<<

Genetics: Published Articles Ahead of Print, published on June 14, 2005 as 10.1534/genetics.105.041517

1

Neurospora strains harboring mitochondrial disease-associated

mutations in iron-sulfur subunits of complex I

Margarida Duarte *, Ulrich Schulte †, Alexandra V. Ushakova * and Arnaldo Videira *,‡

a Instituto de Biologia Molecular e Celular (IBMC) and ‡ Instituto de Ciências

Biomédicas de Abel Salazar (ICBAS), Universidade do Porto, 4150-180 Porto, Portugal

and † Institute of Biochemistry, Heinrich-Heine-Universität, 40225 Düsseldorf,

Germany

2

running head:

Disease-associated mutations in complex I

keywords:

mitochondria; complex I; iron-sulfur; mutations; Neurospora

correspondence address:

Arnaldo Videira

Instituto de Biologia Molecular e Celular

Rua do Campo Alegre 823, 4150-180 Porto, Portugal phone: +351-226074900 fax: +351-226099157 email: [email protected]

3

ABSTRACT

We subjected the genes encoding the 19.3 kDa, 21.3c kDa and 51 kDa iron- sulfur subunits of respiratory chain complex I from Neurospora crassa to site-directed mutagenesis, in order to mimic mutations in human complex I subunits associated with mitochondrial diseases. The V135M substitution was introduced in the 19.3 kDa cDNA, the P88L and R111H substitutions were separately introduced in the 21.3c kDa cDNA and the A353V and T435M alterations were separately introduced in the 51 kDa cDNA.

The altered cDNAs were expressed in the corresponding null-mutants under the control of a heterologous promoter. With exception of the A353V polypeptide, all mutated subunits are able to promote assembly of a functional complex I rescuing the phenotypes of the respective null-mutants. Complex I from these strains display spectroscopic and enzymatic properties similar to those observed in the wild type strain.

A decrease in total complex I amounts may be the major impact of the mutations, although expression levels of mutant genes from the heterologous promoter were sometimes lower and may also account for complex I levels. We discuss these findings in relation to the involvement of complex I deficiencies in mitochondrial disease.

4

INTRODUCTION

The proton-pumping NADH:ubiquinone of the mitochondrial respiratory chain, or complex I (EC.1.6.5.3), is located in the inner membrane of the eukaryotic organelle. It contains up to 46 polypeptide subunits of both nuclear and mitochondrial origin, as well as FMN and several iron-sulfur clusters (CARROLL et al.

2003; WALKER 1992). The is composed of two distinct domains undergoing

independent assembly, the peripheral and membrane arms, which are arranged

perpendicularly to each other in an L-shaped structure (HOFHAUS et al. 1991; TUSCHEN

et al. 1990). Complex I is also found in many prokaryotes (NDH1) with fewer protein constituents but similar constitution of prosthetic groups (YAGI et al. 1998). The prokaryotic contain 14 proteins considered as the minimal assembly required for coupling electron transfer with proton translocation. All subunits are homologous to subunits of the eukaryotic complex I: 7 are homologous to nuclear-encoded and the other 7 are homologous to mitochondrial-encoded polypeptides.

Complex I is important for several cellular processes in different organisms

(VIDEIRA 1998). Deficiencies in complex I activity and/or its mitochondrial or

cytoplasmic-synthesized subunits have been implicated in the development of several

pathogenic human conditions. In fact, disturbances of mitochondrial energy metabolism

occurring with an estimated incidence of 1 in 10,000 live births are often caused by

isolated complex I deficiency (SMEITINK et al. 2001; WALLACE 1992). The relationship

between gene alterations and clinical disease symptoms or, e.g., tissue-specific phenotypes remains rather obscure and there is no effective therapy for these diseases.

As direct studies of human material are subject to strong restrictions and, in particular, many complex I patients die young, the development of non-human models of the 5

diseases is very desirable. Research on mitochondrial complex I deficiency has been

conducted in cell cultures (HOFHAUS and ATTARDI 1993; HOFHAUS et al. 1996), the yeast Yarrowia lipolytica (AHLERS et al. 2000) and Caenorhabditis elegans (GRAD and

LEMIRE 2004). Models of mutations in mitochondrial genes have been generated

especially in prokaryotes (LUNARDI et al. 1998; ZICKERMANN et al. 1998) due to the specific difficulties associated with gene replacement in mitochondria.

A very useful model for investigating the effects of mutations in complex I is

Neurospora crassa. Complex I from N. crassa has been extensively characterized (it is highly similar to that of mammals (VIDEIRA 1998)) and the genetics and physiology of

the fungus are well known (DAVIS and PERKINS 2002). In this article, we describe the

development of Neurospora strains harboring mutations in single subunits relevant to

human diseases.

During the last years, several specific point mutations in nuclear genes have

been discovered, specifically in those coding for the human homologues of the 75 kDa

(BENIT et al. 2001), 51 kDa (SCHUELKE et al. 1999), 49 kDa (LOEFFEN et al. 2001), 30 kDa (BENIT et al. 2004), 24 kDa (BENIT et al. 2003), TYKY (LOEFFEN et al. 1998),

PSST (TRIEPELS et al. 1999), AQDQ (VAN DEN HEUVEL et al. 1998) and IP13 (KIRBY et al. 2004) subunits of complex I. Mimicked in our study are two compound heterozygous mutations in TYKY (P79L and R102H) and a homozygous mutation in

PSST (V122M) in patients suffering from Leigh syndrome (LOEFFEN et al. 1998;

TRIEPELS et al. 1999), as well as mutations in the 51 kDa polypeptide involved in the

development of leukodystrophy and myoclonic epilepsy (associated with a homozygous

A341V mutation and a compound heterozygous T423M and R59X mutation)

(SCHUELKE et al. 1999). These three proteins have homologues in the fungus N. crassa, namely the 21.3c kDa (DUARTE et al. 1996), 19.3 kDa (SOUSA et al. 1999) and 51 kDa 6

(PREIS et al. 1991) subunits, respectively, and are highly conserved from bacteria to mammals. They belong to the peripheral domain of complex I in N. crassa (VIDEIRA

1998) or to IλS subcomplex in bovine (FINEL et al. 1994). The amino acid sequences of

the 21.3c kDa and 19.3 kDa proteins display consensus motives for binding of the two

[4Fe-4S] clusters N6a and N6b (RASMUSSEN et al. 2001) and of the [4Fe-4S] cluster N2

(DUARTE et al. 2002), respectively. The 51 kDa subunit forms the NADH- and FMN-

binding sites of complex I and harbors iron sulfur cluster N3 (FECKE et al. 1994). All these polypeptide subunits have an essential function in electron transfer.

MATERIALS AND METHODS

N. crassa manipulations: We used the N. crassa strain 74-OR23-1A (wild

type), the mutant strains nuo19.3 and nuo21.3c obtained by repeat-induced point-

mutations of the nuo-19.3 (DUARTE et al. 2002) and nuo-21.3c genes (DUARTE and

VIDEIRA 2000), respectively, and mutant nuo51 obtained after disruption of the nuo-51 gene by homologous recombination (FECKE et al. 1994). These mutant strains lack the

19.3 kDa, 21.3c kDa and 51 kDa subunits of complex I, respectively. The former two mutants still display transcription from the mutated genes (a smaller transcript in the case of nuo21.3c, not shown). Anyway, they can be considered as null-mutants because the relevant proteins are not detected. In addition, there are no detectable complex I assembly (by different criteria) and no detectable complex I activity (different activities measured). The failure to undergo homozygous genetic crosses of the nuo19.3 (H.

Pópulo, unpublished results) and nuo21.3c mutants (DUARTE and VIDEIRA 2000) also

support this conclusion. Growth and handling of N. crassa were carried out according to

standard procedures (DAVIS and DE SERRES 1970). Conidia from 7-day-old cultures 7

were used to prepare spheroplasts (DUARTE et al. 1995), which were then transformed

with recombinant pMYX2 vectors and selected on plates containing benomyl. For

expression of cDNAs under the control of the quinic acid promoter of pMYX2

(CAMPBELL et al. 1994), 10 mM quinic acid was added to the medium.

Site-directed mutagenesis: The cDNAs coding for the 21.3c kDa and the 19.3

kDa subunits cloned in pGEM4 (DUARTE et al. 1996; SOUSA et al. 1999), were cleaved with EcoRI, treated with Klenow to create blunt ends and then cloned in the SmaI site of the expression vector pMYX2, downstream of the qa-2 inducible promoter. A N. crassa cDNA mycelial library M-1 cloned in UniZAP XR (obtained from the Fungal Genetics

Stock Center) was screened by hybridization with an incomplete cDNA coding for the

51 kDa subunit of complex I, in order to obtain the complete reading frame. The pBluescript plasmid was excised from a positive phage and a 1.8 kb SmaI/PvuII fragment containing the entire cDNA was also cloned in the SmaI site of pMYX2. Site- directed mutagenesis was achieved using the Quik ChangeTM Site-Directed Mutagenesis kit (Stratagene). Briefly, the pMYX2-recombinant vectors and pairs of synthetic complementary oligonucleotide primers containing the desired mutation were used in

PCR reactions to create mutated plasmids. The mutagenic oligonucleotide primers used

were: P88L, 5’-TTCAGGCCGCTCTACACAATCTATTACC-3’; R111H, 5’-

CACGCCCTTCACCGTTACCCGTCG-3’; V135M, 5’-

GTCATGATTATGGCGGGCACC-3’; A353V, 5’-

GACTTCGATGTCCTCAAGGACAGC-3’; T435M, 5’-

GAAGGTCACATGATTTGCGCTCTC-3’, and their complementary strands. The

underlined nucleotides represent substitutions that change codons within the cDNAs.

They result in the replacement of proline 88 by leucine (P88L) and of arginine 111 by

histidine (R111H) in the 21.3c kDa protein, of valine 135 by methionine (V135M) in 8

the 19.3 kDa subunit and of alanine 353 by valine (A353V) and of threonine 435 by

methionine (T435M) in the 51 kDa protein. The mutagenesis was confirmed by

complete sequencing of the PCR products. The mutated plasmids were transformed into

the respective null-mutant strains. Transformants expressing higher amounts of each

mutated protein were selected for further analysis.

Northern blot analysis: Total RNA was purified from mycelial tissue

(SOKOLOVSKY et al. 1990) and treated with RNase free-DNase I (Boeringher) for 20

min at 37 ºC in 20 mM Tris/HCl, 5 mM MgSO4, pH 7.6. Then, it was extracted once with phenol/chloroform, precipitated with ethanol and stored at -20 ºC. The RNA was resolved by electrophoresis in formaldehyde/formamide denaturing gels before blotting

(SAMBROOK and RUSSELL 2001). Hybridization was conducted with a 51 kDa cDNA probe labeled with Gene Images Labeling kit (Amersham Biosciences).

Characterization of isolated mutant complexes: Mitochondria were isolated from hyphae grown in the presence of 10 mM quinic acid. Ten milliliters of water- suspended organelles (0.6 g protein) were solubilized with 5 ml of 20% Triton X-100.

After centrifugation for 20 min at 20000 x g, the supernatant was applied to 3 sucrose gradients of 30 ml each (10-25% sucrose in 50 mM Tris/HCl pH 7.5, 50 mM NaCl,

0.1% Triton). Mitochondrial proteins were separated at 100000 x g for 22 h. The gradients were fractionated and analyzed for NADH:ferricyanide reductase activity.

Fractions corresponding to the lower quarter of the gradients showing high activity were pooled and applied to an 8 ml source Q anion exchange column (Pharmacia). Complex I was eluted by means of a 50 ml salt gradient: 50-400 mM NaCl in 50 mM Tris/HCl pH

7.5, 0.05% Triton. Fractions with high NADH:ferricyanide reductase activity were pooled. The NADH:decylquinone reductase activity was determined after reconstitution

of the enzyme in phospholipids. To obtain EPR spectra, the enzyme preparation was 9

concentrated five-fold by ultra filtration on a Diaflo XM 300 filter and treated as

described (WANG et al. 1991). Spectra were taken with a Bruker EMX 6/1 EPR spectrometer.

Other protein and enzymatic analysis: The oxidation of dNADH was

-1 -1 measured photometrically (ε340=6.22 mM cm ) in a standard assay mixture (50 mM

Tris/HCl, 0.25 M sucrose, 0.2 mM EDTA, pH 8.0). When dNADH:Q1 reductase was

assayed, 2 mM KCN, 120 µM dNADH and 80 µM Q1 were added to the reaction

mixture. NADH:hexaammineruthenium III (HAR) reductase activity was measured

-1 -1 photometrically by the oxidation of NADH (ε380=1.25 mM cm ) in the presence of 2

mM KCN, 120 µM NADH and 2 mM HAR. NADH:ferricyanide reductase activity was

-1 -1 measured photometrically (ε420=1.05 mM cm ) in the standard assay mixture

containing 2 mM KCN, 120 µM NADH and 0.5 mM ferricyanide.

The techniques used for the preparation of N. crassa mitochondria (WERNER

1977) and mitochondrial membranes (MELO et al. 2001), the development of rabbit

antisera directed to subunits of complex I (VIDEIRA and WERNER 1989), protein determination (BRADFORD 1976), SDS-polyacrylamide gel electrophoresis (ZAUNER et al. 1985), blotting and incubation of blots with antisera, detection of alkaline phosphatase second antibodies and sucrose gradient centrifugation analysis of

detergent-solubilized mitochondrial proteins (ALVES and VIDEIRA 1994) have been

described.

RESULTS

Mutants of the 21.3c, 19.3 and 51 kDa proteins in N. crassa: The null-

mutants nuo21.3c and nuo19.3, which respectively lack the mammalian TYKY- 10

homologous 21.3c kDa subunit and the PSST-homologous 19.3 kDa subunit of fungal

complex I were generated by repeat-induced point mutations in the correspondent genes

(DUARTE et al. 2002; DUARTE and VIDEIRA 2000). Inactivation of the gene encoding the

51 kDa subunit of complex I was obtained by homologous recombination (FECKE et al.

1994). The cDNAs encoding the 21.3c kDa, the 19.3 kDa and the 51 kDa polypeptides

were cloned in the pMYX2 vector, allowing its transformation and controlled

expression in the fungus. Missense mutations were separately introduced in the cDNAs

by site-directed mutagenesis. The mutations P88L or R111H, equivalent to those found

in a patient with Leigh syndrome, were thus applied to the 21.3c kDa subunit. Likewise,

the V135M mutation was generated in the 19.3 kDa cDNA. Finally, mutations A353V

and T435M relevant to the development of leukodystrophy and myoclonic epilepsy

were applied to the 51 kDa subunit. All pMYX2-recombinant plasmids were separately

transformed into the corresponding null-mutants.

Assembly of complex I in strains harboring site-specific mutations: In the

null-mutants nuo21.3c and nuo19.3, assembly of the peripheral arm of complex I is

prevented and only the membrane arm is formed (DUARTE et al. 2002; DUARTE and

VIDEIRA 2000). Lack of the 51 kDa subunit led to the formation of an inactive enzyme

comprising all but the missing subunit (FECKE et al. 1994). In order to find out whether complex I assembly occurs in strains harboring the point mutations separately introduced into the three subunits, we performed sucrose gradient centrifugation analysis of their mitochondrial proteins. Mitochondria from the wild type strain, null- mutants rescued by wild type cDNAs and the different mutants were solubilized with

Triton X-100 and centrifuged in linear sucrose gradients. The NADH:ferricyanide reductase activity as well as the distribution of several complex I subunits throughout the gradients were followed. 11

Figure 1 depicts an experiment with the site-directed 21.3c mutants, where we

immunostained the 12.3 kDa (VIDEIRA et al. 1993) and 20.8 kDa polypeptides (VIDEIRA

et al. 1990a) as markers for the membrane arm, and the 29.9 kDa (VAN DER PAS et al.

1991) and 30.4 kDa polypeptides (VIDEIRA et al. 1990b) as markers for the peripheral

arm of complex I. In the wild type strain, the reductase activity is found in fractions 9-

11 of the gradient, with a peak in fraction 10 (Figure 1A), in agreement with the

sedimentation profile of the complex I proteins (Figure 1B). This represents the typical

behavior of complex I, migrating about two thirds of the gradients. We included as

negative control in this experiment the null-mutant nuo21.3c, which is unable to

assemble the peripheral arm of complex I, as can be deduced from the lack of activity in

the gradients (Figure 1A) and the fact that the 29.9 and 30.4 kDa protein markers of the

peripheral arm of complex I are found in the low molecular weight region of the

gradients (fractions 2-4, Figure 1C). Results similar to wild type were obtained with

mitochondrial proteins from mutants P88L (Figure 1A and D) and R111H (Figure 1A

and E). Similar results were obtained with the V135M mutant as well (results not

shown). We conclude that, in contrast to the respective null-mutants, expression of the mutated 21.3c kDa proteins (carrying either P88L or R111H) as well as expression of the mutated 19.3 kDa protein (carrying V135M) support the assembly of an active complex I. The observation of decreased activity (Figure 1A) and intensity of the protein bands (Figure 1B, D, E) in the gradients of the site-directed mutants reflects mostly that lower amounts of complex I are present in these strains in comparison with wild type (see also below).

Assembly of complex I in strains harboring mutations in the 51 kDa subunit is

shown in Figure 2. The NADH:ferricyanide reductase activity as well as the distribution

of the 30.4 kDa (VIDEIRA et al. 1990b), 51 kDa and 78 kDa subunits (PREIS et al. 1991) 12

of the peripheral arm of complex I were followed in this experiment. In the null-mutant

strain rescued by wild type cDNA, the reductase activity and the complex I proteins

elute in fractions 8-9 of the gradient, with a peak in fraction 9 (Figure 2A and B),

representing entire complex I. As a negative control, we included in these experiments

the mutant strain nuo51, which assembles an inactive complex I (Figure 2A and C). The

A353V mutant behaves like the knockout nuo51 mutant. The complex I subunits elute

in fractions 8 and 9, but the 51 kDa polypeptide is not detected and no reductase activity

is observed (Figure 2A and D). The results obtained with mitochondrial proteins from

mutant T435M (Figure 2A and E) are essentially similar to those of the cDNA-rescued

strain. We conclude that, in contrast to the null-mutant nuo51, expression of a 51 kDa

protein carrying the point-mutation T435M supports the assembly of an active complex

I, quite similar to the wild type enzyme. In fact, analysis of the catalytic properties of the T435M complex I indicate that all activities of the mutant enzyme are present in roughly two-fold lower amounts than in wild type. Moreover, there are no significant differences between mutant and wild type enzymes in terms of sensitivity or affinity to the substrate (table 1).

Gene expression from mutant cDNAs: We were surprised by the fact that

assembly of the 51 kDa protein carrying the A353V substitution into complex I was not

detected. Therefore, we tested additional isolates from the transformation with the

A353V cDNA. The mutated 51 kDa polypeptide could not be detected among 24

benomyl resistant transformants analyzed. This was in contrast to the T435M-mutated

protein, which was found in most of the strains transformed with T435M cDNA. We

subjected 10 of the strains transformed with the A353V cDNA to Southern blot analysis

and found the presence of several copies of the altered cDNA in all of them (not

shown). In order to further clarify this issue, we checked the mRNA expression in 13

different strains. Total RNA was prepared from mycelial tissue of the wild type, cDNA-

rescued, nuo51, A353V and T435M strains and analyzed by Northern blotting with the

51 kDa cDNA as probe (Figure 3). With the exception of nuo51 (obtained by homologous recombination), the transcript encoding the 51 kDa polypeptide could be identified in all other strains. The A353V mutant expresses the cDNA to the same extent as the cDNA-rescued strain, where the 51 kDa subunit of complex I is clearly detected. These results indicate that the point-mutated 51 kDa polypeptide is likely synthesized in the A353V strain but is then degraded during its route to mitochondria and/or assembly into complex I.

We also examined the expression of other cDNAs introduced into the N. crassa

strains (not shown). Expression ranged from about half (nuo21.3c rescued with wild

type cDNA and the P88L strain) to similar (R111H strain) or slightly higher (nuo19.3

rescued with wild type cDNA and the V135M strain) than the expression of the wild

type endogenous gene. Thus, considerable gene expression was observed in all cases,

suggesting that it may not be a limiting factor in complex I assembly. In the case of

P88L, the lower expression may account for the lower levels of complex I observed in

the correspondent strain.

Spectroscopic and enzymatic characteristics of complex I from strains

P88L, R111H and V135M: We studied complex I activities with different electron

donors/acceptors in mitochondrial membrane preparations from wild type and the

mutants. The NADH:HAR reductase activity, which can roughly estimate complex I

amounts (SLED and VINOGRADOV 1993), is about half of the wild type activity in P88L

and V135M and just about 25% of the wild type value in R111H (table 1). These

results, in agreement with the data from the sucrose gradients and the dehydrogenase

activities depicted in table 2 further indicate that the amount of complex I in the mutants 14

is lower than in wild type. To determine the more physiological NADH:ubiquinone

reductase activity of the mutants complex I, we specifically used dNADH as electron

donor because it is less efficiently used by the alternative NADH dehydrogenases present in fungal mitochondria (FRIEDRICH et al. 1994). The dNADH:Q1 reductase

activity in membranes from all three mutants is lower than in wild type, again reflecting

the lower complex I content in these strains. The kinetic parameters of dNADH:Q1

reductase activity, such as the KM for the substrates and the rotenone sensitivity, differ very slightly from the wild type values.

In order to further characterize complex I formed in the mutants, we have

isolated the enzyme from mitochondria of the strains by a two-step procedure involving sucrose gradient centrifugation and anion exchange chromatography. Complex I thus isolated was about 85% pure. The yield in the mutants was 2-4 fold lower compared to complex I isolated from wild type mitochondria. The polypeptide subunit composition as visible on SDS-PAGE gels revealed no significant differences to the wild type enzyme (data not shown).

Table 2 depicts NADH dehydrogenase activities of mitochondria and isolated

complex I from the mutants and the wild type strain. In agreement with the results

obtained with the sucrose gradients, the NADH:ferricyanide reductase activities of mitochondria from mutants P88L, R111H and V135M are only about half, one quarter and half, respectively, of the activities found in wild type. This reflects most likely the facts that lower than wild type amounts of complex I are being formed in the mutants.

In fact, when the same activity is determined in isolated enzyme preparations, the values obtained with the mutants and the wild type strain are not significantly different.

Nevertheless, a slightly decreased NADH:ferricyanide reductase activity was noticed in the mutant strains. We did not find significant differences between the mutants and wild 15

type complex I concerning NADH:quinone reductase activities or its rotenone

sensitivity.

EPR spectra of the isolated enzymes from mutants P88L and R111H show

signals for all detectable iron-sulfur clusters, namely N1, N2, N3 and N4 (Figure 4). The

ratios of the signals show slight variations between both enzymes and compared to the

wild type enzyme. Especially the cluster N2 is somewhat diminished in mutant R111H.

Other variations are in a range usually observed between different complex I

preparations and do not indicate significant alterations in the spectroscopic properties of the EPR detectable iron-sulfur clusters of the mutant enzymes. We obtained similar results with the V135M mutant, namely an EPR spectrum not significantly different from the wild type spectrum (data not shown).

DISCUSSION

We have developed a eukaryotic model to study mutations in respiratory chain complex I subunits associated with mitochondrial disease. In particular, we describe N.

crassa strains with mutations in the nuclear-coded TYKY and PSST polypeptides,

which are associated with Leigh syndrome in humans, and mutations in the 51 kDa

polypeptide, equivalent to mutations associated with leukodystrophy and myoclonic

epilepsy. In human patients, homozygous as well as heterozygous mutations are

associated with disease. The V122M substitution in the PSST subunit (TRIEPELS et al.

1999) and also the A341V mutation in the 51 kDa subunit (SCHUELKE et al. 1999) were

found homozygous in the human patients. Two human patients were compound

heterozygous. One carried two compound heterozygous missense mutations in the

TYKY protein, amino acid substitutions P79L and R102H, respectively arising from the 16

two heterozygous and apparently healthy parents (LOEFFEN et al. 1998). The other

patient is a compound heterozygous expressing both a 51 kDa subunit with a T435M

substitution and a truncated form of the polypeptide (SCHUELKE et al. 1999). The fact that N. crassa is a haploid organism, allowed us to test separately the effects of single mutations in complex I assembly and function.

The null-mutants nuo21.3c (lacking the fungal TYKY homologue) and nuo19.3

(lacking the PSST homologue) are unable to assemble the peripheral arm of complex I

(DUARTE et al. 2002; DUARTE and VIDEIRA 2000). The null mutant nuo51 assembles an

inactive complex I lacking only the 51 kDa subunit (FECKE et al. 1994). When the null-

mutants are complemented with mutant versions of the disrupted genes, four out of the

five single mutations analyzed in this study support assembly of the fungal enzyme.

These are the P88L and the R111H mutations in the 21.3c kDa subunit, the V135M

mutation in the 19.3 kDa subunit and the T435M mutation in the 51 kDa subunit. None

of these mutations seem to affect significantly the specific activity of isolated complex

I. Also the sensitivity to inhibitors is not significantly affected by the mutations. EPR

spectra of the 21.3c kDa and 19.3 kDa site-directed mutants show that all detectable

iron-sulfur clusters are present in the mutant enzymes.

Considerable differences between mutants and wild type are apparent, however,

regarding the specific complex I activities of mitochondria. This is paralleled with

decreased complex I proteins in mitochondria of the N. crassa mutants. Clearly, our data indicate that the reduction in complex I activity in mutant mitochondria is mainly due to a decreased accumulation of active complex I in the membrane and not to a diminished activity of the mutated enzyme. The reduced assembly/stability of the mutant enzymes is also apparent in the low yield during their isolation. Mutation 17

R111H in the 21.3c kDa polypeptide has a stronger effect on both the complex I content

in mitochondria and the stability of the isolated enzyme.

A drastic effect in complex I activity was found in the human situation, where a

marked reduction in NADH:quinone reductase activity was detected in several tissues

of the patient affected in TYKY (LOEFFEN et al. 1998). Reconstruction of the human

pathogenic mutations in the TYKY and PSST subunits of complex I in the aerobic yeast

Y. lipolytica led to complex I catalytic defects. The yeast mutants showed a 50%

decrease in the Vmax of the mutated complex I, elevated Km values and/or elevated I50 values for quinone-analogous inhibitors (AHLERS et al. 2000) and are presumably not

affected in complex I assembly. We have obtained different results in N. crassa suggesting that neither the catalytic activity nor the affinity of complex I to substrates is greatly affected in the correspondent mutant strains. We suggest that the diminished formation and stability of complex I most likely is the major factor in the development of disease in these cases. In fact, analysis by BN-PAGE of mitochondrial proteins from patients carrying the TYKY and PSST mutations revealed that decreased levels of fully assembled complex I correlate with the low enzyme activities, suggesting an assembly/stability defect as the primary pathogenic mechanism (UGALDE et al. 2004).

Only one mutation presented in this study prevents assembly of active complex

I. We were unable to detect any assembly of the A353V-mutated 51 kDa polypeptide into complex I, despite that expression of the protein was clearly visible in Northern blots. We assume that the mutant subunit is degraded at some stage before integration in complex I. Recently, C. elegans transgenic strains comprising the two mutations in the

51 kDa subunit were generated as disease models (GRAD and LEMIRE 2004). The worm

mutants demonstrated hallmark features of complex I dysfunction such as lactic acidosis and decreased NADH-dependent mitochondrial respiration, though the authors did not 18

study the assembly of the mutated enzymes. Furthermore, the C. elegans mutants

displayed a decrease in the assembly and/or activity of complex IV and hypersensitivity

to oxidative stress (GRAD and LEMIRE 2004). In agreement with the results observed in

this system, we found that the A353V mutation has a more drastic effect on N. crassa

complex I than the T435M mutation, which results in circa two-fold reduction in

complex I amounts. It is not clear if a similar situation occurs and is primarily responsible for the phenotype of the human patient. It should be noticed that the patient

is a compound heterozygous expressing a truncated 51 kDa protein in addition to the

T435M polypeptide (SCHUELKE et al. 1999).

It is conceivable that complex I proteins evolved an optimum structure to

interact with other proteins (some amino acid residues are even responsible for species-

specific phenotypes (YADAVA et al. 2002)) and alterations interfere negatively with the

efficiency of synthesis, mitochondrial import or enzyme assembly, limiting the amounts

of complex I. This seems to be particularly evident in the case of the 51 kDa protein

containing a A353V substitution, which seems to be expressed but does not reach

complex I. Our interpretation is supported from the findings that mutations in the

mitochondrial signal sequence of the precursor of the 24 kDa subunit of complex I,

which may influence the levels of the protein in mitochondria, result in increased

susceptibility to Parkinson disease (HATTORI et al. 1998). Other mutations lowering the

mitochondrial amounts of the 24 kDa polypeptide (BENIT et al. 2003) and of other complex I subunits (UGALDE et al. 2004) were found as well.

It has been shown that complex I is essential for different cellular processes in

several unrelated organisms (VIDEIRA 1998). Since mammals do not possess alternative

NADH dehydrogenases as for instance fungi and plants, they are expected to rely

completely on complex I activity. Human diseases have been associated with more or 19

less pronounced deficiencies but not with a complete impairment of complex I function,

indicating that this is incompatible with mammalian life and/or development, as recently

shown in mice by homologous deletion of the GRIM-19 subunit of complex I (HUANG et al. 2004). In fact, point mutations of cysteine ligands of iron-sulfur clusters have more drastic effects in preventing complex I assembly and function (ALMEIDA et al.

1999; CHEVALLET et al. 1997; DUARTE and VIDEIRA 2000) and this type of mutations

has not been found in human disorders (SMEITINK and VAN DEN HEUVEL 1999; UGALDE

et al. 2004). Nevertheless, further work is required to better understand the development

of mitochondrial disease associated with complex I and why different mutations lead to

similar phenotypes. The more obvious explanation is that disease arises from an

energetic deficiency as a result of either altered catalytic or assembly/stability properties

of the mutant enzymes. Despite that our analysis of a few cases suggest that mutations

have a higher impact on complex I formation than on enzyme catalytic activity, we

cannot exclude that subtle changes in catalytic activity represent an important issue in

the mammalian enzyme. Extensive studies performed directly in human material

demonstrated that both the catalytic properties and complex I amounts are associated

with disease (SMEITINK et al. 2004; UGALDE et al. 2004). Complex I disease was also claimed to arise from other causes, like deficiency of other respiratory complexes or an increased oxidative stress (BUDDE et al. 2000; SMEITINK et al. 2004; VAN DER

WESTHUIZEN et al. 2003). Disease development likely arises from a combination of different events. Comparison of the outcome from the study of different systems will certainly contribute to our understanding of the processes involved in disease and our ability to control them.

20

Acknowledgments

We would like to thank Dr. Thorsten Friedrich for EPR spectroscopy. This work

was supported by Fundação para a Ciência e a Tecnologia from Portugal and the POCTI

program of QCA III (co-participated by FEDER) through research grants to AV and by

grants from the Deutsche Forschungsgemeinschaft to US.

LITERATURE CITED

AHLERS, P. M., A. GAROFANO, S. J. KERSCHER and U. BRANDT, 2000 Application of

the obligate aerobic yeast Yarrowia lipolytica as a eucaryotic model to analyse

Leigh syndrome mutations in the complex I core subunits PSST and TYKY.

Biochim. Biophys. Acta 1459: 258-265.

ALMEIDA, T., M. DUARTE, A. M. MELO and A. VIDEIRA, 1999 The 24-kDa iron-sulphur

subunit of complex I is required for enzyme activity. Eur. J. Biochem. 265: 86-

92.

ALVES, P. C., and A. VIDEIRA, 1994 Disruption of the gene coding for the 21.3-kDa

subunit of the peripheral arm of complex I from Neurospora crassa. J. Biol.

Chem. 269: 7777-7784.

BENIT, P., R. BEUGNOT, D. CHRETIEN, I. GIURGEA, P. DE LONLAY-DEBENEY et al.,

2003 Mutant NDUFV2 subunit of mitochondrial complex I causes early onset

hypertrophic cardiomyopathy and encephalopathy. Hum. Mutat. 21: 582-586.

BENIT, P., D. CHRETIEN, N. KADHOM, P. DE LONLAY-DEBENEY, V. CORMIER-DAIRE et

al., 2001 Large-Scale Deletion and Point Mutations of the Nuclear NDUFV1

and NDUFS1 Genes in Mitochondrial Complex I Deficiency. Am. J. Hum.

Genet. 68: 1344-1352. 21

BENIT, P., A. SLAMA, F. CARTAULT, I. GIURGEA, D. CHRETIEN et al., 2004 Mutant

NDUFS3 subunit of mitochondrial complex I causes Leigh syndrome. J. Med.

Genet. 41: 14-17.

BRADFORD, M. M., 1976 A rapid and sensitive method for the quantification of

microgram quantities of protein using the principle of protein-dye binding. Anal.

Biochem. 72: 248-254.

BUDDE, S. M., L. P. VAN DEN HEUVEL, A. J. JANSSEN, R. J. SMEETS, C. A. BUSKENS et

al., 2000 Combined enzymatic complex I and III deficiency associated with

mutations in the nuclear encoded NDUFS4 gene. Biochem. Biophys. Res.

Commun. 275: 63-68.

CAMPBELL, J. W., C. S. ENDERLIN and C. P. SELITRENNIKOFF, 1994 Vectors for

expression and modification of cDNA sequences in Neurospora crassa. Fungal

Genet. Newslett. 41: 20-21.

CARROLL, J., R. J. SHANNON, I. M. FEARNLEY, J. E. WALKER and J. HIRST, 2003

Definition of the nuclear encoded protein composition of bovine heart

mitochondrial complex I: Identification of two new subunits. J. Biol. Chem.

277: 50311-50317.

CHEVALLET, M., A. DUPUIS, J. LUNARDI, R. VAN BELZEN, S. P. ALBRACHT et al., 1997

The NuoI subunit of the Rhodobacter capsulatus

(equivalent to the bovine TYKY subunit) is required for proper assembly of the

membraneous and peripheral domains of the enzyme. Eur. J. Biochem. 250:

451-458.

DAVIS, R. H., and F. J. DE SERRES, 1970 Genetic and microbiological research

techniques for Neurospora crassa. Methods Enzymol. 17A: 79-143. 22

DAVIS, R. H., and D. D. PERKINS, 2002 Timeline: Neurospora: a model of model

microbes. Nat. Rev. Genet. 3: 397-403.

DUARTE, M., M. FINEL and A. VIDEIRA, 1996 Primary structure of a ferredoxin-like

iron-sulfur subunit of complex I from Neurospora crassa. Biochim. Biophys.

Acta 1275: 151-153.

DUARTE, M., H. POPULO, A. VIDEIRA, T. FRIEDRICH and U. SCHULTE, 2002 Disruption

of iron-sulphur cluster N2 from NADH: ubiquinone oxidoreductase by site-

directed mutagenesis. Biochem. J. 364: 833-839.

DUARTE, M., U. SCHULTE and A. VIDEIRA, 1997 Identification of the TYKY

homologous subunit of complex I from Neurospora crassa. Biochim. Biophys.

Acta 1322: 237-241.

DUARTE, M., R. SOUSA and A. VIDEIRA, 1995 Inactivation of genes encoding subunits

of the peripheral and membrane arms of Neurospora mitochondrial complex I

and effects on enzyme assembly. Genetics 139: 1211-1221.

DUARTE, M., and A. VIDEIRA, 2000 Respiratory chain complex I is essential for sexual

development in Neurospora and binding of iron-sulfur clusters are required for

enzyme assembly. Genetics 156: 607-615.

FECKE, W., V. D. SLED, T. OHNISHI and H. WEISS, 1994 Disruption of the gene

encoding the NADH-binding subunit of NADH: ubiquinone oxidoreductase in

Neurospora crassa. Formation of a partially assembled enzyme without FMN

and the iron-sulphur cluster N-3. Eur. J. Biochem. 220: 551-558.

FINEL, M., A. S. MAJANDER, J. TYYNELA, A. M. DE JONG, S. P. ALBRACHT et al., 1994

Isolation and characterisation of subcomplexes of the mitochondrial

NADH:ubiquinone oxidoreductase (complex I). Eur. J. Biochem. 226: 237-242. 23

FRIEDRICH, T., P. VAN HEEK, H. LEIF, T. OHNISHI, E. FORCHE et al., 1994 Two binding

sites of inhibitors in NADH: ubiquinone oxidoreductase (complex I).

Relationship of one site with the ubiquinone- of bacterial

glucose:ubiquinone oxidoreductase. Eur. J. Biochem. 219: 691-698.

GRAD, L. I., and B. D. LEMIRE, 2004 Mitochondrial complex I mutations in

Caenorhabditis elegans produce oxidase deficiency, oxidative

stress and vitamin-responsive lactic acidosis. Hum. Mol. Genet. 13: 303-314.

HATTORI, N., H. YOSHINO, M. TANAKA, H. SUZUKI and Y. MIZUNO, 1998 Genotype in

the 24-kDa subunit gene (NDUFV2) of mitochondrial complex I and

susceptibility to Parkinson disease. Genomics 49: 52-58.

HOFHAUS, G., and G. ATTARDI, 1993 Lack of assembly of mitochondrial DNA-encoded

subunits of respiratory NADH dehydrogenase and loss of enzyme activity in a

human cell mutant lacking the mitochondrial ND4 gene product. EMBO J. 12:

3043-3048.

HOFHAUS, G., D. R. JOHNS, O. HURKO, G. ATTARDI and A. CHOMYN, 1996 Respiration

and growth defects in transmitochondrial cell lines carrying the 11778 mutation

associated with Leber's hereditary optic neuropathy. J. Biol. Chem. 271: 13155-

13161.

HOFHAUS, G., H. WEISS and K. LEONARD, 1991 Electron microscopic analysis of the

peripheral and membrane parts of mitochondrial NADH dehydrogenase

(complex I). J. Mol. Biol. 221: 1027-1043.

HUANG, G., H. LU, A. HAO, D. C. NG, S. PONNIAH et al., 2004 GRIM-19, a cell death

regulatory protein, is essential for assembly and function of mitochondrial

complex I. Mol. Cell. Biol. 24: 8447-8456. 24

KIRBY, D. M., R. SALEMI, C. SUGIANA, A. OHTAKE, L. PARRY et al., 2004 NDUFS6

mutations are a novel cause of lethal neonatal mitochondrial complex I

deficiency. J. Clin. Invest. 114: 837-845.

LOEFFEN, J., O. ELPELEG, J. SMEITINK, R. SMEETS, S. STOCKLER-IPSIROGLU et al., 2001

Mutations in the complex I NDUFS2 gene of patients with cardiomyopathy and

encephalomyopathy. Ann. Neurol. 49: 195-201.

LOEFFEN, J., J. SMEITINK, R. TRIEPELS, R. SMEETS, M. SCHUELKE et al., 1998 The First

Nuclear-Encoded Complex I Mutation in a Patient with Leigh Syndrome. Am. J.

Hum. Genet. 63: 1598-1608.

LUNARDI, J., E. DARROUZET, A. DUPUIS and J. P. ISSARTEL, 1998 The nuoM arg368his

mutation in NADH:ubiquinone oxidoreductase from Rhodobacter capsulatus: a

model for the human nd4-11778 mtDNA mutation associated with Leber's

hereditary optic neuropathy. Biochim. Biophys. Acta 1407: 114-124.

MELO, A. M., M. DUARTE, I. M. MOLLER, H. PROKISCH, P. L. DOLAN et al., 2001 The

external calcium-dependent NADPH dehydrogenase from Neurospora crassa

mitochondria. J. Biol. Chem. 276: 3947-3951.

PREIS, D., U. WEIDNER, C. CONZEN, J. E. AZEVEDO, U. NEHLS et al., 1991 Primary

structures of two subunits of NADH:ubiquinone reductase from Neurospora

crassa concerned with NADH-oxidation. Relationship to a soluble NAD-

reducing hydrogenase of Alcaligenes eutrophus. Biochim. Biophys. Acta 1090:

133-138.

RASMUSSEN, T., D. SCHEIDE, B. BRORS, L. KINTSCHER, H. WEISS et al., 2001

Identification of two tetranuclear FeS clusters on the ferredoxin-type subunit of

NADH:ubiquinone oxidoreductase (complex I). Biochemistry 40: 6124-6131. 25

SAMBROOK, J., and D. W. RUSSELL, 2001 Molecular cloning: a laboratory manual.

Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N. Y.

SCHUELKE, M., J. SMEITINK, E. MARIMAN, J. LOEFFEN, B. PLECKO et al., 1999 Mutant

NDUFV1 subunit of mitochondrial complex I causes leukodystrophy and

myoclonic epilepsy. Nat. Genet. 21: 260-261.

SLED, V. D., and A. D. VINOGRADOV, 1993 Kinetics of the mitochondrial NADH-

ubiquinone oxidoreductase interaction with hexammineruthenium(III). Biochim.

Biophys. Acta 1141: 262-268.

SMEITINK, J., and L. VAN DEN HEUVEL, 1999 Human mitochondrial complex I in health

and disease. Am. J. Hum. Genet. 64: 1505-1510.

SMEITINK, J., L. VAN DEN HEUVEL and S. DIMAURO, 2001 The genetics and pathology

of oxidative phosphorylation. Nat. Rev. Genet. 2: 342-352.

SMEITINK, J. A. M., L. W. P. J. VAN DEN HEUVEL, W. J. H. KOOPMAN, L. G. J.

NIJTMANS, C. UGALDE et al., 2004 Cell biological consequences of

mitochondrial NADH:ubiquinone oxidoreductase deficiency. Curr. Neurovasc.

Res. 1: 29-40.

SOKOLOVSKY, V., R. KALDENHOFF, M. RICCI and V. E. A. RUSSO, 1990 Fast and

reliable mini-prep RNA extraction from Neurospora crassa. Fungal Genet.

Newslett. 37: 41.

SOUSA, R., B. BARQUERA, M. DUARTE, M. FINEL and A. VIDEIRA, 1999

Characterisation of the last Fe-S cluster-binding subunit of Neurospora crassa

complex I. Biochim. Biophys. Acta 1411: 142-146.

TRIEPELS, R. H., L. P. VAN DEN HEUVEL, J. L. LOEFFEN, C. A. BUSKENS, R. J. SMEETS

et al., 1999 Leigh syndrome associated with a mutation in the NDUFS7 (PSST)

nuclear encoded subunit of complex I. Ann. Neurol. 45: 787-790. 26

TUSCHEN, G., U. SACKMANN, U. NEHLS, H. HAIKER, G. BUSE et al., 1990 Assembly of

NADH: ubiquinone reductase (complex I) in Neurospora mitochondria.

Independent pathways of nuclear-encoded and mitochondrially encoded

subunits. J. Mol. Biol. 213: 845-857.

UGALDE, C., R. J. JANSSEN, L. P. VAN DEN HEUVEL, J. A. SMEITINK and L. G.

NIJTMANS, 2004 Differences in assembly or stability of complex I and other

mitochondrial OXPHOS complexes in inherited complex I deficiency. Hum.

Mol. Genet. 13: 659-667.

VAN DEN HEUVEL, L., W. RUITENBEEK, R. SMEETS, Z. GELMAN-KOHAN, O. ELPELEG et

al., 1998 Demonstration of a new pathogenic mutation in human complex I

deficiency: a 5-bp duplication in the nuclear gene encoding the 18-kD (AQDQ)

subunit. Am. J. Hum. Genet. 62: 262-268.

VAN DER PAS, J. C., D. A. ROHLEN, U. WEIDNER and H. WEISS, 1991 Primary structure

of the nuclear-encoded 29.9 kDa subunit of NADH: ubiquinone reductase from

Neurospora crassa mitochondria. Biochim. Biophys. Acta 1089: 389-390.

VAN DER WESTHUIZEN, F. H., L. P. VAN DEN HEUVEL, R. SMEETS, J. A. VELTMAN, R.

PFUNDT et al., 2003 Human mitochondrial complex I deficiency: investigating

transcriptional responses by microarray. Neuropediatrics 34: 14-22.

VIDEIRA, A., 1998 Complex I from the fungus Neurospora crassa. Biochim. Biophys.

Acta 1364: 89-100.

VIDEIRA, A., J. E. AZEVEDO, S. WERNER and P. CABRAL, 1993 The 12.3 kDa subunit of

complex I from Neurospora crassa: cDNA cloning and chromosomal mapping

of the gene. Biochem. J. 291: 729-732.

VIDEIRA, A., M. TROPSCHUG, E. WACHTER, H. SCHNEIDER and S. WERNER, 1990a

Molecular cloning of subunits of complex I from Neurospora crassa. Primary 27

structure and in vitro expression of a 22-kDa polypeptide. J. Biol. Chem. 265:

13060-13065.

VIDEIRA, A., M. TROPSCHUG and S. WERNER, 1990b Primary structure and expression

of a nuclear-coded subunit of complex I homologous to proteins specified by the

chloroplast genome. Biochem. Biophys. Res. Commun. 171: 1168-1174.

VIDEIRA, A., and S. WERNER, 1989 Assembly kinetics and identification of precursor

proteins of complex I from Neurospora crassa. Eur J Biochem 181: 493-502.

WALKER, J. E., 1992 The NADH:ubiquinone oxidoreductase (complex I) of respiratory

chains. Q. Rev. Biophys. 25: 253-324.

WALLACE, D. C., 1992 Diseases of the mitochondrial DNA. Annu. Rev. Biochem. 61:

1175-1212.

WANG, D. C., S. W. MEINHARDT, U. SACKMANN, H. WEISS and T. OHNISHI, 1991 The

iron-sulfur clusters in the two related forms of mitochondrial NADH:

ubiquinone oxidoreductase made by Neurospora crassa. Eur. J. Biochem. 197:

257-264.

WERNER, S., 1977 Preparation of polypeptide subunits of cytochrome oxidase from

Neurospora crassa. Eur. J. Biochem. 79: 103-110.

YADAVA, N., P. POTLURI, E. N. SMITH, A. BISEVAC and I. E. SCHEFFLER, 2002 Species-

specific and mutant MWFE proteins: their effect on the assembly of a functional

mammalian mitochondrial complex I. J. Biol. Chem. 277: 21221-21230.

YAGI, T., T. YANO, S. DI BERNARDO and A. MATSUNO-YAGI, 1998 Procaryotic

complex I (NDH-1), an overview. Biochim. Biophys. Acta 1364: 125-133.

ZAUNER, R., J. CHRISTNER, G. JUNG, U. BORCHART, W. MACHLEIDT et al., 1985

Identification of the polypeptide encoded by the URF-1 gene of Neurospora

crassa mtDNA. Eur. J. Biochem. 150: 447-454. 28

ZICKERMANN, V., B. BARQUERA, M. WIKSTROM and M. FINEL, 1998 Analysis of the

pathogenic human mitochondrial mutation ND1/3460, and mutations of strictly

conserved residues in its vicinity, using the bacterium Paracoccus denitrificans.

Biochemistry 37: 11792-11796.

29

FIGURE LEGENDS

FIGURE 1. Analysis of complex I assembly in 21.3c kDa mutant strains. Mitochondria

(2.5 mg protein) were isolated, solubilized with 5% Triton X-100 and centrifuged

through 12 ml sucrose gradients (7.5-25%). Fractions of the gradients (labeled 1-12

from top to bottom) were collected and assayed for NADH:ferricyanide reductase

activity (panel A: closed circles, wild type; open circles, mutant nuo21.3c; open squares, strain P88L; closed squares, strain R111H). Aliquots of the fractions obtained with material from the wild type strain (B) and mutants nuo21.3c (C), P88L (D) and R111H

(E) were also analyzed by western blotting with antisera against the 30.4, 29.9, 20.8 and

12.3 kDa subunits of complex I, as indicated in the left.

FIGURE 2. Analysis of complex I assembly in 51 kDa mutant strains. The experiment

was performed as described in the legend of fig. 1. NADH:ferricyanide reductase

activities are shown in panel A (closed circles, wild type cDNA-rescued strain; open

circles, mutant nuo51; open squares, A353V strain; closed squares, T435M strain).

Western blots are from the cDNA-rescued strain (B) and from the nuo51 (C), A353V

(D) and T435M (E) mutants.

FIGURE 3. Analysis of gene expression in 51 kDa mutant strains. Total RNA was

prepared from wild type grown for 12– 16 h (E, early exponential phase) or 20– 24 h (L,

late exponential phase) and from the cDNA-rescued, nuo51, A353V and T435M strains

grown to late exponential phase. The transcript encoding the 51 kDa protein (1.7 kb)

was identified by Northern blotting using the correspondent labeled-cDNA as probe.

The lower panel showing the 28S rRNA is a direct photograph of the ethidium bromide

stained gel. 30

FIGURE 4. EPR spectra of complex I from 21.3c kDa mutant strains reduced by NADH.

Spectra from P88L (a), R111H (b) and wild type (c) were recorded at 13 K and 2 mW microwave power. Microwave frequency was 9.46 GHz, modulation amplitude was 6 mT, time constant was 0.032 s and scan rate was 17.9 mT/min.

TABLE 1

Catalytic properties of mitochondrial membranes from different N. crassa mutant strains

N. crassa strain Wild type P88L R111H V135M T435M dNADH:Q1 reductase activity, µmol dNADH/min per mg 0.436±0.115 0.236±0.038 0.148±0.053 0.189±0.045 0.242±0.044

(rotenone sensitivity) (95%) (82%) (82%) (89%) (89%)

dNADH KM , µM 6 7 8 5 9

Q1 KM , µM 24 26 16 22 nd

NADH:HAR reductase activity, µmol NADH/min per mg 1.543±0.285 0.629±0.218 0.365±0.156 0.559±0.042 1.112±0. 215

TABLE 2

NADH dehydrogenase specific activities (U/mg) of N. crassa strains

Mitochondria Isolated enzyme ______NADH:ferricyanide NADH:DecQ

(rotenone sensitivity)

Wild type 0.9 ± 0.1 120 ± 10 0.4 ± 0.1 (70%)

P88L 0.4 ± 0.1 102 ± 20 0.3 ± 0.1 (66%)

R111H 0.2 ± 0.1 95 ± 20 0.4 ± 0.1 (74%)

V135M 0.5 ± 0.1 100 ± 20 0.3 ± 0.1 (70%)

500 A ts)

i 400 n u y r 300 a r t i b 200 (ar ty vi

ti 100 ac

0 123456789101112 29.9 - B - 30.4

20.8 -

12.3 - 29.9 C - 30.4 -

20.8 -

12.3 - 29.9 D - 30.4 -

20.8 -

12.3 - 29.9 E - 30.4 -

20.8 -

12.3 -

Figure 1

140 A 120 s) t i 100

80

60 ty (arbitrary un 40

activi 20

0 1234567891011

B 78 – 51 – 30.4 –

C 78 –

30.4 –

D 78 –

30.4 –

E 78 – 51 –

30.4 –

Figure 2 ed

wt cu

EL cDNA -res nuo51 A353V T435M nuo-51 (1.7)

28S

Figure 3 Figure 4