Pages

Strong Experimental Evidence Implicates the Proton-Motive Force in ATP Synthesis


The transmembrane proton gradient predicted by chemiosmotic theory has been experimentally measured. When intact mitochondria are suspended with an oxidizable substrate such as succinate in a lightly buf fered medium, the addition of O2 results in acidification of the suspending medium (Fig. 18-19). The stoichiometry of proton pumping can be estimated from the pH changes, after correcting for buffering effects. The measurement is difficult, and there is no universal agreement on the stoichiometry; about ten protons are pumped out for each pair of electrons transferred from NADH to O2. The result is a steady state with a difference of about one pH unit between the matrix and the medium, inside (matrix) alkaline. Indirect measurements of the transmembrane electrical potential in respiring mitochondria yield a value of 0.1 to 0.2 V, inside negative. Substituting these values into the equation for proton-motive force (Eqn 18-3) gives a value of 15 to 25 kJ/mol, the free-energy change for the transmembrane movement of one equivalent of protons back into the mitochondrial matrix. This free-energy change is large enough to account for the synthesis of one ATP when two to three protons flow into the matrix.
Uncouplers are hydrophobic weak acids (Fig. 18-14). Their hydrophobicity allows them to diffuse readily across mitochondrial membranes. After entering the mitochondrial matrix in the protonated form, they can release a proton (dissociate), thus dissipating the proton gradient. Ionophores uncouple electron transfer from oxidative phosphorylation by creating electrical short circuits across the mitochondrial membrane (Fig. 18-20). The toxic ionophore valinomycin forms a lipid-soluble complex with K+, which is abundant in the cytosol (see Fig. 10-26). Whereas K+ penetrates the mitochondrial inner membrane only very slowly, the K+-valinomycin complex readily passes through. The influx of positive ions neutralizes the excess of negative charge inside the matrix, diminishing the electrical component of the proton-motive force. Valinomycin slows mitochondrial ATP synthesis without blocking electron transfer to O2
If the role of electron transfer in mitochondrial ATP synthesis is simply to pump protons to create the electrochemical potential of the proton-motive force, an artificially created proton gradient should be able to replace electron transfer in driving ATP synthesis. This prediction of the chemiosmotic theory has been experimentally tested and confirmed (Fig. 18-21). Mitochondria manipulated so as to impose a difference of proton concentration and a separation of charge across the inner membrane (Fig. 18-21) carry out the synthesis of ATP in the absence of an oxidizable substrate; the proton-motive force alone suf fices to drive ATP synthesis.
Chemiosmotic theory also accounts for a third condition that uncouples oxidation from phosphorylation-mechanical disruption of the mitochondrial membrane. Without an intact membrane there can be no proton gradient, hence no energy conservation and no ATP synthesis.

The Detailed Mechanism of ATP Formation ftemains Elusive

Although it is clear that a transmembrane proton gradient provides the energy for ATP synthesis, it is not clear how this energy is transmitted to the ATP synthase. This enzyme catalyzes the conversion of enzyme-bound ADP and Pi into bound ATP even in the absence of a proton gradient. Remarkably, it appears that the reaction ADP + Pi ATP + H2O is readily reversible-that the free-energy change for ATP synthesis on the enzyme surface is close to zero. However, the ATP formed in this reaction remains tightly bound to the active site, preventing further catalysis at that site. The proton-motive force apparently supplies the energy needed to force dissociation of the tightly bound ATP from the enzyme, allowing another catalytic cycle to begin.
How is it possible that ATP synthesis, known to be strongly endergonic in aqueous solution (ΔG°' = 30.5 kJ/mol), is readily reversible on the enzyme surface? The very tight noncovalent binding of enzyme to ATP may supply enough binding energy (Chapter 8) to make bound ATP about as stable as its hydrolysis products (Fig. 18-22). Alternatively, the reaction may occur in a very hydrophobic pocket in the enzyme interior, where the energetics of hydrolysis in water do not apply.

Each F1 complex has the subunit composition α3β3γδε (Fig. 18-15). A tight binding site for ATP, apparently identical to the catalytic site, is located on each β subunit, or perhaps between each β and its associated a subunit. Every F1 complex therefore has three ATP-synthesizing sites, which interact with F0 through the single copies of γ, δ, and ε subunits.
On the basis of detailed kinetic and binding studies of the reactions catalyzed by F0F1, Paul Boyer has suggested a mechanism (Fig. 18-23) in which the three active sites on F1 alternate in catalyzing ATP synthesis. The limiting step in the process is the release of newly synthesized ATP from the enzyme. A conformational transition driven by the proton-motive force reduces the enzyme's affinity for ATP.
The transition induced by the proton-motive force may be envisioned as a 120° rotation of the α3β3 portion of F1 (Fig. 18-23), placing one of the three α-β pairs in a special position relative to the proton channel of F0. However, no physical rotation of F1 has been demonstrated, and the model does not require it; the conformational change that interconverts the three types of sites may be an allosteric transition, in which changes at one α-β pair force compensating changes in the other two pairs.
What makes the mechanism of this enzyme particularly difficult to solve is the Uectorial nature of the process it catalyzes. Somehow, the enzyme must sense not merely a certain concentration of protons, but a difference of proton concentration in two regions of space. Determination of the detailed structure of F1 by x-ray crystallography should shed light on the mechanism of its action.

The Proton-Motive Force Energizes Active Transport

The primary role of electron transfer in mitochondria is to furnish energy for the synthesis of ATP during oxidative phosphorylation, but this energy serves also to drive several transport processes essential to oxidative phosphorylation. We have seen that the inner mitochondrial membrane is generally impermeable to charged species, but two specific systems in the inner mitochondrial membrane transport ADP and Pi into the matrix and ATP out to the cytosol (Fig. 18-24). The adenine nucleotide translocase, which extends across the inner membrane, binds ADP3- on the outside (cytosolic) surface of the inner membrane and transports it inward in exchange for an ATP4- molecule simultaneously transported outward (see Fig. 13-1 for the ionic forms of ATP and ADP). Because this antiporter moves four negative charges out for each three moved in, its activity is favored by the transmembrane electrochemical gradient, which gives the matrix a net negative charge; the proton-motive force drives ATP-ADP exchange.

Adenine nucleotide translocase is specifically inhibited by atractyloside, a toxic glycoside formed by a species of thistle; for centuries it has been known that grazing cattle are poisoned when they ingest this plant. If the transport of ADP into and ATP out of the mitochondria is inhibited, cytosolic ATP cannot be regenerated from ADP, explaining the toxicity of atractyloside
A second membrane transport system essential to oxidative phosphorylation is the phosphate translocase, which promotes symport of one H2P04- and one H+ into the matrix. This transport process, too, is favored by the transmembrane proton gradient (Fig. 18-24).

Shuttle Systems Are Required for Mitochondrial Oxidation of Cytosolic NADH

The NADH dehydrogenase of the inner mitochondrial membrane of animal cells can accept electrons only from NADH in the matrix. Given that the inner membrane is not permeable to cytosolic NADH, how can the NADH generated by glycolysis outside mitochondria be reoxidized to NAD+ by O2 via the respiratory chain? Special shuttle systems carry reducing equivalents from cytosolic NADH into mitochondria by an indirect route. The most active NADH shuttle, which functions in liver, kidney, and heart mitochondria, is the malate-aspartate shuttle (Fig. 18-25). The reducing equivalents of cytosolic NADH are first transferred to cytosolic oxaloacetate to yield malate by the action of cytosolic malate dehydrogenase. The malate thus formed passes through the inner membrane into the matrix via the malatea-ketoglutarate transport system. Within the matrix the reducing equivalents are passed by the action of matrix malate dehydrogenase to matrix NAD+, forming NADH; this NADH can then pass electrons directly to the respiratory chain in the inner membrane. Three molecules of ATP are generated as this pair of electrons passes to 02. Cytosolic oxaloacetate must be regenerated via transamination reactions (see Fig. 17-5) and the activity of membrane transporters (Fig. 18-25) to start another cycle of the shuttle.
In skeletal muscle and brain, another type of NADH shuttle, the glycerol-3-phosphate shuttle, occurs (Fig. 18-26). It differs from the malate-aspartate shuttle in that it delivers the reducing equivalents from NADH into Complex III, not Complex I (Fig. 18-9), providing only enough energy to synthesize two ATP molecules per pair of electrons.
The mitochondria of higher plants have an externally oriented NADH dehydrogenase that is able to transfer electrons directly from cytosolic NADH into the respiratory chain.

Oxidative Phosphorylation Produces Most of the ATP Made in Aerobic Cells

The complete oxidation of a molecule of glucose to CO2 yields two ATP and two NADH from glycolysis in the cytosol (Chapter 14); two NADH from pyruvate oxidation in the mitochondrial matrix  and two ATP, six NADH, and two FADH2 from citric acid cycle reactions in the matrix Each NADH produced in the matrix yields three ATP from mitochondrial oxidative phosphorylation, and for each FADH2, two ATP are generated. Cytosolic NADH, after shuttling into the matrix, yields two or three ATP, depending on which shuttle is used. The total yield of glucose oxidation is therefore 36 or 38 ATP per glucose

By comparison, glycolysis under anaerobic conditions (lactate fermentation) yields only two ATP per glucose. Clearly, the evolution of oxidative phosphorylation provided a tremendous increase in the energetic efficiency of catabolism.
The oxidation of fatty acids and amino acids also takes place within the mitochondrial matrix, and the FADH2 and NADH produced by these oxidative pathways also serve as electron donors for oxidative phosphorylation. Complete oxidation of the 16-carbon saturated fatty acid palmitate to CO2 produces eight acetyl-CoAs, seven FADH2, and seven NADH. The oxidation of each acetyl-CoA via the citric acid cycle produces three NADH, one FADH2, and one ATP (GTP). The gain in ATP is therefore 131 ATP but, because the activation of palmitate to palmitoyl-CoA costs two ATP equivalents, the net gain is 129 ATP per palmitate. A similar calculation may be made for the ATP yield upon oxidation of each of the amino acids. The important conclusion here is that aerobic oxidative pathways that result in electron transfer to O2 accompanied by oxidative phosphorylation account for the vast majority of the ATP produced in catabolism.