Critical Review Mitochondrial Energetic Metabolism—Some General Principles

October 13, 2017 | Autor: Jean-Pierre Mazat | Categoría: Biochemistry, Molecular Biology, Membrane Proteins, Mitochondria
Share Embed


Descripción

Critical Review Mitochondrial Energetic Metabolism—Some General Principles

Jean-Pierre Mazat1,2* phane Ransac1,2 Ste Margit Heiske1,2,3 Anne Devin1,2 Michel Rigoulet1,2

1

CNRS-UMR5095, Institute of Biochemistry and Genetics of the Cell, 1 € ns, 33077 Bordeaux cedex, France Rue Camille Saint Sae 2 Univ. Bordeaux, IBGC, UMR 5095, 33077 Bordeaux cedex, France 3 Institut fu¨r Biologie, Theoretische Biophysik, Humboldt-Universita€t zu Berlin, Invalidenstrabe 42, Berlin, Germany

Abstract Summary: In nonphotosynthetic organisms, mitochondria are the power plant of the cell, emphasizing their great potentiality for adenosine triphosphate (ATP) synthesis from the redox span between nutrients and oxygen. Also of great importance is their role in the maintenance of the cell redox balance. Even though crystallographic structures of respiratory complexes, ATP synthase, and ATP/adenosine diphosphate (ADP) carrier are now

quite well known, the coupling between ATP synthesis and cell redox state remains a controversial issue. In this review, we will present some of the processes that allow a modular coupling between ATP synthesis and redox state. Furthermore, we will present some theoretical approaches of this highly integrated C 2013 IUBMB Life, 65(3):171–179, 2013 system. V

Keywords: mitochondria; membrane proteins; membrane permeability; mitochondrial disorders; electron transfer in proteins

Introduction During aerobic growth, a great part of the adenosine triphosphate (ATP) produced in nonphotosynthetic cells originates from mitochondrial oxidative phosphorylation (OxPhos). The respiratory chain (Fig. 1) supports a sequence of redox reactions in which electrons passing along a series of enzymes located in the inner mitochondrial membrane release free energy that is used

Abbreviations DGox, (Joule) Gibbs energy difference of a redox reactions: DGox ¼ nF  DE, where n is the number of electron involved, DE (volt) ¼ Eacceptor  Edonor and E ¼ Eh,7  RT/nF  ln[(Red)/(Ox)]; DGP, phosphorylation potential: DGP ¼ DGp,o þ RTln[(ATP)/(ADP)(Pi)], DGp,o is the free energy of ATP synthesis at standard conditions. DlHþ , gradient of electrochemical potential of protons across a membrane: DlHþ ¼ F  DW  2,3RT. DpH (J/mole); DW, membrane potential. Dp, proton-motive force: Dp ¼ DlHþ /F (volt); ROS, reactive oxygen species. C 2013 International Union of Biochemistry and Molecular Biology, Inc. V Volume 65, Number 3, March 2013, Pages 171–179 *Address for correspondence to: Jean-Pierre Mazat, CNRS-UMR5095, Institute of Biochemistry and Genetics of the Cell, 1 Rue Camille Saint € ns, 33077 Bordeaux cedex, France. Tel: +33 5 56 99 90 41; Sae Fax: +33 5 56 99 90 40 E-mail: [email protected]. Received 22 November 2012; accepted 23 December 2012 DOI: 10.1002/iub.1138 Published online in Wiley Online Library (wileyonlinelibrary.com)

IUBMB Life

for the translocation of protons across this membrane. This proton flux from the matrix to the intermembrane space establishes a difference in proton electrochemical potential (called protonmotive force), which is used by different energy transducers and particularly by membrane bound ATP-synthase to convert adenosine diphosphate (ADP) and phosphate to ATP. This sequence of reactions called OxPhos ensures two important cell functions, the redox state maintenance (mainly nicotinamide adenine dinucleotide (NADH) reoxidation) and ATP synthesis (2). Figure 1 is an idealized representation of OxPhos. Indeed, several other proteins exist in the inner mitochondrial membrane, which can feed (usually to a lesser extent) the respiratory chain with electrons such as glycerol-3-phosphate (G3P) dehydrogenase, electron transfer flavoprotein dehydrogenase, dihydroorotate dehydrogenase, and choline dehydrogenase. In the yeast Saccharomyces cerevisiae and in some fungi, there are both external and internal NADH dehydrogenases that reduce the quinone pool without extruding protons from the matrix. For space restrictions, we will not approach the diversity of respiratory chains in this short review (3), but rather we will review the thermodynamics of the respiratory complexes, the link (stoichiometry) between the conservation of the cell redox state and ATP synthesis, the control of OxPhos in the light of the metabolic control theory (MCT), and finally we will present some OxPhos mathematical models.

171

IUBMB LIFE

FIG 1

Mitochondrial OxPhos. Pumped protons are in red. Chemical protons are in brown. Electrons transfers are in blue. The scheme is designed with transfer of two electrons, even for complex IV (half reaction). PDB codes are: 2FUG and 3RKO for complex I (in blue); 3SFE for complex II (in red); 1PP9 for complex III (dimer in magenta and cyan); 1OCR for complex IV (in green); 2B4Z for cyt c (in grey); 2C3E for ADP/ATP carrier (in cyan) and Pi carrier (in purple); (1) for ATPase. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

Thermodynamics of Respiratory Chain Complexes The enzymatic reactions catalyzed by the four complexes of the respiratory chain are as follows: Complex I: NADH þ Q þ 5Hþin $ NADþ þ QH2 þ 4Hþout Complex II: Succinate þ Q $ fumarate þ QH2 Complex III: QH2 þ 2 cyt cox þ 2Hþin $ Q þ 2 cyt cred þ 4Hþout Complex IV: 2 cyt cred þ 1/2 O2 þ 4Hþin $ 2 cyt cox þ H2O þ 2Hþout I þ III þ IV: NADH þ 1/2 O2 þ 11Hþin $ NADþ þ H2O þ 10Hþout II þ III þ IV: Succinate þ 1/2 O2 þ 6Hþin $ fumarate þ H2O þ 6Hþout The DG0 of these reactions are listed in Table 1. Important differences in proton transfers exist between the mitochondrial respiratory chain complexes. First, the energy produced by redox reaction at complex II (DG0 ¼ 5.8 kJ/mol, see Table 1) is too low to allow proton transfer across the membrane. Second, complexes I and IV have special protons pathways to transfer protons across the membrane, whereas complex III transfers protons according to the modified Q cycle mechanism (5,6). Complex III is not a pump because the protons released on the intermembrane side are not the same protons taken up on the matrix side. Instead, chemical protons are taken from the matrix at the Qi site and transferred into the Q pool; other chemical protons are expelled into the intermembrane space at the Qo site from the Q pool, taking advantage of the position of Qo and Qi sites close to the intermembrane space and the matrix side, respectively. Thus, the stoichiometry at complex III is strictly 2Hþ/2e. On the contrary, the coupling mechanism between redox reactions and proton transfers operating in complex I is the

172

object of great debates, recently revived by the crystallographic structure of this complex (7,8). Complex I has an L shape with the redox reactions in the hydrophilic arm of the L well separated from three proton-pumping antiporter-like subunits L, M, and N (Escherichia coli nomenclature) situated in the membrane (the membranous arm of the L) linked by a long a-helix parallel to the membrane. Subunits N, K, J, and A form the ‘‘heel’’ of the ‘‘boot’’ structure of complex I and are probably involved in the coupling mechanism and perhaps in a putative fourth Hþ transfer (Fig. 1). The fact that only three subunits clearly involved in proton transfer are apparent in this structure suggests that the number of protons transferred per 2e could be three contrary to the well accepted values of 4 (discussed in ref. 9). Regarding the number of protons pumped through complex I, thermodynamics imposes the constraint that:

DEh  n 0 Dp where DEh ¼ EhUQ  EhNADH (Eh ¼ Em þ RT/nF  ln([Ox]/[Red])) is the redox span of complex I reaction and Dp (mV) ¼ Dl ~Hþ /F (F ¼ 96,500 coulomb/mole) is the proton-motive force (Dl ~Hþ (kJ/mole) is the electrochemical proton gradient), n is the number of electrons transferred, and n0 is the number of protons transferred. The Eh values under state 3 are given in Table 2 and represented in states 3 and 4 in Fig. 2. In state 3, the mitochondrial membrane electrochemical potential is approximately 175 mV, which necessitates at least a redox input of 2  175 ¼ 350 mV to transfer 2Hþ/e, that is, 4Hþ/2e by complex I. The matrix NADþ/NADH ratio is known to be about 7.15 (11); hence, EhNADH ¼ 294 mV (as compared to EmNADH ¼ 320 mV ). Assuming complex I is operated at equilibrium, which means that EhUQ ¼ þ56 mV (as compared to a midpoint

Mitochondrial Energetic Metabolism—Some General Principles

Mazat et al.

173

Q/QH2

2 Cyt c (ox/red)

Succinate/ fumarate

QH2/Q

2 Cyt c (red/ox) 1=2O2/H2O

I

II

III

IV

II þ III þ IV Succinate/ fumarate

1=2O /H O 2 2

1=2O /H O 2 2

Q/QH2

NADH/NADþ

complex

I þ III þ IV NADH/NADþ

Electron acceptor

2

6

2

4

10

6

4 Hþ (compound ! intermembrane space) 4 Hþ (matrix ! compound)

2

2 Hþ (matrix ! compound)

4 Hþ (compound ! intermembrane space) 5 Hþ (matrix ! compound)

2

4

1 less qþ in matrix

2 less qþ in matrix 2 less q in intermembrane space

2 qþ and 2 q to intermembrane space

1 less qþ in matrix

Charge transfer other than Total Hþ transferred transferred Hþ

4 Hþ (compound ! intermembrane space) 2 Hþ (matrix ! compound)

1 Hþ (matrix ! compound)

Pumping Hþ matrix ! intermembrane space Chemical Hþ

Energies, standard midpoint potentials, and stoichiometry in mitochondrial respiratory chain complexes

Electron donor

TABLE 1

þ30 (4)

320 (4)

þ220 (4)

30.9

5.8

73.3

þ820 (4) 152.4

þ820 (4) 220.0

þ820 (4) 115.8

þ220 (4)

þ60 (4)

þ30 (4) þ60 (4)

þ60 (4)

320 (4)

Em,7 Em,7 donor acceptor DG0 0 (mV) (mV) (kJ/mol)

IUBMB LIFE

TABLE 2

Mitochondrial redox ratio for substrates of complexes in mitochondrial respiratory chain and associated midpoint potentials and energies under state 3 (Dp ¼ 175 mV)a

Mitochondrial electron acceptor concentrations ratio

complex

Electron Electron donor acceptor

Mitochondrial electron donor concentrations ratio

I

NADH/NADþ

Q/QH2

0.14 (12% reduced) (11) 0.76 (57% reduced) (10)

II

Succinate/ fumarate

Q/QH2

2 (67% reduced)

III

QH2/Q

2 Cytochrome 1.3 (57% reduced) (10) c (ox/red)

IV

2 Cyt. c (red/ox)

1=2O /H O 2 2

I þ III þ IV

NADH/NADþ

II þ III þ IV Succinate/ fumarate a

Eh,7 donor (mV)

Eh,7 acceptor DG0 (mV) (kJ/mol)

294

56

67.5

0.76 (57% reduced) (10)

21

56

6.9

1.5 (40% reduced) (10)

56

231

33.8

0.66 (40% reduced) (10) 5.4 105 (100% reduced)

231

754

100.9

1=2O /H O 2 2

0.14 (12% reduced) (11) 5.4 105 (100% reduced)

267

754

202.2

1=2O /H O 2 2

2 (67% reduced)

5.4 105 (100% reduced)

21

754

141.6

Concentration ratios without reference have been estimated.

potential of þ60 mV) that corresponds to an ubiquinone pool 57% reduced. Benard et al. (10) measured a reduction state of this pool of 60% in liver and kidney, in agreement with this value. Hence, the redox span of complex I is sufficient in all tissues to provide the energy needed for the transfer of 4Hþ under state 3. These authors (10) also determined the percentage of reduced cytochrome c (about 40% in liver, kidney, and brain, i.e., Ehcytc ¼ þ231 mV and about 65% in muscle and heart, i.e., Ehcytc ¼ þ204 mV). Contrary to complex I, complex III has a well-established stoichiometry of 2Hþ/2e that necessitates a minimum redox span of 175 mV in state 3, which is satisfied in liver and kidney (231  56 ¼ 175 mV). However, Benard et al. (10) determined that the ubiquinone pool is only 2% reduced, EhUQ ¼ þ110 mV, in heart, muscle, and brain, which does not satisfy the minimum redox span required, reinforcing the idea of several Q pools in these tissues. It also means that at state 3, complexes I and III operate at nearequilibrium. Moreover, things are more constrained in state 4, because in this case, Dp is approximately 220 mV. Figure 2 shows that taking the same central value of EhUQ ¼ þ56 mV, we need EhNADH ¼ 384 mV (99% reduced) and EhCytc ¼ þ276 mV (10% reduced). This has been experimentally shown: Kim et al. (12) measured all the thermodynamical parameters in whole cells and found at state 4 in the presence of oligomycin with a Dp  200 mV associated with EhUQ ¼ 70 mV and EhCytc ¼ þ280 mV, which gives 210 mV as redox span for complex III and places EhNADH at at least 330 mV, that is, 68% of reduction (NADþ/NADH ¼ 0.46). Thus, although only three antiporter-like subunits exist in complex I, in principle, the transfer of four protons is thermo-

174

dynamically possible up to a Dp  200mV. The fourth pathway for protons could be situated at the interface linking the hydrophilic to the hydrophobic arms (13) (see also Fig. 1). However, for Dp > 200 mV, the transfer of 4 Hþ will be rather difficult €m and Hummer (9) and shown in Fig. 2. as stressed by Wikstro The role of the Q/QH2 ratio as an adjusting variable must be mentioned. It allows distributing the full (NADþ/NADH – cyt cox/cyt cred) span between complexes I and III to permit an optimal transfer of protons.

Link Between Redox and Energy State: are the Stoichiometry Fixed? We have shown in the previous section how OxPhos links two important cell fluxes, the ATP synthesis and NADH (and some other reduced cofactors) reoxidation. Indeed, keeping both a high phosphate potential and an intracellular redox balance is crucial for all living organisms and a necessity for sustained metabolic function. However, the specific turnovers of either NADH or ATP are not necessarily always the same. They are largely dependent on a number of variables such as cell types, metabolic activities, and phases of growth. Thus, it is necessary that the coupling between both fluxes is adaptable to different physiological situations. More precisely, it means that mechanisms must exist to vary the ratio ATP/O (ATP synthesis over oxygen consumption and thus NADH reoxidation) in response to different demands in energy and/or in reducing power. We will describe below the five mechanisms that are

Mitochondrial Energetic Metabolism—Some General Principles

FIG 2

Thermodynamics of proton transfers in complexes I and III. The Eh redox potentials of the couples of electron transporters are given according to the equation Eh ¼ Em þ RT/nF  ln([Ox]/[Red]). The Em midpoint potentials are represented with dashed line. Left part: at state 3, where the Dp is approximately 175 mV, the transfers of 4Hþ by complex I and 2Hþ by complex III are possible. Middle part: at state 4, if the Dp is approximately 220 mV, the transfers of 4Hþ by complex I and 2Hþ by complex III are still possible, but by enlarging the DEh span at the maximum width with highly reduced NAD/NADH couple and highly oxidized cytochrome c. Right part: a possible situation where complex I transfers only 3 Hþ per e at state 4. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

known to affect the ATP/O ratio and thus allow a modulation between ATP synthesis and NADH reoxydation. The first mechanism decreasing the coupling efficiency, the passive proton leak, is a direct consequence of the nature of the energetic intermediary, the proton-motive force. Indeed, biological membranes present some proton conductance (LH), and the resulting proton (in)flux is strictly dependent on the proton-motive force (JH ¼ LH  Dp). The membrane conductance is a specific property of the membrane itself, but is not entirely independent from the proton-motive force: at high value of this force, the proton membrane conductance increases. It means that LH is a function of Dp itself, LH ¼ f(Dp). This determines a non-ohmic relationship between passive proton flux (proton leak) and proton-motive force that has been observed in mitochondria of various origins (14,15) and lead to a maximal value of Dp in which the leak is equal to the flux of proton transduction by the respiratory chain (state 4). Obviously, the size of this passive proton leak may modulate the yield of OxPhos (ATP/O). More precisely, it will decrease the rate of ATP synthesis for a given oxygen consumption because the proton flux giving rise to leak cannot be used to

Mazat et al.

synthesize ATP. Passive proton leak, for a given type of mitochondria, is only dependent on the proton-motive force and consequently cannot be widely modulated. Another mitochondrial process allows modulation of the membrane proton permeability: the proton transport catalyzed by the uncoupling proteins (UCP). The expression of these proteins is tissue specific. In the case of UCP1, which is mainly expressed in brown adipose tissue, the Hþ transport function is well-established and highly regulated by guanosine triphosphate (GTP) (inhibition) and fatty acids (stimulation) (16). The participation of the other UCP’s in proton transport is controversial (17). However, it has been proposed that even if they participate only marginally to Hþ permeability, their effect may be crucial in decreasing reactive oxygen species (ROS) production simply by a slight decrease in a high proton-motive force (18). Another mechanism, which can be called facilitated Hþ transport (called ‘‘active proton leak’’ in ref. 19) has been observed in yeast, where a decrease in the proton-motive force occurs because of the activity of external NADH or G3P dehydrogenases. It should be stressed here that in the yeast

175

IUBMB LIFE

ATP or Hþ/2e) varies oppositely when the forces inducing the slip are increased. This was not observed in the case of Almitrine effect on ATP synthase for which an increase in Hþ/ATP is observed both in the ATP synthesis and in the ATP hydrolysis. It is concluded in this case that almitrine induces an actual change in the mechanistic stoichiometry of the ATPase/ATPsynthase activity (23). The respiratory chains of bacteria and plants possess alternative oxidases in which the electron transfer is not linked to a Hþ extrusion and which allows an uncoupling of ATP synthesis from reoxidation mechanisms that could decrease the ROS production (24).

Control of Oxphos FIG 3

Threshold curves. Decrease in the flux in a network at steady-state dF as a function of the inhibition of a step dv. The control coefficient is defined as CFv ¼ dF/dv. A threshold is usually associated to low control coefficients (0.2), the decrease is more gradual. A control coefficient of one means that any change in the activity of a given reaction is repercuted on the flux (to the same extend). It mainly occurs when the quantity of enzyme is low, that is, the activity is low compared with the activities of the other enzymes which appear as in excess.

Saccharomyces cerevisiae, there is no complex I and the NADH and other dehydrogenases are not coupled to proton extrusion. This observed increase in proton permeability associated with a high activity of these dehydrogenases is independent of the rest of the respiratory chain and the ATP synthase proton pump. This mechanism could permit a decrease in redox pressure with little effect on ATP synthesis. In the examples described above, the uncoupling of ATP synthesis (decrease in VATP) from the respiration (increase in VO2) is mediated by a decrease in proton-motive force (decrease in Dl ~Hþ ). There are experimental data showing that VATP can be decreased and VO2 increased without changing the Dl ~Hþ (20). This led Azzone’s group (21) to propose another mechanism causing a loss of OxPhos yield called slip or intrinsic uncoupling. It is a decrease in the efficiency of a proton pump because of a partial and variable decoupling of the chemical reaction from the proton transport, that is, a decrease in the Hþ/2e stoichiometry of a respiratory chain complex or an increase in the Hþ/ATP stoichiometry of the ATP synthase. A kinetic model for proton pump functioning has been proposed by Pietrobon and Caplan (22). This model is only possible with Hþ pumps (complexes I and IV) and not with Mitchell loops (complex III, which has a fixed stoichiometry). In the case of a reversible pump, the stoichiometry (Hþ/

176

In the previous section, we analyzed the structure of the OxPhos that both reoxidizes NADH and FADH2 and generates ATP. The problem is now: what triggers the changes in the rate of OxPhos? What is the target when ATP or a readjustment of NADH/NAD ratio is necessary? The control of OxPhos was the object of a great deal of discussion with the idea in mind that a unique ‘‘rate-limiting’’ step should exist. However, opinions differed regarding the limiting step: cytochrome oxidase, ATP synthase, and ATP/ADP carrier, which were presented as good candidates (25), most of the time because their inhibition decreased mitochondrial respiration. This long standing riddle was solved by the Tager and coworkers’ group in Amsterdam (25) and the Kunz’s group in Magdeburg (26) showing that the control was not restricted to a unique step as it was supposed but was shared between many steps according to the MCT (27–30) with control coefficients all 0.2 – 0.3; CVF ¼ 0.4 in Fig. 3) or no effect until a high deficiency associated with a low control coefficient (1 was taken as an evidence of a possible association of respiratory complexes in super complexes (36) and (37) in certain conditions, which was experimentally demonstrated later (38).

Models of Oxphos There is a long tradition of modeling OxPhos to integrate all aspects, kinetic and thermodynamic, of chemiosmotic theory (2). The first approach was in the framework of nonequilibrium thermodynamic model involving a linear approximation of the coupled fluxes on the thermodynamic forces, DGox of the redox reactions, DGP, the phosphate potential, and Dp, the proton-motive force (39–42). In this framework, all the states of OxPhos (state 4, state 3, uncoupled state, etc.) were described as phenomenological relations in terms of energy conversion with the use of phenomenological coefficients (Lij):

JOX ¼ L11 DGOX þ L12 DGP JP ¼ L21 DGOX þ L22 DGP with

0q¼

L12 L12 pffiffiffiffiffiffiffiffiffiffi L11 L22

¼

L21,

Onsager

 1, the degree of coupling, and Z ¼

the phenomenological stoichiometry.

Mazat et al.

relationships,

and

qffiffiffiffiffi

L22 L11 ,

The optimal efficiency of the system and the degree of coupling in these conditions were defined and analyzed. Even though one can argue that OxPhos could be out of the linear domain around equilibrium, this description is simple and indicates the fundamental parameters involved: degree of coupling, thermodynamic forces, rates, optimal efficiency, phenomenological stoichiometry, and so on. Furthermore, as evidenced in ref. 43, the linear domain might be extended away from equilibrium because of kinetics regulations, Similar models were derived by Pietrobon and coworkers (22,44) to describe redox-driven proton pumps and ATP synthesis in mitochondria. These models are kinetic models but with the calculation of thermodynamic parameters, evidencing the relationships between kinetics and thermodynamics. Bohnensack was probably the first to derive a quantitative model involving nearly all the components of OxPhos. To do so, he used approximate rate laws of near equilibrium reaction detailed in ref. 45 (first appendix in ref. 45). With the help of this model, the Magdeburg group (26,46) was able to demonstrate that the control of OxPhos was shared by several steps as predicted by MCT (see above). Korzeniewski and Froncisz (47) applied, to complexes I and III of the respiratory chain, the principles of linear dependency on the thermodynamic force, that is, DGox – n0 Dl ~Hþ , where n0 is the number of protons extruded by the complex. Different versions of the model, which include other type of rate equations, were applied to isolated mitochondria or to intact tissues (muscle, heart, and liver). The model was used to calculate the control coefficient in OxPhos (48), to fit threshold curves in muscle and to predict the shape of threshold curves at low oxygen pressure (49), to study the transition from rest to intensive work in muscle (50), leading to the concept of parallel activation. We used this model (51) to compare the threshold curves obtained with mitochondrial or nuclear DNA mutations. One of the first models taking into account the organization and compartmentalization of oxidative-phosphorylation inside the cell was that of Aliev and Saks (52) describing heart bioenergetics and creatine/creatine phosphate shuttle. This model was refined by Vendelin et al. (53,54). More recently, Beard (55) proposed ‘‘A biophysical model of the mitochondrial respiratory system and oxidative phosphorylation’’ mainly applied to cardiac mitochondria. The rate equations of the complex are based on mass-action law with the introduction of the Dl ~Hþ . More recently, the model was extended to a larger model encompassing the mitochondrial energy metabolism (56). Several models are now available, which try to represent cell energy metabolism. The model of Cortassa et al. (57) with one equation for the whole respiratory chain and the one of Holzhu¨tter and coworkers (58) with a very detailed modeling of the respiratory complexes must be cited among others. In the area of Systems Biology and high throughput methods in biology, the Oxphos models increase in size, now incorporating tricarboxylic acid (TCA) cycle, b-oxidation of fatty

177

IUBMB LIFE

acids, and several other metabolic pathways. Because in these large models not all the kinetic parameters are known (and are different in different tissues and organisms), a detailed approach based on ordinary differential equations is difficult (59). To circumvent this problem, a new theoretical approach was developed by the Pallson’s laboratory called Flux Balance Analysis and applied to mitochondrial metabolism (60–62). It uses linear programming, the stoichiometric coefficients for each reaction and biological constraints to optimize a function (objective function) supposed to represent the biological phenotype (maximizing ATP production, biomass, growth rate, etc.). This interesting method does not require the knowledge of the kinetic parameters but relies on optimized functions, which can reflect a simplified picture of the biological reality.

Conclusion In nonphotosynthetic organisms, mitochondria are the power plant of the cell, emphasizing their great potentiality for ATP synthesis from the redox span between nutriments and oxygen. More important is probably their role in the maintenance of cell redox balance as it is well demonstrated in mitochondrial diseases where this function is also affected. These two functions are linked at several levels: in the respiratory complexes and in the global functioning of OxPhos through the a˜ Dl ~Hþ (or the Dp). However, the link/links between these two important functions has/have to be modulated because of the differences in cell demand in energy and redox state maintenance. If the coupling between respiratory chain and ATP synthesis is well understood thanks to Mitchell’s theory (2), the molecular intimate mechanism through which the redox energy is used for Hþ transport by the respiratory complexes and ATP synthase is still a matter of debates and investigation. The fact that the crystallographic structures are now known does not solve these problems but shifts it to a more molecular and precise level and raises new issues. This is particularly the case with the coupling in complex I with the quinone site out of the membrane. It is also the case when one compares the stoichiometry of ATPase with the number of c subunits in the Fo. These old fundamental questions are not only a matter of basic research. Mitochondria appear yet as a great player in cell life with newly evidenced role such as in apoptosis, autophagy, calcium cell signaling, ROS production, aging, cancer, and so on. All these functions are dependent on a correct metabolic function that has to be comprehensively understood. In this matter, use of models is of great help to take into account the highly structured integration of these systems.

Acknowledgements The authors acknowledge Dr Roger Springett for constructive comments and editing the manuscript and Dr Alain Dautant for help in Fig. 1.

178

REFERENCES [1] Habersetzer, J., Ziani, W., Larrieu, I., Stines-Chaumeil, C., Giraud, M.-F., et al. (2013) ATP synthase oligomerization: from the enzyme models to the mitochondrial morphology. Int. J. Biochem. Cell Biol. 45, 99–105 [2] Mitchell, P. (1961) Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 191, 144–148. [3] Mu¨ller, M., Mentel, M., Van Hellemond, J. J., Henze, K., Woehle, C., et al. (2012) Biochemistry and evolution of anaerobic energy metabolism in eukaryotes. Microbiol. Mol. Biol. Rev. 76, 444–495. [4] Nicholls, D. G., and Ferguson, S. J. (2002) Bioenergetics, 3rd edn, Academic Press, London, San-Diego. [5] Mitchell, P. (1975) Protonmotive redox mechanism of the cytochrome b-c1 complex in the respiratory chain: protonmotive ubiquinone cycle. FEBS Lett. 56, 1–6. [6] Crofts, A. R. (2004) The cytochrome bc1 complex: function in the context of structure. Annu. Rev. Physiol. 66, 689–733. [7] Hunte, C., Zickermann, V., and Brandt, U. (2010) Functional modules and structural basis of conformational coupling in mitochondrial complex I. Science 329, 448–451. [8] Efremov, R. G., Baradaran, R., and Sazanov, L. A. (2010) The architecture of respiratory complex I. Nature 465, 441–445. € m, M., and Hummer, G. (2012) Stoichiometry of proton translocation [9] Wikstro by respiratory complex I and its mechanistic implications. Proc. Natl. Acad. Sci. USA 109, 4431–4436. [10] Benard, G., Faustin, B., Passerieux, E., Galinier, A., Rocher, C., et al. (2006) Physiological diversity of mitochondrial oxidative phosphorylation. Am. J. Physiol. Cell Physiol. 291, C1172–1182. [11] Stubbs, M., Veech, R. L., and Krebs, H. A. (1972) Control of the redox state of the nicotinamide–adenine dinucleotide couple in rat liver cytoplasm. Biochem. J. 126, 59–65. [12] Kim, N., Ripple, M. O., and Springett, R. (2012) Measurement of the mitochondrial membrane potential and pH gradient from the redox poise of the hemes of the bc1 complex. Biophys. J. 102, 1194–1203. [13] Efremov, R. G., and Sazanov, L. A. (2012) The coupling mechanism of respiratory complex I – a structural and evolutionary perspective. Biochim. Biophys. Acta 1817, 1785–1795. [14] Nicholls, D. G. (1977) The effective proton conductance of the inner membrane of mitochondria from brown adipose tissue. Dependency on proton electrochemical potential gradient. Eur. J. Biochem. 77, 349–356. [15] Nobes, C. D., Brown, G. C., Olive, P. N., and Brand, M. D. (1990) Non-ohmic proton conductance of the mitochondrial inner membrane in hepatocytes. J. Biol. Chem. 265, 12903–12909. [16] Nicholls, D. G., and Locke, R. M. (1984) Thermogenic mechanisms in brown fat. Physiol. Rev. 64, 1–64. [17] Nicholls, D. G. (2006) The physiological regulation of uncoupling proteins. Biochim. Biophys. Acta 1757, 459–466. nicaud, L. (2001) Mitochondrial ROS me[18] Casteilla, L., Rigoulet, M., and Pe tabolism: modulation by uncoupling proteins. IUBMB Life 52, 181–188. [19] Mourier, A., Devin, A., and Rigoulet, M. (2010) Active proton leak in mitochondria: a new way to regulate substrate oxidation. Biochim. Biophys. Acta 1797, 255–261. [20] Rottenberg, H. (1983) Uncoupling of oxidative phosphorylation in rat liver mitochondria by general anesthetics. Proc. Natl. Acad. Sci. USA 80, 3313–3317. [21] Luvisetto, S., Conti, E., Buso, M., and Azzone, G. F. (1991) Flux ratios and pump stoichiometries at sites II and III in liver mitochondria. Effect of slips and leaks. J. Biol. Chem. 266, 1034–1042. [22] Pietrobon, D., and Caplan, S. R. (1985) Flow-force relationships for a six-state proton pump model: intrinsic uncoupling, kinetic equivalence of input and output forces, and domain of approximate linearity. Biochemistry 24, 5764–5776. rin, B., Fontaine, E., et al. (1990) [23] Rigoulet, M., Fraisse, L., Ouhabi, R., Gue Flux-dependent increase in the stoichiometry of charge translocation by mitochondrial ATPase/ATP synthase induced by almitrine. Biochim. Biophys. Acta 1018, 91–97. [24] Maxwell, D. P., Wang, Y., and McIntosh, L. (1999) The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proc. Natl. Acad. Sci. USA 96, 8271–8276.

Mitochondrial Energetic Metabolism—Some General Principles

[25] Groen, A. K., Wanders, R. J., Westerhoff, H. V., Van der Meer, R., and Tager, J. M. (1982) Quantification of the contribution of various steps to the control of mitochondrial respiration. J. Biol. Chem. 257, 2754–2757. [26] Bohnensack, R., Ku¨ster, U., and Letko, G. (1982) Rate-controlling steps of oxidative phosphorylation in rat liver mitochondria. A synoptic approach of model and experiment. Biochim. Biophys. Acta 680, 271–280. [27] Kacser, H., and Burns, J. A. (1973) The control of flux. Symp. Soc. Exp. Biol. 27, 65–104. [28] Heinrich, R., and Rapoport, T. A. (1974) A linear steady-state treatment of enzymatic chains. General properties, control and effector strength. Eur. J. Biochem. 42, 89–95. [29] Reder, C. (1988) Metabolic control theory: a structural approach. J. Theor. Biol. 135, 175–201. [30] Fell, D. A. (1992) Metabolic control analysis: a survey of its theoretical and experimental development. Biochem. J. 286 (Pt 2),313–330. rin, B. (1986) Control of oxi[31] Mazat, J.-P., Jean-Bart, E., Rigoulet, M., and Gue dative phosphorylations in yeast mitochondria. Role of the phosphate carrier. Biochim. Biophys. Acta Bioenerg. 849, 7–15. [32] Letellier, T., Malgat, M., and Mazat, J. P. (1993) Control of oxidative phosphorylation in rat muscle mitochondria: implications for mitochondrial myopathies. Biochim. Biophys. Acta 1141, 58–64. [33] Rossignol, R., Letellier, T., Malgat, M., Rocher, C., and Mazat, J. P. (2000) Tissue variation in the control of oxidative phosphorylation: implication for mitochondrial diseases. Biochem. J. 347Pt 1,45–53. [34] Rossignol, R., Malgat, M., Mazat, J. P., and Letellier, T. (1999) Threshold effect and tissue specificity. Implication for mitochondrial cytopathies. J. Biol. Chem. 274, 33426–33432. [35] Kholodenko, B. N., and Westerhoff, H. V. (1993) Metabolic channelling and control of the flux. FEBS Lett. 320, 71–74. [36] Boumans, H., Grivell, L. A., and Berden, J. A. (1998) The respiratory chain in yeast behaves as a single functional unit. J. Biol. Chem. 273, 4872–4877. [37] Bianchi, C., Genova, M. L., Parenti Castelli, G., and Lenaz, G. (2004) The mitochondrial respiratory chain is partially organized in a supercomplex assembly: kinetic evidence using flux control analysis. J. Biol. Chem. 279, 36562–36569. €gger, H. (2001) Respiratory chain supercomplexes. IUBMB Life 52, [38] Scha 119–128. [39] Rottenberg, H. (1979) Non-equilibrium thermodynamics of energy conversion in bioenergetics. Biochim. Biophys. Acta 549, 225–253. [40] Westerhoff, H. V., and Van Dam, K. (1987) Thermodynamics and Control of Free-Energy Transduction,Elsevier,Amsterdam. [41] Stucki, J. W. (1980) The thermodynamic-buffer enzymes. Eur. J. Biochem. 109, 257–267. [42] Stucki, J. W. (1980) The optimal efficiency and the economic degrees of coupling of oxidative phosphorylation. Eur. J. Biochem. 109, 269–283. [43] Rigoulet, M., Guerin, B., and Denis, M. (1987) Modification of flow-force relationships by external ATP in yeast mitochondria. Eur. J. Biochem. 168, 275–279. [44] Pietrobon, D., Zoratti, M., Azzone, G. F., Stucki, J. W., and Walz, D. (1982) Non-equilibrium thermodynamic assessment of redox-driven Hþ pumps in mitochondria. Eur. J. Biochem. 127, 483–494. [45] Bohnensack, R. (1981) Control of energy transformation of mitochondria. Analysis by a quantitative model. Biochim. Biophys. Acta 634, 203–218.

Mazat et al.

[46] Gellerich, F. N., Bohnensack, R., and Kunz, W. (1983) Control of mitochondrial respiration. The contribution of the adenine nucleotide translocator depends on the ATP- and ADP-consuming enzymes. Biochim. Biophys. Acta 722, 381–391. [47] Korzeniewski, B., and Froncisz, W. (1991) An extended dynamic model of oxidative phosphorylation. Biochim. Biophys. Acta 1060, 210–223. [48] Korzeniewski, B., and Froncisz, W. (1992) Theoretical studies on the control of the oxidative phosphorylation system. Biochim. Biophys. Acta 1102, 67–75. [49] Korzeniewski, B., and Mazat, J. P. (1996) Theoretical studies on control of oxidative phosphorylation in muscle mitochondria at different energy demands and oxygen concentrations. Acta Biotheor. 44, 263–269. [50] Korzeniewski, B. (2000) Regulation of ATP supply in mammalian skeletal muscle during resting state: intensive work transition. Biophys. Chem. 83, 19–34. [51] Korzeniewski, B., Malgat, M., Letellier, T., and Mazat, J. P. (2001) Effect of ‘‘binary mitochondrial heteroplasmy’’ on respiration and ATP synthesis: implications for mitochondrial diseases. Biochem. J. 357, 835–842. [52] Aliev, M. K., and Saks, V. A. (1997) Compartmentalized energy transfer in cardiomyocytes: use of mathematical modeling for analysis of in vivo regulation of respiration. Biophys. J. 73, 428–445. [53] Vendelin, M., Kongas, O., and Saks, V. (2000) Regulation of mitochondrial respiration in heart cells analyzed by reaction-diffusion model of energy transfer. Am. J. Physiol., Cell Physiol. 278, C747–764. [54] Vendelin, M., Lemba, M., and Saks, V. A. (2004) Analysis of functional coupling: mitochondrial creatine kinase and adenine nucleotide translocase. Biophys. J. 87, 696–713. [55] Beard, D. A. (2005) A biophysical model of the mitochondrial respiratory system and oxidative phosphorylation. PLoS Comput. Biol. 1, e36. [56] Wu, F., Yang, F., Vinnakota, K. C., and Beard, D. A. (2007) Computer modeling of mitochondrial tricarboxylic acid cycle, oxidative phosphorylation, metabolite transport, and electrophysiology. J. Biol. Chem. 282, 24525–24537. n, E., Winslow, R. L., and O’Rourke, B. [57] Cortassa, S., Aon, M. A., Marba (2003) An integrated model of cardiac mitochondrial energy metabolism and calcium dynamics. Biophys. J. 84, 2734–2755. [58] Berndt, N., Bulik, S., and Holzhu¨tter, H.-G. (2012) Kinetic modeling of the mitochondrial energy metabolism of neuronal cells: the impact of reduced a-ketoglutarate dehydrogenase activities on ATP production and generation of reactive oxygen species. Int. J. Cell Biol.2012,757594. [59] Yugi, K., and Tomita, M. (2004) A general computational model of mitochondrial metabolism in a whole organelle scale. Bioinformatics 20, 1795–1796. [60] Ramakrishna, R., Edwards, J. S., McCulloch, A., and Palsson, B. O. (2001) Flux-balance analysis of mitochondrial energy metabolism: consequences of systemic stoichiometric constraints. Am. J. Physiol. Regul. Integr. Comp. Physiol. 280, R695–704. [61] Vo, T. D., Greenberg, H. J., and Palsson, B. O. (2004) Reconstruction and functional characterization of the human mitochondrial metabolic network based on proteomic and biochemical data. J. Biol. Chem. 279, 39532–39540. [62] Thiele, I., Price, N. D., Vo, T. D., and Palsson, B. A¨. (2005) Candidate metabolic network states in human mitochondria. Impact of diabetes, ischemia, and diet. J. Biol. Chem. 280, 11683–11695.

179

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.