Get 20M+ Full-Text Papers For Less Than $1.50/day. Start a 14-Day Trial for You or Your Team.

Learn More →

Transmission of mitochondrial DNA following assisted reproduction and nuclear transfer

Transmission of mitochondrial DNA following assisted reproduction and nuclear transfer Abstract Mitochondria are the organelles responsible for producing the majority of a cell’s ATP and also play an essential role in gamete maturation and embryo development. ATP production within the mitochondria is dependent on proteins encoded by both the nuclear and the mitochondrial genomes, therefore co-ordination between the two genomes is vital for cell survival. To assist with this co-ordination, cells normally contain only one type of mitochondrial DNA (mtDNA) termed homoplasmy. Occasionally, however, two or more types of mtDNA are present termed heteroplasmy. This can result from a combination of mutant and wild-type mtDNA molecules or from a combination of wild-type mtDNA variants. As heteroplasmy can result in mitochondrial disease, various mechanisms exist in the natural fertilization process to ensure the maternal-only transmission of mtDNA and the maintenance of homoplasmy in future generations. However, there is now an increasing use of invasive oocyte reconstruction protocols, which tend to bypass mechanisms for the maintenance of homoplasmy, potentially resulting in the transmission of either form of mtDNA heteroplasmy. Indeed, heteroplasmy caused by combinations of wild-type variants has been reported following cytoplasmic transfer (CT) in the human and following nuclear transfer (NT) in various animal species. Other techniques, such as germinal vesicle transfer and pronuclei transfer, have been proposed as methods of preventing transmission of mitochondrial diseases to future generations. However, resulting embryos and offspring may contain mtDNA heteroplasmy, which itself could result in mitochondrial disease. It is therefore essential that uniparental transmission of mtDNA is ensured before these techniques are used therapeutically. assisted reproduction technology, mitochondrial DNA, nuclear transfer Submitted on November 25, 2005; accepted on February 28, 2006 Introduction Mitochondria are found in all eukaryotic cells and are essential to cellular function because of their ability to generate ATP, the cell’s major source of energy. They comprise a double membrane with the outer membrane separating the mitochondrial matrix from the cytoplasm of the cell, whereas the inner membrane forms the cristae. Mitochondrial cristae are the site of the electron transfer chain (ETC), the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place (Sherratt, 1991). This aerobic process allows further metabolism of the products of anaerobic glycolysis and the citric acid cycle to produce carbon dioxide and water, with the subsequent release of 32 molecules of ATP. This is a far more economical return than provided by anaerobic generation of ATP (Pfeiffer et al., 2001). The ETC consists of five enzyme complexes, each containing multiple polypeptide subunits (Figure 1). Electrons flow through the first four complexes of the ETC releasing protons from the mitochondria which set up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in the generation of ATP. Figure 1. View largeDownload slide The electron transfer chain (ETC) is the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place. The ETC consists of five enzyme complexes which, with the exception of complex II, all consist of polypeptides encoded by both the nuclear and the mitochondrial genomes. Electrons flow through the first four complexes of the ETC releasing protons (H+) from the mitochondria and thereby setting up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in ATP generation. CoQ, Coenzyme Q; Cyt C, Cytochrome C; ANT, adenosine nucleoside transporter; KCN, potassium cyanide; ADP, adenosine diphosphate; ATP, adenosine triphosphate. Figure 1. View largeDownload slide The electron transfer chain (ETC) is the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place. The ETC consists of five enzyme complexes which, with the exception of complex II, all consist of polypeptides encoded by both the nuclear and the mitochondrial genomes. Electrons flow through the first four complexes of the ETC releasing protons (H+) from the mitochondria and thereby setting up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in ATP generation. CoQ, Coenzyme Q; Cyt C, Cytochrome C; ANT, adenosine nucleoside transporter; KCN, potassium cyanide; ADP, adenosine diphosphate; ATP, adenosine triphosphate. The importance of ATP in oocytes and early embryos has been demonstrated in cattle, as assessed by morphology. Here, the higher quality oocytes contained significantly higher levels of ATP than poorer quality oocytes (Stojkovic et al., 2001). Those containing the highest ATP levels also progressed to produce a significantly higher proportion of blastocysts. A similar pattern has been demonstrated in human oocytes, with those containing at least 2ρmol ATP at fertilization being more capable of embryonic development and implantation (Van Blerkom et al., 1995). The correlation between mitochondrial distribution in individual blastomeres and blastomere ATP content further emphasizes the importance of OX-PHOS for early preimplantation development (Van Blerkom et al., 2000). Mitochondrial DNA Components of the ETC are encoded by both chromosomal and mitochondrial DNA (mtDNA), with all components being required for efficient function of the ETC. The human mitochondrial genome is a double-stranded circular DNA consisting of 16 569 base pairs (Anderson et al., 1981). Using a slightly different genetic code to that of nuclear DNA, mtDNA encodes 13 polypeptides involved in the ETC, 22 transfer (t)RNAs and 2 ribosomal (r)RNAs (Anderson et al., 1981). MtDNA codes for seven subunits of nicotinamide adenine dinucleotide (NADH) dehydrogenase (ND; complex I), one component of complex III, three subunits of cytochrome C oxidase (complex IV) and two subunits of ATP synthase (complex V). All other components of the ETC are encoded by the nuclear genome (Figure 1). The mitochondrial genome is highly compact, with no introns contained between the coding regions (Anderson et al., 1981) and some of the coding regions even overlap (ATPase 6 and ATPase 8, ND4 and ND4L; Anderson et al., 1981). Others lack termination codons, with these sequences being provided by post-transcriptional polyadenylation (Ojala et al., 1981). Despite its small size in comparison with the nuclear genome, the host cell depends on the mitochondrial genome for aerobic respiration as much as the mitochondria rely on proteins encoded by the host cell’s nuclear DNA. In fact, as much as one third of the total DNA content of an oocyte consists of mtDNA (Piko and Matsumoto, 1976). Somatic cells have between 103 and 104 copies of mtDNA, with 2–10 copies per organelle (Satoh and Kuroiwa, 1991), correlating to each particular cell’s requirement for OX-PHOS. For example, it has been reported that human subcutaneous fat cells contain more mtDNA copies than peripheral blood mononuclear cells (Gahan et al., 2001) and that liver tissue contains more mtDNA than various muscle tissues (Wiesner et al., 1992). MtDNA transcription and replication Transcription of mtDNA occurs following interaction between nuclear-encoded regulatory proteins and regions within the d-loop of mtDNA (Figure 2). Requirements for transcription include the mitochondrial RNA polymerase (Tiranti et al., 1997), mitochondrial transcription factor A (TFAM; Fisher and Clayton, 1985; 1988), and one of the recently identified TFB1M and TFB2M (Falkenberg et al., 2002; Table I). Once initiated, transcription generates a polycistronic precursor RNA transcript, allowing co-ordinated transcription of all genes on the same strand. Excision of the polycistronic precursor by endonucleases produces precursor rRNAs and tRNAs, which are then processed further to allow them to translate the precursor mRNAs (Ojala et al., 1981). Figure 2. View largeDownload slide Mitochondrial DNA (mtDNA) transcription and replication origins in the d-loop. Mitochondrial transcription factor A (TFAM) binds and separates the two DNA strands. The accessory transcription factor, either TFB1M or TFB2M, allows mitochondrial RNA polymerase (mtRNAPol) to bind to the correct part of the DNA to produce the transcripts and the RNA primer for replication. Mitochondrial DNA polymerase gamma (PolG) is then able to use the primer to produce the daughter DNA strand. HSP, heavy strand promoter; LSP, light strand promoter; CSB, conserved sequence block; OH, origin of heavy strand replication; d-loop, displacement loop, the non-coding region that is the binding site for nuclear-encoded transcription and replication factors; Phe tRNA, phenylalanine transfer RNA; Pro tRNA, proline transfer RNA. Figure 2. View largeDownload slide Mitochondrial DNA (mtDNA) transcription and replication origins in the d-loop. Mitochondrial transcription factor A (TFAM) binds and separates the two DNA strands. The accessory transcription factor, either TFB1M or TFB2M, allows mitochondrial RNA polymerase (mtRNAPol) to bind to the correct part of the DNA to produce the transcripts and the RNA primer for replication. Mitochondrial DNA polymerase gamma (PolG) is then able to use the primer to produce the daughter DNA strand. HSP, heavy strand promoter; LSP, light strand promoter; CSB, conserved sequence block; OH, origin of heavy strand replication; d-loop, displacement loop, the non-coding region that is the binding site for nuclear-encoded transcription and replication factors; Phe tRNA, phenylalanine transfer RNA; Pro tRNA, proline transfer RNA. The replication of mtDNA is also dependent on nuclear-encoded regulatory proteins (Table I). The replication process begins with light strand transcription from the light strand promoter region to the second conserved sequence block (CSB) within the d-loop region, producing an RNA sequence complementary to the light strand DNA (Xu and Clayton, 1995). Binding of the mtDNA-specific polymerase gamma (PolG) to the resulting RNA–DNA hybrid allows replication of the heavy strand to begin from the origin of heavy strand replication (OH) contained within the d-Loop (Figure 2). After the replication of approximately two thirds of the genome, the formation of the new heavy strand exposes the origin of light strand replication (OL) on the heavy strand parent. Light strand replication is then able to commence in the opposite direction. Table I. Major nuclear-encoded regulatory factors required for mitochondrial DNA (mtDNA) transcription and replication Protein  Binding site  Role  TFAM  Enhancer region  Causes unwinding of mtDNA, exposing the binding site for mtRNA polymerase  TFB1M or TFB2M  RNA polymerase  Form heterodimer with RNA polymerase allowing specific transcription initiation  MtRNAPol  Transcription initiation region  Produces RNA copy of mtDNA sequence  PolG  RNA–DNA hybrid  Produces replicate copy of mtDNA  MtSSB  Single-stranded mtDNA  Prevents re-annealing of single-stranded mtDNA allowing PolG polymerization  MTERF  RNA polymerase  Termination of transcription  RNase MRP  MtDNA primary light strand transcript  Processing of primary light strand transcripts to produce RNA primer for replication  Protein  Binding site  Role  TFAM  Enhancer region  Causes unwinding of mtDNA, exposing the binding site for mtRNA polymerase  TFB1M or TFB2M  RNA polymerase  Form heterodimer with RNA polymerase allowing specific transcription initiation  MtRNAPol  Transcription initiation region  Produces RNA copy of mtDNA sequence  PolG  RNA–DNA hybrid  Produces replicate copy of mtDNA  MtSSB  Single-stranded mtDNA  Prevents re-annealing of single-stranded mtDNA allowing PolG polymerization  MTERF  RNA polymerase  Termination of transcription  RNase MRP  MtDNA primary light strand transcript  Processing of primary light strand transcripts to produce RNA primer for replication  MTERF, mitochondrial transcription termination factor; mtRNAPol, mitochondrial RNA polymerase; mtSSB, mitochondrial single-stranded binding protein; PolG, polymerase gamma; RNase MRP, mitochondrial RNA processing endoribonuclease; TFAM, mitochondrial transcription factor A; TFB1M/TFB2M, mitochondrial transcription factor B1 or B2. View Large As mtDNA transcription and replication, and therefore cellular ATP-generating capacity, are so dependent on proteins encoded by the nuclear genome, mutations in these genes can result in mtDNA depletion syndromes (reviewed by Suomalainen and Kaukonen, 2001). Unlike the vast majority of mtDNA diseases, these diseases can be transmitted in a Mendelian fashion and include chronic progressive external ophthalmoplegia (CPEO; Van Goethem et al., 2001; Graziewicz et al., 2004) and Alpers syndrome (Naviaux et al., 1999). This effect is also mimicked by nucleoside antiretroviral analogues such as 3′-azido-3′-deoxythymidine (AZT) and 2′3′-dideoxycytidine (ddC). For example, men treated with these drugs for 12 months had increased numbers of deleted mtDNA molecules in their sperm samples (White et al., 2001). Furthermore, although heterozygous PolG knockout mice develop normally, despite slightly reduced mtDNA levels, homozygous knockouts have severe mtDNA depletion and do not survive beyond embryonic day 8.5 (Hance et al., 2005). TFAM plays an important role in both transcription and replication of the mitochondrial genome. It is therefore not surprising that low levels of TFAM have been associated with a variety of mitochondrial depletion syndromes, such as infantile mitochondrial myopathy (Poulton et al., 1994), familial mtDNA-associated liver disease (Spelbrink et al., 1998), fatal childhood myopathy (Larsson et al., 1994), skeletal muscle and mitochondrial encephalomyopathy disorders (Siciliano et al., 2000) and ocular myopathy, exercise intolerance and muscle wasting (Tessa et al., 2000). In contrast, there is a decrease in TFAM expression localized to the mitochondria during spermatogenesis. This has been proposed as a functional mechanism for specifically reducing mtDNA copy number in maturing sperm and therefore paternal mtDNA transmission to future generations, in both the human (Larsson et al., 1997) and the mouse (Larsson et al., 1996). This arises at the late spermatocyte/early spermatid stage where an isoform is expressed lacking the mitochondrial targeting sequence. In the mouse, TFAM was also detected in the nucleus of developing sperm cells, indicating a possible further role as a nuclear transcription factor (Larsson et al., 1996). The significance of TFAM expression as a major regulator of mtDNA copy number is exemplified in homozygous knockout mouse embryos that are completely depleted of mtDNA and do not survive after embryonic day 10.5 (Larsson et al., 1998). The replication of embryonic mtDNA, at least in the mouse, does not occur during early preimplantation development (Piko and Taylor, 1987). However, a small window of replication at the two-cell stage was recently reported (McConnell and Petrie, 2004). The mtDNA molecules present in the mature oocyte at fertilization are therefore largely relied upon to provide sufficient ETC components to allow OX-PHOS activity during the preimplantation phase of embryonic development. In order to meet this demand, mtDNA is amplified during oocyte maturation within the ovary, with a 45-fold increase being reported in cattle from the primordial germ cells stage to pre-ovulating oocytes (Smith and Alcivar, 1993), resulting in a mean of 260 000 copies being present in mature bovine oocytes (Michaels et al., 1982). Similar amplifications have been reported in other species. For example, human premigratory primordial germ cells possess approximately 10 mitochondria per cell (Jansen and De Boer, 1998). Mature mouse oocytes contain mean mtDNA copy number values between 119 000 (Piko and Taylor, 1987) and 159 000 (Steuerwald et al., 2000). In the human, copy number ranges from 10 000 to 700 000 (Almeida-Santos et al., 2006), with other reported mean values of 314 000 (Steuerwald et al., 2000) and 193 000 (Reynier et al., 2001). Oocytes with higher copy numbers of mtDNA are known to be associated with improved fertilization rates. For example, human oocytes failing to fertilize because of poor sperm quality, and therefore assumed to be of good quality, had significantly higher mtDNA copy number than poor quality oocytes which were fertilized by good quality sperm (mean 255 000 versus 152 000 copies; Reynier et al., 2001). Furthermore, low-quality oocytes from patients with a variety of ovarian disorders contained an average of 100 000 mtDNA molecules compared with an average of 256 000 copies in oocytes from healthy individuals (May-Panloup et al., 2005). We have also recently reported similar results in porcine (El Shourbagy et al., 2006) and human (Almeida-Santos et al., 2006) oocytes. However, we also saw significantly different copy number between arrested two-cell human embryos and those that developed further (Almeida-Santos et al., 2006). A putative threshold number of mtDNA molecules has been proposed for the mature unfertilized oocyte to allow embryo development to take place after fertilization. In the mouse, this has been predicted at approximately 100 000 copies per oocyte (Piko and Taylor, 1987), with a similar value being appropriate for the pig (El Shourbagy et al., 2006). The relevance of mtDNA copy number threshold to fertilization outcome and embryo development The number of mtDNA copies per organelle is indicative of the number of mitochondria in an oocyte. In the mouse, each mitochondrion has just one copy (Piko and Taylor, 1987). Consequently, the minimum number of mtDNA copies would indicate the minimum number of mitochondria required for fertilization to take place. These mitochondria would support embryonic development to cavitation, after which glycolysis becomes increasingly important in many species (human – Leese et al., 1993; pig – Sturmey and Leese, 2003; cow – Thompson et al., 1996; rat – Brison and Leese, 1991; mouse – Houghton et al., 1996). Furthermore, this minimum number may be essential as it has been hypothesized that a threshold level of ATP is required to energize cell division (Brenner et al., 2000) and may also synchronize mitochondrial dynamics. The distribution of mitochondrial networks throughout the cell varies with different stages of the cell cycle (Margineantu et al., 2002). Mature and immature oocytes also have distinct patterns of mitochondrial distribution (Stojkovic et al., 2001; Nishi et al., 2003). Those oocytes with mtDNA copy number lower than the required threshold may be unable to form the networks required for developmental competence. In support of this, developmental competence appears to be related to the numbers of mitochondria present (El Shourbagy et al., 2006). In sea urchin oocytes, mitochondria act as calcium sinks (Eisen and Reynolds, 1985). The Ca2+ wave pacemaker that is necessary for the completion of meiosis at fertilization is maintained by mitochondrial uptake of Ca2+ (Dumollard et al., 2003). Failure to maintain the Ca2+ wave pacemaker results in apoptotic cell death of the oocyte (Liu et al., 2001). Therefore, oocytes with low levels of ATP production, because of low levels of mitochondria and mtDNA, may be unable to maintain the Ca2+ wave pacemaker and complete meiosis, and instead, undergo apoptosis. We have previously hypothesized that oocytes without sufficient wild-type mtDNA, and therefore the capacity to generate ATP would not normally be ovulated (St. John, 2002). This is reflected in the low numbers of mtDNA observed in degenerate oocytes (mean = 44 629 ± 40 729; Almeida-Santos et al., 2006). Consequently, the use of superovulation protocols during IVF allows these low-quality oocytes to undergo the maturation process and the inadequacies of insufficient mtDNA molecules to be observed (St. John, 2002). Mitochondrial morphology changes after oocyte maturation to the inactive state (Bavister and Squirrell, 2000). This may be because of the lack of mtDNA replication and the import of nuclear-encoded regulatory factors from the cytoplasm. However OX-PHOS is still functional during early embryo development, and mitochondria are dynamic organelles, moving to provide ATP where it is most needed (Stojkovic et al., 2001). Staining of mitochondria has shown that during embryo development, they cluster around the nucleus of blastomeres during the cleavage stages, providing the energy required for cell division processes such as spindle formation. However, when cell division is not taking place, they tend to spread throughout the cytoplasm of the cell (Stojkovic et al., 2001; Sun et al., 2001; Wilding et al., 2001). Equal segregation of mitochondria following cell division reduces the possibility of resulting tissues of the offspring lacking the ability to produce sufficient ATP through OX-PHOS. However, disproportionate numbers of mitochondria among blastomeres has been reported in the pig (El Shourbagy et al., 2006) and human (Van Blerkom et al., 2000; Lin et al., 2004). This may result in some blastomeres with reduced ATP-generating capacity. If this occurs early on in development, blastomeres may fragment, resulting in embryo arrest or less competent blastocysts with fewer cell numbers. In the human, mitochondrial morphology changes back to the active state at the eight-cell stage, coinciding with the activation of the embryonic genome (Sathananthan and Trounson, 2000). Furthermore, during cavitation, shortly after embryonic genome activation, there is an increased dependence on glycolysis in most mammalian species (see Van Blerkom et al., 2002 for discussion). This would suggest that nuclear embryonic genome activation provides a signal to the mitochondria to reduce OX-PHOS activity. Increased use of glycolysis may allow embryonic cells that are no longer able to generate sufficient ATP through OX-PHOS to survive for long enough to enable production and import of nuclear-encoded regulatory factors necessary for the onset of mtDNA replication post-implantation (El Shourbagy et al., 2006). Indeed, studies in pigs using OX-PHOS inhibitors have indicated that the inhibition of OX-PHOS function at compaction can improve embryo viability (Machaty et al., 2001). This hypothesis is supported by reports of glycolysis rather than OX-PHOS metabolism occurring in other cell types with a functioning nuclear genome but low mtDNA copy number, including somatic cells being depleted of their mtDNA (King and Attardi, 1989). However, ATP production through glycolysis is far less efficient than through OX-PHOS. It is therefore vital that sufficient mtDNA molecules are present at fertilization to maintain embryo survival until mtDNA replication resumes. The relevance of mtDNA replication to maternal transmission of mtDNA MtDNA is further unlike nuclear DNA as it does not follow the normal Mendelian pattern of inheritance but is maternally inherited (Giles et al., 1980). Owing to its size, a mature oocyte can contain far more mitochondria in its cytoplasm than the fertilizing sperm cell. Coupled with replication of the mitochondrial genome during oocyte maturation in the ovary, this results in an oocyte : sperm mtDNA ratio of 1000:1 (Ankel-Simons and Cummins, 1996), with sperm mtDNA contributing approximately only 100 molecules at fertilization (Hecht and Liem, 1984). MtDNA is usually homoplasmically transmitted, i.e. only one mtDNA genotype being present (Monnat et al., 1985). Occasionally, two or more mtDNA genotypes can be present, often because of either a pathological or a non-pathological rearrangement, resulting in heteroplasmy, as in the oocytes of women harbouring a mtDNA disease (Blok et al., 1997). MtDNA has a much higher mutation rate than nuclear DNA (Wallace et al., 1987). This is a consequence of the production of mutagenic free radicals within the mitochondria, which are further confounded by limited mtDNA repair mechanisms (Yakes and Van Houten, 1997). Furthermore, mtDNA is not stabilized with histones as is the case for nuclear DNA, although TFAM is now thought to be involved in mtDNA packaging (Alam et al., 2003). Despite the high mutation rate, mtDNA has been highly conserved throughout evolution, and when it arises, mtDNA heteroplasmy tends to revert back to homoplasmy sometimes within one generation (Koehler et al., 1991). MtDNA replication does not occur in synchrony with nuclear DNA replication, and therefore some mtDNA genomes are not replicated during one-cell cycle, whereas others are copied many times (Sazer and Sherwood, 1990). It has been reported that there is a preferential replication of those mtDNA molecules that are closer to the nucleus (Davis and Clayton, 1996), perhaps because of the faster diffusion of nuclear-encoded regulatory factors. The introduction of exogenous mtDNA into zygotes either close to the nucleus or at the periphery influences the amount of exogenous mtDNA observed at the blastocyst stage, with higher proportions arising from karyoplast injections (Meirelles and Smith, 1998). Consequently, the position of mitochondria in the early embryo may play a role in determining which mtDNA genomes are selected for amplification. Contradictory data have recently shown that mtDNA replication is not just restricted to mitochondria in proximity to the nucleus (Magnusson et al., 2003). Here, it was demonstrated that all components of the replication machinery were available in sufficient quantities throughout the cell, allowing replication to take place throughout the entire cytoplasm. This included PolG and TFAM. The previously reported preferential labelling of mtDNA near cell nuclei was suggested to be because of denser aggregation of mitochondria around nuclei (Magnusson et al., 2003). MtDNA replication throughout the cell cytoplasm is supported by a lack of evidence for selection against pathogenic mtDNA mutations in oogenesis or early embryonic development (reviewed by Shoubridge, 2000). When mtDNA replication resumes following implantation (Piko and Taylor, 1987), each cell type of the developing embryo contains a few copies of mtDNA, as indicated by the few mitochondria present in undifferentiated embryonic stem cells (ESCs; Sathananthan et al., 2002; St. John et al., 2005a). Consequently, the combination of segregation followed by the amplification of selected templates produces an mtDNA purification process (Hauswirth and Laipis, 1982), which is important for the maintenance of mtDNA homoplasmy. If blastomeres inherit mutated genomes or are depleted of mitochondrial genomes, then the resulting adult tissues would be deficient in OX-PHOS capacity, which could result in the onset of mitochondrial disease (Wallace, 1992). There are many diseases caused by mutations in mtDNA and symptoms of these diseases generally appear in tissues highly dependent on OX-PHOS (Wallace, 1992). These include Leber’s hereditary optic neuropathy (LHON), which can result from mutations in the gene for the NADH-Q oxidoreductase component of complexes I and III (Brown et al., 1992). Myoclonic epilepsy with ragged-red fibres (MERRF) is also transmitted by maternal inheritance and results in myoclonus, epilepsy and ataxia. This can result from mutations in mitochondrially encoded tRNA genes (Hanna et al., 1995; Nakamura et al., 1995; Nishigaki et al., 2003; Mancuso et al., 2004) and protein genes (Naini et al., 2005). As well as point mutations, some mitochondrial diseases are caused by large-scale deletions to the genome. The most well known in this group is Kearns–Sayre Syndrome (KSS), which includes symptoms of PEO, pigment retinopathy and disorders of cardiac conduction (Klopstock et al., 1995; Lertrit et al., 1999; De Block et al., 2004). The clinical phenotypes resulting from mtDNA mutations are dependent on the proportion of mutated mtDNAs (Wallace, 1992). In the case of LHON, >60% mutant mtDNA load is required before the disease phenotype presents (Chinnery et al., 2001). In some mitochondrial diseases, including a case of MERRF, over 85% mutant mtDNAs need to be present before symptoms are observed (Boulet et al., 1992). Paternal mtDNA inheritance There is increasing evidence demonstrating that maternal-only mtDNA transmission is not merely because of dilution of the paternal mitochondria beyond the level of detection. It has been suggested that paternal mtDNA is actively eliminated by ubiquitination in the oocyte cytoplasm and subsequent proteolysis during embryonic development (Sutovsky et al., 1999; Nishimura et al., 2006). The elimination hypothesis is supported by reports demonstrating reduced paternal mtDNA content as embryonic development progresses (Cummins et al., 1997). Paternal mtDNA was detected only as far as the pronucleus stage following intra-specific mouse crosses compared with detection in neonates following inter-specific mouse crosses (Kaneda et al., 1995). This suggests that the elimination mechanism is species specific. However, the elimination mechanism is not only species specific but also tissue specific, as donor sperm mtDNA was eliminated, whereas exogenous liver mtDNA injected into the oocyte persisted at least until the neonatal period (Shitara et al., 2000). The destruction of paternal mtDNA (reviewed in Sutovsky, 2003) and the presence of the genetic bottleneck should usually result in homoplasmic individuals. Recently, the accepted theory of uniparental mtDNA inheritance has been questioned. This is especially so in the case of a male patient presenting with a mitochondrial myopathy who was found to have inherited his muscle mtDNA from his father. However, the 2bp deletion giving rise to his myopathy was de novo (Schwartz and Vissing, 2002). This outcome is further supported by the evidence of sperm mtDNA persistence in human polyploid IVF-generated blastocysts (St. John et al., 2000). Paternal leakage of mtDNA has also been reported following inter-specific crosses in fruitflies (Kondo et al., 1990), mice (Gyllensten et al., 1991), honeybees (Meusel and Moritz, 1993), birds (Kvist et al., 2003), sheep (Zhao et al., 2004) and non-human primates (St. John and Schatten, 2004) and following intra-specific crosses in mussels (Zouros et al., 1992) and fruitflies (Kondo et al., 1992). The ubiquitin-related paternal mtDNA destruction process has therefore proven ineffective in these cases. It is as yet unknown how paternal mtDNA might escape the destruction process and how the paternal mtDNA might be preferentially replicated in a particular tissue to such an extent that it resulted in a mitochondrial disease (Schwartz and Vissing, 2002). In this latter instance, a considerable selective advantage must have taken place to out-compete the maternal source of mtDNA in the affected tissue. A possible mechanism for bypassing the paternal mtDNA destruction process is the fusion of parental mitochondria followed by recombination of the two genomes. Recombination of mammalian mtDNA was originally thought not to arise, although it is known to occur in yeast. However, more recent evidence suggests that it may also occur in mammals. For example, Ono and colleagues created two respiration-deficient cell lines from patients with mitochondrial diseases caused by different mutations (Ono et al., 2001). Fusion of the two-cell lines resulted in morphology and respiratory enzyme activity returning to normal within 10–14 days. These authors proposed that the outcome resulted from mtDNA recombination and genome compensation. Mammalian mtDNA recombination has since been confirmed in human cells using PCR-based techniques (D’Aurelio et al., 2004) and recent sequence analysis of invertebrate and vertebrate species, including primates (Tsaousis et al., 2005). These authors suggested that recombination may have been previously missed because of the purification process during oocyte production prohibiting the generation of new haplotypes and their transmission to the next generation. Recombination of sperm and oocyte mtDNA is now thought to explain the presence of the 2bp deletion in the male myopathy patient (Kraytsberg et al., 2004). Invasive oocyte reconstruction protocols The introduction of more advanced assisted reproductive technologies (ART) and their current use in the clinical setting and in animal production further confounds the unimaternal transmission of mtDNA. These technologies include (i) ICSI which is used clinically; (ii) cytoplasmic transfer (CT) which has been used clinically but is now, for example, banned in the United Kingdom and United States and (iii) germinal vesicle transfer (GVT), pronuclear transfer (PNT) and NT, which have not yet been used to treat humans. These more invasive ART protocols present more opportunities for the violation of the normal pattern of mtDNA inheritance. If uniparental mtDNA inheritance is not maintained following these techniques, offspring may harbour various forms of mtDNA heteroplasmy. Indeed, the introduction of these new technologies results in heteroplasmy being redefined. In this instance, either a wild-type molecule can co-exist with a rearranged molecule or two wild-type molecules may persist in the reconstructed oocyte. ICSI Unlike IVF, the ICSI procedure varies dramatically from the natural fertilization process. For example, the intact, non-capacitated, non-acrosome-reacted sperm is injected into the oocyte. There are also delays and alterations in post-fertilization processes following ICSI, including male pronucleus formation (Luetjens et al., 1999) and the calcium oscillations essential for the activation process (reviewed in Ludwig et al., 2001). This delayed response after sperm entry means that ICSI also has the potential to delay degradation of sperm mtDNA. Furthermore, couples who have male factor infertility and require ICSI are at a higher risk of using sperm containing abnormal mtDNA as some human mtDNA haplogroups are associated with high or reduced sperm motility (Ruiz-Pesini et al., 2000). This is demonstrated in those men harbouring large numbers of mtDNA deletions who are more likely to exhibit poorer semen quality (Lestienne et al., 1997; Reynier et al., 1997; St. John et al., 2001) as are those exhibiting point mutations (Spiropoulos et al., 2002). Despite the many diversions from the natural fertilization processes following ICSI, to date, it appears that sperm mtDNA is eliminated in offspring generated using this technique (Danan et al., 1999; Marchington et al., 2002). Furthermore, it appears to be eliminated before the blastocyst stage in ICSI-derived polyploid human embryos (St. John et al., 2000). CT CT is an extension of ICSI and involves the injection of supplementary cytoplasm, as well as a sperm, into the oocyte (Figure 3). The rationale for using CT is that donor cytoplasm will contain supplementary, often younger, non-mutated mtDNA and will possibly contain other important cytoplasmic components to compensate for the defective cytoplasm in the oocyte (Cohen et al., 1997). In the human, it has resulted in higher pregnancy rates and live births (Cohen et al., 1997; Barritt et al., 2001). CT can also result in heteroplasmy as demonstrated by the analysis of various tissues of human CT offspring (Brenner et al., 2000). Consequently, it remains to be demonstrated whether the donor mtDNA out-competes the recipient mtDNA in the heteroplasmic offspring. We also need to ascertain whether the coupling of two distinct genomes accounted for the two cases of Turner’s syndrome, one electively aborted and the other spontaneously aborted, and the one case of pervasive development disorder (Barritt et al., 2001). Figure 3. View largeDownload slide Assisted reproductive techniques. Predicted outcomes associated with mitochondrial DNA (mtDNA) transmission following IVF, ICSI, cytoplasmic transfer (CT), germinal vesicle transfer (GVT), pronuclei transfer (PNT) and somatic cell nuclear transfer (SCNT). Figure 3. View largeDownload slide Assisted reproductive techniques. Predicted outcomes associated with mitochondrial DNA (mtDNA) transmission following IVF, ICSI, cytoplasmic transfer (CT), germinal vesicle transfer (GVT), pronuclei transfer (PNT) and somatic cell nuclear transfer (SCNT). To avoid the possible harmful effects of heteroplasmy in the offspring, cytoplasm from granulosa cells of the mother rather than cytoplasm from donor oocytes has been used (Kong et al., 2003; 2004). These are likely to have the same mtDNA sequence as the recipient oocyte. However, one possible problem with this procedure is that mtDNA deletions in granulosa cells are more common in older women (Seifer et al., 2002), i.e. those that are likely to have defective oocytes and require CT. Therefore, the treatment may be unsuccessful because of continued lack of wild-type mtDNA and may even result in the transmission of mitochondrial disease to the offspring that would not have occurred without CT. Another likely problem is that granulosa cells are fully differentiated, whereas newly fertilized oocytes are pluripotent. Cells at different stages of differentiation are known to be different in terms of their mitochondrial activity and content. For example, undifferentiated human (h) ESCs (Sathananthan et al., 2002; St. John et al., 2005a) and primordial germ cells (Jansen and De Boer, 1998) have been observed to contain low numbers of mitochondria. However, differentiated cells contain higher numbers of more active mitochondria, depending on OX-PHOS requirements of specific cell types (Moyes et al., 1998). Mitochondria donated from granulosa cells would therefore need to be ‘reprogrammed’ to match those of the oocyte. We have supplemented developmentally incompetent pig oocytes with pure populations of mitochondria from developmentally competent oocytes from maternal relatives. Interestingly, we observed improved fertilization outcome and the maintenance of unimaternal inheritance (El Shourbagy et al., 2006). GVT and PNT GVT and PNT have also been proposed as techniques to overcome problems associated with aged oocytes, whereby the GV or PN is transferred into an enucleated recipient oocyte. Ideally, the enucleated recipient should be from the same stage of development as the donor nucleus (Liu et al., 1999). In aged oocytes, there is likely to be a larger number and higher proportion of mutant mtDNA (Keefe et al., 1995). It has been hypothesized that mitochondria are important not only for the production of ATP but also for correct chromosomal segregation during meiosis (Schon et al., 2000). Furthermore, it has been suggested that aberrant mitochondrial function in aged oocytes may have a role in oocyte aneuploidy (Eichenlaub-Ritter et al., 2004), a common cause of infertility in older women (Munne et al., 1995; Dailey et al., 1996). The chances of success may therefore increase if the nuclear component of the aged oocyte was transferred into the enucleated ooplasm of a younger, healthier oocyte, rather than supplement the existing ooplasm. This transfer can be carried out at either the GV or the PN stage (Figure 3). Indeed, studies using human oocytes have shown GVT from aged oocytes to the enucleated ooplasm of young oocytes can overcome oocyte aneuploidy, with the majority of reconstructions having normal karyotypes (Zhang et al., 1999; Takeuchi et al., 2001). There may also be non-cytoplasmic components of the GV that influence chromosome segregation. This has been indicated in a series of mouse experiments mixing GVs and ooplasms of varying ages (Cui et al., 2005). A significantly greater number (57.1 %) of chromosomal abnormalities were found in reconstructions when the GV was transferred from older females to the enucleated ooplasm of younger females. This is in contrast to abnormalities found in reconstructions derived from the transfer of younger GVs into aged ooplasms (16.7 %). Mitochondria are known to be involved in chromosome organization and movement. For example, the induction of mitochondrial damage in mouse oocytes prevented oocyte maturation, chromosomal segregation and spindle formation (Takeuchi et al., 2005). This damage was overcome by GVT of the karyoplast derived from a damaged oocyte into the ooplasm of a healthy enucleated oocyte. Cytogenetic analysis showed that 20/21 of these reconstructions had a normal number of chromosomes. Furthermore, performing ICSI on GVT reconstructions and transfer of the resulting two-cell embryos led to the generation of live mouse offspring. It has also been demonstrated that an intact mouse ooplasm could rescue mitochondrial-damaged karyoplasts induced by photoirradiation with 62% of these reconstructions maturing to metaphase II (Palermo et al., 2002). PNT is essentially the same procedure, except for the nuclear material being removed after fertilization. This technique has also been used in mice with considerable success as determined by the birth of live offspring (Meirelles and Smith, 1997; 1998;He et al., 2003; Sato et al., 2005). This procedure, however, involves the destruction of a zygote, which may restrict its usage because of ethical and moral considerations. GVT and PNT have been proposed as techniques for treating mtDNA disease (Cummins 1998; Trounson, 2001). However, potential problems could arise as the transferred GV or PN is still surrounded by mitochondria, which will also be carried over into the donor ooplasm. These mitochondria remain close to the centre of the immature reconstruction at first but then disperse throughout the cytoplasm as maturation ensues (Fulka, 2004). Owing to segregation during embryogenesis, this provides an opportunity for karyoplast mitochondria to be present in all cell types of the offspring. Any disease-associated mitochondria transferred with the nuclear material may therefore continue to be transmitted. It is further possible that, due to its initial proximity to the nucleus, karyoplast mtDNA would be preferentially replicated in the reconstructions. Indeed, fusion of enucleated somatic cells with mtDNA-depleted somatic cells can lead to the preferential expansion of one particular type of mtDNA such as the mutant molecule (Dunbar et al., 1995). Although, to our knowledge, there has been no analysis of mtDNA transmission following GVT, results are likely to be similar to those reported following PNT. In this instance, exogenous mtDNA accompanying the murine zygotic karyoplast is found in higher concentrations at blastocyst than mtDNA from transferred cytoplasts (Meirelles and Smith, 1998). This has been further highlighted where zygotic karyoplast transfer into zygotes resulted in varying amounts of mtDNA being transmitted to offspring (0–69%; Meirelles and Smith, 1997). Consequently, it would be difficult to predict to what extent a mutated or deleted mtDNA molecule would be selected. However, a recent report suggests that the varying levels of mtDNA deletion in mouse zygotes could be reduced to levels lower than those expected for the phenotypic onset of mtDNA disease when the zygotic nucleus is transferred into an enucleated healthy oocyte (Sato et al., 2005). The subsequent offspring also remained under the level for pathological onset even with the accumulative increase of the deletion with time. These authors state however that, in the human, this form of nuclear transplantation could not be applied to those mtDNA diseases where the pathogenic mutation had significant replicative advantage over wild-type mtDNA. This is most likely as in humans the prenatal period is 13 times longer than in the mouse, and this represents the window in which mutant mtDNA replication is most proliferative (Sato et al., 2005). NT NT involves the injection of a donor nucleus or a whole cell into an enucleated recipient oocyte (Figure 3). Mostly commonly, either embryonic cell (EC) NT, using a blastomere, or somatic cell NT (SCNT) are performed. Resulting embryos have then been used to generate blastocysts and offspring in a variety of species (Campbell et al., 2005). However, NT is renowned for its low efficiency, with low proportions of embryos developing to the blastocyst stage (White et al., 1999) and even fewer progressing far enough to implant, survive pregnancy and produce live offspring (Wilmut et al., 1997). One of the major abnormalities associated with NT offspring is the production of abnormal placentae which are thought to be the cause of many fetal deaths (Cibelli et al., 2002). Incorrect reprogramming of imprinted genes has also been cited as one of the major factors which leads to early fetal death. This was highlighted by non-surviving NT calves having incorrect expression of imprinted genes, namely Igf2, Igf2r and H19 in some tissues, whereas surviving adult clones had normal expression at these loci except for Igf2 in muscle (Yang et al., 2005). Furthermore, in mouse NT embryos and offspring irregular patterns of gene expression have been observed most likely as a result of inappropriate reprogramming (Humpherys et al., 2002). Although the process of NT for human reproductive purposes is ethically unacceptable and is banned in many countries (UK – Human Fertilisation and Embryology Act, 1990; USA – H.R. 534 Office of Legislative Policy and Analysis, 2003; South Korea – see Normile and Mann, 2005), its use as a means of producing histocompatible stem cells for treatment of various diseases is likely to have a significant impact. Cell replacement therapy has great potential for the treatment of certain conditions such as myocardial infarction (Rubart and Field, 2006) and Parkinson’s disease (Taylor and Minger, 2005) and as a replacement for tissue/organ transplantation (Cortesini, 2005). To date, encouraging results have been obtained in animal studies, whereby hESC-derived neural precursors (Wernig et al., 2004) and cardiomyocytes (Laflamme et al., 2005) were successfully transplanted into animal organs. However, considerable characterization and validation of precursor cells derived from hESCs still needs to be undertaken before cell replacement can be used as an effective therapy in humans. Some success at generating NT-derived ESCs has so far been achieved in mouse models (Munsie et al., 2000; Wakayama et al., 2001), and its use as a means of delivering the patient’s own chromosomal DNA with the corrected gene for a single gene disorder has been demonstrated in the mouse (Rideout et al., 2002). One of the rate-limiting factors in generating autologous ESCs, i.e. stem cells possessing the patient’s own chromosomal and mtDNA compositions is the regulation of the mtDNA composition. In this respect, a considerable amount can be learnt from those offspring derived from NT through either ECNT or SCNT. Here, transmission results in either homoplasmy derived from the recipient oocyte mtDNA or heteroplasmy resulting from transfer of both donor and recipient mtDNA (reviewed in St. John et al., 2004a,b). This is probably because of the random pattern of mtDNA segregation during embryogenesis (reviewed in Shoubridge, 2000). Consequently, it is apparent that no real mechanism is currently understood that could account for the irregular patterns of mtDNA transmission observed. One parallel that can be drawn from both ECNT and SCNT is that the range of donor mtDNA transmitted is relatively similar at 0–57% (Hiendleder et al., 1999) and 0–59% (Takeda et al., 2003), respectively. However, two explanations could exist for those studies not detecting donor mtDNA in live offspring: (i) no donor mtDNA is actually present or (ii) too few tissues have been analysed. Consequently, the true extent of heteroplasmy is difficult to determine. This is perhaps exemplified by the recent studies of pig (Takeda et al., 2006), cow (Takeda et al., 2003) and mouse (Inoue et al., 2004) NT-derived offspring, where varying levels of heteroplasmy has been observed between the tissues. This would indicate that if all tissues are not analysed, then these studies are not truly subjective. Furthermore, investigations should include analysis of the surrogate mother’s mtDNA as leakage through the placenta into fetal blood has been reported in cattle (Hiendleder et al., 2004). The initial number of donor mtDNA molecules introduced into the recipient oocyte also does not necessarily influence whether donor mtDNA is actually transmitted. Using blastomeres from more advanced embryonic stages in bovine NT has resulted in reduced levels of heteroplasmy being reported (Steinborn et al., 1998). However, this was not a uniform pattern as two offspring derived from 52-cell morula blastomeres contained less donor mtDNA than offspring derived from 92-cell morula blastomeres. A further issue is the stage of embryonic development of the donor cell. A recent analysis of two NT-derived monkey offspring (St. John and Schatten, 2004) demonstrated mtDNA transmission from both the recipient oocyte and the donor cell. In this instance, the early IVF blastomere (8- to 12-cell stage) still contained sperm mtDNA. Consequently, the offspring possessed three different populations of mtDNA. Furthermore, the depletion of donor cell mtDNA to residual levels (<0.02% of its original content) still resulted in the persistence of some mtDNA at the blastocyst stage in sheep NT (Lloyd et al., 2006). The use of a somatic donor cell with accompanying mitochondria not only raises interesting questions with regard to the differentiated status of nucleus but also of the mitochondria. During development, from the inner cell mass to a fully differentiated cell, mitochondria increase in number to meet the metabolic requirements of that specialized cell. The proliferation of mitochondria further discriminates between the number of mtDNA copies per mitochondrion and cell type, with higher ATP users possessing greater copy numbers (Moyes et al., 1998). Therefore, for SCNT, the mitochondria accompanying the donor cell should also be reprogrammed to the embryonic state, as is the case for the donor nucleus. Failure to do this may result in embryos containing mitochondria with insufficient plasticity to suit a multitude of cell types that they will contribute to. The persistence of donor mitochondria could also influence the metabolic pathways used during early development and explain why improved developmental rates are observed in those embryos cultured in media that enhance anaerobic respiration (Chung et al., 2002). It is further evident that the injection of somatic donor mitochondria into oocytes which are then parthenogenetically activated is detrimental to embryonic development though this is not observed with donor ooplasm (Takeda et al., 2005). Interaction between multiple wild-type mitochondrial genomes It has not been clearly demonstrated whether heteroplasmy resulting from two wild-type mtDNA molecules can be detrimental to the offspring and result in mitochondrial disease. This is partly because many post-mortem results have not been published, and the cause of many deaths following NT is not documented. Consequently, we are unsure whether those offspring that survive are the exception to the rule. However, from one report, some common causes of death in sheep NT offspring are similar to symptoms of mitochondrial diseases, including brain, kidney and liver defects (McCreath et al., 2000) and in a series of other species (Cibelli et al., 2002). Sequence differences due to mtDNA from a ‘foreign’ source can give rise to proteins with altered amino acid sequences. This has been demonstrated in both cattle (Steinborn et al., 2002) and pigs (St. John et al., 2005b) and may result in inadequate interaction between the individual subunits of the ETC. This would result in reduced energy production capacity and symptoms common to mitochondrial disease and influence embryo and fetal survival. The question is whether handmade cloning, the fusion of one or more oocyte cytoplasts with a somatic cell (Vajta et al., 2001), would resolve such issues or complicate them. The significance of nucleo–mitochondrial interaction Intergenomic communication is vital for efficient cellular function as much interaction occurs between nuclear-encoded proteins and those encoded by the mtDNA genome (reviewed by Scarpulla, 2002). The ETC requires nuclear-encoded proteins to be transported to the mitochondria and to assemble appropriately (Stojanovski et al., 2003). Failure of TFAM, TFB1M and TFB2M to co-ordinate transcription would also have serious implications for a functional ETC (Falkenberg et al., 2002). Furthermore, failure of TFAM and PolG to co-ordinate replication would have similar outcomes for replication (Larsson et al., 1998; Hance et al., 2005). However, NT appears to result in the continued expression of both these factors during embryo development and could explain why even residual levels of donor mtDNA persist up to the blastocyst stage (Lloyd et al., 2006). Failure of TFAM or PolG to interact with the mtDNA genome, post-implantation, through the diversity of the donor nucleus and the recipient oocyte’s mtDNA would result in compromised transcription or replication and mimic mtDNA-depletion syndromes (Poulton et al., 1994). Patient-specific ESCs derived by NT would require large numbers of good-quality human oocytes which, due to medical, ethical, legal and moral issues, are a limited resource. Added to this, the methods of recruitment of donor oocytes for the derivation of such ESCs have caused considerable outrage recently (Holden, 2005). Cross-species NT offers the opportunity to derive ESCs that would carry the chromosomal DNA from a patient propagated in a surrogate ooplasm of another species. Encouragingly, there is a single report of the generation of a hybrid ESC line following the transfer of a human nucleus into a rabbit oocyte (Chen et al., 2003). However, outcomes from cross-species NT further exemplify the importance of nucleo-mitochondrial compatibility for ESC derivation. Use of both human cord fibroblasts with bovine oocytes (Chang et al., 2003) and chicken blastodermal cells with rabbit oocytes (Liu et al., 2004) resulted in elimination of the donor mitochondrial genome by the morula or blastocyst stage of development. Rhesus macaque donor cells injected into rabbit oocytes resulted in heteroplasmic embryos but with only the rabbit mtDNA being replicated after the blastocyst stage (Yang et al., 2003), suggesting that these too could have soon become homoplasmic for the recipient oocyte mtDNA. This suggests either preferential replication of the oocyte mtDNA or degradation of the mtDNA from the nuclear donor cell. Analysis of cross-species NT offspring provides further indications as to potential outcomes for cross-species ESCs. In two combinations of cross subspecies cow NT experiments, the donor nucleus was capable of interaction with both types of mtDNA with the production of apparently healthy heteroplasmic calves (Steinborn et al., 2002; Han et al., 2004). However, the combination of Bos indicus donor cells with Bos taurus oocytes resulted in homoplasmic embryos and one offspring (Meirelles et al., 2001). There has also been one report of preferential replication of mtDNA from the nuclear donor cell following cross-species NT (Chen et al., 2002). Here, nuclei from giant panda somatic cells were injected into rabbit oocytes, and preferential replication of the panda mtDNA was observed after analysis of the fetuses. This suggests that some combinations of cross-species NT, particularly where the species are closely related or a subspecies, may result in a functional ETC and efficient mitochondrial respiration. However, more distantly related species might preferentially replicate their own mtDNA genome because of incompatibility with that in the recipient oocyte. Much can also be learnt from somatic cell–cell fusion studies and xenomitochondrial offspring in relation to outcomes associated with cross-species NT. For example, one study involved combinations of cells depleted of their mtDNA and exogenous mitochondria from rats and two subspecies of mouse (McKenzie and Trounce, 2000). Transfer of Mus spretus mtDNA into Mus musculus cells resulted in heteroplasmy, and no obvious OX-PHOS defects, probably due to the two types of mice, being closely related. However, when rat mtDNA was transferred into Mus musculus cells, its replication and transcription were unaffected, but OX-PHOS capacity was reduced. Other studies have also demonstrated increased respiratory chain defects following crosses of greater evolutionary distance both in mice (McKenzie et al., 2003) and in primates (Kenyon and Moraes, 1997). Effects of species differences, and also mitochondrial disease, can now be studied in vivo following the production of xenomitochondrial mice (McKenzie et al., 2004; Sokolova et al., 2004). Xenomitochondrial mice can be produced by (i) direct injection of foreign mitochondria into zygotes (Irwin et al., 1999), (ii) pronuclei fusion with enucleated oocytes (Meirelles and Smith, 1997) followed by the transfer to surrogate mothers and (iii) the production of ESC cybrids for injection into blastocysts (Levy et al., 1999). Again, it has been demonstrated that greater evolutionary distance between species results in offspring with increased respiratory chain defects (Trounce et al., 2004). It is most likely that OX-PHOS defects arose from abnormal combinations of nuclear and mitochondrially encoded ETC proteins resulting in abnormal structure of ETC complexes, a scenario very similar to outcomes that arise through cross-species NT. There are many cases where interspecies NT embryos implant and survive pregnancy but die shortly after birth (Hammer et al., 2001; Han et al., 2004). These may be a result of inefficient ETC activity due to species differences. However, should these combinations result in apparently healthy live offspring, it is possible that phenotypic alterations may still be present. Such outcomes could affect those mtDNA haplotypes that influence, for example, milk quality in cattle (Brown et al., 1989), growth and physical performance in mice (Nagao et al., 1998), a reduction of fertility in beef cattle (Sutarno et al., 2002) and sperm motility in men (Ruiz-Pesini et al., 2000). Perhaps, the biggest note of caution is exemplified by the result of introducing a common carp nucleus into an enucleated goldfish oocyte, which resulted in the offspring developing phenotypic features of the goldfish (Sun et al., 2005). Consequently, the use of cross-species NT for the derivation of hESCs needs considerable investigation although it offers great potential. Are these technologies safe? Although some experiments have been carried out in animals and a small number of groups practise invasive ART protocols in humans, little is known about the likely outcomes of these techniques in relation to the long-term effects on offspring and future generations. It is essential that we understand the risks of these techniques before their therapeutic use becomes widespread. The use of preimplantation genetic diagnosis (PGD) has recently been proposed for early diagnosis of mitochondrial disease (Dean et al., 2003). The use of PGD for chromosomal genetic analysis has been implemented in many clinics. However, its effectiveness needs to be demonstrated for the detection of potential mtDNA disease transmission. Some work has indicated its usefulness in this respect. In the mouse, mtDNA genotypes in polar bodies were compared with their respective oocytes and in different blastomeres within the same early embryo (Dean et al., 2003). The level of heteroplasmy present in polar bodies and blastomeres was representative of that in the associated oocytes and embryos, with the authors therefore concluding that PGD was feasible for the diagnosis of mitochondrial disease. However, cloned mouse offspring (Inoue et al., 2004), various farm animal species such as pigs (Takeda et al., 2006) and cattle (Hiendleder et al., 2003) and patients with mitochondrial mutations (Kirches et al., 2001) and diseases (Ponzetto et al., 1990) often demonstrate varying levels of heteroplasmy in different tissues. Furthermore, considerable variation in oocyte-derived mtDNA copy number among intra-embryo blastomeres has been demonstrated in both the pig (El Shourbagy et al., 2006) and the human (Lin et al., 2004). Therefore, even if the proportion of heteroplasmic mtDNA molecules were the same in each blastomere, the actual number of mutant molecules is likely to be different. Blastomeres containing lower numbers of mtDNA molecules will have reduced ATP-generating capacities compared with their counterparts with higher overall numbers of mtDNA. Therefore, the effects in tissues resulting from these blastomeres will be more severe than those resulting from blastomeres with a larger mtDNA complement. Furthermore, different tissues require different ATP-generating capacity. Low levels of wild-type mtDNA in some blastomeres may therefore go unnoticed if that particular blastomere develops into a low ATP-requiring tissue. However, this could result in severe mitochondrial disease should that blastomere develop into a tissue highly dependent on OX-PHOS. It is therefore unlikely that PGD would be a suitable technique to determine the risk of mitochondrial disease following oocyte reconstruction, especially following GVT, PNT and NT as all have reported preferential replication of donor mtDNA. Consequently, an alternative is required. Conclusion Mitochondria are vitally important during embryonic development and tissue differentiation, as demonstrated by the changes in their localization, activity and morphology in the early embryo. The propagation of mitochondria before metaphase II and the maternal transmission of mtDNA ensure that sufficient numbers of the organelle and its genome are invested into the process of early embryogenesis. Consequently, these pre-metaphase II events are under the control of the oocyte nuclear genome, which regulates transcription and replication, and ensure that transmission is not affected during a period when there is no nuclear control, i.e. before activation of the embryonic genome. In the process of trying to treat various diseases through the generation of patient-specific stem cells and the use of GVT and PNT to prevent the transmission of mtDNA-type disease, we would be taking inappropriate risks if current strategies were introduced into the clinic. In many ways, our strategies are confounding and our ingenuity could turn into our downfall. We are attempting to provide the nucleus with a new environment in which it can prosper but, for GVT and PNT, we have forgotten that this karyoplast is still accompanied by the mutant mtDNA that we are trying to eradicate. For NT-derived patient-specific stem cells, GVT and PNT, we are asking the transferred nucleus to interact with a cytoplasm which may or may not be compatible. We also expect two or more populations of mtDNA to happily cohabit. We are however ignoring how the cytoplasm will adjust to its new nuclear companion in favour of concentrating on the nucleus and its requirement to achieve pluripotency. Cohabitation requires considerable foresight, and in this respect we need to understand considerably more about nucleo-mitochondrial harmony before we exploit such technologies. The question still remaining to be answered is: Can heteroplasmy be tolerated at a functional level even if it compromises genetic identity? Paradoxically, inheritance and therefore analysis of recipient oocyte mtDNA does, however, allow us to confirm whether oocyte reconstruction has taken place. This is a lesson we have hopefully learnt from the discredited Korean attempts at generating patient-specific ESCs. References Alam TI, Kanki T, Muta T, Ukaji K, Abe Y, Nakayama H, Takio K, Hamasaki N and Kang D ( 2003) Human mitochondrial DNA is packaged with TFAM. Nucleic Acids Res  31, 1640–1645. CrossRef Search ADS PubMed  Almeida-Santos T, El Shourbagy S and St John JC ( 2006) Mitochondrial content reflects oocyte variability and fertilization outcome. Fertil Steril 85, 584–591. Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F et al. ( 1981) Sequence and organization of the human mitochondrial genome. Nature  290, 457–465. CrossRef Search ADS PubMed  Ankel-Simons F and Cummins JM ( 1996) Misconceptions about mitochondria and mammalian fertilization: implications for theories on human evolution. Proc Natl Acad Sci USA  93, 13859–13863. CrossRef Search ADS PubMed  Barritt JA, Brenner CA, Malter HE and Cohen J ( 2001) Rebuttal: interooplasmic transfers in humans. Reprod Biomed Online  3, 47–48. CrossRef Search ADS PubMed  Bavister BD and Squirrell JM ( 2000) Mitochondrial distribution and function in oocytes and early embryos. Hum Reprod  15, 189–198. CrossRef Search ADS PubMed  Blok RB, Gook DA, Thorburn DR and Dahl HH ( 1997) Skewed segregation of the mtDNA nt 8993 (T –-> G) mutation in human oocytes. Am J Hum Genet  60, 1495–1501. CrossRef Search ADS PubMed  Boulet L, Karpati G and Shoubridge EA ( 1992) Distribution and threshold expression of the tRNA (Lys) mutation in skeletal muscle of patients with myoclonic epilepsy and ragged-red fibers (MERRF). Am J Hum Genet  51, 1187–1200. PubMed  Brenner CA, Barritt JA, Willadsen S and Cohen J ( 2000) Mitochondrial DNA heteroplasmy after human ooplasmic transplantation. Fertil Steril  74, 573–578. CrossRef Search ADS PubMed  Brison DR and Leese HJ ( 1991) Energy metabolism in late preimplantation rat embryos. J Reprod Fertil  93, 245–251. CrossRef Search ADS PubMed  Brown DR, Koehler CM, Lindberg GL, Freeman AE, Mayfield JE, Myers AM, Schutz MM and Beitz DC ( 1989) Molecular analysis of cytoplasmic genetic variation in Holstein cows. J Anim Sci  67, 1926–1932. CrossRef Search ADS PubMed  Brown MD, Voljavec AS, Lott MT, Torroni A, Yang CC and Wallace DC ( 1992) Mitochondrial DNA complex I and III mutations associated with Leber’s hereditary optic neuropathy. Genetics  130, 163–173. PubMed  Campbell K, Alberio R, Choi I, Fisher P, Kelly R, Lee J-H and Maalouf W ( 2005) Cloning: eight years after Dolly. Reprod Domest Anim  40, 256–268. CrossRef Search ADS PubMed  Chang KH, Lim JM, Kang SK, Lee BC, Moon SY and Hwang WS ( 2003) Blastocyst formation, karyotype, and mitochondrial DNA of interspecies embryos derived from nuclear transfer of human cord fibroblasts into enucleated bovine oocytes. Fertil Steril  80, 1380–1387. CrossRef Search ADS PubMed  Chen DY, Wen DC, Zhang YP, Sun QY, Han ZM, Liu ZH, Shi P, Li JS, Xiangyu JG, Lian L et al. ( 2002) Interspecies implantation and mitochondria fate of panda-rabbit cloned embryos. Biol Reprod  67, 637–642. CrossRef Search ADS PubMed  Chen Y, He ZX, Liu A, Wang K, Mao WW, Chu JX, Lu Y, Fang ZF, Shi YT, Yang QZ et al. ( 2003) Embryonic stem cells generated by nuclear transfer of human somatic nuclei into rabbit oocytes. Cell Res  13, 251–263. CrossRef Search ADS PubMed  Chinnery PF, Andrews RM, Turnbull DM and Howell NN ( 2001) Leber hereditary optic neuropathy: Does heteroplasmy influence the inheritance and expression of the G11778A mitochondrial DNA mutation? Am J Med Genet  98, 235–243. CrossRef Search ADS PubMed  Chung Y, Mann MRW, Bartolomei MS and Latham KE ( 2002) Nuclear-cytoplasmic tug of war during cloning: effects of somatic cell nuclei on culture medium preferences of preimplantation cloned mouse embryos. Biol Reprod  66, 1178–1184. CrossRef Search ADS PubMed  Cibelli JB, Campbell KH, Seidel GE, West MD and Lanza RP ( 2002) The health profile of cloned animals. Nat Biotechnol  20, 13–14. CrossRef Search ADS PubMed  Cohen J, Scott R, Schimmel T, Levron J and Willadsen S ( 1997) Birth of infant after transfer of anucleate donor oocyte cytoplasm into recipient eggs. Lancet  350, 186–187. CrossRef Search ADS PubMed  Cortesini R ( 2005) Stem cells, tissue engineering and organogenesis in transplantation. Transpl Immunol  15, 81–89. CrossRef Search ADS PubMed  Cui LB, Huang XY and Sun FZ ( 2005) Transfer of germinal vesicle to ooplasm of young mice could not rescue ageing-associated chromosome misalignment in meiosis of oocytes from aged mice. Hum Reprod  20, 1624–1631. CrossRef Search ADS PubMed  Cummins J ( 1998) Mitochondrial DNA in mammalian reproduction. Rev Reprod  3, 172–182. CrossRef Search ADS PubMed  Cummins JM, Wakayama T and Yanagimachi R ( 1997) Fate of microinjected sperm components in the mouse oocyte and embryo. Zygote  5, 301–308. CrossRef Search ADS PubMed  D’Aurelio M, Gajewski CD, Lin MT, Mauck WM, Shao LZ, Lenaz G, Moraes CT and Manfredi G ( 2004) Heterologous mitochondrial DNA recombination in human cells. Hum Mol Genet  13, 3171–3179. CrossRef Search ADS PubMed  Dailey T, Dale B, Cohen J and Munne S ( 1996) Association between nondisjunction and maternal age in meiosis-II human oocytes. Am J Hum Genet  59, 176–184. PubMed  Danan C, Sternberg D, Van Steirteghem A, Cazeneuve C, Duquesnoy P, Besmond C, Goossens M, Lissens W and Amselem S ( 1999) Evaluation of parental mitochondrial inheritance in neonates born after intracytoplasmic sperm injection. Am J Hum Genet  65, 463–473. CrossRef Search ADS PubMed  Davis AF and Clayton DA ( 1996) In situ localization of mitochondrial DNA replication in intact mammalian cells. J Cell Biol  135, 883–893. CrossRef Search ADS PubMed  De Block CE, De Leeuw IH, Maassen JA, Ballaux D and Martin JJ ( 2004) A novel 7301-bp deletion in mitochondrial DNA in a patient with Kearns–Sayre syndrome, diabetes mellitus, and primary amenorrhoea. Exp Clin Endocrinol Diabetes  112, 80–83. CrossRef Search ADS PubMed  Dean NL, Battersby BJ, Ao A, Gosden RG, Tan SL, Shoubridge EA and Molnar MJ ( 2003) Prospect of preimplantation genetic diagnosis for heritable mitochondrial DNA diseases. Mol Hum Reprod  9, 631–638. CrossRef Search ADS PubMed  Dumollard R, Hammar K, Porterfield M, Smith PJ, Cibert C, Rouviere C and Sardet C ( 2003) Mitochondrial respiration and Ca2+ waves are linked during fertilization and meiosis completion. Development  130, 683–692. CrossRef Search ADS PubMed  Dunbar DR, Moonie PA, Jacobs HT and Holt IJ ( 1995) Different cellular backgrounds confer a marked advantage to either mutant or wild-type mitochondrial genomes. Proc Natl Acad Sci USA  92, 6562–6566. CrossRef Search ADS PubMed  Eichenlaub-Ritter U, Vogt E, Yin H and Gosden R ( 2004) Spindles, mitochondria and redox potential in ageing oocytes. Reprod Biomed Online  8, 45–58. CrossRef Search ADS PubMed  Eisen A and Reynolds GT ( 1985) Source and sinks for the calcium released during fertilization of single sea urchin eggs. J Cell Biol  100, 1522–1527. CrossRef Search ADS PubMed  El Shourbagy SH, Spikings EC, Freitas M and St. John JC ( 2006) Mitochondria directly influence fertilisation outcome in the pig. Reproduction  131, 233–245. CrossRef Search ADS PubMed  Falkenberg M, Gaspari M, Rantanen A, Trifunovic A, Larsson NG and Gustafsson CM ( 2002) Mitochondrial transcription factors B1 and B2 activate transcription of human mtDNA. Nat Genet  31, 289–294. CrossRef Search ADS PubMed  Fisher RP and Clayton DA ( 1985) A transcription factor required for promoter recognition by human mitochondrial RNA polymerase. Accurate initiation at the heavy- and light-strand promoters dissected and reconstituted in vitro. J Biol Chem  260, 11330–11338. PubMed  Fisher RP and Clayton DA ( 1988) Purification and characterization of human mitochondrial transcription factor 1. Mol Cell Biol  8, 3496–3509. CrossRef Search ADS PubMed  Fulka H ( 2004) Distribution of mitochondria in reconstructed mouse oocytes. Reproduction  127, 195–200. CrossRef Search ADS PubMed  Gahan ME, Miller F, Lewin SR, Cherry CL, Hoy JF, Mijch A, Rosenfeldt F and Wesselingh SL ( 2001) Quantification of mitochondrial DNA in peripheral blood mononuclear cells and subcutaneous fat using real-time polymerase chain reaction. J Clin Virol  22, 241–247. CrossRef Search ADS PubMed  Giles RE, Blanc H, Cann HM and Wallace DC ( 1980) Maternal inheritance of human mitochondrial DNA. Proc Natl Acad Sci USA  77, 6715–6719. CrossRef Search ADS PubMed  Graziewicz MA, Longley MJ, Bienstock RJ, Zeviani M and Copeland WC ( 2004) Structure-function defects of human mitochondrial DNA polymerase in autosomal dominant progressive external ophthalmoplegia. Nat Struct Mol Biol  11, 770–776. CrossRef Search ADS PubMed  Gyllensten U, Wharton D, Josefsson A and Wilson AC ( 1991) Paternal inheritance of mitochondrial DNA in mice. Nature  352, 255–257. CrossRef Search ADS PubMed  Hammer CJ, Tyler HD, Loskutoff NM, Armstrong DL, Funk DJ, Lindsey BR and Simmons LG ( 2001) Compromised development of calves (Bos gaurus) derived from vitro-generated embryos transferred interspecifically into domestic cattle (Bos taurus). Theriogenology  55, 1447–1455. CrossRef Search ADS PubMed  Han ZM, Chen DY, Li JS, Sun QY, Wan QH, Kou ZH, Rao G, Lei L, Liu ZH and Fang SG ( 2004) Mitochondrial DNA heteroplasmy in calves cloned by using adult somatic cell. Mol Reprod Dev  67, 207–214. CrossRef Search ADS PubMed  Hance N, Ekstrand MI and Trifunovic A ( 2005) Mitochondrial DNA polymerase gamma is essential for mammalian embryogenesis. Hum Mol Genet  14, 1775–1783. CrossRef Search ADS PubMed  Hanna MG, Nelson IP, Morgan-Hughes JA and Harding AE ( 1995) Impaired mitochondrial translation in human myoblasts harbouring the mitochondrial DNA tRNA lysine 8344 A→G (MERRF) mutation: relationship to proportion of mutant mitochondrial DNA. J Neurol Sci  130, 154–160. CrossRef Search ADS PubMed  Hauswirth WW and Laipis PJ ( 1982) Mitochondrial DNA polymorphism in a maternal lineage of Holstein cows. Proc Natl Acad Sci USA  79, 4686–4690. CrossRef Search ADS PubMed  He Z, Liu HC and Rosenwaks Z ( 2003) Cryopreservation of nuclear material as a potential method of fertility preservation. Fertil Steril  79, 347–354. CrossRef Search ADS PubMed  Hecht NB and Liem H ( 1984) Mitochondrial DNA is synthesized during meiosis and spermiogenesis in the mouse. Exp Cell Res  154, 293–298. CrossRef Search ADS PubMed  Hiendleder S, Schmutz SM, Erhardt G, Green RD and Plante Y ( 1999) Transmitochondrial differences and varying levels of heteroplasmy in nuclear transfer cloned cattle. Mol Reprod Dev  54, 24–31. CrossRef Search ADS PubMed  Hiendleder S, Zakhartchenko V, Wenigerkind H, Reichenbach HD, Bruggerhoff K, Prelle K, Brem G, Stojkovic M and Wolf E ( 2003) Heteroplasmy in bovine fetuses produced by intra- and inter-subspecific somatic cell nuclear transfer: neutral segregation of nuclear donor mitochondrial DNA in various tissues and evidence for recipient cow mitochondria in fetal blood. Biol Reprod  68, 159–166. CrossRef Search ADS PubMed  Hiendleder S, Bebbere D, Zakhartchenko V, Reichenbach HD, Wenigerkind H, Ledda S and Wolf E ( 2004) Maternal-fetal transplacental leakage of mitochondrial DNA in bovine nuclear transfer pregnancies: potential implications for offspring and recipients. Cloning Stem Cells  6, 150–156. CrossRef Search ADS PubMed  Holden C ( 2005) Stem cell research. Korean cloner admits lying about oocyte donations. Science  310, 1402–1403. CrossRef Search ADS PubMed  Houghton FD, Thompson JG, Kennedy CJ and Leese HJ ( 1996) Oxygen consumption and energy metabolism of the early mouse embryo. Mol Reprod Dev  44, 476–485. CrossRef Search ADS PubMed  Human Fertilisation and Embryology Act (c. 37) ( 1990) Available from http://www.opsi.gov.uk/acts/acts1990/Ukpga_19900037_en_1.htm#tcon. Google Scholar Humpherys D, Eggan K, Akutsu H, Friedman A, Hochedlinger K, Yanagimachi R, Lander ES, Golub TR and Jaenisch R ( 2002) Abnormal gene expression in cloned mice derived from embryonic stem cell and cumulus cell nuclei. Proc Natl Acad Sci USA  99, 12889–12894. CrossRef Search ADS PubMed  Inoue K, Ogonuki N, Yamamoto Y, Takano K, Miki H, Mochida K and Ogura A ( 2004) Tissue-specific distribution of donor mitochondrial DNA in cloned mice produced by somatic cell nuclear transfer. Genesis  39, 79–83. CrossRef Search ADS PubMed  Irwin MH, Johnson LW and Pinkert CA ( 1999) Isolation and microinjection of somatic cell derived mitochondria and germline heteroplasmy in transmitochondrial mice. Transgenic Res  8, 119–123. CrossRef Search ADS PubMed  Jansen RP and de Boer K ( 1998) The bottleneck: mitochondrial imperatives in oogenesis and ovarian follicular fate. Mol Cell Endocrinol  145, 81–88. CrossRef Search ADS PubMed  Kaneda H, Hayashi J, Takahama S, Taya C, Lindahl KF and Yonekawa H ( 1995) Elimination of paternal mitochondrial DNA in intraspecific crosses during early mouse embryogenesis. Proc Natl Acad Sci USA  92, 4542–4546. CrossRef Search ADS PubMed  Keefe DL, Niven-Fairchild T, Powell S and Buradagunta S ( 1995) Mitochondrial deoxyribonucleic acid deletions in oocytes and reproductive aging in women. Fertil Steril  64, 577–583. CrossRef Search ADS PubMed  Kenyon L and Moraes CT ( 1997) Expanding the functional human mitochondrial DNA database by the establishment of primate xenomitochondrial cybrids. Proc Natl Acad Sci USA  94, 9131–9135. CrossRef Search ADS PubMed  King MP and Attardi G ( 1989) Human cells lacking mtDNA: repopulation with exogenous mitochondria by complementation. Science  246, 500–503. CrossRef Search ADS PubMed  Kirches E, Michael M, Warich-Kirches M, Schneider T, Weis S, Krause G, Mawrin C and Dietzmann K ( 2001) Heterogeneous tissue distribution of a mitochondrial DNA polymorphism in heteroplasmic subjects without mitochondrial disorders. J Med Genet  38, 312–317. CrossRef Search ADS PubMed  Klopstock T, Bischof F, Gerok K, Deuschl G, Seibel P, Ketelsen UP and Reichmann H ( 1995) 3.1-kb deletion of mitochondrial DNA in a patient with Kearns–Sayre syndrome. Acta Neuropathol  90, 126–129. CrossRef Search ADS PubMed  Koehler CM, Lindberg GL, Brown DR, Beitz DC, Freeman AE, Mayfield JE and Myers AM ( 1991) Replacement of bovine mitochondrial DNA by a sequence variant within one generation. Genetics  129, 247–255. PubMed  Kondo R, Satta Y, Matsuura ET, Ishiwa H, Takahata N and Chigusa SI ( 1990) Incomplete maternal transmission of mitochondrial DNA in Drosophila. Genetics  126, 657–663. PubMed  Kondo R, Matsuura ET and Chigusa SI ( 1992) Further observation of paternal transmission of Drosophila mitochondrial DNA by PCR selective amplification method. Genet Res  59, 81–84. CrossRef Search ADS PubMed  Kong LH, Liu Z, Li H, Zhu L and Xing FQ ( 2003) Pregnancy in a 46-year-old woman after autologous granular cell mitochondria transfer. Di Yi Jun Yi Da Xue Xue Bao  23, 743–747. Kong LH, Liu Z, Li H, Zhu L, Chen SM, Chen SL and Xing FQ ( 2004) Mitochondria transfer from self-granular cells to improve embryos’ quality. Zhonghua Fu Chan Ke Za Zhi  39, 105–107. PubMed  Kraytsberg Y, Schwartz M, Brown TA, Ebralidse K, Kunz WS, Clayton DA, Vissing J and Khrapko K ( 2004) Recombination of human mitochondrial DNA. Science  304, 981. CrossRef Search ADS PubMed  Kvist L, Martens J, Nazarenko AA and Orell M ( 2003) Paternal leakage of mitochondrial DNA in the great tit (Parus major) Mol Biol Evol  20, 243–247. CrossRef Search ADS PubMed  Laflamme MA, Gold J, Xu C, Hassanipour M, Rosler E, Police S, Muskheli V and Murry CE ( 2005) Formation of human myocardium in the rat heart from human embryonic stem cells. Am J Pathol  167, 663–671. CrossRef Search ADS PubMed  Larsson N-G, Garman JD, Oldfors A, Barsh GS and Clayton DA ( 1996) A single mouse gene encodes the mitochondrial transcription factor A and a testis-specific nuclear HMG-box protein. Nat Genet  13, 296–302. CrossRef Search ADS PubMed  Larsson NG, Oldfors A, Holme E and Clayton DA ( 1994) Low levels of mitochondrial transcription factor A in mitochondrial DNA depletion. Biochem Biophys Res Commun  200, 1374–1381. CrossRef Search ADS PubMed  Larsson N, Oldfors A, Garman J, Barsh G and Clayton D ( 1997) Down-regulation of mitochondrial transcription factor A during spermatogenesis in humans. Hum Mol Genet  6, 185–191. CrossRef Search ADS PubMed  Larsson NG, Wang J, Wilhelmsson H, Oldfors A, Rustin P, Lewandoski M, Barsh GS and Clayton DA ( 1998) Mitochondrial transcription factor A is necessary for mtDNA maintenance and embryogenesis in mice. Nat Genet  18, 231–236. CrossRef Search ADS PubMed  Leese HJ, Conaghan J, Martin KL and Hardy K ( 1993) Early human embryo metabolism. Bioessays  15, 259–264. CrossRef Search ADS PubMed  Lertrit P, Imsumran A, Karnkirawattana P, Devahasdin V, Sangruchi T, Atchaneeyasakul L, Mungkornkarn C and Neungton N ( 1999) A unique 3.5-kb deletion of the mitochondrial genome in Thai patients with Kearns–Sayre syndrome. Hum Genet  105, 127–131. CrossRef Search ADS PubMed  Lestienne P, Reynier P, Chretien MF, Penisson-Besnier I, Malthiery Y and Rohmer V ( 1997) Oligoasthenospermia associated with multiple mitochondrial DNA rearrangements. Mol Hum Reprod  3, 811–814. CrossRef Search ADS PubMed  Levy SE, Waymire KG, Kim YL, MacGregor GR and Wallace DC ( 1999) Transfer of chloramphenicol-resistant mitochondrial DNA into the chimeric mouse. Transgenic Res  8, 137–145. CrossRef Search ADS PubMed  Lin DP, Huang CC, Wu HM, Cheng TC, Chen CI and Lee MS ( 2004) Comparison of mitochondrial DNA contents in human embryos with good or poor morphology at the 8-cell stage. Fertil Steril  81, 73–79. CrossRef Search ADS PubMed  Liu H, Wang CW, Grifo JA, Krey LC and Zhang J ( 1999) Reconstruction of mouse oocytes by germinal vesicle transfer: maturity of host oocyte cytoplasm determines meiosis. Hum Reprod  14, 2357–2361. CrossRef Search ADS PubMed  Liu L, Hammar K, Smith PJ, Inoue S and Keefe DL ( 2001) Mitochondrial modulation of calcium signaling at the initiation of development. Cell Calcium  30, 423–433. CrossRef Search ADS PubMed  Liu SZ, Zhou ZM, Chen T, Zhang YL, Wen DC, Kou ZH, Li ZD, Sun QY and Chen DY ( 2004) Blastocysts produced by nuclear transfer between chicken blastodermal cells and rabbit oocytes. Mol Reprod Dev  69, 296–302. CrossRef Search ADS PubMed  Lloyd RE, Lee JH, Alberio R, Bowles EJ, Ramalho-Santos J, Campbell KH and St. John JC ( 2006) Aberrant nucleo-cytoplasmic cross-talk results in donor cell mtDNA persistence in cloned embryos. Genetics. Epub ahead of print February 1, 2006. Ludwig M, Schroder AK and Diedrich K ( 2001) Impact of intracytoplasmic sperm injection on the activation and fertilization process of oocytes. Reprod Biomed Online  3, 230–240. CrossRef Search ADS PubMed  Luetjens CM, Payne C and Schatten G ( 1999) Non-random chromosome positioning in human sperm and sex chromosome anomalies following intracytoplasmic sperm injection. Lancet  353, 1240. CrossRef Search ADS PubMed  Machaty Z, Thompson JG, Abeydeera LR, Day BN and Prather RS ( 2001) Inhibitors of mitochondrial ATP production at the time of compaction improve development of in vitro produced porcine embryos. Mol Reprod Dev  58, 39–44. CrossRef Search ADS PubMed  Magnusson J, Orth M, Lestienne P and Taanman JW ( 2003) Replication of mitochondrial DNA occurs throughout the mitochondria of cultured human cells. Exp Cell Res  289, 133–142. CrossRef Search ADS PubMed  Mancuso M, Filosto M, Mootha VK, Rocchi A, Pistolesi S, Murri L, DiMauro S and Siciliano G ( 2004) A novel mitochondrial tRNAPhe mutation causes MERRF syndrome. Neurology  62, 2119–2121. CrossRef Search ADS PubMed  Marchington DR, Brown MSS, Lamb VK, van Golde RJ, Kremer JA, Tuerlings JH, Mariman EC, Balen AH and Poulton J ( 2002) No evidence for paternal mtDNA transmission to offspring or extra-embryonic tissues after ICSI. Mol Hum Reprod  8, 1046–1049. CrossRef Search ADS PubMed  Margineantu DH, Cox WG, Sundell L, Sherwood SW, Beechem JM and Capaldi RA ( 2002) Cell cycle dependent morphology changes and associated mitochondrial DNA redistribution in mitochondria of human cell lines. Mitochondrion  1, 425–435. CrossRef Search ADS PubMed  May-Panloup P, Chretien MF, Jacques C, Vasseur C, Malthiery Y and Reynier P ( 2005) Low oocyte mitochondrial DNA content in ovarian insufficiency. Hum Reprod  20, 593–597. CrossRef Search ADS PubMed  McConnell JM and Petrie L ( 2004) Mitochondrial DNA turnover occurs during preimplantation development and can be modulated by environmental factors. Reprod Biomed Online  9, 418–424. CrossRef Search ADS PubMed  McCreath KJ, Howcroft J, Campbell KH, Colman A, Schnieke AE and Kind AJ ( 2000) Production of gene-targeted sheep by nuclear transfer from cultured somatic cells. Nature  405, 1066–1069. CrossRef Search ADS PubMed  McKenzie M and Trounce I ( 2000) Expression of Rattus norvegicus mtDNA in Mus musculus cells results in multiple respiratory chain defects. J Biol Chem  275, 31514–31519. CrossRef Search ADS PubMed  McKenzie M, Chiotis M, Pinkert CA and Trounce IA ( 2003) Functional respiratory chain analyses in murid xenomitochondrial cybrids expose coevolutionary constraints of cytochrome b and nuclear subunits of complex III. Mol Biol Evol  20, 1117–1124. CrossRef Search ADS PubMed  McKenzie M, Trounce IA, Cassar CA and Pinkert CA ( 2004) Production of homoplasmic xenomitochondrial mice. Proc Natl Acad Sci USA  101, 1685–1690. CrossRef Search ADS PubMed  Meirelles FV and Smith LC ( 1997) Mitochondrial genotype segregation in a mouse heteroplasmic lineage produced by embryonic karyoplast transplantation. Genetics  145, 445–451. PubMed  Meirelles FV and Smith LC ( 1998) Mitochondrial genotype segregation during preimplantation development in mouse heteroplasmic embryos. Genetics  148, 877–883. PubMed  Meirelles FV, Bordignon V, Watanabe Y, Watanabe M, Dayan A, Lobo RB, Garcia JM and Smith LC ( 2001) Complete replacement of the mitochondrial genotype in a Bos indicus calf reconstructed by nuclear transfer to a Bos taurus oocyte. Genetics  158, 351–356. PubMed  Meusel MS and Moritz RF ( 1993) Transfer of paternal mitochondrial DNA during fertilization of honeybee (Apis mellifera L.) eggs. Curr Genet  24, 539–543. CrossRef Search ADS PubMed  Michaels GS, Hauswirth WW and Laipis PJ ( 1982) Mitochondrial DNA copy number in bovine oocytes and somatic cells. Dev Biol  94, 246–251. CrossRef Search ADS PubMed  Monnat RJ Jr, Maxwell CL and Loeb LA ( 1985) Nucleotide sequence preservation of human leukemic mitochondrial DNA. Cancer Res  45, 1809–1814. PubMed  Moyes CD, Battersby BJ and Leary SC ( 1998) Regulation of muscle mitochondrial design. J Exp Biol  201, 299–307. Munne S, Alikani M, Tomkin G, Grifo J and Cohen J ( 1995) Embryo morphology, developmental rates, and maternal age are correlated with chromosome abnormalities. Fertil Steril  64, 382–391. CrossRef Search ADS PubMed  Munsie MJ, Michalska AE, O’Brien CM, Trounson AO, Pera MF and Mountford PS ( 2000) Isolation of pluripotent embryonic stem cells from reprogrammed adult mouse somatic cell nuclei. Curr Biol  10, 989–992. CrossRef Search ADS PubMed  Nagao Y, Totsuka Y, Atomi Y, Kaneda H, Lindahl KF, Imai H and Yonekawa H ( 1998) Decreased physical performance of congenic mice with mismatch between the nuclear and the mitochondrial genome. Genes Genet Syst  73, 21–27. CrossRef Search ADS PubMed  Naini AB, Lu J, Kaufmann P, Bernstein RA, Mancuso M, Bonilla E, Hirano M and DiMauro S ( 2005) Novel mitochondrial DNA ND5 mutation in a patient with clinical features of MELAS and MERRF. Arch Neurol  62, 473–476. CrossRef Search ADS PubMed  Nakamura M, Nakano S, Goto Y, Ozawa M, Nagahama Y, Fukuyama H, Akiguchi I, Kaji R and Kimura J ( 1995) A novel point mutation in the mitochondrial tRNA (Ser (UCN)) gene detected in a family with MERRF/MELAS overlap syndrome. Biochem Biophys Res Commun  214, 86–93. CrossRef Search ADS PubMed  Naviaux RK, Nyhan WL, Barshop BA, Poulton J, Markusic D, Karpinski NC and Haas RH ( 1999) Mitochondrial DNA polymerase gamma deficiency and mtDNA depletion in a child with Alpers’ syndrome. Ann Neurol  45, 54–58. CrossRef Search ADS PubMed  Nishi Y, Takeshita T, Sato K and Araki T ( 2003) Change of the mitochondrial distribution in mouse ooplasm during in vitro maturation. J Nippon Med Sch  70, 408–415. CrossRef Search ADS PubMed  Nishigaki Y, Tadesse S, Bonilla E, Shungu D, Hersh S, Keats BJ, Berlin CI, Goldberg MF, Vockley J, DiMauro S et al. ( 2003) A novel mitochondrial tRNA (Leu (UUR)) mutation in a patient with features of MERRF and Kearns–Sayre syndrome. Neuromuscul Disord  13, 334–340. CrossRef Search ADS PubMed  Nishimura Y, Yoshinari T, Naruse K, Yamada T, Sumi K, Mitani H, Higashiyama T and Kuroiwa T ( 2006) Active digestion of sperm mitochondrial DNA in single living sperm revealed by optical tweezers. Proc Natl Acad Sci USA  103, 1382–1387. CrossRef Search ADS PubMed  Normile D and Mann C ( 2005) Asian countries permit research, with safeguards. Science  307, 664. CrossRef Search ADS PubMed  Office of Legislative Policy and Analysis ( 2003) Human Cloning Prohibition Act of 2003. Available from http://olpa.od.nih.gov/tracking/house_bills/session1/hr-534.asp. Google Scholar Ojala D, Montoya J and Attardi G ( 1981) tRNA punctuation model of RNA processing in human mitochondria. Nature  290, 470–474. CrossRef Search ADS PubMed  Ono T, Isobe K, Nakada K and Hayashi JI ( 2001) Human cells are protected from mitochondrial dysfunction by complementation of DNA products in fused mitochondria. Nat Genet  28, 272–275. CrossRef Search ADS PubMed  Palermo GD, Takeuchi T and Rosenwaks Z ( 2002) Technical approaches to correction of oocyte aneuploidy. Hum Reprod  17, 2165–2173. CrossRef Search ADS PubMed  Pfeiffer T, Schuster S and Bonhoeffer S ( 2001) Cooperation and competition in the evolution of ATP-producing pathways. Science  292, 504–507. CrossRef Search ADS PubMed  Piko L and Matsumoto L ( 1976) Number of mitochondria and some properties of mitochondrial DNA in the mouse egg. Dev Biol  49, 1–10. CrossRef Search ADS PubMed  Piko L and Taylor KD ( 1987) Amounts of mitochondrial DNA and abundance of some mitochondrial gene transcripts in early mouse embryos. Dev Biol  123, 364–374. CrossRef Search ADS PubMed  Ponzetto C, Bresolin N, Bordoni A, Moggio M, Meola G, Bet L, Prelle A and Scarlato G ( 1990) Kearns–Sayre syndrome: different amounts of deleted mitochondrial DNA are present in several autoptic tissues. J Neurol Sci  96, 207–210. CrossRef Search ADS PubMed  Poulton J, Morten K, Freeman-Emmerson C, Potter C, Sewry C, Dubowitz V, Kidd H, Stephenson J, Whitehouse W, Hansen FJ et al. ( 1994) Deficiency of the human mitochondrial transcription factor h-mtTFA in infantile mitochondrial myopathy is associated with mtDNA depletion. Hum Mol Genet  3, 1763–1769. CrossRef Search ADS PubMed  Reynier P, Chretien MF, Penisson-Besnier I, Malthiery Y, Rohmer V and Lestienne P ( 1997) Male infertility associated with multiple mitochondrial DNA rearrangements. C R Acad Sci III  320, 629–636. CrossRef Search ADS PubMed  Reynier P, May-Panloup P, Chretien MF, Morgan CJ, Jean M, Savagner F, Barriere P and Malthiery Y ( 2001) Mitochondrial DNA content affects the fertilizability of human oocytes. Mol Hum Reprod  7, 425–429. CrossRef Search ADS PubMed  Rideout WM 3rd, Hochedlinger K, Kyba M, Daley GQ and Jaenisch R ( 2002) Correction of a genetic defect by nuclear transplantation and combined cell and gene therapy. Cell  109, 17–27. CrossRef Search ADS PubMed  Rubart M and Field LJ ( 2006) Cardiac repair by embryonic stem-derived cells. Handb Exp Pharmacol  174, 73–100. PubMed  Ruiz-Pesini E, Lapena AC, Diez-Sanchez C, Perez-Martos A, Montoya J, Alvarez E, Diaz M, Urries A, Montoro L, Lopez-Perez MJ et al. ( 2000) Human mtDNA haplogroups associated with high or reduced spermatozoa motility. Am J Hum Genet  67, 682–696. CrossRef Search ADS PubMed  Sathananthan AH and Trounson AO ( 2000) Mitochondrial morphology during preimplantational human embryogenesis. Hum Reprod  15, 148–159. CrossRef Search ADS PubMed  Sathananthan H, Pera M and Trounson A ( 2002) The fine structure of human embryonic stem cells. Reprod Biomed Online  4, 56–61. CrossRef Search ADS PubMed  Sato A, Kono T, Nakada K, Ishikawa K, Inoue SI, Yonekawa H and Hayashi JI ( 2005) Gene therapy for progeny of mito-mice carrying pathogenic mtDNA by nuclear transplantation. Proc Natl Acad Sci USA  102, 16765–16770. CrossRef Search ADS PubMed  Satoh M and Kuroiwa T ( 1991) Organization of multiple nucleoids and DNA molecules in mitochondria of a human cell. Exp Cell Res  196, 137–140. CrossRef Search ADS PubMed  Sazer S and Sherwood SW ( 1990) Mitochondrial growth and DNA synthesis occur in the absence of nuclear DNA replication in fission yeast. J Cell Sci  97, 509–516. PubMed  Scarpulla RC ( 2002) Nuclear activators and coactivators in mammalian mitochondrial biogenesis. Biochim Biophys Acta  1576, 1–14. CrossRef Search ADS PubMed  Schon EA, Kim SH, Ferreira JC, Magalhaes P, Grace M, Warburton D and Gross SJ ( 2000) Chromosomal non-disjunction in human oocytes: is there a mitochondrial connection? Hum Reprod  15, 160–172. CrossRef Search ADS PubMed  Schwartz M and Vissing J ( 2002) Paternal inheritance of mitochondrial DNA. N Engl J Med  347, 576–580. CrossRef Search ADS PubMed  Seifer DB, DeJesus V and Hubbard K ( 2002) Mitochondrial deletions in luteinized granulosa cells as a function of age in women undergoing in vitro fertilization. Fertil Steril  78, 1046–1048. CrossRef Search ADS PubMed  Sherratt HS ( 1991) Mitochondria: structure and function. Rev Neurol (Paris)  147, 417–430. PubMed  Shitara H, Kaneda H, Sato A, Inoue K, Ogura A, Yonekawa H and Hayashi JI ( 2000) Selective and continuous elimination of mitochondria microinjected into mouse eggs from spermatids, but not from liver cells, occurs throughout embryogenesis. Genetics  156, 1277–1284. PubMed  Shoubridge EA ( 2000) Mitochondrial DNA segregation in the developing embryo. Hum Reprod  15, 229–234. CrossRef Search ADS PubMed  Siciliano G, Mancuso M, Pasquali L, Manca ML, Tessa A and Iudice A ( 2000) Abnormal levels of human mitochondrial transcription factor A in skeletal muscle in mitochondrial encephalomyopathies. Neurol Sci  21,S 985–S987. CrossRef Search ADS   Smith LC and Alcivar AA ( 1993) Cytoplasmic inheritance and its effects on development and performance. J Reprod Fertil Suppl  48, 31–43. PubMed  Sokolova VA, Kustova ME, Arbuzova NI, Sorokin AV, Moskaliova OS, Bass MG and Vasilyev VB ( 2004) Obtaining mice that carry human mitochondrial DNA transmitted to the progeny. Mol Reprod Dev  68, 299–307. CrossRef Search ADS PubMed  Spelbrink JN, Van Galen MJ, Zwart R, Bakker HD, Rovio A, Jacobs HT and Van den Bogert C ( 1998) Familial mitochondrial DNA depletion in liver: haplotype analysis of candidate genes. Hum Genet  102, 327–331. CrossRef Search ADS PubMed  Spiropoulos J, Turnbull DM and Chinnery PF ( 2002) Can mitochondrial DNA mutations cause sperm dysfunction? Mol Hum Reprod  8, 719–721. CrossRef Search ADS PubMed  St. John JC ( 2002) Ooplasm donation in humans: the need to investigate the transmission of mitochondrial DNA following cytoplasmic transfer. Hum Reprod  17, 1954–1958. CrossRef Search ADS PubMed  St. John JC and Schatten G ( 2004) Paternal mitochondrial DNA transmission during nonhuman primate nuclear transfer. Genetics  167, 897–905. CrossRef Search ADS PubMed  St. John J, Sakkas D, Dimitriadi K, Barnes A, Maclin V, Ramey J, Barratt C and De Jonge C ( 2000) Failure of elimination of paternal mitochondrial DNA in abnormal embryos. Lancet  355, 200. CrossRef Search ADS   St. John JC, Jokhi RP and Barratt CL ( 2001) Men with oligoasthenoteratozoospermia harbour higher numbers of multiple mitochondrial DNA deletions in their spermatozoa, but individual deletions are not indicative of overall aetiology. Mol Hum Reprod  7, 103–111. CrossRef Search ADS PubMed  St. John JC, Lloyd R and El Shourbagy S ( 2004) The potential risks of abnormal transmission of mtDNA through assisted reproductive technologies. Reprod Biomed Online  8, 34–44. CrossRef Search ADS PubMed  St. John JC, Lloyd RE, Bowles EJ, Thomas EC and El Shourbagy S ( 2004) The consequences of nuclear transfer for mammalian foetal development and offspring survival. A mitochondrial DNA perspective. Reproduction  127, 631–641. CrossRef Search ADS PubMed  St. John JC, Ramalho-Santos J, Gray HL, Petrosko P, Rawe VY, Navara CS, Simerly CR and Schatten GP ( 2005) The expression of mitochondrial DNA transcription factors during early cardiomyocyte in vitro differentiation from human embryonic stem cells. Cloning Stem Cells  7, 141–153. CrossRef Search ADS PubMed  St. John J, Moffatt O and D’Souza N ( 2005) Aberrant heteroplasmic transmission of mtDNA in cloned pigs arising from double nuclear transfer. Mol Reprod Dev  72, 450–460. CrossRef Search ADS PubMed  Steinborn R, Zakhartchenko V, Jelyazkov J, Klein D, Wolf E, Muller M and Brem G ( 1998) Composition of parental mitochondrial DNA in cloned bovine embryos. FEBS Lett  426, 352–356. CrossRef Search ADS PubMed  Steinborn R, Schinogl P, Wells DN, Bergthaler A, Muller M and Brem G ( 2002) Coexistence of Bos taurus and B. indicus mitochondrial DNAs in nuclear transfer-derived somatic cattle clones. Genetics  162, 823–829. PubMed  Steuerwald N, Barritt JA, Adler R, Malter H, Schimmel T, Cohen J and Brenner CA ( 2000) Quantification of mtDNA in single oocytes, polar bodies and subcellular components by real-time rapid cycle fluorescence monitored PCR. Zygote  8, 209–215. CrossRef Search ADS PubMed  Stojanovski D, Johnston AJ, Streimann I, Hoogenraad NJ and Ryan MT ( 2003) Import of nuclear-encoded proteins into mitochondria. Exp Physiol  88, 57–64. CrossRef Search ADS PubMed  Stojkovic M, Machado SA, Stojkovic P, Zakhartchenko V, Hutzler P, Goncalves PB and Wolf E ( 2001) Mitochondrial distribution and adenosine triphosphate content of bovine oocytes before and after in vitro maturation: correlation with morphological criteria and developmental capacity after in vitro fertilization and culture. Biol Reprod  64, 904–909. CrossRef Search ADS PubMed  Sturmey RG and Leese HJ ( 2003) Energy metabolism in pig oocytes and early embryos. Reproduction  126, 197–204. CrossRef Search ADS PubMed  Sun QY, Wu GM, Lai L, Park KW, Cabot R, Cheong HT, Day BN, Prather RS and Schatten H ( 2001) Translocation of active mitochondria during pig oocyte maturation, fertilization and early embryo development in vitro. Reproduction  122, 155–163. CrossRef Search ADS PubMed  Sun YH, Chen SP, Wang YP, Hu W and Zhu ZY ( 2005) Cytoplasmic impact on cross-genus cloned fish derived from transgenic common carp (Cyprinus carpio) nuclei and goldfish (Carassius auratus) enucleated eggs. Biol Reprod  72, 510–515. CrossRef Search ADS PubMed  Suomalainen A and Kaukonen J ( 2001) Diseases caused by nuclear genes affecting mtDNA stability. Am J Med Genet  106, 53–61. CrossRef Search ADS PubMed  Sutarno, Cummins JM, Greeff J and Lymbery AJ ( 2002) Mitochondrial DNA polymorphisms and fertility in beef cattle. Theriogenology  57, 1603–1610. CrossRef Search ADS PubMed  Google Scholar Sutovsky P ( 2003) Ubiquitin-dependent proteolysis in mammalian spermatogenesis, fertilization, and sperm quality control: killing three birds with one stone. Microsc Res Tech  61, 88–102. CrossRef Search ADS PubMed  Sutovsky P, Moreno RD, Ramalho-Santos J, Dominko T, Simerly C and Schatten G ( 1999) Ubiquitin tag for sperm mitochondria. Nature  402, 371–372. CrossRef Search ADS PubMed  Takeda K, Akagi S, Kaneyama K, Kojima T, Takahashi S, Imai H, Yamanaka M, Onishi A and Hanada H ( 2003) Proliferation of donor mitochondrial DNA in nuclear transfer calves (Bos taurus) derived from cumulus cells. Mol Reprod Dev  64, 429–437. CrossRef Search ADS PubMed  Takeda K, Tasai M, Iwamoto M, Onishi A, Tagami T, Nirasawa K, Hanada H and Pinkert CA ( 2005) Microinjection of cytoplasm or mitochondria derived from somatic cells affects parthenogenetic development of murine oocytes. Biol Reprod  72, 1397–1404. CrossRef Search ADS PubMed  Takeda K, Tasai M, Iwamoto M, Akita T, Tagami T, Nirasawa K, Hanada H and Onishi A ( 2006) Transmission of mitochondrial DNA in pigs and progeny derived from nuclear transfer of Meishan pig fibroblast cells. Mol Reprod Dev  73, 306–312. CrossRef Search ADS PubMed  Takeuchi T, Gong J, Veeck LL, Rosenwaks Z and Palermo GD ( 2001) Preliminary findings in germinal vesicle transplantation of immature human oocytes. Hum Reprod  16, 730–736. CrossRef Search ADS PubMed  Takeuchi T, Neri QV, Katagiri Y, Rosenwaks Z and Palermo GD ( 2005) Effect of treating induced mitochondrial damage on embryonic development and epigenesis. Biol Reprod  72, 584–592. CrossRef Search ADS PubMed  Taylor H and Minger SL ( 2005) Regenerative medicine in Parkinson’s disease: generation of mesencephalic dopaminergic cells from embryonic stem cells. Curr Opin Biotechnol  16, 487–492. CrossRef Search ADS PubMed  Tessa A, Manca ML, Mancuso M, Renna MR, Murri L, Martini B, Santorelli FM and Siciliano G ( 2000) Abnormal H-Tfam in a patient harboring a single mtDNA deletion. Funct Neurol  15, 211–214. PubMed  Thompson JG, Partridge RJ, Houghton FD, Cox CI and Leese HJ ( 1996) Oxygen uptake and carbohydrate metabolism by in vitro derived bovine embryos. J Reprod Fertil  106, 299–306. CrossRef Search ADS PubMed  Tiranti V, Savoia A, Forti F, D’Apolito MF, Centra M, Rocchi M and Zeviani M ( 1997) Identification of the gene encoding the human mitochondrial RNA polymerase (h-mtRPOL) by cyberscreening of the Expressed Sequence Tags database. Hum Mol Genet  6, 615–625. CrossRef Search ADS PubMed  Trounce IA, McKenzie M, Cassar CA, Ingraham CA, Lerner CA, Dunn DA, Donegan CL, Takeda K, Pogozelski WK, Howell RL et al. ( 2004) Development and initial characterization of xenomitochondrial mice. J Bioenerg Biomembr  36, 421–427. CrossRef Search ADS PubMed  Trounson A ( 2001) Nuclear transfer in human medicine and animal breeding. Reprod Fertil Dev  13, 31–39. CrossRef Search ADS PubMed  Tsaousis AD, Martin DP, Ladoukakis ED, Posada D and Zouros E ( 2005) Widespread recombination in published animal mtDNA sequences. Mol Biol Evol  22, 925–933. CrossRef Search ADS PubMed  Vajta G, Lewis IM, Hyttel P, Thouas GA and Trounson AO ( 2001) Somatic cell cloning without micromanipulators. Cloning  3, 89–95. CrossRef Search ADS PubMed  Van Blerkom J, Davis PW and Lee J ( 1995) ATP content of human oocytes and developmental potential and outcome after in-vitro fertilization and embryo transfer. Hum Reprod  10, 415–424. PubMed  Van Blerkom J, Davis P and Alexander S ( 2000) Differential mitochondrial distribution in human pronuclear embryos leads to disproportionate inheritance between blastomeres: relationship to microtubular organization, ATP content and competence. Hum Reprod  15, 2621–2633. CrossRef Search ADS PubMed  Van Blerkom J, Davis P, Mathwig V and Alexander S ( 2002) Domains of high-polarized and low-polarized mitochondria may occur in mouse and human oocytes and early embryos. Hum Reprod  17, 393–406. CrossRef Search ADS PubMed  Van Goethem G, Dermaut B, Lofgren A, Martin JJ and Van Broeckhoven C ( 2001) Mutation of POLG is associated with progressive external ophthalmoplegia characterized by mtDNA deletions. Nat Genet  28, 211–212. CrossRef Search ADS PubMed  Wakayama T, Tabar V, Rodriguez I, Perry AC, Studer L and Mombaerts P ( 2001) Differentiation of embryonic stem cell lines generated from adult somatic cells by nuclear transfer. Science  292, 740–743. CrossRef Search ADS PubMed  Wallace DC ( 1992) Diseases of the mitochondrial DNA. Annu Rev Biochem  61, 1175–1212. CrossRef Search ADS PubMed  Wallace DC, Ye JH, Neckelmann SN, Singh G, Webster KA and Greenberg BD ( 1987) Sequence analysis of cDNAs for the human and bovine ATP synthase beta subunit: mitochondrial DNA genes sustain seventeen times more mutations. Curr Genet  12, 81–90. CrossRef Search ADS PubMed  Wernig M, Benninger F, Schmandt T, Rade M, Tucker KL, Bussow H, Beck H and Brustle O ( 2004) Functional integration of embryonic stem cell-derived neurons in vivo. J Neurosci  24, 5258–5268. CrossRef Search ADS PubMed  White KL, Bunch TD, Mitalipov S and Reed WA ( 1999) Establishment of pregnancy after the transfer of nuclear transfer embryos produced from the fusion of argali (Ovis ammon) nuclei into domestic sheep (Ovis aries) enucleated oocytes. Cloning  1, 47–54. CrossRef Search ADS PubMed  White DJ, Mital D, Taylor S and St John JC ( 2001) Sperm mitochondrial DNA deletions as a consequence of long term highly active antiretroviral therapy. Aids  15, 1061–1062. CrossRef Search ADS PubMed  Wiesner RJ, Ruegg JC and Morano I ( 1992) Counting target molecules by exponential polymerase chain reaction: copy number of mitochondrial DNA in rat tissues. Biochem Biophys Res Commun  183, 553–559. CrossRef Search ADS PubMed  Wilding M, Dale B, Marino M, di Matteo L, Alviggi C, Pisaturo ML, Lombardi L and De Placido G ( 2001) Mitochondrial aggregation patterns and activity in human oocytes and preimplantation embryos. Hum Reprod  16, 909–917. CrossRef Search ADS PubMed  Wilmut I, Schnieke AE, McWhir J, Kind AJ and Campbell KH ( 1997) Viable offspring derived from fetal and adult mammalian cells. Nature  385, 810–813. CrossRef Search ADS PubMed  Xu B and Clayton DA ( 1995) A persistent RNA-DNA hybrid is formed during transcription at a phylogenetically conserved mitochondrial DNA sequence. Mol Cell Biol  15, 580–589. CrossRef Search ADS PubMed  Yakes FM and Van Houten B ( 1997) Mitochondrial DNA damage is more extensive and persists longer than nuclear DNA damage in human cells following oxidative stress. Proc Natl Acad Sci USA  94, 514–519. CrossRef Search ADS PubMed  Yang CX, Han ZM, Wen DC, Sun QY, Zhang KY, Zhang LS, Wu YQ, Kou ZH and Chen DY ( 2003) In vitro development and mitochondrial fate of macaca-rabbit cloned embryos. Mol Reprod Dev  65, 396–401. CrossRef Search ADS PubMed  Yang L, Chavatte-Palmer P, Kubota C, O’Neill M, Hoagland T, Renard JP, Taneja M, Yang X and Tian C ( 2005) Expression of imprinted genes is aberrant in deceased newborn cloned calves and relatively normal in surviving adult clones. Mol Reprod Dev  71, 431–438. CrossRef Search ADS PubMed  Zhang J, Wang CW, Krey L, Liu H, Meng L, Blaszczyk A, Adler A and Grifo J ( 1999) In vitro maturation of human preovulatory oocytes reconstructed by germinal vesicle transfer. Fertil Steril  71, 726–731. CrossRef Search ADS PubMed  Zhao X, Li N, Guo W, Hu X, Liu Z, Gong G, Wang A, Feng J and Wu C ( 2004) Further evidence for paternal inheritance of mitochondrial DNA in the sheep (Ovis aries) Heredity  93, 399–403. Zouros E, Freeman KR, Ball AO and Pogson GH ( 1992) Direct evidence for extensive paternal mitochondrial DNA inheritance in the marine mussel Mytilus. Nature  359, 412–414. CrossRef Search ADS PubMed  © The Author 2006. Published by Oxford University Press on behalf of the European Society of Human Reproduction and Embryology. All rights reserved. For Permissions, please email: journals.permissions@oxfordjournals.org http://www.deepdyve.com/assets/images/DeepDyve-Logo-lg.png Human Reproduction Update Oxford University Press

Transmission of mitochondrial DNA following assisted reproduction and nuclear transfer

Loading next page...
 
/lp/oxford-university-press/transmission-of-mitochondrial-dna-following-assisted-reproduction-and-X2hR6ctrzc

References (208)

Publisher
Oxford University Press
Copyright
© The Author 2006. Published by Oxford University Press on behalf of the European Society of Human Reproduction and Embryology. All rights reserved. For Permissions, please email: journals.permissions@oxfordjournals.org
ISSN
1355-4786
eISSN
1460-2369
DOI
10.1093/humupd/dml011
pmid
16581809
Publisher site
See Article on Publisher Site

Abstract

Abstract Mitochondria are the organelles responsible for producing the majority of a cell’s ATP and also play an essential role in gamete maturation and embryo development. ATP production within the mitochondria is dependent on proteins encoded by both the nuclear and the mitochondrial genomes, therefore co-ordination between the two genomes is vital for cell survival. To assist with this co-ordination, cells normally contain only one type of mitochondrial DNA (mtDNA) termed homoplasmy. Occasionally, however, two or more types of mtDNA are present termed heteroplasmy. This can result from a combination of mutant and wild-type mtDNA molecules or from a combination of wild-type mtDNA variants. As heteroplasmy can result in mitochondrial disease, various mechanisms exist in the natural fertilization process to ensure the maternal-only transmission of mtDNA and the maintenance of homoplasmy in future generations. However, there is now an increasing use of invasive oocyte reconstruction protocols, which tend to bypass mechanisms for the maintenance of homoplasmy, potentially resulting in the transmission of either form of mtDNA heteroplasmy. Indeed, heteroplasmy caused by combinations of wild-type variants has been reported following cytoplasmic transfer (CT) in the human and following nuclear transfer (NT) in various animal species. Other techniques, such as germinal vesicle transfer and pronuclei transfer, have been proposed as methods of preventing transmission of mitochondrial diseases to future generations. However, resulting embryos and offspring may contain mtDNA heteroplasmy, which itself could result in mitochondrial disease. It is therefore essential that uniparental transmission of mtDNA is ensured before these techniques are used therapeutically. assisted reproduction technology, mitochondrial DNA, nuclear transfer Submitted on November 25, 2005; accepted on February 28, 2006 Introduction Mitochondria are found in all eukaryotic cells and are essential to cellular function because of their ability to generate ATP, the cell’s major source of energy. They comprise a double membrane with the outer membrane separating the mitochondrial matrix from the cytoplasm of the cell, whereas the inner membrane forms the cristae. Mitochondrial cristae are the site of the electron transfer chain (ETC), the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place (Sherratt, 1991). This aerobic process allows further metabolism of the products of anaerobic glycolysis and the citric acid cycle to produce carbon dioxide and water, with the subsequent release of 32 molecules of ATP. This is a far more economical return than provided by anaerobic generation of ATP (Pfeiffer et al., 2001). The ETC consists of five enzyme complexes, each containing multiple polypeptide subunits (Figure 1). Electrons flow through the first four complexes of the ETC releasing protons from the mitochondria which set up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in the generation of ATP. Figure 1. View largeDownload slide The electron transfer chain (ETC) is the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place. The ETC consists of five enzyme complexes which, with the exception of complex II, all consist of polypeptides encoded by both the nuclear and the mitochondrial genomes. Electrons flow through the first four complexes of the ETC releasing protons (H+) from the mitochondria and thereby setting up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in ATP generation. CoQ, Coenzyme Q; Cyt C, Cytochrome C; ANT, adenosine nucleoside transporter; KCN, potassium cyanide; ADP, adenosine diphosphate; ATP, adenosine triphosphate. Figure 1. View largeDownload slide The electron transfer chain (ETC) is the final stage in cellular respiration where oxidative phosphorylation (OX-PHOS) takes place. The ETC consists of five enzyme complexes which, with the exception of complex II, all consist of polypeptides encoded by both the nuclear and the mitochondrial genomes. Electrons flow through the first four complexes of the ETC releasing protons (H+) from the mitochondria and thereby setting up the mitochondrial membrane potential. Pumping these protons back into the mitochondria through the fifth complex results in ATP generation. CoQ, Coenzyme Q; Cyt C, Cytochrome C; ANT, adenosine nucleoside transporter; KCN, potassium cyanide; ADP, adenosine diphosphate; ATP, adenosine triphosphate. The importance of ATP in oocytes and early embryos has been demonstrated in cattle, as assessed by morphology. Here, the higher quality oocytes contained significantly higher levels of ATP than poorer quality oocytes (Stojkovic et al., 2001). Those containing the highest ATP levels also progressed to produce a significantly higher proportion of blastocysts. A similar pattern has been demonstrated in human oocytes, with those containing at least 2ρmol ATP at fertilization being more capable of embryonic development and implantation (Van Blerkom et al., 1995). The correlation between mitochondrial distribution in individual blastomeres and blastomere ATP content further emphasizes the importance of OX-PHOS for early preimplantation development (Van Blerkom et al., 2000). Mitochondrial DNA Components of the ETC are encoded by both chromosomal and mitochondrial DNA (mtDNA), with all components being required for efficient function of the ETC. The human mitochondrial genome is a double-stranded circular DNA consisting of 16 569 base pairs (Anderson et al., 1981). Using a slightly different genetic code to that of nuclear DNA, mtDNA encodes 13 polypeptides involved in the ETC, 22 transfer (t)RNAs and 2 ribosomal (r)RNAs (Anderson et al., 1981). MtDNA codes for seven subunits of nicotinamide adenine dinucleotide (NADH) dehydrogenase (ND; complex I), one component of complex III, three subunits of cytochrome C oxidase (complex IV) and two subunits of ATP synthase (complex V). All other components of the ETC are encoded by the nuclear genome (Figure 1). The mitochondrial genome is highly compact, with no introns contained between the coding regions (Anderson et al., 1981) and some of the coding regions even overlap (ATPase 6 and ATPase 8, ND4 and ND4L; Anderson et al., 1981). Others lack termination codons, with these sequences being provided by post-transcriptional polyadenylation (Ojala et al., 1981). Despite its small size in comparison with the nuclear genome, the host cell depends on the mitochondrial genome for aerobic respiration as much as the mitochondria rely on proteins encoded by the host cell’s nuclear DNA. In fact, as much as one third of the total DNA content of an oocyte consists of mtDNA (Piko and Matsumoto, 1976). Somatic cells have between 103 and 104 copies of mtDNA, with 2–10 copies per organelle (Satoh and Kuroiwa, 1991), correlating to each particular cell’s requirement for OX-PHOS. For example, it has been reported that human subcutaneous fat cells contain more mtDNA copies than peripheral blood mononuclear cells (Gahan et al., 2001) and that liver tissue contains more mtDNA than various muscle tissues (Wiesner et al., 1992). MtDNA transcription and replication Transcription of mtDNA occurs following interaction between nuclear-encoded regulatory proteins and regions within the d-loop of mtDNA (Figure 2). Requirements for transcription include the mitochondrial RNA polymerase (Tiranti et al., 1997), mitochondrial transcription factor A (TFAM; Fisher and Clayton, 1985; 1988), and one of the recently identified TFB1M and TFB2M (Falkenberg et al., 2002; Table I). Once initiated, transcription generates a polycistronic precursor RNA transcript, allowing co-ordinated transcription of all genes on the same strand. Excision of the polycistronic precursor by endonucleases produces precursor rRNAs and tRNAs, which are then processed further to allow them to translate the precursor mRNAs (Ojala et al., 1981). Figure 2. View largeDownload slide Mitochondrial DNA (mtDNA) transcription and replication origins in the d-loop. Mitochondrial transcription factor A (TFAM) binds and separates the two DNA strands. The accessory transcription factor, either TFB1M or TFB2M, allows mitochondrial RNA polymerase (mtRNAPol) to bind to the correct part of the DNA to produce the transcripts and the RNA primer for replication. Mitochondrial DNA polymerase gamma (PolG) is then able to use the primer to produce the daughter DNA strand. HSP, heavy strand promoter; LSP, light strand promoter; CSB, conserved sequence block; OH, origin of heavy strand replication; d-loop, displacement loop, the non-coding region that is the binding site for nuclear-encoded transcription and replication factors; Phe tRNA, phenylalanine transfer RNA; Pro tRNA, proline transfer RNA. Figure 2. View largeDownload slide Mitochondrial DNA (mtDNA) transcription and replication origins in the d-loop. Mitochondrial transcription factor A (TFAM) binds and separates the two DNA strands. The accessory transcription factor, either TFB1M or TFB2M, allows mitochondrial RNA polymerase (mtRNAPol) to bind to the correct part of the DNA to produce the transcripts and the RNA primer for replication. Mitochondrial DNA polymerase gamma (PolG) is then able to use the primer to produce the daughter DNA strand. HSP, heavy strand promoter; LSP, light strand promoter; CSB, conserved sequence block; OH, origin of heavy strand replication; d-loop, displacement loop, the non-coding region that is the binding site for nuclear-encoded transcription and replication factors; Phe tRNA, phenylalanine transfer RNA; Pro tRNA, proline transfer RNA. The replication of mtDNA is also dependent on nuclear-encoded regulatory proteins (Table I). The replication process begins with light strand transcription from the light strand promoter region to the second conserved sequence block (CSB) within the d-loop region, producing an RNA sequence complementary to the light strand DNA (Xu and Clayton, 1995). Binding of the mtDNA-specific polymerase gamma (PolG) to the resulting RNA–DNA hybrid allows replication of the heavy strand to begin from the origin of heavy strand replication (OH) contained within the d-Loop (Figure 2). After the replication of approximately two thirds of the genome, the formation of the new heavy strand exposes the origin of light strand replication (OL) on the heavy strand parent. Light strand replication is then able to commence in the opposite direction. Table I. Major nuclear-encoded regulatory factors required for mitochondrial DNA (mtDNA) transcription and replication Protein  Binding site  Role  TFAM  Enhancer region  Causes unwinding of mtDNA, exposing the binding site for mtRNA polymerase  TFB1M or TFB2M  RNA polymerase  Form heterodimer with RNA polymerase allowing specific transcription initiation  MtRNAPol  Transcription initiation region  Produces RNA copy of mtDNA sequence  PolG  RNA–DNA hybrid  Produces replicate copy of mtDNA  MtSSB  Single-stranded mtDNA  Prevents re-annealing of single-stranded mtDNA allowing PolG polymerization  MTERF  RNA polymerase  Termination of transcription  RNase MRP  MtDNA primary light strand transcript  Processing of primary light strand transcripts to produce RNA primer for replication  Protein  Binding site  Role  TFAM  Enhancer region  Causes unwinding of mtDNA, exposing the binding site for mtRNA polymerase  TFB1M or TFB2M  RNA polymerase  Form heterodimer with RNA polymerase allowing specific transcription initiation  MtRNAPol  Transcription initiation region  Produces RNA copy of mtDNA sequence  PolG  RNA–DNA hybrid  Produces replicate copy of mtDNA  MtSSB  Single-stranded mtDNA  Prevents re-annealing of single-stranded mtDNA allowing PolG polymerization  MTERF  RNA polymerase  Termination of transcription  RNase MRP  MtDNA primary light strand transcript  Processing of primary light strand transcripts to produce RNA primer for replication  MTERF, mitochondrial transcription termination factor; mtRNAPol, mitochondrial RNA polymerase; mtSSB, mitochondrial single-stranded binding protein; PolG, polymerase gamma; RNase MRP, mitochondrial RNA processing endoribonuclease; TFAM, mitochondrial transcription factor A; TFB1M/TFB2M, mitochondrial transcription factor B1 or B2. View Large As mtDNA transcription and replication, and therefore cellular ATP-generating capacity, are so dependent on proteins encoded by the nuclear genome, mutations in these genes can result in mtDNA depletion syndromes (reviewed by Suomalainen and Kaukonen, 2001). Unlike the vast majority of mtDNA diseases, these diseases can be transmitted in a Mendelian fashion and include chronic progressive external ophthalmoplegia (CPEO; Van Goethem et al., 2001; Graziewicz et al., 2004) and Alpers syndrome (Naviaux et al., 1999). This effect is also mimicked by nucleoside antiretroviral analogues such as 3′-azido-3′-deoxythymidine (AZT) and 2′3′-dideoxycytidine (ddC). For example, men treated with these drugs for 12 months had increased numbers of deleted mtDNA molecules in their sperm samples (White et al., 2001). Furthermore, although heterozygous PolG knockout mice develop normally, despite slightly reduced mtDNA levels, homozygous knockouts have severe mtDNA depletion and do not survive beyond embryonic day 8.5 (Hance et al., 2005). TFAM plays an important role in both transcription and replication of the mitochondrial genome. It is therefore not surprising that low levels of TFAM have been associated with a variety of mitochondrial depletion syndromes, such as infantile mitochondrial myopathy (Poulton et al., 1994), familial mtDNA-associated liver disease (Spelbrink et al., 1998), fatal childhood myopathy (Larsson et al., 1994), skeletal muscle and mitochondrial encephalomyopathy disorders (Siciliano et al., 2000) and ocular myopathy, exercise intolerance and muscle wasting (Tessa et al., 2000). In contrast, there is a decrease in TFAM expression localized to the mitochondria during spermatogenesis. This has been proposed as a functional mechanism for specifically reducing mtDNA copy number in maturing sperm and therefore paternal mtDNA transmission to future generations, in both the human (Larsson et al., 1997) and the mouse (Larsson et al., 1996). This arises at the late spermatocyte/early spermatid stage where an isoform is expressed lacking the mitochondrial targeting sequence. In the mouse, TFAM was also detected in the nucleus of developing sperm cells, indicating a possible further role as a nuclear transcription factor (Larsson et al., 1996). The significance of TFAM expression as a major regulator of mtDNA copy number is exemplified in homozygous knockout mouse embryos that are completely depleted of mtDNA and do not survive after embryonic day 10.5 (Larsson et al., 1998). The replication of embryonic mtDNA, at least in the mouse, does not occur during early preimplantation development (Piko and Taylor, 1987). However, a small window of replication at the two-cell stage was recently reported (McConnell and Petrie, 2004). The mtDNA molecules present in the mature oocyte at fertilization are therefore largely relied upon to provide sufficient ETC components to allow OX-PHOS activity during the preimplantation phase of embryonic development. In order to meet this demand, mtDNA is amplified during oocyte maturation within the ovary, with a 45-fold increase being reported in cattle from the primordial germ cells stage to pre-ovulating oocytes (Smith and Alcivar, 1993), resulting in a mean of 260 000 copies being present in mature bovine oocytes (Michaels et al., 1982). Similar amplifications have been reported in other species. For example, human premigratory primordial germ cells possess approximately 10 mitochondria per cell (Jansen and De Boer, 1998). Mature mouse oocytes contain mean mtDNA copy number values between 119 000 (Piko and Taylor, 1987) and 159 000 (Steuerwald et al., 2000). In the human, copy number ranges from 10 000 to 700 000 (Almeida-Santos et al., 2006), with other reported mean values of 314 000 (Steuerwald et al., 2000) and 193 000 (Reynier et al., 2001). Oocytes with higher copy numbers of mtDNA are known to be associated with improved fertilization rates. For example, human oocytes failing to fertilize because of poor sperm quality, and therefore assumed to be of good quality, had significantly higher mtDNA copy number than poor quality oocytes which were fertilized by good quality sperm (mean 255 000 versus 152 000 copies; Reynier et al., 2001). Furthermore, low-quality oocytes from patients with a variety of ovarian disorders contained an average of 100 000 mtDNA molecules compared with an average of 256 000 copies in oocytes from healthy individuals (May-Panloup et al., 2005). We have also recently reported similar results in porcine (El Shourbagy et al., 2006) and human (Almeida-Santos et al., 2006) oocytes. However, we also saw significantly different copy number between arrested two-cell human embryos and those that developed further (Almeida-Santos et al., 2006). A putative threshold number of mtDNA molecules has been proposed for the mature unfertilized oocyte to allow embryo development to take place after fertilization. In the mouse, this has been predicted at approximately 100 000 copies per oocyte (Piko and Taylor, 1987), with a similar value being appropriate for the pig (El Shourbagy et al., 2006). The relevance of mtDNA copy number threshold to fertilization outcome and embryo development The number of mtDNA copies per organelle is indicative of the number of mitochondria in an oocyte. In the mouse, each mitochondrion has just one copy (Piko and Taylor, 1987). Consequently, the minimum number of mtDNA copies would indicate the minimum number of mitochondria required for fertilization to take place. These mitochondria would support embryonic development to cavitation, after which glycolysis becomes increasingly important in many species (human – Leese et al., 1993; pig – Sturmey and Leese, 2003; cow – Thompson et al., 1996; rat – Brison and Leese, 1991; mouse – Houghton et al., 1996). Furthermore, this minimum number may be essential as it has been hypothesized that a threshold level of ATP is required to energize cell division (Brenner et al., 2000) and may also synchronize mitochondrial dynamics. The distribution of mitochondrial networks throughout the cell varies with different stages of the cell cycle (Margineantu et al., 2002). Mature and immature oocytes also have distinct patterns of mitochondrial distribution (Stojkovic et al., 2001; Nishi et al., 2003). Those oocytes with mtDNA copy number lower than the required threshold may be unable to form the networks required for developmental competence. In support of this, developmental competence appears to be related to the numbers of mitochondria present (El Shourbagy et al., 2006). In sea urchin oocytes, mitochondria act as calcium sinks (Eisen and Reynolds, 1985). The Ca2+ wave pacemaker that is necessary for the completion of meiosis at fertilization is maintained by mitochondrial uptake of Ca2+ (Dumollard et al., 2003). Failure to maintain the Ca2+ wave pacemaker results in apoptotic cell death of the oocyte (Liu et al., 2001). Therefore, oocytes with low levels of ATP production, because of low levels of mitochondria and mtDNA, may be unable to maintain the Ca2+ wave pacemaker and complete meiosis, and instead, undergo apoptosis. We have previously hypothesized that oocytes without sufficient wild-type mtDNA, and therefore the capacity to generate ATP would not normally be ovulated (St. John, 2002). This is reflected in the low numbers of mtDNA observed in degenerate oocytes (mean = 44 629 ± 40 729; Almeida-Santos et al., 2006). Consequently, the use of superovulation protocols during IVF allows these low-quality oocytes to undergo the maturation process and the inadequacies of insufficient mtDNA molecules to be observed (St. John, 2002). Mitochondrial morphology changes after oocyte maturation to the inactive state (Bavister and Squirrell, 2000). This may be because of the lack of mtDNA replication and the import of nuclear-encoded regulatory factors from the cytoplasm. However OX-PHOS is still functional during early embryo development, and mitochondria are dynamic organelles, moving to provide ATP where it is most needed (Stojkovic et al., 2001). Staining of mitochondria has shown that during embryo development, they cluster around the nucleus of blastomeres during the cleavage stages, providing the energy required for cell division processes such as spindle formation. However, when cell division is not taking place, they tend to spread throughout the cytoplasm of the cell (Stojkovic et al., 2001; Sun et al., 2001; Wilding et al., 2001). Equal segregation of mitochondria following cell division reduces the possibility of resulting tissues of the offspring lacking the ability to produce sufficient ATP through OX-PHOS. However, disproportionate numbers of mitochondria among blastomeres has been reported in the pig (El Shourbagy et al., 2006) and human (Van Blerkom et al., 2000; Lin et al., 2004). This may result in some blastomeres with reduced ATP-generating capacity. If this occurs early on in development, blastomeres may fragment, resulting in embryo arrest or less competent blastocysts with fewer cell numbers. In the human, mitochondrial morphology changes back to the active state at the eight-cell stage, coinciding with the activation of the embryonic genome (Sathananthan and Trounson, 2000). Furthermore, during cavitation, shortly after embryonic genome activation, there is an increased dependence on glycolysis in most mammalian species (see Van Blerkom et al., 2002 for discussion). This would suggest that nuclear embryonic genome activation provides a signal to the mitochondria to reduce OX-PHOS activity. Increased use of glycolysis may allow embryonic cells that are no longer able to generate sufficient ATP through OX-PHOS to survive for long enough to enable production and import of nuclear-encoded regulatory factors necessary for the onset of mtDNA replication post-implantation (El Shourbagy et al., 2006). Indeed, studies in pigs using OX-PHOS inhibitors have indicated that the inhibition of OX-PHOS function at compaction can improve embryo viability (Machaty et al., 2001). This hypothesis is supported by reports of glycolysis rather than OX-PHOS metabolism occurring in other cell types with a functioning nuclear genome but low mtDNA copy number, including somatic cells being depleted of their mtDNA (King and Attardi, 1989). However, ATP production through glycolysis is far less efficient than through OX-PHOS. It is therefore vital that sufficient mtDNA molecules are present at fertilization to maintain embryo survival until mtDNA replication resumes. The relevance of mtDNA replication to maternal transmission of mtDNA MtDNA is further unlike nuclear DNA as it does not follow the normal Mendelian pattern of inheritance but is maternally inherited (Giles et al., 1980). Owing to its size, a mature oocyte can contain far more mitochondria in its cytoplasm than the fertilizing sperm cell. Coupled with replication of the mitochondrial genome during oocyte maturation in the ovary, this results in an oocyte : sperm mtDNA ratio of 1000:1 (Ankel-Simons and Cummins, 1996), with sperm mtDNA contributing approximately only 100 molecules at fertilization (Hecht and Liem, 1984). MtDNA is usually homoplasmically transmitted, i.e. only one mtDNA genotype being present (Monnat et al., 1985). Occasionally, two or more mtDNA genotypes can be present, often because of either a pathological or a non-pathological rearrangement, resulting in heteroplasmy, as in the oocytes of women harbouring a mtDNA disease (Blok et al., 1997). MtDNA has a much higher mutation rate than nuclear DNA (Wallace et al., 1987). This is a consequence of the production of mutagenic free radicals within the mitochondria, which are further confounded by limited mtDNA repair mechanisms (Yakes and Van Houten, 1997). Furthermore, mtDNA is not stabilized with histones as is the case for nuclear DNA, although TFAM is now thought to be involved in mtDNA packaging (Alam et al., 2003). Despite the high mutation rate, mtDNA has been highly conserved throughout evolution, and when it arises, mtDNA heteroplasmy tends to revert back to homoplasmy sometimes within one generation (Koehler et al., 1991). MtDNA replication does not occur in synchrony with nuclear DNA replication, and therefore some mtDNA genomes are not replicated during one-cell cycle, whereas others are copied many times (Sazer and Sherwood, 1990). It has been reported that there is a preferential replication of those mtDNA molecules that are closer to the nucleus (Davis and Clayton, 1996), perhaps because of the faster diffusion of nuclear-encoded regulatory factors. The introduction of exogenous mtDNA into zygotes either close to the nucleus or at the periphery influences the amount of exogenous mtDNA observed at the blastocyst stage, with higher proportions arising from karyoplast injections (Meirelles and Smith, 1998). Consequently, the position of mitochondria in the early embryo may play a role in determining which mtDNA genomes are selected for amplification. Contradictory data have recently shown that mtDNA replication is not just restricted to mitochondria in proximity to the nucleus (Magnusson et al., 2003). Here, it was demonstrated that all components of the replication machinery were available in sufficient quantities throughout the cell, allowing replication to take place throughout the entire cytoplasm. This included PolG and TFAM. The previously reported preferential labelling of mtDNA near cell nuclei was suggested to be because of denser aggregation of mitochondria around nuclei (Magnusson et al., 2003). MtDNA replication throughout the cell cytoplasm is supported by a lack of evidence for selection against pathogenic mtDNA mutations in oogenesis or early embryonic development (reviewed by Shoubridge, 2000). When mtDNA replication resumes following implantation (Piko and Taylor, 1987), each cell type of the developing embryo contains a few copies of mtDNA, as indicated by the few mitochondria present in undifferentiated embryonic stem cells (ESCs; Sathananthan et al., 2002; St. John et al., 2005a). Consequently, the combination of segregation followed by the amplification of selected templates produces an mtDNA purification process (Hauswirth and Laipis, 1982), which is important for the maintenance of mtDNA homoplasmy. If blastomeres inherit mutated genomes or are depleted of mitochondrial genomes, then the resulting adult tissues would be deficient in OX-PHOS capacity, which could result in the onset of mitochondrial disease (Wallace, 1992). There are many diseases caused by mutations in mtDNA and symptoms of these diseases generally appear in tissues highly dependent on OX-PHOS (Wallace, 1992). These include Leber’s hereditary optic neuropathy (LHON), which can result from mutations in the gene for the NADH-Q oxidoreductase component of complexes I and III (Brown et al., 1992). Myoclonic epilepsy with ragged-red fibres (MERRF) is also transmitted by maternal inheritance and results in myoclonus, epilepsy and ataxia. This can result from mutations in mitochondrially encoded tRNA genes (Hanna et al., 1995; Nakamura et al., 1995; Nishigaki et al., 2003; Mancuso et al., 2004) and protein genes (Naini et al., 2005). As well as point mutations, some mitochondrial diseases are caused by large-scale deletions to the genome. The most well known in this group is Kearns–Sayre Syndrome (KSS), which includes symptoms of PEO, pigment retinopathy and disorders of cardiac conduction (Klopstock et al., 1995; Lertrit et al., 1999; De Block et al., 2004). The clinical phenotypes resulting from mtDNA mutations are dependent on the proportion of mutated mtDNAs (Wallace, 1992). In the case of LHON, >60% mutant mtDNA load is required before the disease phenotype presents (Chinnery et al., 2001). In some mitochondrial diseases, including a case of MERRF, over 85% mutant mtDNAs need to be present before symptoms are observed (Boulet et al., 1992). Paternal mtDNA inheritance There is increasing evidence demonstrating that maternal-only mtDNA transmission is not merely because of dilution of the paternal mitochondria beyond the level of detection. It has been suggested that paternal mtDNA is actively eliminated by ubiquitination in the oocyte cytoplasm and subsequent proteolysis during embryonic development (Sutovsky et al., 1999; Nishimura et al., 2006). The elimination hypothesis is supported by reports demonstrating reduced paternal mtDNA content as embryonic development progresses (Cummins et al., 1997). Paternal mtDNA was detected only as far as the pronucleus stage following intra-specific mouse crosses compared with detection in neonates following inter-specific mouse crosses (Kaneda et al., 1995). This suggests that the elimination mechanism is species specific. However, the elimination mechanism is not only species specific but also tissue specific, as donor sperm mtDNA was eliminated, whereas exogenous liver mtDNA injected into the oocyte persisted at least until the neonatal period (Shitara et al., 2000). The destruction of paternal mtDNA (reviewed in Sutovsky, 2003) and the presence of the genetic bottleneck should usually result in homoplasmic individuals. Recently, the accepted theory of uniparental mtDNA inheritance has been questioned. This is especially so in the case of a male patient presenting with a mitochondrial myopathy who was found to have inherited his muscle mtDNA from his father. However, the 2bp deletion giving rise to his myopathy was de novo (Schwartz and Vissing, 2002). This outcome is further supported by the evidence of sperm mtDNA persistence in human polyploid IVF-generated blastocysts (St. John et al., 2000). Paternal leakage of mtDNA has also been reported following inter-specific crosses in fruitflies (Kondo et al., 1990), mice (Gyllensten et al., 1991), honeybees (Meusel and Moritz, 1993), birds (Kvist et al., 2003), sheep (Zhao et al., 2004) and non-human primates (St. John and Schatten, 2004) and following intra-specific crosses in mussels (Zouros et al., 1992) and fruitflies (Kondo et al., 1992). The ubiquitin-related paternal mtDNA destruction process has therefore proven ineffective in these cases. It is as yet unknown how paternal mtDNA might escape the destruction process and how the paternal mtDNA might be preferentially replicated in a particular tissue to such an extent that it resulted in a mitochondrial disease (Schwartz and Vissing, 2002). In this latter instance, a considerable selective advantage must have taken place to out-compete the maternal source of mtDNA in the affected tissue. A possible mechanism for bypassing the paternal mtDNA destruction process is the fusion of parental mitochondria followed by recombination of the two genomes. Recombination of mammalian mtDNA was originally thought not to arise, although it is known to occur in yeast. However, more recent evidence suggests that it may also occur in mammals. For example, Ono and colleagues created two respiration-deficient cell lines from patients with mitochondrial diseases caused by different mutations (Ono et al., 2001). Fusion of the two-cell lines resulted in morphology and respiratory enzyme activity returning to normal within 10–14 days. These authors proposed that the outcome resulted from mtDNA recombination and genome compensation. Mammalian mtDNA recombination has since been confirmed in human cells using PCR-based techniques (D’Aurelio et al., 2004) and recent sequence analysis of invertebrate and vertebrate species, including primates (Tsaousis et al., 2005). These authors suggested that recombination may have been previously missed because of the purification process during oocyte production prohibiting the generation of new haplotypes and their transmission to the next generation. Recombination of sperm and oocyte mtDNA is now thought to explain the presence of the 2bp deletion in the male myopathy patient (Kraytsberg et al., 2004). Invasive oocyte reconstruction protocols The introduction of more advanced assisted reproductive technologies (ART) and their current use in the clinical setting and in animal production further confounds the unimaternal transmission of mtDNA. These technologies include (i) ICSI which is used clinically; (ii) cytoplasmic transfer (CT) which has been used clinically but is now, for example, banned in the United Kingdom and United States and (iii) germinal vesicle transfer (GVT), pronuclear transfer (PNT) and NT, which have not yet been used to treat humans. These more invasive ART protocols present more opportunities for the violation of the normal pattern of mtDNA inheritance. If uniparental mtDNA inheritance is not maintained following these techniques, offspring may harbour various forms of mtDNA heteroplasmy. Indeed, the introduction of these new technologies results in heteroplasmy being redefined. In this instance, either a wild-type molecule can co-exist with a rearranged molecule or two wild-type molecules may persist in the reconstructed oocyte. ICSI Unlike IVF, the ICSI procedure varies dramatically from the natural fertilization process. For example, the intact, non-capacitated, non-acrosome-reacted sperm is injected into the oocyte. There are also delays and alterations in post-fertilization processes following ICSI, including male pronucleus formation (Luetjens et al., 1999) and the calcium oscillations essential for the activation process (reviewed in Ludwig et al., 2001). This delayed response after sperm entry means that ICSI also has the potential to delay degradation of sperm mtDNA. Furthermore, couples who have male factor infertility and require ICSI are at a higher risk of using sperm containing abnormal mtDNA as some human mtDNA haplogroups are associated with high or reduced sperm motility (Ruiz-Pesini et al., 2000). This is demonstrated in those men harbouring large numbers of mtDNA deletions who are more likely to exhibit poorer semen quality (Lestienne et al., 1997; Reynier et al., 1997; St. John et al., 2001) as are those exhibiting point mutations (Spiropoulos et al., 2002). Despite the many diversions from the natural fertilization processes following ICSI, to date, it appears that sperm mtDNA is eliminated in offspring generated using this technique (Danan et al., 1999; Marchington et al., 2002). Furthermore, it appears to be eliminated before the blastocyst stage in ICSI-derived polyploid human embryos (St. John et al., 2000). CT CT is an extension of ICSI and involves the injection of supplementary cytoplasm, as well as a sperm, into the oocyte (Figure 3). The rationale for using CT is that donor cytoplasm will contain supplementary, often younger, non-mutated mtDNA and will possibly contain other important cytoplasmic components to compensate for the defective cytoplasm in the oocyte (Cohen et al., 1997). In the human, it has resulted in higher pregnancy rates and live births (Cohen et al., 1997; Barritt et al., 2001). CT can also result in heteroplasmy as demonstrated by the analysis of various tissues of human CT offspring (Brenner et al., 2000). Consequently, it remains to be demonstrated whether the donor mtDNA out-competes the recipient mtDNA in the heteroplasmic offspring. We also need to ascertain whether the coupling of two distinct genomes accounted for the two cases of Turner’s syndrome, one electively aborted and the other spontaneously aborted, and the one case of pervasive development disorder (Barritt et al., 2001). Figure 3. View largeDownload slide Assisted reproductive techniques. Predicted outcomes associated with mitochondrial DNA (mtDNA) transmission following IVF, ICSI, cytoplasmic transfer (CT), germinal vesicle transfer (GVT), pronuclei transfer (PNT) and somatic cell nuclear transfer (SCNT). Figure 3. View largeDownload slide Assisted reproductive techniques. Predicted outcomes associated with mitochondrial DNA (mtDNA) transmission following IVF, ICSI, cytoplasmic transfer (CT), germinal vesicle transfer (GVT), pronuclei transfer (PNT) and somatic cell nuclear transfer (SCNT). To avoid the possible harmful effects of heteroplasmy in the offspring, cytoplasm from granulosa cells of the mother rather than cytoplasm from donor oocytes has been used (Kong et al., 2003; 2004). These are likely to have the same mtDNA sequence as the recipient oocyte. However, one possible problem with this procedure is that mtDNA deletions in granulosa cells are more common in older women (Seifer et al., 2002), i.e. those that are likely to have defective oocytes and require CT. Therefore, the treatment may be unsuccessful because of continued lack of wild-type mtDNA and may even result in the transmission of mitochondrial disease to the offspring that would not have occurred without CT. Another likely problem is that granulosa cells are fully differentiated, whereas newly fertilized oocytes are pluripotent. Cells at different stages of differentiation are known to be different in terms of their mitochondrial activity and content. For example, undifferentiated human (h) ESCs (Sathananthan et al., 2002; St. John et al., 2005a) and primordial germ cells (Jansen and De Boer, 1998) have been observed to contain low numbers of mitochondria. However, differentiated cells contain higher numbers of more active mitochondria, depending on OX-PHOS requirements of specific cell types (Moyes et al., 1998). Mitochondria donated from granulosa cells would therefore need to be ‘reprogrammed’ to match those of the oocyte. We have supplemented developmentally incompetent pig oocytes with pure populations of mitochondria from developmentally competent oocytes from maternal relatives. Interestingly, we observed improved fertilization outcome and the maintenance of unimaternal inheritance (El Shourbagy et al., 2006). GVT and PNT GVT and PNT have also been proposed as techniques to overcome problems associated with aged oocytes, whereby the GV or PN is transferred into an enucleated recipient oocyte. Ideally, the enucleated recipient should be from the same stage of development as the donor nucleus (Liu et al., 1999). In aged oocytes, there is likely to be a larger number and higher proportion of mutant mtDNA (Keefe et al., 1995). It has been hypothesized that mitochondria are important not only for the production of ATP but also for correct chromosomal segregation during meiosis (Schon et al., 2000). Furthermore, it has been suggested that aberrant mitochondrial function in aged oocytes may have a role in oocyte aneuploidy (Eichenlaub-Ritter et al., 2004), a common cause of infertility in older women (Munne et al., 1995; Dailey et al., 1996). The chances of success may therefore increase if the nuclear component of the aged oocyte was transferred into the enucleated ooplasm of a younger, healthier oocyte, rather than supplement the existing ooplasm. This transfer can be carried out at either the GV or the PN stage (Figure 3). Indeed, studies using human oocytes have shown GVT from aged oocytes to the enucleated ooplasm of young oocytes can overcome oocyte aneuploidy, with the majority of reconstructions having normal karyotypes (Zhang et al., 1999; Takeuchi et al., 2001). There may also be non-cytoplasmic components of the GV that influence chromosome segregation. This has been indicated in a series of mouse experiments mixing GVs and ooplasms of varying ages (Cui et al., 2005). A significantly greater number (57.1 %) of chromosomal abnormalities were found in reconstructions when the GV was transferred from older females to the enucleated ooplasm of younger females. This is in contrast to abnormalities found in reconstructions derived from the transfer of younger GVs into aged ooplasms (16.7 %). Mitochondria are known to be involved in chromosome organization and movement. For example, the induction of mitochondrial damage in mouse oocytes prevented oocyte maturation, chromosomal segregation and spindle formation (Takeuchi et al., 2005). This damage was overcome by GVT of the karyoplast derived from a damaged oocyte into the ooplasm of a healthy enucleated oocyte. Cytogenetic analysis showed that 20/21 of these reconstructions had a normal number of chromosomes. Furthermore, performing ICSI on GVT reconstructions and transfer of the resulting two-cell embryos led to the generation of live mouse offspring. It has also been demonstrated that an intact mouse ooplasm could rescue mitochondrial-damaged karyoplasts induced by photoirradiation with 62% of these reconstructions maturing to metaphase II (Palermo et al., 2002). PNT is essentially the same procedure, except for the nuclear material being removed after fertilization. This technique has also been used in mice with considerable success as determined by the birth of live offspring (Meirelles and Smith, 1997; 1998;He et al., 2003; Sato et al., 2005). This procedure, however, involves the destruction of a zygote, which may restrict its usage because of ethical and moral considerations. GVT and PNT have been proposed as techniques for treating mtDNA disease (Cummins 1998; Trounson, 2001). However, potential problems could arise as the transferred GV or PN is still surrounded by mitochondria, which will also be carried over into the donor ooplasm. These mitochondria remain close to the centre of the immature reconstruction at first but then disperse throughout the cytoplasm as maturation ensues (Fulka, 2004). Owing to segregation during embryogenesis, this provides an opportunity for karyoplast mitochondria to be present in all cell types of the offspring. Any disease-associated mitochondria transferred with the nuclear material may therefore continue to be transmitted. It is further possible that, due to its initial proximity to the nucleus, karyoplast mtDNA would be preferentially replicated in the reconstructions. Indeed, fusion of enucleated somatic cells with mtDNA-depleted somatic cells can lead to the preferential expansion of one particular type of mtDNA such as the mutant molecule (Dunbar et al., 1995). Although, to our knowledge, there has been no analysis of mtDNA transmission following GVT, results are likely to be similar to those reported following PNT. In this instance, exogenous mtDNA accompanying the murine zygotic karyoplast is found in higher concentrations at blastocyst than mtDNA from transferred cytoplasts (Meirelles and Smith, 1998). This has been further highlighted where zygotic karyoplast transfer into zygotes resulted in varying amounts of mtDNA being transmitted to offspring (0–69%; Meirelles and Smith, 1997). Consequently, it would be difficult to predict to what extent a mutated or deleted mtDNA molecule would be selected. However, a recent report suggests that the varying levels of mtDNA deletion in mouse zygotes could be reduced to levels lower than those expected for the phenotypic onset of mtDNA disease when the zygotic nucleus is transferred into an enucleated healthy oocyte (Sato et al., 2005). The subsequent offspring also remained under the level for pathological onset even with the accumulative increase of the deletion with time. These authors state however that, in the human, this form of nuclear transplantation could not be applied to those mtDNA diseases where the pathogenic mutation had significant replicative advantage over wild-type mtDNA. This is most likely as in humans the prenatal period is 13 times longer than in the mouse, and this represents the window in which mutant mtDNA replication is most proliferative (Sato et al., 2005). NT NT involves the injection of a donor nucleus or a whole cell into an enucleated recipient oocyte (Figure 3). Mostly commonly, either embryonic cell (EC) NT, using a blastomere, or somatic cell NT (SCNT) are performed. Resulting embryos have then been used to generate blastocysts and offspring in a variety of species (Campbell et al., 2005). However, NT is renowned for its low efficiency, with low proportions of embryos developing to the blastocyst stage (White et al., 1999) and even fewer progressing far enough to implant, survive pregnancy and produce live offspring (Wilmut et al., 1997). One of the major abnormalities associated with NT offspring is the production of abnormal placentae which are thought to be the cause of many fetal deaths (Cibelli et al., 2002). Incorrect reprogramming of imprinted genes has also been cited as one of the major factors which leads to early fetal death. This was highlighted by non-surviving NT calves having incorrect expression of imprinted genes, namely Igf2, Igf2r and H19 in some tissues, whereas surviving adult clones had normal expression at these loci except for Igf2 in muscle (Yang et al., 2005). Furthermore, in mouse NT embryos and offspring irregular patterns of gene expression have been observed most likely as a result of inappropriate reprogramming (Humpherys et al., 2002). Although the process of NT for human reproductive purposes is ethically unacceptable and is banned in many countries (UK – Human Fertilisation and Embryology Act, 1990; USA – H.R. 534 Office of Legislative Policy and Analysis, 2003; South Korea – see Normile and Mann, 2005), its use as a means of producing histocompatible stem cells for treatment of various diseases is likely to have a significant impact. Cell replacement therapy has great potential for the treatment of certain conditions such as myocardial infarction (Rubart and Field, 2006) and Parkinson’s disease (Taylor and Minger, 2005) and as a replacement for tissue/organ transplantation (Cortesini, 2005). To date, encouraging results have been obtained in animal studies, whereby hESC-derived neural precursors (Wernig et al., 2004) and cardiomyocytes (Laflamme et al., 2005) were successfully transplanted into animal organs. However, considerable characterization and validation of precursor cells derived from hESCs still needs to be undertaken before cell replacement can be used as an effective therapy in humans. Some success at generating NT-derived ESCs has so far been achieved in mouse models (Munsie et al., 2000; Wakayama et al., 2001), and its use as a means of delivering the patient’s own chromosomal DNA with the corrected gene for a single gene disorder has been demonstrated in the mouse (Rideout et al., 2002). One of the rate-limiting factors in generating autologous ESCs, i.e. stem cells possessing the patient’s own chromosomal and mtDNA compositions is the regulation of the mtDNA composition. In this respect, a considerable amount can be learnt from those offspring derived from NT through either ECNT or SCNT. Here, transmission results in either homoplasmy derived from the recipient oocyte mtDNA or heteroplasmy resulting from transfer of both donor and recipient mtDNA (reviewed in St. John et al., 2004a,b). This is probably because of the random pattern of mtDNA segregation during embryogenesis (reviewed in Shoubridge, 2000). Consequently, it is apparent that no real mechanism is currently understood that could account for the irregular patterns of mtDNA transmission observed. One parallel that can be drawn from both ECNT and SCNT is that the range of donor mtDNA transmitted is relatively similar at 0–57% (Hiendleder et al., 1999) and 0–59% (Takeda et al., 2003), respectively. However, two explanations could exist for those studies not detecting donor mtDNA in live offspring: (i) no donor mtDNA is actually present or (ii) too few tissues have been analysed. Consequently, the true extent of heteroplasmy is difficult to determine. This is perhaps exemplified by the recent studies of pig (Takeda et al., 2006), cow (Takeda et al., 2003) and mouse (Inoue et al., 2004) NT-derived offspring, where varying levels of heteroplasmy has been observed between the tissues. This would indicate that if all tissues are not analysed, then these studies are not truly subjective. Furthermore, investigations should include analysis of the surrogate mother’s mtDNA as leakage through the placenta into fetal blood has been reported in cattle (Hiendleder et al., 2004). The initial number of donor mtDNA molecules introduced into the recipient oocyte also does not necessarily influence whether donor mtDNA is actually transmitted. Using blastomeres from more advanced embryonic stages in bovine NT has resulted in reduced levels of heteroplasmy being reported (Steinborn et al., 1998). However, this was not a uniform pattern as two offspring derived from 52-cell morula blastomeres contained less donor mtDNA than offspring derived from 92-cell morula blastomeres. A further issue is the stage of embryonic development of the donor cell. A recent analysis of two NT-derived monkey offspring (St. John and Schatten, 2004) demonstrated mtDNA transmission from both the recipient oocyte and the donor cell. In this instance, the early IVF blastomere (8- to 12-cell stage) still contained sperm mtDNA. Consequently, the offspring possessed three different populations of mtDNA. Furthermore, the depletion of donor cell mtDNA to residual levels (<0.02% of its original content) still resulted in the persistence of some mtDNA at the blastocyst stage in sheep NT (Lloyd et al., 2006). The use of a somatic donor cell with accompanying mitochondria not only raises interesting questions with regard to the differentiated status of nucleus but also of the mitochondria. During development, from the inner cell mass to a fully differentiated cell, mitochondria increase in number to meet the metabolic requirements of that specialized cell. The proliferation of mitochondria further discriminates between the number of mtDNA copies per mitochondrion and cell type, with higher ATP users possessing greater copy numbers (Moyes et al., 1998). Therefore, for SCNT, the mitochondria accompanying the donor cell should also be reprogrammed to the embryonic state, as is the case for the donor nucleus. Failure to do this may result in embryos containing mitochondria with insufficient plasticity to suit a multitude of cell types that they will contribute to. The persistence of donor mitochondria could also influence the metabolic pathways used during early development and explain why improved developmental rates are observed in those embryos cultured in media that enhance anaerobic respiration (Chung et al., 2002). It is further evident that the injection of somatic donor mitochondria into oocytes which are then parthenogenetically activated is detrimental to embryonic development though this is not observed with donor ooplasm (Takeda et al., 2005). Interaction between multiple wild-type mitochondrial genomes It has not been clearly demonstrated whether heteroplasmy resulting from two wild-type mtDNA molecules can be detrimental to the offspring and result in mitochondrial disease. This is partly because many post-mortem results have not been published, and the cause of many deaths following NT is not documented. Consequently, we are unsure whether those offspring that survive are the exception to the rule. However, from one report, some common causes of death in sheep NT offspring are similar to symptoms of mitochondrial diseases, including brain, kidney and liver defects (McCreath et al., 2000) and in a series of other species (Cibelli et al., 2002). Sequence differences due to mtDNA from a ‘foreign’ source can give rise to proteins with altered amino acid sequences. This has been demonstrated in both cattle (Steinborn et al., 2002) and pigs (St. John et al., 2005b) and may result in inadequate interaction between the individual subunits of the ETC. This would result in reduced energy production capacity and symptoms common to mitochondrial disease and influence embryo and fetal survival. The question is whether handmade cloning, the fusion of one or more oocyte cytoplasts with a somatic cell (Vajta et al., 2001), would resolve such issues or complicate them. The significance of nucleo–mitochondrial interaction Intergenomic communication is vital for efficient cellular function as much interaction occurs between nuclear-encoded proteins and those encoded by the mtDNA genome (reviewed by Scarpulla, 2002). The ETC requires nuclear-encoded proteins to be transported to the mitochondria and to assemble appropriately (Stojanovski et al., 2003). Failure of TFAM, TFB1M and TFB2M to co-ordinate transcription would also have serious implications for a functional ETC (Falkenberg et al., 2002). Furthermore, failure of TFAM and PolG to co-ordinate replication would have similar outcomes for replication (Larsson et al., 1998; Hance et al., 2005). However, NT appears to result in the continued expression of both these factors during embryo development and could explain why even residual levels of donor mtDNA persist up to the blastocyst stage (Lloyd et al., 2006). Failure of TFAM or PolG to interact with the mtDNA genome, post-implantation, through the diversity of the donor nucleus and the recipient oocyte’s mtDNA would result in compromised transcription or replication and mimic mtDNA-depletion syndromes (Poulton et al., 1994). Patient-specific ESCs derived by NT would require large numbers of good-quality human oocytes which, due to medical, ethical, legal and moral issues, are a limited resource. Added to this, the methods of recruitment of donor oocytes for the derivation of such ESCs have caused considerable outrage recently (Holden, 2005). Cross-species NT offers the opportunity to derive ESCs that would carry the chromosomal DNA from a patient propagated in a surrogate ooplasm of another species. Encouragingly, there is a single report of the generation of a hybrid ESC line following the transfer of a human nucleus into a rabbit oocyte (Chen et al., 2003). However, outcomes from cross-species NT further exemplify the importance of nucleo-mitochondrial compatibility for ESC derivation. Use of both human cord fibroblasts with bovine oocytes (Chang et al., 2003) and chicken blastodermal cells with rabbit oocytes (Liu et al., 2004) resulted in elimination of the donor mitochondrial genome by the morula or blastocyst stage of development. Rhesus macaque donor cells injected into rabbit oocytes resulted in heteroplasmic embryos but with only the rabbit mtDNA being replicated after the blastocyst stage (Yang et al., 2003), suggesting that these too could have soon become homoplasmic for the recipient oocyte mtDNA. This suggests either preferential replication of the oocyte mtDNA or degradation of the mtDNA from the nuclear donor cell. Analysis of cross-species NT offspring provides further indications as to potential outcomes for cross-species ESCs. In two combinations of cross subspecies cow NT experiments, the donor nucleus was capable of interaction with both types of mtDNA with the production of apparently healthy heteroplasmic calves (Steinborn et al., 2002; Han et al., 2004). However, the combination of Bos indicus donor cells with Bos taurus oocytes resulted in homoplasmic embryos and one offspring (Meirelles et al., 2001). There has also been one report of preferential replication of mtDNA from the nuclear donor cell following cross-species NT (Chen et al., 2002). Here, nuclei from giant panda somatic cells were injected into rabbit oocytes, and preferential replication of the panda mtDNA was observed after analysis of the fetuses. This suggests that some combinations of cross-species NT, particularly where the species are closely related or a subspecies, may result in a functional ETC and efficient mitochondrial respiration. However, more distantly related species might preferentially replicate their own mtDNA genome because of incompatibility with that in the recipient oocyte. Much can also be learnt from somatic cell–cell fusion studies and xenomitochondrial offspring in relation to outcomes associated with cross-species NT. For example, one study involved combinations of cells depleted of their mtDNA and exogenous mitochondria from rats and two subspecies of mouse (McKenzie and Trounce, 2000). Transfer of Mus spretus mtDNA into Mus musculus cells resulted in heteroplasmy, and no obvious OX-PHOS defects, probably due to the two types of mice, being closely related. However, when rat mtDNA was transferred into Mus musculus cells, its replication and transcription were unaffected, but OX-PHOS capacity was reduced. Other studies have also demonstrated increased respiratory chain defects following crosses of greater evolutionary distance both in mice (McKenzie et al., 2003) and in primates (Kenyon and Moraes, 1997). Effects of species differences, and also mitochondrial disease, can now be studied in vivo following the production of xenomitochondrial mice (McKenzie et al., 2004; Sokolova et al., 2004). Xenomitochondrial mice can be produced by (i) direct injection of foreign mitochondria into zygotes (Irwin et al., 1999), (ii) pronuclei fusion with enucleated oocytes (Meirelles and Smith, 1997) followed by the transfer to surrogate mothers and (iii) the production of ESC cybrids for injection into blastocysts (Levy et al., 1999). Again, it has been demonstrated that greater evolutionary distance between species results in offspring with increased respiratory chain defects (Trounce et al., 2004). It is most likely that OX-PHOS defects arose from abnormal combinations of nuclear and mitochondrially encoded ETC proteins resulting in abnormal structure of ETC complexes, a scenario very similar to outcomes that arise through cross-species NT. There are many cases where interspecies NT embryos implant and survive pregnancy but die shortly after birth (Hammer et al., 2001; Han et al., 2004). These may be a result of inefficient ETC activity due to species differences. However, should these combinations result in apparently healthy live offspring, it is possible that phenotypic alterations may still be present. Such outcomes could affect those mtDNA haplotypes that influence, for example, milk quality in cattle (Brown et al., 1989), growth and physical performance in mice (Nagao et al., 1998), a reduction of fertility in beef cattle (Sutarno et al., 2002) and sperm motility in men (Ruiz-Pesini et al., 2000). Perhaps, the biggest note of caution is exemplified by the result of introducing a common carp nucleus into an enucleated goldfish oocyte, which resulted in the offspring developing phenotypic features of the goldfish (Sun et al., 2005). Consequently, the use of cross-species NT for the derivation of hESCs needs considerable investigation although it offers great potential. Are these technologies safe? Although some experiments have been carried out in animals and a small number of groups practise invasive ART protocols in humans, little is known about the likely outcomes of these techniques in relation to the long-term effects on offspring and future generations. It is essential that we understand the risks of these techniques before their therapeutic use becomes widespread. The use of preimplantation genetic diagnosis (PGD) has recently been proposed for early diagnosis of mitochondrial disease (Dean et al., 2003). The use of PGD for chromosomal genetic analysis has been implemented in many clinics. However, its effectiveness needs to be demonstrated for the detection of potential mtDNA disease transmission. Some work has indicated its usefulness in this respect. In the mouse, mtDNA genotypes in polar bodies were compared with their respective oocytes and in different blastomeres within the same early embryo (Dean et al., 2003). The level of heteroplasmy present in polar bodies and blastomeres was representative of that in the associated oocytes and embryos, with the authors therefore concluding that PGD was feasible for the diagnosis of mitochondrial disease. However, cloned mouse offspring (Inoue et al., 2004), various farm animal species such as pigs (Takeda et al., 2006) and cattle (Hiendleder et al., 2003) and patients with mitochondrial mutations (Kirches et al., 2001) and diseases (Ponzetto et al., 1990) often demonstrate varying levels of heteroplasmy in different tissues. Furthermore, considerable variation in oocyte-derived mtDNA copy number among intra-embryo blastomeres has been demonstrated in both the pig (El Shourbagy et al., 2006) and the human (Lin et al., 2004). Therefore, even if the proportion of heteroplasmic mtDNA molecules were the same in each blastomere, the actual number of mutant molecules is likely to be different. Blastomeres containing lower numbers of mtDNA molecules will have reduced ATP-generating capacities compared with their counterparts with higher overall numbers of mtDNA. Therefore, the effects in tissues resulting from these blastomeres will be more severe than those resulting from blastomeres with a larger mtDNA complement. Furthermore, different tissues require different ATP-generating capacity. Low levels of wild-type mtDNA in some blastomeres may therefore go unnoticed if that particular blastomere develops into a low ATP-requiring tissue. However, this could result in severe mitochondrial disease should that blastomere develop into a tissue highly dependent on OX-PHOS. It is therefore unlikely that PGD would be a suitable technique to determine the risk of mitochondrial disease following oocyte reconstruction, especially following GVT, PNT and NT as all have reported preferential replication of donor mtDNA. Consequently, an alternative is required. Conclusion Mitochondria are vitally important during embryonic development and tissue differentiation, as demonstrated by the changes in their localization, activity and morphology in the early embryo. The propagation of mitochondria before metaphase II and the maternal transmission of mtDNA ensure that sufficient numbers of the organelle and its genome are invested into the process of early embryogenesis. Consequently, these pre-metaphase II events are under the control of the oocyte nuclear genome, which regulates transcription and replication, and ensure that transmission is not affected during a period when there is no nuclear control, i.e. before activation of the embryonic genome. In the process of trying to treat various diseases through the generation of patient-specific stem cells and the use of GVT and PNT to prevent the transmission of mtDNA-type disease, we would be taking inappropriate risks if current strategies were introduced into the clinic. In many ways, our strategies are confounding and our ingenuity could turn into our downfall. We are attempting to provide the nucleus with a new environment in which it can prosper but, for GVT and PNT, we have forgotten that this karyoplast is still accompanied by the mutant mtDNA that we are trying to eradicate. For NT-derived patient-specific stem cells, GVT and PNT, we are asking the transferred nucleus to interact with a cytoplasm which may or may not be compatible. We also expect two or more populations of mtDNA to happily cohabit. We are however ignoring how the cytoplasm will adjust to its new nuclear companion in favour of concentrating on the nucleus and its requirement to achieve pluripotency. Cohabitation requires considerable foresight, and in this respect we need to understand considerably more about nucleo-mitochondrial harmony before we exploit such technologies. The question still remaining to be answered is: Can heteroplasmy be tolerated at a functional level even if it compromises genetic identity? Paradoxically, inheritance and therefore analysis of recipient oocyte mtDNA does, however, allow us to confirm whether oocyte reconstruction has taken place. This is a lesson we have hopefully learnt from the discredited Korean attempts at generating patient-specific ESCs. References Alam TI, Kanki T, Muta T, Ukaji K, Abe Y, Nakayama H, Takio K, Hamasaki N and Kang D ( 2003) Human mitochondrial DNA is packaged with TFAM. Nucleic Acids Res  31, 1640–1645. CrossRef Search ADS PubMed  Almeida-Santos T, El Shourbagy S and St John JC ( 2006) Mitochondrial content reflects oocyte variability and fertilization outcome. Fertil Steril 85, 584–591. Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F et al. ( 1981) Sequence and organization of the human mitochondrial genome. Nature  290, 457–465. CrossRef Search ADS PubMed  Ankel-Simons F and Cummins JM ( 1996) Misconceptions about mitochondria and mammalian fertilization: implications for theories on human evolution. Proc Natl Acad Sci USA  93, 13859–13863. CrossRef Search ADS PubMed  Barritt JA, Brenner CA, Malter HE and Cohen J ( 2001) Rebuttal: interooplasmic transfers in humans. Reprod Biomed Online  3, 47–48. CrossRef Search ADS PubMed  Bavister BD and Squirrell JM ( 2000) Mitochondrial distribution and function in oocytes and early embryos. Hum Reprod  15, 189–198. CrossRef Search ADS PubMed  Blok RB, Gook DA, Thorburn DR and Dahl HH ( 1997) Skewed segregation of the mtDNA nt 8993 (T –-> G) mutation in human oocytes. Am J Hum Genet  60, 1495–1501. CrossRef Search ADS PubMed  Boulet L, Karpati G and Shoubridge EA ( 1992) Distribution and threshold expression of the tRNA (Lys) mutation in skeletal muscle of patients with myoclonic epilepsy and ragged-red fibers (MERRF). Am J Hum Genet  51, 1187–1200. PubMed  Brenner CA, Barritt JA, Willadsen S and Cohen J ( 2000) Mitochondrial DNA heteroplasmy after human ooplasmic transplantation. Fertil Steril  74, 573–578. CrossRef Search ADS PubMed  Brison DR and Leese HJ ( 1991) Energy metabolism in late preimplantation rat embryos. J Reprod Fertil  93, 245–251. CrossRef Search ADS PubMed  Brown DR, Koehler CM, Lindberg GL, Freeman AE, Mayfield JE, Myers AM, Schutz MM and Beitz DC ( 1989) Molecular analysis of cytoplasmic genetic variation in Holstein cows. J Anim Sci  67, 1926–1932. CrossRef Search ADS PubMed  Brown MD, Voljavec AS, Lott MT, Torroni A, Yang CC and Wallace DC ( 1992) Mitochondrial DNA complex I and III mutations associated with Leber’s hereditary optic neuropathy. Genetics  130, 163–173. PubMed  Campbell K, Alberio R, Choi I, Fisher P, Kelly R, Lee J-H and Maalouf W ( 2005) Cloning: eight years after Dolly. Reprod Domest Anim  40, 256–268. CrossRef Search ADS PubMed  Chang KH, Lim JM, Kang SK, Lee BC, Moon SY and Hwang WS ( 2003) Blastocyst formation, karyotype, and mitochondrial DNA of interspecies embryos derived from nuclear transfer of human cord fibroblasts into enucleated bovine oocytes. Fertil Steril  80, 1380–1387. CrossRef Search ADS PubMed  Chen DY, Wen DC, Zhang YP, Sun QY, Han ZM, Liu ZH, Shi P, Li JS, Xiangyu JG, Lian L et al. ( 2002) Interspecies implantation and mitochondria fate of panda-rabbit cloned embryos. Biol Reprod  67, 637–642. CrossRef Search ADS PubMed  Chen Y, He ZX, Liu A, Wang K, Mao WW, Chu JX, Lu Y, Fang ZF, Shi YT, Yang QZ et al. ( 2003) Embryonic stem cells generated by nuclear transfer of human somatic nuclei into rabbit oocytes. Cell Res  13, 251–263. CrossRef Search ADS PubMed  Chinnery PF, Andrews RM, Turnbull DM and Howell NN ( 2001) Leber hereditary optic neuropathy: Does heteroplasmy influence the inheritance and expression of the G11778A mitochondrial DNA mutation? Am J Med Genet  98, 235–243. CrossRef Search ADS PubMed  Chung Y, Mann MRW, Bartolomei MS and Latham KE ( 2002) Nuclear-cytoplasmic tug of war during cloning: effects of somatic cell nuclei on culture medium preferences of preimplantation cloned mouse embryos. Biol Reprod  66, 1178–1184. CrossRef Search ADS PubMed  Cibelli JB, Campbell KH, Seidel GE, West MD and Lanza RP ( 2002) The health profile of cloned animals. Nat Biotechnol  20, 13–14. CrossRef Search ADS PubMed  Cohen J, Scott R, Schimmel T, Levron J and Willadsen S ( 1997) Birth of infant after transfer of anucleate donor oocyte cytoplasm into recipient eggs. Lancet  350, 186–187. CrossRef Search ADS PubMed  Cortesini R ( 2005) Stem cells, tissue engineering and organogenesis in transplantation. Transpl Immunol  15, 81–89. CrossRef Search ADS PubMed  Cui LB, Huang XY and Sun FZ ( 2005) Transfer of germinal vesicle to ooplasm of young mice could not rescue ageing-associated chromosome misalignment in meiosis of oocytes from aged mice. Hum Reprod  20, 1624–1631. CrossRef Search ADS PubMed  Cummins J ( 1998) Mitochondrial DNA in mammalian reproduction. Rev Reprod  3, 172–182. CrossRef Search ADS PubMed  Cummins JM, Wakayama T and Yanagimachi R ( 1997) Fate of microinjected sperm components in the mouse oocyte and embryo. Zygote  5, 301–308. CrossRef Search ADS PubMed  D’Aurelio M, Gajewski CD, Lin MT, Mauck WM, Shao LZ, Lenaz G, Moraes CT and Manfredi G ( 2004) Heterologous mitochondrial DNA recombination in human cells. Hum Mol Genet  13, 3171–3179. CrossRef Search ADS PubMed  Dailey T, Dale B, Cohen J and Munne S ( 1996) Association between nondisjunction and maternal age in meiosis-II human oocytes. Am J Hum Genet  59, 176–184. PubMed  Danan C, Sternberg D, Van Steirteghem A, Cazeneuve C, Duquesnoy P, Besmond C, Goossens M, Lissens W and Amselem S ( 1999) Evaluation of parental mitochondrial inheritance in neonates born after intracytoplasmic sperm injection. Am J Hum Genet  65, 463–473. CrossRef Search ADS PubMed  Davis AF and Clayton DA ( 1996) In situ localization of mitochondrial DNA replication in intact mammalian cells. J Cell Biol  135, 883–893. CrossRef Search ADS PubMed  De Block CE, De Leeuw IH, Maassen JA, Ballaux D and Martin JJ ( 2004) A novel 7301-bp deletion in mitochondrial DNA in a patient with Kearns–Sayre syndrome, diabetes mellitus, and primary amenorrhoea. Exp Clin Endocrinol Diabetes  112, 80–83. CrossRef Search ADS PubMed  Dean NL, Battersby BJ, Ao A, Gosden RG, Tan SL, Shoubridge EA and Molnar MJ ( 2003) Prospect of preimplantation genetic diagnosis for heritable mitochondrial DNA diseases. Mol Hum Reprod  9, 631–638. CrossRef Search ADS PubMed  Dumollard R, Hammar K, Porterfield M, Smith PJ, Cibert C, Rouviere C and Sardet C ( 2003) Mitochondrial respiration and Ca2+ waves are linked during fertilization and meiosis completion. Development  130, 683–692. CrossRef Search ADS PubMed  Dunbar DR, Moonie PA, Jacobs HT and Holt IJ ( 1995) Different cellular backgrounds confer a marked advantage to either mutant or wild-type mitochondrial genomes. Proc Natl Acad Sci USA  92, 6562–6566. CrossRef Search ADS PubMed  Eichenlaub-Ritter U, Vogt E, Yin H and Gosden R ( 2004) Spindles, mitochondria and redox potential in ageing oocytes. Reprod Biomed Online  8, 45–58. CrossRef Search ADS PubMed  Eisen A and Reynolds GT ( 1985) Source and sinks for the calcium released during fertilization of single sea urchin eggs. J Cell Biol  100, 1522–1527. CrossRef Search ADS PubMed  El Shourbagy SH, Spikings EC, Freitas M and St. John JC ( 2006) Mitochondria directly influence fertilisation outcome in the pig. Reproduction  131, 233–245. CrossRef Search ADS PubMed  Falkenberg M, Gaspari M, Rantanen A, Trifunovic A, Larsson NG and Gustafsson CM ( 2002) Mitochondrial transcription factors B1 and B2 activate transcription of human mtDNA. Nat Genet  31, 289–294. CrossRef Search ADS PubMed  Fisher RP and Clayton DA ( 1985) A transcription factor required for promoter recognition by human mitochondrial RNA polymerase. Accurate initiation at the heavy- and light-strand promoters dissected and reconstituted in vitro. J Biol Chem  260, 11330–11338. PubMed  Fisher RP and Clayton DA ( 1988) Purification and characterization of human mitochondrial transcription factor 1. Mol Cell Biol  8, 3496–3509. CrossRef Search ADS PubMed  Fulka H ( 2004) Distribution of mitochondria in reconstructed mouse oocytes. Reproduction  127, 195–200. CrossRef Search ADS PubMed  Gahan ME, Miller F, Lewin SR, Cherry CL, Hoy JF, Mijch A, Rosenfeldt F and Wesselingh SL ( 2001) Quantification of mitochondrial DNA in peripheral blood mononuclear cells and subcutaneous fat using real-time polymerase chain reaction. J Clin Virol  22, 241–247. CrossRef Search ADS PubMed  Giles RE, Blanc H, Cann HM and Wallace DC ( 1980) Maternal inheritance of human mitochondrial DNA. Proc Natl Acad Sci USA  77, 6715–6719. CrossRef Search ADS PubMed  Graziewicz MA, Longley MJ, Bienstock RJ, Zeviani M and Copeland WC ( 2004) Structure-function defects of human mitochondrial DNA polymerase in autosomal dominant progressive external ophthalmoplegia. Nat Struct Mol Biol  11, 770–776. CrossRef Search ADS PubMed  Gyllensten U, Wharton D, Josefsson A and Wilson AC ( 1991) Paternal inheritance of mitochondrial DNA in mice. Nature  352, 255–257. CrossRef Search ADS PubMed  Hammer CJ, Tyler HD, Loskutoff NM, Armstrong DL, Funk DJ, Lindsey BR and Simmons LG ( 2001) Compromised development of calves (Bos gaurus) derived from vitro-generated embryos transferred interspecifically into domestic cattle (Bos taurus). Theriogenology  55, 1447–1455. CrossRef Search ADS PubMed  Han ZM, Chen DY, Li JS, Sun QY, Wan QH, Kou ZH, Rao G, Lei L, Liu ZH and Fang SG ( 2004) Mitochondrial DNA heteroplasmy in calves cloned by using adult somatic cell. Mol Reprod Dev  67, 207–214. CrossRef Search ADS PubMed  Hance N, Ekstrand MI and Trifunovic A ( 2005) Mitochondrial DNA polymerase gamma is essential for mammalian embryogenesis. Hum Mol Genet  14, 1775–1783. CrossRef Search ADS PubMed  Hanna MG, Nelson IP, Morgan-Hughes JA and Harding AE ( 1995) Impaired mitochondrial translation in human myoblasts harbouring the mitochondrial DNA tRNA lysine 8344 A→G (MERRF) mutation: relationship to proportion of mutant mitochondrial DNA. J Neurol Sci  130, 154–160. CrossRef Search ADS PubMed  Hauswirth WW and Laipis PJ ( 1982) Mitochondrial DNA polymorphism in a maternal lineage of Holstein cows. Proc Natl Acad Sci USA  79, 4686–4690. CrossRef Search ADS PubMed  He Z, Liu HC and Rosenwaks Z ( 2003) Cryopreservation of nuclear material as a potential method of fertility preservation. Fertil Steril  79, 347–354. CrossRef Search ADS PubMed  Hecht NB and Liem H ( 1984) Mitochondrial DNA is synthesized during meiosis and spermiogenesis in the mouse. Exp Cell Res  154, 293–298. CrossRef Search ADS PubMed  Hiendleder S, Schmutz SM, Erhardt G, Green RD and Plante Y ( 1999) Transmitochondrial differences and varying levels of heteroplasmy in nuclear transfer cloned cattle. Mol Reprod Dev  54, 24–31. CrossRef Search ADS PubMed  Hiendleder S, Zakhartchenko V, Wenigerkind H, Reichenbach HD, Bruggerhoff K, Prelle K, Brem G, Stojkovic M and Wolf E ( 2003) Heteroplasmy in bovine fetuses produced by intra- and inter-subspecific somatic cell nuclear transfer: neutral segregation of nuclear donor mitochondrial DNA in various tissues and evidence for recipient cow mitochondria in fetal blood. Biol Reprod  68, 159–166. CrossRef Search ADS PubMed  Hiendleder S, Bebbere D, Zakhartchenko V, Reichenbach HD, Wenigerkind H, Ledda S and Wolf E ( 2004) Maternal-fetal transplacental leakage of mitochondrial DNA in bovine nuclear transfer pregnancies: potential implications for offspring and recipients. Cloning Stem Cells  6, 150–156. CrossRef Search ADS PubMed  Holden C ( 2005) Stem cell research. Korean cloner admits lying about oocyte donations. Science  310, 1402–1403. CrossRef Search ADS PubMed  Houghton FD, Thompson JG, Kennedy CJ and Leese HJ ( 1996) Oxygen consumption and energy metabolism of the early mouse embryo. Mol Reprod Dev  44, 476–485. CrossRef Search ADS PubMed  Human Fertilisation and Embryology Act (c. 37) ( 1990) Available from http://www.opsi.gov.uk/acts/acts1990/Ukpga_19900037_en_1.htm#tcon. Google Scholar Humpherys D, Eggan K, Akutsu H, Friedman A, Hochedlinger K, Yanagimachi R, Lander ES, Golub TR and Jaenisch R ( 2002) Abnormal gene expression in cloned mice derived from embryonic stem cell and cumulus cell nuclei. Proc Natl Acad Sci USA  99, 12889–12894. CrossRef Search ADS PubMed  Inoue K, Ogonuki N, Yamamoto Y, Takano K, Miki H, Mochida K and Ogura A ( 2004) Tissue-specific distribution of donor mitochondrial DNA in cloned mice produced by somatic cell nuclear transfer. Genesis  39, 79–83. CrossRef Search ADS PubMed  Irwin MH, Johnson LW and Pinkert CA ( 1999) Isolation and microinjection of somatic cell derived mitochondria and germline heteroplasmy in transmitochondrial mice. Transgenic Res  8, 119–123. CrossRef Search ADS PubMed  Jansen RP and de Boer K ( 1998) The bottleneck: mitochondrial imperatives in oogenesis and ovarian follicular fate. Mol Cell Endocrinol  145, 81–88. CrossRef Search ADS PubMed  Kaneda H, Hayashi J, Takahama S, Taya C, Lindahl KF and Yonekawa H ( 1995) Elimination of paternal mitochondrial DNA in intraspecific crosses during early mouse embryogenesis. Proc Natl Acad Sci USA  92, 4542–4546. CrossRef Search ADS PubMed  Keefe DL, Niven-Fairchild T, Powell S and Buradagunta S ( 1995) Mitochondrial deoxyribonucleic acid deletions in oocytes and reproductive aging in women. Fertil Steril  64, 577–583. CrossRef Search ADS PubMed  Kenyon L and Moraes CT ( 1997) Expanding the functional human mitochondrial DNA database by the establishment of primate xenomitochondrial cybrids. Proc Natl Acad Sci USA  94, 9131–9135. CrossRef Search ADS PubMed  King MP and Attardi G ( 1989) Human cells lacking mtDNA: repopulation with exogenous mitochondria by complementation. Science  246, 500–503. CrossRef Search ADS PubMed  Kirches E, Michael M, Warich-Kirches M, Schneider T, Weis S, Krause G, Mawrin C and Dietzmann K ( 2001) Heterogeneous tissue distribution of a mitochondrial DNA polymorphism in heteroplasmic subjects without mitochondrial disorders. J Med Genet  38, 312–317. CrossRef Search ADS PubMed  Klopstock T, Bischof F, Gerok K, Deuschl G, Seibel P, Ketelsen UP and Reichmann H ( 1995) 3.1-kb deletion of mitochondrial DNA in a patient with Kearns–Sayre syndrome. Acta Neuropathol  90, 126–129. CrossRef Search ADS PubMed  Koehler CM, Lindberg GL, Brown DR, Beitz DC, Freeman AE, Mayfield JE and Myers AM ( 1991) Replacement of bovine mitochondrial DNA by a sequence variant within one generation. Genetics  129, 247–255. PubMed  Kondo R, Satta Y, Matsuura ET, Ishiwa H, Takahata N and Chigusa SI ( 1990) Incomplete maternal transmission of mitochondrial DNA in Drosophila. Genetics  126, 657–663. PubMed  Kondo R, Matsuura ET and Chigusa SI ( 1992) Further observation of paternal transmission of Drosophila mitochondrial DNA by PCR selective amplification method. Genet Res  59, 81–84. CrossRef Search ADS PubMed  Kong LH, Liu Z, Li H, Zhu L and Xing FQ ( 2003) Pregnancy in a 46-year-old woman after autologous granular cell mitochondria transfer. Di Yi Jun Yi Da Xue Xue Bao  23, 743–747. Kong LH, Liu Z, Li H, Zhu L, Chen SM, Chen SL and Xing FQ ( 2004) Mitochondria transfer from self-granular cells to improve embryos’ quality. Zhonghua Fu Chan Ke Za Zhi  39, 105–107. PubMed  Kraytsberg Y, Schwartz M, Brown TA, Ebralidse K, Kunz WS, Clayton DA, Vissing J and Khrapko K ( 2004) Recombination of human mitochondrial DNA. Science  304, 981. CrossRef Search ADS PubMed  Kvist L, Martens J, Nazarenko AA and Orell M ( 2003) Paternal leakage of mitochondrial DNA in the great tit (Parus major) Mol Biol Evol  20, 243–247. CrossRef Search ADS PubMed  Laflamme MA, Gold J, Xu C, Hassanipour M, Rosler E, Police S, Muskheli V and Murry CE ( 2005) Formation of human myocardium in the rat heart from human embryonic stem cells. Am J Pathol  167, 663–671. CrossRef Search ADS PubMed  Larsson N-G, Garman JD, Oldfors A, Barsh GS and Clayton DA ( 1996) A single mouse gene encodes the mitochondrial transcription factor A and a testis-specific nuclear HMG-box protein. Nat Genet  13, 296–302. CrossRef Search ADS PubMed  Larsson NG, Oldfors A, Holme E and Clayton DA ( 1994) Low levels of mitochondrial transcription factor A in mitochondrial DNA depletion. Biochem Biophys Res Commun  200, 1374–1381. CrossRef Search ADS PubMed  Larsson N, Oldfors A, Garman J, Barsh G and Clayton D ( 1997) Down-regulation of mitochondrial transcription factor A during spermatogenesis in humans. Hum Mol Genet  6, 185–191. CrossRef Search ADS PubMed  Larsson NG, Wang J, Wilhelmsson H, Oldfors A, Rustin P, Lewandoski M, Barsh GS and Clayton DA ( 1998) Mitochondrial transcription factor A is necessary for mtDNA maintenance and embryogenesis in mice. Nat Genet  18, 231–236. CrossRef Search ADS PubMed  Leese HJ, Conaghan J, Martin KL and Hardy K ( 1993) Early human embryo metabolism. Bioessays  15, 259–264. CrossRef Search ADS PubMed  Lertrit P, Imsumran A, Karnkirawattana P, Devahasdin V, Sangruchi T, Atchaneeyasakul L, Mungkornkarn C and Neungton N ( 1999) A unique 3.5-kb deletion of the mitochondrial genome in Thai patients with Kearns–Sayre syndrome. Hum Genet  105, 127–131. CrossRef Search ADS PubMed  Lestienne P, Reynier P, Chretien MF, Penisson-Besnier I, Malthiery Y and Rohmer V ( 1997) Oligoasthenospermia associated with multiple mitochondrial DNA rearrangements. Mol Hum Reprod  3, 811–814. CrossRef Search ADS PubMed  Levy SE, Waymire KG, Kim YL, MacGregor GR and Wallace DC ( 1999) Transfer of chloramphenicol-resistant mitochondrial DNA into the chimeric mouse. Transgenic Res  8, 137–145. CrossRef Search ADS PubMed  Lin DP, Huang CC, Wu HM, Cheng TC, Chen CI and Lee MS ( 2004) Comparison of mitochondrial DNA contents in human embryos with good or poor morphology at the 8-cell stage. Fertil Steril  81, 73–79. CrossRef Search ADS PubMed  Liu H, Wang CW, Grifo JA, Krey LC and Zhang J ( 1999) Reconstruction of mouse oocytes by germinal vesicle transfer: maturity of host oocyte cytoplasm determines meiosis. Hum Reprod  14, 2357–2361. CrossRef Search ADS PubMed  Liu L, Hammar K, Smith PJ, Inoue S and Keefe DL ( 2001) Mitochondrial modulation of calcium signaling at the initiation of development. Cell Calcium  30, 423–433. CrossRef Search ADS PubMed  Liu SZ, Zhou ZM, Chen T, Zhang YL, Wen DC, Kou ZH, Li ZD, Sun QY and Chen DY ( 2004) Blastocysts produced by nuclear transfer between chicken blastodermal cells and rabbit oocytes. Mol Reprod Dev  69, 296–302. CrossRef Search ADS PubMed  Lloyd RE, Lee JH, Alberio R, Bowles EJ, Ramalho-Santos J, Campbell KH and St. John JC ( 2006) Aberrant nucleo-cytoplasmic cross-talk results in donor cell mtDNA persistence in cloned embryos. Genetics. Epub ahead of print February 1, 2006. Ludwig M, Schroder AK and Diedrich K ( 2001) Impact of intracytoplasmic sperm injection on the activation and fertilization process of oocytes. Reprod Biomed Online  3, 230–240. CrossRef Search ADS PubMed  Luetjens CM, Payne C and Schatten G ( 1999) Non-random chromosome positioning in human sperm and sex chromosome anomalies following intracytoplasmic sperm injection. Lancet  353, 1240. CrossRef Search ADS PubMed  Machaty Z, Thompson JG, Abeydeera LR, Day BN and Prather RS ( 2001) Inhibitors of mitochondrial ATP production at the time of compaction improve development of in vitro produced porcine embryos. Mol Reprod Dev  58, 39–44. CrossRef Search ADS PubMed  Magnusson J, Orth M, Lestienne P and Taanman JW ( 2003) Replication of mitochondrial DNA occurs throughout the mitochondria of cultured human cells. Exp Cell Res  289, 133–142. CrossRef Search ADS PubMed  Mancuso M, Filosto M, Mootha VK, Rocchi A, Pistolesi S, Murri L, DiMauro S and Siciliano G ( 2004) A novel mitochondrial tRNAPhe mutation causes MERRF syndrome. Neurology  62, 2119–2121. CrossRef Search ADS PubMed  Marchington DR, Brown MSS, Lamb VK, van Golde RJ, Kremer JA, Tuerlings JH, Mariman EC, Balen AH and Poulton J ( 2002) No evidence for paternal mtDNA transmission to offspring or extra-embryonic tissues after ICSI. Mol Hum Reprod  8, 1046–1049. CrossRef Search ADS PubMed  Margineantu DH, Cox WG, Sundell L, Sherwood SW, Beechem JM and Capaldi RA ( 2002) Cell cycle dependent morphology changes and associated mitochondrial DNA redistribution in mitochondria of human cell lines. Mitochondrion  1, 425–435. CrossRef Search ADS PubMed  May-Panloup P, Chretien MF, Jacques C, Vasseur C, Malthiery Y and Reynier P ( 2005) Low oocyte mitochondrial DNA content in ovarian insufficiency. Hum Reprod  20, 593–597. CrossRef Search ADS PubMed  McConnell JM and Petrie L ( 2004) Mitochondrial DNA turnover occurs during preimplantation development and can be modulated by environmental factors. Reprod Biomed Online  9, 418–424. CrossRef Search ADS PubMed  McCreath KJ, Howcroft J, Campbell KH, Colman A, Schnieke AE and Kind AJ ( 2000) Production of gene-targeted sheep by nuclear transfer from cultured somatic cells. Nature  405, 1066–1069. CrossRef Search ADS PubMed  McKenzie M and Trounce I ( 2000) Expression of Rattus norvegicus mtDNA in Mus musculus cells results in multiple respiratory chain defects. J Biol Chem  275, 31514–31519. CrossRef Search ADS PubMed  McKenzie M, Chiotis M, Pinkert CA and Trounce IA ( 2003) Functional respiratory chain analyses in murid xenomitochondrial cybrids expose coevolutionary constraints of cytochrome b and nuclear subunits of complex III. Mol Biol Evol  20, 1117–1124. CrossRef Search ADS PubMed  McKenzie M, Trounce IA, Cassar CA and Pinkert CA ( 2004) Production of homoplasmic xenomitochondrial mice. Proc Natl Acad Sci USA  101, 1685–1690. CrossRef Search ADS PubMed  Meirelles FV and Smith LC ( 1997) Mitochondrial genotype segregation in a mouse heteroplasmic lineage produced by embryonic karyoplast transplantation. Genetics  145, 445–451. PubMed  Meirelles FV and Smith LC ( 1998) Mitochondrial genotype segregation during preimplantation development in mouse heteroplasmic embryos. Genetics  148, 877–883. PubMed  Meirelles FV, Bordignon V, Watanabe Y, Watanabe M, Dayan A, Lobo RB, Garcia JM and Smith LC ( 2001) Complete replacement of the mitochondrial genotype in a Bos indicus calf reconstructed by nuclear transfer to a Bos taurus oocyte. Genetics  158, 351–356. PubMed  Meusel MS and Moritz RF ( 1993) Transfer of paternal mitochondrial DNA during fertilization of honeybee (Apis mellifera L.) eggs. Curr Genet  24, 539–543. CrossRef Search ADS PubMed  Michaels GS, Hauswirth WW and Laipis PJ ( 1982) Mitochondrial DNA copy number in bovine oocytes and somatic cells. Dev Biol  94, 246–251. CrossRef Search ADS PubMed  Monnat RJ Jr, Maxwell CL and Loeb LA ( 1985) Nucleotide sequence preservation of human leukemic mitochondrial DNA. Cancer Res  45, 1809–1814. PubMed  Moyes CD, Battersby BJ and Leary SC ( 1998) Regulation of muscle mitochondrial design. J Exp Biol  201, 299–307. Munne S, Alikani M, Tomkin G, Grifo J and Cohen J ( 1995) Embryo morphology, developmental rates, and maternal age are correlated with chromosome abnormalities. Fertil Steril  64, 382–391. CrossRef Search ADS PubMed  Munsie MJ, Michalska AE, O’Brien CM, Trounson AO, Pera MF and Mountford PS ( 2000) Isolation of pluripotent embryonic stem cells from reprogrammed adult mouse somatic cell nuclei. Curr Biol  10, 989–992. CrossRef Search ADS PubMed  Nagao Y, Totsuka Y, Atomi Y, Kaneda H, Lindahl KF, Imai H and Yonekawa H ( 1998) Decreased physical performance of congenic mice with mismatch between the nuclear and the mitochondrial genome. Genes Genet Syst  73, 21–27. CrossRef Search ADS PubMed  Naini AB, Lu J, Kaufmann P, Bernstein RA, Mancuso M, Bonilla E, Hirano M and DiMauro S ( 2005) Novel mitochondrial DNA ND5 mutation in a patient with clinical features of MELAS and MERRF. Arch Neurol  62, 473–476. CrossRef Search ADS PubMed  Nakamura M, Nakano S, Goto Y, Ozawa M, Nagahama Y, Fukuyama H, Akiguchi I, Kaji R and Kimura J ( 1995) A novel point mutation in the mitochondrial tRNA (Ser (UCN)) gene detected in a family with MERRF/MELAS overlap syndrome. Biochem Biophys Res Commun  214, 86–93. CrossRef Search ADS PubMed  Naviaux RK, Nyhan WL, Barshop BA, Poulton J, Markusic D, Karpinski NC and Haas RH ( 1999) Mitochondrial DNA polymerase gamma deficiency and mtDNA depletion in a child with Alpers’ syndrome. Ann Neurol  45, 54–58. CrossRef Search ADS PubMed  Nishi Y, Takeshita T, Sato K and Araki T ( 2003) Change of the mitochondrial distribution in mouse ooplasm during in vitro maturation. J Nippon Med Sch  70, 408–415. CrossRef Search ADS PubMed  Nishigaki Y, Tadesse S, Bonilla E, Shungu D, Hersh S, Keats BJ, Berlin CI, Goldberg MF, Vockley J, DiMauro S et al. ( 2003) A novel mitochondrial tRNA (Leu (UUR)) mutation in a patient with features of MERRF and Kearns–Sayre syndrome. Neuromuscul Disord  13, 334–340. CrossRef Search ADS PubMed  Nishimura Y, Yoshinari T, Naruse K, Yamada T, Sumi K, Mitani H, Higashiyama T and Kuroiwa T ( 2006) Active digestion of sperm mitochondrial DNA in single living sperm revealed by optical tweezers. Proc Natl Acad Sci USA  103, 1382–1387. CrossRef Search ADS PubMed  Normile D and Mann C ( 2005) Asian countries permit research, with safeguards. Science  307, 664. CrossRef Search ADS PubMed  Office of Legislative Policy and Analysis ( 2003) Human Cloning Prohibition Act of 2003. Available from http://olpa.od.nih.gov/tracking/house_bills/session1/hr-534.asp. Google Scholar Ojala D, Montoya J and Attardi G ( 1981) tRNA punctuation model of RNA processing in human mitochondria. Nature  290, 470–474. CrossRef Search ADS PubMed  Ono T, Isobe K, Nakada K and Hayashi JI ( 2001) Human cells are protected from mitochondrial dysfunction by complementation of DNA products in fused mitochondria. Nat Genet  28, 272–275. CrossRef Search ADS PubMed  Palermo GD, Takeuchi T and Rosenwaks Z ( 2002) Technical approaches to correction of oocyte aneuploidy. Hum Reprod  17, 2165–2173. CrossRef Search ADS PubMed  Pfeiffer T, Schuster S and Bonhoeffer S ( 2001) Cooperation and competition in the evolution of ATP-producing pathways. Science  292, 504–507. CrossRef Search ADS PubMed  Piko L and Matsumoto L ( 1976) Number of mitochondria and some properties of mitochondrial DNA in the mouse egg. Dev Biol  49, 1–10. CrossRef Search ADS PubMed  Piko L and Taylor KD ( 1987) Amounts of mitochondrial DNA and abundance of some mitochondrial gene transcripts in early mouse embryos. Dev Biol  123, 364–374. CrossRef Search ADS PubMed  Ponzetto C, Bresolin N, Bordoni A, Moggio M, Meola G, Bet L, Prelle A and Scarlato G ( 1990) Kearns–Sayre syndrome: different amounts of deleted mitochondrial DNA are present in several autoptic tissues. J Neurol Sci  96, 207–210. CrossRef Search ADS PubMed  Poulton J, Morten K, Freeman-Emmerson C, Potter C, Sewry C, Dubowitz V, Kidd H, Stephenson J, Whitehouse W, Hansen FJ et al. ( 1994) Deficiency of the human mitochondrial transcription factor h-mtTFA in infantile mitochondrial myopathy is associated with mtDNA depletion. Hum Mol Genet  3, 1763–1769. CrossRef Search ADS PubMed  Reynier P, Chretien MF, Penisson-Besnier I, Malthiery Y, Rohmer V and Lestienne P ( 1997) Male infertility associated with multiple mitochondrial DNA rearrangements. C R Acad Sci III  320, 629–636. CrossRef Search ADS PubMed  Reynier P, May-Panloup P, Chretien MF, Morgan CJ, Jean M, Savagner F, Barriere P and Malthiery Y ( 2001) Mitochondrial DNA content affects the fertilizability of human oocytes. Mol Hum Reprod  7, 425–429. CrossRef Search ADS PubMed  Rideout WM 3rd, Hochedlinger K, Kyba M, Daley GQ and Jaenisch R ( 2002) Correction of a genetic defect by nuclear transplantation and combined cell and gene therapy. Cell  109, 17–27. CrossRef Search ADS PubMed  Rubart M and Field LJ ( 2006) Cardiac repair by embryonic stem-derived cells. Handb Exp Pharmacol  174, 73–100. PubMed  Ruiz-Pesini E, Lapena AC, Diez-Sanchez C, Perez-Martos A, Montoya J, Alvarez E, Diaz M, Urries A, Montoro L, Lopez-Perez MJ et al. ( 2000) Human mtDNA haplogroups associated with high or reduced spermatozoa motility. Am J Hum Genet  67, 682–696. CrossRef Search ADS PubMed  Sathananthan AH and Trounson AO ( 2000) Mitochondrial morphology during preimplantational human embryogenesis. Hum Reprod  15, 148–159. CrossRef Search ADS PubMed  Sathananthan H, Pera M and Trounson A ( 2002) The fine structure of human embryonic stem cells. Reprod Biomed Online  4, 56–61. CrossRef Search ADS PubMed  Sato A, Kono T, Nakada K, Ishikawa K, Inoue SI, Yonekawa H and Hayashi JI ( 2005) Gene therapy for progeny of mito-mice carrying pathogenic mtDNA by nuclear transplantation. Proc Natl Acad Sci USA  102, 16765–16770. CrossRef Search ADS PubMed  Satoh M and Kuroiwa T ( 1991) Organization of multiple nucleoids and DNA molecules in mitochondria of a human cell. Exp Cell Res  196, 137–140. CrossRef Search ADS PubMed  Sazer S and Sherwood SW ( 1990) Mitochondrial growth and DNA synthesis occur in the absence of nuclear DNA replication in fission yeast. J Cell Sci  97, 509–516. PubMed  Scarpulla RC ( 2002) Nuclear activators and coactivators in mammalian mitochondrial biogenesis. Biochim Biophys Acta  1576, 1–14. CrossRef Search ADS PubMed  Schon EA, Kim SH, Ferreira JC, Magalhaes P, Grace M, Warburton D and Gross SJ ( 2000) Chromosomal non-disjunction in human oocytes: is there a mitochondrial connection? Hum Reprod  15, 160–172. CrossRef Search ADS PubMed  Schwartz M and Vissing J ( 2002) Paternal inheritance of mitochondrial DNA. N Engl J Med  347, 576–580. CrossRef Search ADS PubMed  Seifer DB, DeJesus V and Hubbard K ( 2002) Mitochondrial deletions in luteinized granulosa cells as a function of age in women undergoing in vitro fertilization. Fertil Steril  78, 1046–1048. CrossRef Search ADS PubMed  Sherratt HS ( 1991) Mitochondria: structure and function. Rev Neurol (Paris)  147, 417–430. PubMed  Shitara H, Kaneda H, Sato A, Inoue K, Ogura A, Yonekawa H and Hayashi JI ( 2000) Selective and continuous elimination of mitochondria microinjected into mouse eggs from spermatids, but not from liver cells, occurs throughout embryogenesis. Genetics  156, 1277–1284. PubMed  Shoubridge EA ( 2000) Mitochondrial DNA segregation in the developing embryo. Hum Reprod  15, 229–234. CrossRef Search ADS PubMed  Siciliano G, Mancuso M, Pasquali L, Manca ML, Tessa A and Iudice A ( 2000) Abnormal levels of human mitochondrial transcription factor A in skeletal muscle in mitochondrial encephalomyopathies. Neurol Sci  21,S 985–S987. CrossRef Search ADS   Smith LC and Alcivar AA ( 1993) Cytoplasmic inheritance and its effects on development and performance. J Reprod Fertil Suppl  48, 31–43. PubMed  Sokolova VA, Kustova ME, Arbuzova NI, Sorokin AV, Moskaliova OS, Bass MG and Vasilyev VB ( 2004) Obtaining mice that carry human mitochondrial DNA transmitted to the progeny. Mol Reprod Dev  68, 299–307. CrossRef Search ADS PubMed  Spelbrink JN, Van Galen MJ, Zwart R, Bakker HD, Rovio A, Jacobs HT and Van den Bogert C ( 1998) Familial mitochondrial DNA depletion in liver: haplotype analysis of candidate genes. Hum Genet  102, 327–331. CrossRef Search ADS PubMed  Spiropoulos J, Turnbull DM and Chinnery PF ( 2002) Can mitochondrial DNA mutations cause sperm dysfunction? Mol Hum Reprod  8, 719–721. CrossRef Search ADS PubMed  St. John JC ( 2002) Ooplasm donation in humans: the need to investigate the transmission of mitochondrial DNA following cytoplasmic transfer. Hum Reprod  17, 1954–1958. CrossRef Search ADS PubMed  St. John JC and Schatten G ( 2004) Paternal mitochondrial DNA transmission during nonhuman primate nuclear transfer. Genetics  167, 897–905. CrossRef Search ADS PubMed  St. John J, Sakkas D, Dimitriadi K, Barnes A, Maclin V, Ramey J, Barratt C and De Jonge C ( 2000) Failure of elimination of paternal mitochondrial DNA in abnormal embryos. Lancet  355, 200. CrossRef Search ADS   St. John JC, Jokhi RP and Barratt CL ( 2001) Men with oligoasthenoteratozoospermia harbour higher numbers of multiple mitochondrial DNA deletions in their spermatozoa, but individual deletions are not indicative of overall aetiology. Mol Hum Reprod  7, 103–111. CrossRef Search ADS PubMed  St. John JC, Lloyd R and El Shourbagy S ( 2004) The potential risks of abnormal transmission of mtDNA through assisted reproductive technologies. Reprod Biomed Online  8, 34–44. CrossRef Search ADS PubMed  St. John JC, Lloyd RE, Bowles EJ, Thomas EC and El Shourbagy S ( 2004) The consequences of nuclear transfer for mammalian foetal development and offspring survival. A mitochondrial DNA perspective. Reproduction  127, 631–641. CrossRef Search ADS PubMed  St. John JC, Ramalho-Santos J, Gray HL, Petrosko P, Rawe VY, Navara CS, Simerly CR and Schatten GP ( 2005) The expression of mitochondrial DNA transcription factors during early cardiomyocyte in vitro differentiation from human embryonic stem cells. Cloning Stem Cells  7, 141–153. CrossRef Search ADS PubMed  St. John J, Moffatt O and D’Souza N ( 2005) Aberrant heteroplasmic transmission of mtDNA in cloned pigs arising from double nuclear transfer. Mol Reprod Dev  72, 450–460. CrossRef Search ADS PubMed  Steinborn R, Zakhartchenko V, Jelyazkov J, Klein D, Wolf E, Muller M and Brem G ( 1998) Composition of parental mitochondrial DNA in cloned bovine embryos. FEBS Lett  426, 352–356. CrossRef Search ADS PubMed  Steinborn R, Schinogl P, Wells DN, Bergthaler A, Muller M and Brem G ( 2002) Coexistence of Bos taurus and B. indicus mitochondrial DNAs in nuclear transfer-derived somatic cattle clones. Genetics  162, 823–829. PubMed  Steuerwald N, Barritt JA, Adler R, Malter H, Schimmel T, Cohen J and Brenner CA ( 2000) Quantification of mtDNA in single oocytes, polar bodies and subcellular components by real-time rapid cycle fluorescence monitored PCR. Zygote  8, 209–215. CrossRef Search ADS PubMed  Stojanovski D, Johnston AJ, Streimann I, Hoogenraad NJ and Ryan MT ( 2003) Import of nuclear-encoded proteins into mitochondria. Exp Physiol  88, 57–64. CrossRef Search ADS PubMed  Stojkovic M, Machado SA, Stojkovic P, Zakhartchenko V, Hutzler P, Goncalves PB and Wolf E ( 2001) Mitochondrial distribution and adenosine triphosphate content of bovine oocytes before and after in vitro maturation: correlation with morphological criteria and developmental capacity after in vitro fertilization and culture. Biol Reprod  64, 904–909. CrossRef Search ADS PubMed  Sturmey RG and Leese HJ ( 2003) Energy metabolism in pig oocytes and early embryos. Reproduction  126, 197–204. CrossRef Search ADS PubMed  Sun QY, Wu GM, Lai L, Park KW, Cabot R, Cheong HT, Day BN, Prather RS and Schatten H ( 2001) Translocation of active mitochondria during pig oocyte maturation, fertilization and early embryo development in vitro. Reproduction  122, 155–163. CrossRef Search ADS PubMed  Sun YH, Chen SP, Wang YP, Hu W and Zhu ZY ( 2005) Cytoplasmic impact on cross-genus cloned fish derived from transgenic common carp (Cyprinus carpio) nuclei and goldfish (Carassius auratus) enucleated eggs. Biol Reprod  72, 510–515. CrossRef Search ADS PubMed  Suomalainen A and Kaukonen J ( 2001) Diseases caused by nuclear genes affecting mtDNA stability. Am J Med Genet  106, 53–61. CrossRef Search ADS PubMed  Sutarno, Cummins JM, Greeff J and Lymbery AJ ( 2002) Mitochondrial DNA polymorphisms and fertility in beef cattle. Theriogenology  57, 1603–1610. CrossRef Search ADS PubMed  Google Scholar Sutovsky P ( 2003) Ubiquitin-dependent proteolysis in mammalian spermatogenesis, fertilization, and sperm quality control: killing three birds with one stone. Microsc Res Tech  61, 88–102. CrossRef Search ADS PubMed  Sutovsky P, Moreno RD, Ramalho-Santos J, Dominko T, Simerly C and Schatten G ( 1999) Ubiquitin tag for sperm mitochondria. Nature  402, 371–372. CrossRef Search ADS PubMed  Takeda K, Akagi S, Kaneyama K, Kojima T, Takahashi S, Imai H, Yamanaka M, Onishi A and Hanada H ( 2003) Proliferation of donor mitochondrial DNA in nuclear transfer calves (Bos taurus) derived from cumulus cells. Mol Reprod Dev  64, 429–437. CrossRef Search ADS PubMed  Takeda K, Tasai M, Iwamoto M, Onishi A, Tagami T, Nirasawa K, Hanada H and Pinkert CA ( 2005) Microinjection of cytoplasm or mitochondria derived from somatic cells affects parthenogenetic development of murine oocytes. Biol Reprod  72, 1397–1404. CrossRef Search ADS PubMed  Takeda K, Tasai M, Iwamoto M, Akita T, Tagami T, Nirasawa K, Hanada H and Onishi A ( 2006) Transmission of mitochondrial DNA in pigs and progeny derived from nuclear transfer of Meishan pig fibroblast cells. Mol Reprod Dev  73, 306–312. CrossRef Search ADS PubMed  Takeuchi T, Gong J, Veeck LL, Rosenwaks Z and Palermo GD ( 2001) Preliminary findings in germinal vesicle transplantation of immature human oocytes. Hum Reprod  16, 730–736. CrossRef Search ADS PubMed  Takeuchi T, Neri QV, Katagiri Y, Rosenwaks Z and Palermo GD ( 2005) Effect of treating induced mitochondrial damage on embryonic development and epigenesis. Biol Reprod  72, 584–592. CrossRef Search ADS PubMed  Taylor H and Minger SL ( 2005) Regenerative medicine in Parkinson’s disease: generation of mesencephalic dopaminergic cells from embryonic stem cells. Curr Opin Biotechnol  16, 487–492. CrossRef Search ADS PubMed  Tessa A, Manca ML, Mancuso M, Renna MR, Murri L, Martini B, Santorelli FM and Siciliano G ( 2000) Abnormal H-Tfam in a patient harboring a single mtDNA deletion. Funct Neurol  15, 211–214. PubMed  Thompson JG, Partridge RJ, Houghton FD, Cox CI and Leese HJ ( 1996) Oxygen uptake and carbohydrate metabolism by in vitro derived bovine embryos. J Reprod Fertil  106, 299–306. CrossRef Search ADS PubMed  Tiranti V, Savoia A, Forti F, D’Apolito MF, Centra M, Rocchi M and Zeviani M ( 1997) Identification of the gene encoding the human mitochondrial RNA polymerase (h-mtRPOL) by cyberscreening of the Expressed Sequence Tags database. Hum Mol Genet  6, 615–625. CrossRef Search ADS PubMed  Trounce IA, McKenzie M, Cassar CA, Ingraham CA, Lerner CA, Dunn DA, Donegan CL, Takeda K, Pogozelski WK, Howell RL et al. ( 2004) Development and initial characterization of xenomitochondrial mice. J Bioenerg Biomembr  36, 421–427. CrossRef Search ADS PubMed  Trounson A ( 2001) Nuclear transfer in human medicine and animal breeding. Reprod Fertil Dev  13, 31–39. CrossRef Search ADS PubMed  Tsaousis AD, Martin DP, Ladoukakis ED, Posada D and Zouros E ( 2005) Widespread recombination in published animal mtDNA sequences. Mol Biol Evol  22, 925–933. CrossRef Search ADS PubMed  Vajta G, Lewis IM, Hyttel P, Thouas GA and Trounson AO ( 2001) Somatic cell cloning without micromanipulators. Cloning  3, 89–95. CrossRef Search ADS PubMed  Van Blerkom J, Davis PW and Lee J ( 1995) ATP content of human oocytes and developmental potential and outcome after in-vitro fertilization and embryo transfer. Hum Reprod  10, 415–424. PubMed  Van Blerkom J, Davis P and Alexander S ( 2000) Differential mitochondrial distribution in human pronuclear embryos leads to disproportionate inheritance between blastomeres: relationship to microtubular organization, ATP content and competence. Hum Reprod  15, 2621–2633. CrossRef Search ADS PubMed  Van Blerkom J, Davis P, Mathwig V and Alexander S ( 2002) Domains of high-polarized and low-polarized mitochondria may occur in mouse and human oocytes and early embryos. Hum Reprod  17, 393–406. CrossRef Search ADS PubMed  Van Goethem G, Dermaut B, Lofgren A, Martin JJ and Van Broeckhoven C ( 2001) Mutation of POLG is associated with progressive external ophthalmoplegia characterized by mtDNA deletions. Nat Genet  28, 211–212. CrossRef Search ADS PubMed  Wakayama T, Tabar V, Rodriguez I, Perry AC, Studer L and Mombaerts P ( 2001) Differentiation of embryonic stem cell lines generated from adult somatic cells by nuclear transfer. Science  292, 740–743. CrossRef Search ADS PubMed  Wallace DC ( 1992) Diseases of the mitochondrial DNA. Annu Rev Biochem  61, 1175–1212. CrossRef Search ADS PubMed  Wallace DC, Ye JH, Neckelmann SN, Singh G, Webster KA and Greenberg BD ( 1987) Sequence analysis of cDNAs for the human and bovine ATP synthase beta subunit: mitochondrial DNA genes sustain seventeen times more mutations. Curr Genet  12, 81–90. CrossRef Search ADS PubMed  Wernig M, Benninger F, Schmandt T, Rade M, Tucker KL, Bussow H, Beck H and Brustle O ( 2004) Functional integration of embryonic stem cell-derived neurons in vivo. J Neurosci  24, 5258–5268. CrossRef Search ADS PubMed  White KL, Bunch TD, Mitalipov S and Reed WA ( 1999) Establishment of pregnancy after the transfer of nuclear transfer embryos produced from the fusion of argali (Ovis ammon) nuclei into domestic sheep (Ovis aries) enucleated oocytes. Cloning  1, 47–54. CrossRef Search ADS PubMed  White DJ, Mital D, Taylor S and St John JC ( 2001) Sperm mitochondrial DNA deletions as a consequence of long term highly active antiretroviral therapy. Aids  15, 1061–1062. CrossRef Search ADS PubMed  Wiesner RJ, Ruegg JC and Morano I ( 1992) Counting target molecules by exponential polymerase chain reaction: copy number of mitochondrial DNA in rat tissues. Biochem Biophys Res Commun  183, 553–559. CrossRef Search ADS PubMed  Wilding M, Dale B, Marino M, di Matteo L, Alviggi C, Pisaturo ML, Lombardi L and De Placido G ( 2001) Mitochondrial aggregation patterns and activity in human oocytes and preimplantation embryos. Hum Reprod  16, 909–917. CrossRef Search ADS PubMed  Wilmut I, Schnieke AE, McWhir J, Kind AJ and Campbell KH ( 1997) Viable offspring derived from fetal and adult mammalian cells. Nature  385, 810–813. CrossRef Search ADS PubMed  Xu B and Clayton DA ( 1995) A persistent RNA-DNA hybrid is formed during transcription at a phylogenetically conserved mitochondrial DNA sequence. Mol Cell Biol  15, 580–589. CrossRef Search ADS PubMed  Yakes FM and Van Houten B ( 1997) Mitochondrial DNA damage is more extensive and persists longer than nuclear DNA damage in human cells following oxidative stress. Proc Natl Acad Sci USA  94, 514–519. CrossRef Search ADS PubMed  Yang CX, Han ZM, Wen DC, Sun QY, Zhang KY, Zhang LS, Wu YQ, Kou ZH and Chen DY ( 2003) In vitro development and mitochondrial fate of macaca-rabbit cloned embryos. Mol Reprod Dev  65, 396–401. CrossRef Search ADS PubMed  Yang L, Chavatte-Palmer P, Kubota C, O’Neill M, Hoagland T, Renard JP, Taneja M, Yang X and Tian C ( 2005) Expression of imprinted genes is aberrant in deceased newborn cloned calves and relatively normal in surviving adult clones. Mol Reprod Dev  71, 431–438. CrossRef Search ADS PubMed  Zhang J, Wang CW, Krey L, Liu H, Meng L, Blaszczyk A, Adler A and Grifo J ( 1999) In vitro maturation of human preovulatory oocytes reconstructed by germinal vesicle transfer. Fertil Steril  71, 726–731. CrossRef Search ADS PubMed  Zhao X, Li N, Guo W, Hu X, Liu Z, Gong G, Wang A, Feng J and Wu C ( 2004) Further evidence for paternal inheritance of mitochondrial DNA in the sheep (Ovis aries) Heredity  93, 399–403. Zouros E, Freeman KR, Ball AO and Pogson GH ( 1992) Direct evidence for extensive paternal mitochondrial DNA inheritance in the marine mussel Mytilus. Nature  359, 412–414. CrossRef Search ADS PubMed  © The Author 2006. Published by Oxford University Press on behalf of the European Society of Human Reproduction and Embryology. All rights reserved. For Permissions, please email: journals.permissions@oxfordjournals.org

Journal

Human Reproduction UpdateOxford University Press

Published: Mar 31, 2006

There are no references for this article.