing". Nature. 408 (6809): 239&ndash, 47. PMID 11089981.
  81. 81.0 81.1 Joshi S, Huang YG (1991). "ATP synthase complex from bovine heart mitochondria: the oligomycin sensitivity conferring protein is essential for dicyclohexyl carbodiimide-sensitive ATPase". Biochim. Biophys. Acta. 1067 (2): 255&ndash, 8. PMID 1831660.
  82. Tsubaki M (1993). "Fourier-transform infrared study of cyanide binding to the Fea3-CuB binuclear site of bovine heart cytochrome c oxidase: implication of the redox-linked conformational change at the binuclear site". Biochemistry. 32 (1): 164&ndash, 73. PMID 8380331.
  83. Heytler PG (1979). "Uncouplers of oxidative phosphorylation". Meth. Enzymol. 55: 462&ndash, 42. PMID 156853.
  84. Lambert AJ, Brand MD (2004). "Inhibitors of the quinone-binding site allow rapid superoxide production from mitochondrial NADH:ubiquinone oxidoreductase (complex I)". J. Biol. Chem. 279 (38): 39414&ndash, 20. PMID 15262965.
  85. Ricquier D, Bouillaud F (2000). "The uncoupling protein homologues: UCP1, UCP2, UCP3, StUCP and AtUCP". Biochem. J. 345 Pt 2: 161–79. PMID 10620491.
  86. Borecký J, Vercesi AE (2005). "Plant uncoupling mitochondrial protein and alternative oxidase: energy metabolism and stress". Biosci. Rep. 25 (3–4): 271–86. PMID 16283557.
  87. Harden A, Young WJ. (1906). "The alcoholic ferment of yeast-juice". Proc. R. Soc. (Lond.). B (77): 405&ndash, 20.
  88. Kalckar HM (1974). "Origins of the concept oxidative phosphorylation". Mol. Cell. Biochem. 5 (1–2): 55&ndash, 63. PMID 4279328.
  89. Lipmann F, (1941). "Metabolic generation and utilization of phosphate bond energy". Adv Enzymol. 1: 99&ndash, 162.
  90. Friedkin M, Lehninger AL. (1949). "Esterification of inorganic phosphate coupled to electron transport between dihydrodiphosphopyridine nucleotide and oxygen". J. Biol. Chem. 178 (2): 611&ndash, 23.
  91. Slater EC. (1953). "Mechanism of Phosphorylation in the Respiratory Chain". Nature. 172 (4387): 975. doi:10.1038/172975a0.
  92. Mitchell P. (1961). "Coupling of Phosphorylation to Electron and Hydrogen Transfer by a Chemi-Osmotic type of Mechanism". Nature. 191 (4784): 144. PMID 13771349.
  93. Milton H. Saier Jr. "Peter Mitchell and the Vital Force". Retrieved 2007-08-23.
  94. Mitchell, Peter (1978). "David Keilin's Respiratory Chain Concept and Its Chemiosmotic Consequences" (Pdf). Nobel lecture. Nobel Foundation. Retrieved 2007-07-21.
  95. Pullman ME, Penefsky HS, Datta A, and Racker E. (1960). "Partial Resolution of the Enzymes Catalyzing Oxidative Phosphorylation. I. Purification and Properties of Soluble, Dinitrophenol-stimulated Adenosine Triphosphatase". J. Biol. Chem. 235 (11): 3322&ndash, 3329.
  96. Boyer PD, Cross RL, Momsen W (1973). "A new concept for energy coupling in oxidative phosphorylation based on a molecular explanation of the oxygen exchange reactions". Proc. Natl. Acad. Sci. U.S.A. 70 (10): 2837&ndash, 9. PMID 4517936.
  97. "The Nobel Prize in Chemistry 1997". Nobel Foundation. Retrieved 2007-07-21.

Further reading

Introductory

  • Nelson DL (2004). Lehninger Principles of Biochemistry (4th ed ed.). W. H. Freeman. ISBN 0-716-74339-6. Unknown parameter |coauthors= ignored (help)
  • Schneider ED (2006). Into the Cool: Energy Flow, Thermodynamics and Life (1st ed ed.). University of Chicago Press. ISBN 0-226-73937-6. Unknown parameter |coauthors= ignored (help)

Advanced

  • Nicholls DG (2002). Bioenergetics 3 (1st ed ed.). Academic Press. ISBN 0-125-18121-3. Unknown parameter |coauthors= ignored (help)
  • Haynie D (2001). Biological Thermodynamics (1st ed ed.). Cambridge University Press. ISBN 0-521-79549-4.
  • Rajan SS (2003). Introduction to Bioenergetics (1st ed ed.). Anmol. ISBN 8-126-11364-2.
  • Wikstrom M (Ed) (2005). Biophysical and Structural Aspects of Bioenergetics (1st ed ed.). Royal Society of Chemistry. ISBN 0-854-04346-2.

External links

General resources

Structural resources

ar:فسفرة أكسيدية de:Oxidative Phosphorylierung he:זרחון חמצוני nl:Oxidatieve fosforylatie fi:Oksidatiivinen fosforylaatio sv:Oxidativ fosforylering

Template:WH Template:WikiDoc Sources