Mitochondrial Genomes Anything Goes Trends Genet 1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/8987066

Mitochondrial genomes: anything goes. Trends Genet 19:709-716

Article  in  Trends in Genetics · January 2004


DOI: 10.1016/j.tig.2003.10.012 · Source: PubMed

CITATIONS READS
441 424

3 authors, including:

Gertraud Burger
Université de Montréal
274 PUBLICATIONS   12,639 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Perfection of eccentricity: Mitochondrial genomes of diplonemids View project

Mitochondrial Genome Structure View project

All content following this page was uploaded by Gertraud Burger on 30 June 2018.

The user has requested enhancement of the downloaded file.


Review TRENDS in Genetics Vol.19 No.12 December 2003 709

Mitochondrial genomes: anything goes


Gertraud Burger1, Michael W. Gray2 and B. Franz Lang1
1
Canadian Institute for Advanced Research, Programme in Evolutionary Biology, Départment de Biochimie, Université de Montréal,
2900 Boulevard Edouard-Montpetit, Montréal, Québec, Canada H3T 1J4
2
Canadian Institute for Advanced Research, Programme in Evolutionary Biology, Department of Biochemistry and Molecular
Biology, Dalhousie University, Halifax, Nova Scotia, Canada B3H 1X5

Mitochondria have their own genetic system – a vesti- The limited genetic function of mtDNA is well-con-
gial genome originating from an endosymbiotic served across eukaryotes: mtDNA encodes a small number
a-proteobacterial ancestor. The genetic function of of proteins whose mRNAs are translated by a distinctive
mitochondrial DNA (mtDNA) is well-conserved, being mitochondrial protein-synthesizing system, some of whose
involved in a maximum of five mitochondrial processes: components are always (rRNAs), usually (tRNAs) or
invariantly in respiration and/or oxidative phosphoryl- occasionally (ribosomal proteins) specified by the mito-
ation and translation, and also in transcription, RNA chondrial genome. In marked contrast to this genetic
maturation and protein import. Recent data from mito- conservatism (as this review will emphasize), the mito-
chondrial genomics research demonstrate that this gen- chondrial genome is characterized by an extreme and often
etic conservatism contrasts with an often-perplexing bewildering diversity in structure and gene expression
diversity in structure and gene expression mechanisms. mechanisms [5,6]. This dichotomy – genetic conservatism
In addition to outlining the extraordinary diversity of versus structural flamboyance – has long obscured the
mtDNA, this review highlights the divergent trends in origin of mitochondria and raised the question of whether
mitochondrial genome evolution in the various eukary- this organelle arose more than once during eukaryotic cell
otic lineages, and examines the relationship between evolution. The weight of evidence now firmly argues that
mitochondrial and nuclear genome evolution in a given the mitochondrion had a single origin, arising from a
organism. eubacterial symbiont whose closest contemporary rela-
tives are found within the a-proteobacteria.
Mitochondria, the morphologically distinctive double- This review summarizes recent data from mitochon-
membrane organelles of most cells, use electron transport drial genomics research. It aims to highlight the extra-
coupled with oxidative phosphorylation to generate ATP ordinary structural diversity shown by mtDNA across the
[1]. Although mitochondria have a central role in energy eukaryotic domain (Figure 2), and the exceptionally
transduction (the powerhouse of eukaryotic cells), these divergent trends in mitochondrial genome evolution in
organelles participate in several other important func- various eukaryotic lineages. Chloroplasts, a second type of
tions, including ion homeostasis, intermediary metab- DNA-containing organelle in algae and plants, will not be
olism and apoptosis. However, in only a few instances do discussed. It is worth noting that chloroplasts, which are
these pathways have components encoded by mitochon- thought to be the product of a more recent endosymbiotic
drial DNA (mtDNA) (Figure 1); rather, most mitochondrial event than mitochondria, have genomes that show
proteins are specified by nuclear genes, synthesized in the significantly less structural diversity than mitochondrial
cytosol and imported into the organelle [2]. genomes.
Mitochondria are ubiquitous throughout the eukaryotic
domain: with very few exceptions, eukaryotes cannot exist The many faces of mitochondrial genome architecture
without them. Eukaryotic lineages that lack both func- Given that the mitochondrial genome of several animals
tional mitochondria and a mitochondrial genome do exist, exists in the form of super-coiled DNA [7], and that mtDNA
but several of these ‘amitochondriate’ eukaryotes contain in most eukaryotes maps as a circle in restriction analysis
DERIVED (see Glossary) mitochondria (e.g. hydrogeno- and sequence assembly, it was assumed, until recently,
somes) that generate ATP anaerobically [3]. Other that mtDNAs are generally circular molecules. However,
amitochondriate species contain what appear to be there is now strong evidence that many, if not most,
remnant mitochondrial structures (of currently undefined
function), and/or express nucleus-encoded proteins typi-
Glossary
cally targeted to, and functioning in, mitochondria [4].
These observations argue that the last common ancestor of Derived: taxa or characters that have evolved far away from the primitive state.
In molecular phylogeny, long tree-branches indicate highly derived taxa.
extant eukaryotes was a mitochondria-containing organ- Homologous: similar in sequence owing to common evolutionary origin.
ism, implying that PRIMITIVE amitochondriate eukaryotes Primitive: original or ancient state.
Monophyly: unique origin in a single ancestral species (a monophyletic
(eukaryotes whose ancestors never had mitochondria)
grouping includes the ancestor and all of its descendants).
might not exist. Basal (early diverging): emerging early in the evolution of a clade (i.e. a deep
branch in a phylogenetic tree).
Corresponding author: Gertraud Burger (gertraud.burger@umontreal.ca).

http://tigs.trends.com 0168-9525/$ - see front matter q 2003 Elsevier Ltd. All rights reserved. doi:10.1016/j.tig.2003.10.012
710 Review TRENDS in Genetics Vol.19 No.12 December 2003

Lipid metabolism Heme synthesis Proteases


Nucleotide metabolism Fe-S synthesis Chaperones
Amino acid metabolism Ubiquinone synthesis Signalling pathways
Carbohydrate metabolism Co-factor synthesis DNA repair, replication, etc.

RNA Heme lyase RNAse P


polymerase

EF-Tu Ribosome

TIM Oxa1
translocases translocase

Sec Tat
translocase translocase

Complex I Complex V
Complex II Complex IV
Complex III
Cytochrome c
TRENDS in Genetics

Figure 1. Biological processes in mitochondria. The majority of mitochondrial functions is shown; only the mitochondrial matrix and inner membrane are depicted.
(For more details on mitochondrial metabolism and transport see [79].) Most mitochondrial components are encoded by the nucleus (blue); those highlighted in pink are
encoded by mtDNA in some eukaryotes but by the nuclear genome in other eukaryotes, whereas a small portion is invariably specified by mtDNA (orange). Processes that
have exclusively nucleus-encoded constituents are listed. Complexes I– V are involved in electron-transport and oxidative phosphorylation. TIM translocases are involved
in protein import and insertion into the inner membrane. Tat, Sec and Oxa1 translocases are involved in protein export from matrix and insertion into the inner membrane
(for a review see [2]). RNase P is a ribozyme that processes the 50 -end of tRNAs (for recent results on fungal RNase P RNAs see [42]). It should be noted that the majority of
eukaryotes has a multi-subunit, rotenone-sensitive NADH dehydrogenase as depicted, with nad1, nad4 and nad5 invariantly encoded by mtDNA. However, in a few organ-
isms (yeast, Schizosaccharomyces and Plasmodium), this complex is replaced by a nucleus-encoded, single-polypeptide enzyme [80]. Abbreviation: EF-Tu, elongation
factor Tu.

Holozoa
Amoebidium
Monosiga Saccharomyces
Metridium
Fugu Rhizopus Fungi
Homo Allomyces
Arabidopsis Harpochytrium
Green algae Marchantia Dictyostelium
and plants Pedinomonas
Amoebozoa
Acanthamoeba
Scenedesmus Plasmodium Apicomplexa
Chlamydomonas Tetrahymena Ciliates
Cyanidioschyzon Paramecium
Red algae
Porphyra Cafeteria
Cryptomonas Ochromonas Stramenopiles
Cryptophytes Phytophthora
Rhodomonas
Reclinomonas Trypanosoma
Jakobids Euglena
Euglenozoa
Jakoba

TRENDS in Genetics

Figure 2. Eukaryotic phylogeny. Schematic phylogenetic tree depicting relationships among eukaryotes based on evidence from both statistically well-supported molecular
phylogenies with different datasets [81], and from conjecture. Only the better-known, major organismal groups are shown, with taxa belonging to the same clade sharing
the same color shade. Broken lines indicate uncertain branching orders and tree topologies. For example, the Monophyly of apicomplexa plus ciliates and stramenopiles is
not well supported. Euglenozoa and jakobids have been suggested to be the most Basal (Early Diverging) lineages; however, the relative branching order at the base of the
tree (in excess of 1 billion years) remains largely unresolved.

http://tigs.trends.com
Review TRENDS in Genetics Vol.19 No.12 December 2003 711

circular-mapping mtDNAs consist primarily of linear, Size versus content


multimeric head-to-tail concatamers [8,9]. Linear map- The size of mtDNA in most eukaryotic phyla ranges from
ping (i.e. linear monomeric) mtDNA molecules have 15 – 60 kbp; however, there are some notable exceptions.
also been detected in a few unrelated organisms, such Among those organisms whose mtDNA has been comple-
as ciliates, apicomplexa (Plasmodium and relatives), tely sequenced, the two extreme outliers are the apicom-
fungi, chlorophycean green algae (Chlamydomonas and plexan protists Plasmodium sp. (obligate parasites
relatives), and several cnidarian animals. Notably, causing malaria), which have a minuscule mitochondrial
these linear molecules contain various specialized end- genome of 6 kbp [17], and rice (Oryza sativa), whose
structures, such as covalently closed single-stranded mtDNA (at 490 kbp [18]) is , 80 times larger than that of
DNA termini and terminally attached proteins; in Plasmodium. The mapped, but as-yet unsequenced
addition, they tend to have telomere-like repeats of mtDNAs of cucurbit plants could be even larger than
varying length [10]. 2000 kbp [19].
Mitochondrial genomes typically consist of a single The coding capacity of mtDNA ranges from nearly
‘chromosome’. Cases of mitochondrial genomes comprising 100 genes in jakobid flagellates [20] to only five in
multiple, circular-mapping molecules have been reported Plasmodium, with the average across eukaryotes being
40 – 50 genes. In contrast to their up-to-20-fold differ-
in Spizellomyecs punctatus [11], a chytridiomycete fungus,
ence in number, mitochondrial genes are involved, in at
and in the animals Globodera [12] and Dicyema [13]. A
most, five basic processes: invariantly in respiration
most unusual and, indeed, unprecedented situation has
and/or oxidative phosphorylation and translation, and
been identified in a single-celled protistan relative of
occasionally also in transcription, RNA maturation and
animals, Amoebidium parasiticum [14], whose mtDNA
protein import (Figure 4).
comprises a complex set of several hundred distinct types
Somewhat counter-intuitively, there is no correlation
of linear molecules [15] (Figure 3). These molecules carry between mtDNA size and gene content (Figure 5). In fact,
at their termini a common pattern of short terminal the differences in size of mitochondrial genomes are mostly
repeats, consisting invariably of the same ,40-bp term- caused by marked variations in the length and organiz-
inal sequence motif in inverted orientation, followed ation of intergenic regions, which, in some cases, consist of
internally by an , 100-bp repeat at one end, and a extensive tandem-repeat arrays [21,22] or stem-loop
different , 65-bp repeat at the other end. The large motifs [23]. Some of the repeat elements have been
number of distinct ‘chromosomes’ and the apparent proposed to be mobile and ‘selfish’ (e.g. [24]); others, to
absence of centromeric structures characteristic of nuclear play a role in recombination [25]. Another determinant of
chromosomes raise questions about the mechanism of mtDNA size is gene structure: a given gene product can be
concerted replication and equal segregation of mtDNA encoded by stretches of DNA of variable length owing to
during mitochondrial division and subsequent cell the presence of introns (group I and II) of various size
division in A. parasiticum. A similar dilemma exists in (0.15 –4.00 kbp) and number (0 to .30). In the filamentous
kinetoplastid protists (e.g. Trypanosoma), whose mito- fungus Podospora anserina, introns account for up to 75%
chondria contain up to a few dozen gene-coding maxicircles of the overall size of mtDNA [26].
and up to several thousand minicircles that specify the
guide RNAs involved in editing of mitochondrial Unusual mitochondrial gene structures
mRNAs [16]. In addition to genes split by introns, mitochondria show
other, more unusual gene organizations, one of which is
gene fusion. In two amoebozoan protists, Acanthamoeba
castellanii [27] and Dictyostelium discoideum [28], the
Tetrahymena mtDNA
genes coding for subunits 1 and 2 of cytochrome oxidase
(cox1 and cox2, respectively) are immediately adjacent to
one another and in the same transcriptional orientation in
the mtDNA, forming a single contiguous ORF (Figure 6a).
The existence of a bi-cistronic mRNA has been shown in
Acanthamoeba, but it is not clear at this time whether a
Human mtDNA fusion protein is initially synthesized as well [29].
Chimeric genes, often containing recognizable pieces of a
known gene (or genes) within an otherwise unassigned
10 kbp
Spizellomyces mtDNA Amoebidium mtDNA ORF, have been associated with cytoplasmic male sterility
TRENDS in Genetics
in plants, a phenotype that is probably attributable to
alterations in electron-transport-coupled oxidative phos-
Figure 3. Mitochondrial genome architectures. DNA molecules are mostly circular- phorylation in pollen-producing cells [30]. In other cases
supercoiled in humans, and linear-monomeric in the ciliate Tetrahymena pyrifor- (e.g. [31,32]), the functional consequences (if any) of
mis (reviewed in [10]). In the fungus Spizellomyces punctatus, mtDNA consists of
three types of circular-mapping molecules [6], whereas several hundred types of
reported gene fusions are unknown.
linear-monomeric molecules comprise the mitochondrial genome in the ichthyos- Genes that exist in pieces, which were first described for
porean protist Amoebidium parasiticum [15]. Triangles represent terminal repeat the mitochondrial rRNA genes in the ciliate protozoon
motifs of 32 bp in T. pyriformis mtDNA and 40 bp in A. parasiticum mtDNA. Filled
circles and diamonds represent sub-terminal repeats, which are ,100-bp and Tetrahymena pyriformis [33,34] and the green alga
, 65-bp long, respectively, in A. parasiticum mtDNA. Chlamydomonas reinhardtii [35], represent another type
http://tigs.trends.com
712 Review TRENDS in Genetics Vol.19 No.12 December 2003

100 400

90
350
80

Mitochondrial genome size (kbp)


70
200
Number of genes

60

50 150

40
100
30

20 50

10
0
0 is m tia ba es as as ria um mo as um
ps diu n ko yc n n te di n i
as as tia ria on sis iga mo um nas ces ium ido ebi rcha Ja om omo omo afe etri Ho mo od
on on han fete hyz dop os Ho ridi o y od ab o a r n od C M o
yd las
m
om dom arc Ca osc abi on et nom rom sm Ar Am M ha cli
n cc Re Rh am P
cli ho M id
i
Ar
M M di ha Pla a l
Re R an Pe acc
S Ch
Cy S Key:

Key: Authentic mitochondrial genes


Introns, ORFs, plasmid-derived genes
Respiration and oxidative phosphorylation
Non-coding regions
Ribosomal RNAs Protein import and maturation
Transfer RNAs RNA maturation TRENDS in Genetics
Ribosomal proteins and EF-Tu Transcription
Figure 5. Mitochondrial genome size and coding content across eukaryotes.
TRENDS in Genetics Length of coding regions of authentic mitochondrial genes (purple), introns, intro-
nic ORFs, phage-like reverse transcriptases and DNA polymerases (blue), and
Figure 4. Mitochondrial gene classes and their representation across eukaryotes. intergenic regions (green). Species as in Figure 4, plus Amoebidium parasiticum
Genes included in the functional classes are as follows. Respiration and oxidative (ichthyosporean protist); Jakoba libera (jakobid flagellate); and Chlamydomonas
phosphorylation: atp1, atp3, atp4, atp6, atp8, atp9; cob; cox1–3; nad1 –4, nad4L, reinhardtii (green alga, chlorophyte).
nad6 –11, sdh2 –4. rRNAs: rnl, rns, rrn5. tRNAs: trnA–Y (among others). Ribosomal
proteins and EF-Tu: rps1– 4, rps7, rps8, rps10– 14, rps19; rpl1, rpl5, rpl6, rpl10, translated into two separate polypeptides (Figure 6b);
rpl11, rpl14, rpl16, rpl18– 20, rpl27, rpl31, rpl32, rpl34, rpl36; tufA. RNA maturation:
rnpB. Protein import and maturation: secY, tatC, yejR (ccmF), yejU (ccmC), yejV
however, post-translational protein fusion can not be ruled
(ccmB), yejW (ccmA); cox11. Transcription: rpoA-D. Species names are: Reclino- out at present [37].
monas americana (jakobid flagellate); Rhodomonas salina (cryptophyte alga); A stunning variation on this theme is illustrated in two
Marchantia polymorpha (liverwort, bryophyte); Cafeteria roenbergensis (strame-
nopile flagellate); Cyanidioschyzon merolae (red alga); Arabidopsis thaliana (flow-
recent reports. In the green alga Scenedesmus obliquus,
ering plant, angiosperm); Monosiga brevicollis (choanozoan flagellate); Homo the mitochondrial cytochrome oxidase subunit 2 is
sapiens (vertebrate animal); Metridium senile (cnidarian animal); Pedinomonas apparently encoded by two gene pieces; that specifying
minor (green alga, chlorophyte); Saccharomyces cerevisiae (ascomycete fungus);
and Plasmodium falciparum (apicomplexan protist). Data taken from the Organelle
the N-terminal portion is located on mtDNA [38], whereas
Genome Database GOBASE (http://megasun.bch.umontreal.ca/gobase/).

(a) (b) (c)


of unorthodox gene structure in mitochondria. In some cox2-b
instances, these genes are broken up into as many as ncDNA
eight modules that are scrambled in the genome,
cox1 cox2 nad1-a cox2-a
interspersed with other genes and located on both mtDNA
strands of the mtDNA [35]. Separate transcription of nad1-b

these subgenic modules yields discrete pieces of rRNA Acanthamoeba Tetrahymena Scenedesmus
that are held together via base pairing of complemen- Dictyostelium Paramecium
tary sequence stretches. In P. falciparum, the number TRENDS in Genetics
of mtDNA-encoded rRNA pieces is even higher than in
Chlamydomonas [36]. Figure 6. Unusual structures of mitochondrial genes. (a) cox1 and cox2 genes spe-
cifying cytochrome oxidase subunits 1 and 2. (b) nad1 encodes NADH dehydro-
The genes of a few mtDNA-encoded proteins also genase subunit 1. Mitochondrial DNA (mtDNA) and nuclear DNA (ncDNA) are
display a ‘genes-in-pieces’ type of arrangement. In two shown in blue and purple, respectively. Suffixes -a and -b indicate the gene pieces
ciliate species, Tetrahymena pyriformis and Paramecium encoding the N- or C-terminal portion of a given protein, respectively. (c) cox2
gene, specifying cytochrome oxidase subunit 2. Species are: Acanthamoeba
aurelia, the nad1 gene is split into two fragments that are castellanii; Dictyostelium discoideum (amoebozoa); Tetrahymena pyriformis; Para-
transcribed into two separate mRNAs, which are probably mecium aurelia (ciliates); and Scenedesmus obliquus (chlorophycean green alga).

http://tigs.trends.com
Review TRENDS in Genetics Vol.19 No.12 December 2003 713

the C-terminal portion is believed to be nucleus-encoded eukaryotic groups, mitochondrial genomes are distributed
[39] (Figure 6c). In the green algae Chlamydomonas and over more than one mitochondrion per cell, and the
Polytomella, the corresponding two protein portions are mtDNA copy number ranges from around ten to several
encoded by separate nuclear genes, which are synthesized thousand per cell (often varying with growth conditions).
in the cytoplasm and imported into the mitochondrion [39]. The multiplicity of mtDNA molecules within the cell
A second recent example of a split gene with one-half in implies that mitochondrial mutations might accumulate
mtDNA and the other half in nuclear DNA is rpl2 in gradually, without immediate deleterious impact. This
flowering plants [40]. could partly explain the accelerated evolutionary rates of
Finally, gene simplification is another eccentric mtDNA relative to nuclear DNA that are primarily
phenomenon seen in mitochondrial genomes. Examples observed in animals (e.g. vertebrates and insects),
include the truncated tRNAs (lacking one, or more, helical fungi (e.g. yeasts and fission yeasts) and certain protists
arms) found in the mitochondria of several animal lineages (e.g. ciliates and slime moulds).
[7], the structurally streamlined mitochondrial rRNAs in In human somatic cells, it has been well-documented
most animals [7] and some protists [41], and mitochondrial that mtDNA accumulates point mutations and large
RNase P RNAs in fungi [42]. In certain organisms deletions, predominantly in brain and heart muscle of
(e.g. in the fungus Harpochytrium [43]), a similar older individuals (. 40 years) [54]. This high incidence of
trend is observed for mitochondrion-encoded protein mutations has been attributed to several factors, including
genes, whose coding regions are often shortened in the slipped mispairing at direct sequence repeats, topoisome-
C-terminal portion compared with their more primitive rase cleavage of mtDNA, the absence of excision repair,
counterparts. and incomplete repair of damage caused by free-radicals
produced as a by-product of oxidative phosphorylation
Events implicated in changes in the overall organization [55]. However, as there are few comparative biochemical
of mitochondrial genomes studies in other eukaryotes, the specific mechanisms
Changes in mtDNA architecture have been associated underlying the particularly high occurrence of mitochon-
with autonomously replicating mitochondrial plasmids, drial mutations in human versus other eukaryotes
which are common in fungi but which also occur in several remains largely unknown.
plant and protist taxa. The well-characterized mitochon-
drial plasmids in fungi are primarily monomeric linear Exceptional horizontal transfer of mitochondrial genes
molecules with inverted terminal repeats [44]. Insertion of Although the evolutionary transfer of genes from organelle
such linear plasmids into mtDNA probably triggers genomes is well-documented (Figure 7), there are few
conversion of genome conformation from circular mapping reports to-date of the transfer of genetic material into the
to monomeric linear, as observed in maize cytoplasmic mitochondrial genome and, especially, between mitochon-
male sterility [45]. Stable transitions of mtDNA topology drial genomes. Only in flowering plants (angiosperms)
from circular to linear have occurred within relatively does incorporation of nuclear and, particularly, chloroplast
short phylogenetic distances, as seen in yeasts [10], golden DNA sequences into mtDNA occur frequently; such
algae [46] and chytridiomycete fungi [43,47]. transfer is not evident among non-angiosperm land plants
Changes in mtDNA that lead to size expansion are or in green algae [49]. Exchange of mobile elements (group
mainly caused by an accumulation of repeat sequences and I and group II introns) between mtDNA and non-mtDNA
AT-rich intergenic spacers, mobile elements and introns. sources has been inferred on the basis of sequence
Even within a genus, the length of non-coding intergenic comparisons [56– 58], whereas mtDNA-to-mtDNA trans-
regions in mtDNA can vary considerably. For example, fer is suggested by the presence of highly similar introns at
mtDNA size in fission yeast species varies between 17 and identical positions in Homologous mitochondrial genes of
$ 80 kbp [48], with multiple arrays of intergenic repeats distantly related species [59 – 61]. A few cases have been
accounting for most of the variation. By contrast, rampant described that involve transfer of coding regions between
size-expansion of plant mtDNAs seems to have arisen mitochondrial genomes; one of these examples involves
principally through acquisition of foreign DNA of chlor- fungi (probably from two different classes) and results in a
oplast, nuclear, viral and as-yet-unknown origin [49], hybrid atp6 gene [62].
involving active DNA uptake [50] with master circles and Recently, widespread horizontal transfer of mitochon-
subcircles constantly recombining at a low level to create drial genes (both respiratory chain and ribosomal protein)
mixed populations of subgenomic DNA molecules (‘sub- has been reported to occur between distantly related
limons’ [25]). In turn, size reduction of mtDNAs often angiosperms, even between dicotyledons and monocotyle-
occurs by gene transfer to the nucleus (Figure 7); examples dons [63]. The genomic consequences of these mtDNA-to-
of this phenomenon in flowering plants are especially mtDNA transfers include gene duplication, recapture of
notable in terms of both number and frequency [51– 53]. genes previously lost through transfer to the nucleus and,
Other strategies of mtDNA size reduction include intron in one particularly notable case, a chimeric, half-monocot –
loss, and elimination of intergenic spacers, sometimes half-dicot rps11 gene. It is estimated that these transfer
to such a degree that coding regions of adjacent genes events occurred between 30 and 140 million years ago in
overlap (e.g. [7,27]). individual cases. It would seem that transfer of genetic
Rearrangements through insertions, deletions or material into angiosperm mtDNA occurs relatively fre-
recombination follow special rules owing to the multi- quently on an evolutionary timescale, regardless of
copy character of mitochondrial genomes. In several whether such transfer emanates from the chloroplast
http://tigs.trends.com
714 Review TRENDS in Genetics Vol.19 No.12 December 2003

(a) Gene export and loss


α-Proteobacterial
endosymbiont Mitochondrion

Primitive Modern
eukaryote eukaryote

(b) Primitive nucleus Nucleus


Acquisition of
symbiont genes and foreign genes;
gene substitutions
TRENDS in Genetics

Figure 7. Fate of proto-mitochondrial genes. Early gene migration has taken two main routes. (a) The a-proteobacterial symbiont underwent massive gene loss during tran-
sition to an organelle. Among the genes essential for mitochondrial functions, the majority was transferred to the nucleus, with a small portion remaining in the evolving
mitochondrial DNA (mtDNA). (b) The hypothetical primitive nucleus of the host [either an amitochondriate eukaryote (see text) [82] or an Archaea-related organism [83])
acquired several hundred symbiont genes, some of which were substituted for nuclear genes. In some instances, nucleus-acquired DNA from other sources replaced a
mitochondrial function initially encoded by the a-proteobacterial symbiont. For example, mtDNA in jakobid flagellates contains genes specifying the components of a typi-
cal eubacteria-like mitochondrial RNA polymerase, whereas in all other eukaryotes, this enzyme has a T3-phage-like structure and is encoded by the nuclear genome [84].
Genes are colour-coded as follows: red, a-proteobacterial genes; blue, nuclear genes of the original host; yellow, a-proteobacterial genes that replaced host genes; green,
foreign genes acquired from sources other than the genome of the mitochondrial ancestor.

genome of the same species or from the mtDNA of a direction) at a notably high frequency [72]. Moreover, in
different plant species. Successful functional transfer of primates, nuclear genomes contain substantial numbers of
genes into angiosperm mtDNA might be facilitated by the mtDNA insertions [73], but the mitochondrial genome in
expansive, largely non-coding nature of this genome [64], these cases is free of foreign DNA. These contrasting
whose expression seems little constrained by gene location examples illustrate the concept that evolution of the
and overall organization. nuclear and organellar genomes of a given eukaryote can
proceed in very different directions.
Are evolutionary trends similar in mitochondrial and
nuclear genomes? Does a parasitic lifestyle accelerate mtDNA evolution?
With the recent availability of nuclear genome data from It is tempting to speculate that mtDNA might be under
several eukaryotes, we are now in a position to compare more relaxed selection in parasitic lineages, leading to the
evolutionary trends for the mitochondrial and nuclear emergence of exotic mitochondrial genome features, as
genomes of a given organism. One readily measurable described earlier. Indeed, mtDNA with the smallest
parameter is the proportion of non-coding sequence in the number of genes is found in Plasmodium, and the most
two genomes. For example, human mtDNA is tightly extreme case of mRNA editing occurs in trypanosomes,
packed with intron-less genes [65], whereas human both of which are obligate parasites. Somewhat surpris-
nuclear genes contain a large number of introns separated ingly, this correlation does not withstand careful scrutiny.
by extensive intergenic regions [66,67]. By contrast, genes For example, Prototheca wickerhamii, a non-photosyn-
in yeast mtDNA are intron-rich and separated by long non- thetic chlorophycean alga and (mild) human pathogen, has
coding stretches [68], whereas yeast nuclear DNA har- one of the most primitive mitochondrial genomes
bours closely spaced and mostly intron-less genes [69]. within this group of green algae [38,74,75]. By
Finally, in the pufferfish Fugu, gene density is high in both contrast, several free-living, photosynthetic relatives
genomes [70,71]. Comparison of the relative gene numbers have highly derived mtDNAs: Scenedesmus obliquus
in the two genomes shows similarly unrelated variations, has an unusual translation code and substantially
implying that the evolutionary trends shaping coding fewer mtDNA-encoded genes than P. wickerhamii
capacity and density in the nuclear versus mitochondrial mtDNA [38]. In addition, Pedinomonas minor [21]
genome proceed independently from one another in a given and Chlamydomonas reinhardtii (and its close rela-
organism. tives) have mtDNAs that are typical cases of reduced
Another issue that can now be investigated is whether mitochondrial genomes [76]. Other examples of non-
the propensity to insert foreign DNA is the same for the pathogens with highly derived mtDNAs are Euglena
mitochondrial genome and nuclear genome of a given gracilis [77], baker’s and fission yeast (and its close
organism. In flowering plants, acquisition of foreign DNA relatives), animals in general, and Amoebidium para-
is rampant both in the nucleus and in mitochondria (but siticum [15], a saprotrophic ectosymbiont that is
not in chloroplasts), whereas in yeast, DNA escapes from frequently found on crustaceans and insects that
mitochondria to the nucleus (but not in the opposite inhabit fresh water ponds.
http://tigs.trends.com
Review TRENDS in Genetics Vol.19 No.12 December 2003 715

Why do we see the highly derived mtDNAs in some forms of the same fundamental organelle. Philos. Trans. R. Soc. Lond.
species but not in others? One possible explanation is that B Biol. Sci. 358, 191 – 201
4 Roger, A.J. and Silberman, J.D. (2002) Cell evolution: mitochondria in
derived mitochondrial genomes are a consequence of hiding. Nature 418, 827– 829
accelerated growth rates (i.e. they tend to occur in 5 Gray, M.W. et al. (1999) Mitochondrial evolution. Science 283,
‘weeds’, that is, organisms with short generation times). 1476– 1481
Some such species (e.g. yeasts) might be highly specialized 6 Lang, B.F. et al. (1999) Mitochondrial genome evolution and the origin
of eukaryotes. Annu. Rev. Genet. 33, 351 – 397
to thrive in specific environments, whereas other organ-
7 Boore, J.L. (1999) Animal mitochondrial genomes. Nucleic Acids Res.
isms of this type (such as chlamydomonad algae) occur 27, 1767 – 1780
more widely, dominating over other species through more 8 Bendich, A.J. (1993) Reaching for the ring: the study of mitochondrial
efficient propagation. genome structure. Curr. Genet. 24, 279 – 290
9 Bendich, A.J. (1996) Structural analysis of mitochondrial DNA
molecules from fungi and plants using moving pictures and pulsed-
Mitochondrial DNA variability: implications for gene
field gel electrophoresis. J. Mol. Biol. 255, 564 – 588
expression 10 Nosek, J. and Tomaska, L. Mitochondrial genome diversity: evolution
In our quest to understand the role of genome organization of the molecular architecture and replication strategy. Curr. Genet.
in gene expression, a comparative mitochondrial genomics (in press)
approach is only the first step. Complete sequencing of 11 Burger, G. and Lang, B.F. (2003) Parallels in genome evolution in
mitochondria and bacterial symbionts. IUBMB Life 55, 205 – 212
mtDNA enables the identification of genes and permits one 12 Armstrong, M.R. et al. (2000) A multipartite mitochondrial genome in
to make inferences about unusual modes of gene the potato cyst nematode Globodera pallida. Genetics 154, 181 – 192
expression (e.g. split and divergently transcribed genes, 13 Watanabe, K.I. et al. (1999) Mitochondrial genes are found on
or fused ORFs). However, only systematic and compre- minicircle DNA molecules in the mesozoan animal Dicyema. J. Mol.
Biol. 286, 645 – 650
hensive analysis of the transcriptome (the portion of the
14 Lang, B.F. et al. (2002) The closest unicellular relatives of animals.
genome that is transcribed) will reveal: (i) whether all Curr. Biol. 12, 1773– 1778
identified genes are active, (ii) whether co- or post- 15 Burger, G. et al. (2003) Unique mitochondrial genome architecture in
transcriptional processes, such as RNA editing, alter unicellular relatives of animals. Proc. Natl. Acad. Sci. U. S. A. 100,
encoded genetic information on the way to its expression, 892– 897
16 Morris, J.C. et al. (2001) Replication of kinetoplast DNA: an update for
and (iii) whether RNA species (perhaps having regulatory the new millennium. Int. J. Parasitol. 31, 453 – 458
functions) are synthesized in addition to the usual 17 Feagin, J.E. (2000) Mitochondrial genome diversity in parasites. Int.
informational and structural RNAs. Identification of a J. Parasitol. 30, 371 – 390
gene, even one having all the hallmarks of a functional 18 Notsu, Y. et al. (2002) The complete sequence of the rice (Oryza sativa
unit, is no guarantee that it is active, as exemplified by the L.) mitochondrial genome: frequent DNA sequence acquisition and
loss during the evolution of flowering plants. Mol. Genet. Genomics
silent cox2 gene in soybean mtDNA [78]. Even where 268, 434 – 445
transcription of a protein-coding gene is shown, definitive 19 Ward, B.L. et al. (1981) The mitochondrial genome is large and
evidence of function ultimately requires identification of variable in a family of plants (cucurbitaceae). Cell 25, 793 – 803
the corresponding protein. This criterion has been 20 Lang, B.F. et al. (1997) An ancestral mitochondrial DNA resembling a
eubacterial genome in miniature. Nature 387, 493 – 497
satisfied in very few cases, although direct analysis of
21 Turmel, M. et al. (1999) The complete mitochondrial DNA sequences of
the mitochondrial proteome is rapidly gaining favour. Nephroselmis olivacea and Pedinomonas minor. Two radically
different evolutionary patterns within green algae. Plant Cell 11,
Concluding remarks 1717– 1730
Comparative mitochondrial genomics, especially in pro- 22 Lunt, D.H. et al. (1998) Mitochondrial DNA variable number tandem
repeats (VNTRs): utility and problems in molecular ecology. Molec.
tists, has revealed an often perplexing diversity in overall
Ecol. 7, 1441– 1455
organization and mode of expression of mtDNA. The 23 Paquin, B. and Lang, B.F. (1996) The mitochondrial DNA of Allomyces
comparative approach has also amply demonstrated that macrogynus: the complete genomic sequence from an ancestral fungus.
this extreme structural divergence co-exists with the J. Mol. Biol. 255, 688– 701
conserved functional role of the mitochondrial genome. 24 Paquin, B. et al. (2000) Double-hairpin elements in the mitochondrial
DNA of Allomyces: evidence for mobility. Mol. Biol. Evol. 17,
In this respect, mtDNA is an evolutionary playground par 1760– 1768
excellence. No correlation is evident in the mode and tempo 25 Andre, C. et al. (1992) Small repeated sequences and the structure of
of evolutionary change in the mitochondrial and nuclear plant mitochondrial genomes. Trends Genet. 8, 128– 132
genomes across the broad spectrum of eukaryotic lineages. 26 Cummings, D.J. et al. (1990) The complete DNA sequence of the
mitochondrial genome of Podospora anserina. Curr. Genet. 17,
375– 402
Acknowledgements
27 Burger, G. et al. (1995) The mitochondrial DNA of the amoeboid
This project was supported by grants from the Canadian Institutes for
protozoon, Acanthamoeba castellanii: complete sequence, gene content
Health Research (MSP14226, MOP42475 and MOP4124). Salary and
and genome organization. J. Mol. Biol. 245, 522 – 537
interaction support from the Canadian Institute for Advanced Research
28 Ogawa, S. et al. (2000) The mitochondrial DNA of Dictyostelium
(G.B., M.W.G. and B.F.L.) and the Canada Research Chairs Programme
discoideum: complete sequence, gene content and genome organiz-
(M.W.G. and B.F.L.) is gratefully acknowledged.
ation. Mol. Gen. Genet. 263, 514– 519
29 Lonergan, K.M. and Gray, M.W. (1996) Expression of a continuous
References open reading frame encoding subunits 1 and 2 of cytochrome c oxidase
1 Saraste, M. (1999) Oxidative phosphorylation at the fin de siecle. in the mitochondrial DNA of Acanthamoeba castellanii. J. Mol. Biol.
Science 283, 1488 – 1493 257, 1019– 1030
2 Herrmann, J.M. (2003) Converting bacteria to organelles: evolution of 30 Hanson, M.R. (1991) Plant mitochondrial mutations and male sterility.
mitochondrial protein sorting. Trends Microbiol. 11, 74– 79 Annu. Rev. Genet. 25, 461 – 486
3 Embley, T.M. et al. (2003) Mitochondria and hydrogenosomes are two 31 Oudot-Le Secq, M.P. et al. (2001) The complete sequence of a brown
http://tigs.trends.com
716 Review TRENDS in Genetics Vol.19 No.12 December 2003

algal mitochondrial genome, the ectocarpale Pylaiella littoralis (L.) 57 Turmel, M. et al. (1995) Evolutionary transfer of ORF-containing
Kjellm. J. Mol. Evol. 53, 80 – 88 group I introns between different subcellular compartments (chlor-
32 Bullerwell, C.E. et al. (2000) A novel motif for identifying rps3 homologs oplast and mitochondrion). Mol. Biol. Evol. 12, 533 – 545
in fungal mitochondrial genomes. Trends Biochem. Sci. 25, 363–365 58 Burger, G. et al. (1999) Complete sequence of the mitochondrial DNA of
33 Heinonen, T.Y. et al. (1987) Rearranged coding segments, separated by the red alga Porphyra purpurea. Cyanobacterial introns and shared
a transfer RNA gene, specify the two parts of a discontinuous large ancestry of red and green algae. Plant Cell 11, 1675– 1694
subunit ribosomal RNA in Tetrahymena pyriformis mitochondria. 59 Lang, B.F. (1984) The mitochondrial genome of the fission yeast
J. Biol. Chem. 262, 2879– 2887 Schizosaccharomyces pombe: highly homologous introns are inserted
34 Schnare, M. et al. (1986) A discontinuous subunit ribosomal RNA in at the same position of the otherwise less conserved cox1 genes in
Tetrahymena pyriformis mitochondria. J. Biol. Chem. 261, 5187– 5193 Schizosaccharomyces pombe and Aspergillus nidulans. EMBO J. 3,
35 Boer, P.H. and Gray, M.W. (1988) Scrambled ribosomal RNA gene 2129–2136
pieces in Chlamydomonas reinhardtii mitochondrial DNA. Cell 55, 60 Cho, Y. et al. (1998) Explosive invasion of plant mitochondria by a
399 – 411 group I intron. Proc. Natl. Acad. Sci. U. S. A. 95, 14244 – 14249
36 Gillespie, D.E. et al. (1999) The fragmented mitochondrial ribosomal 61 Cho, Y. and Palmer, J.D. (1999) Multiple acquisitions via horizontal
RNAs of Plasmodium falciparum have short A tails. Nucleic Acids Res. transfer of a group I intron in the mitochondrial cox1 gene during
27, 2416 – 2422 evolution of the Araceae family. Mol. Biol. Evol. 16, 1155 – 1165
37 Edqvist, J. et al. (2000) Expression of mitochondrial protein-coding 62 Paquin, B. et al. (1994) Interspecific transfer of mitochondrial genes in
genes in Tetrahymena pyriformis. J. Mol. Biol. 297, 381– 393 fungi and creation of a homologous hybrid gene. Proc. Natl. Acad. Sci.
38 Nedelcu, A.M. et al. (2000) The complete mitochondrial DNA sequence of U. S. A. 91, 11807 – 11810
Scenedesmus obliquus reflects an intermediate stage in the evolution of 63 Bergthorsson, U. et al. (2003) Widespread horizontal transfer of
the green algal mitochondrial genome. Genome Res. 10, 819–831 mitochondrial genes in flowering plants. Nature 424, 197 – 201
39 Perez-Martinez, X. et al. (2001) Subunit II of cytochrome c oxidase in 64 Hanson, M.R. and Folkerts, O. (1992) Structure and function of the
chlamydomonad algae is a heterodimer encoded by two independent higher plant mitochondrial genome. Int. J. Cytol. 141, 129– 172
nuclear genes. J. Biol. Chem. 276, 11302 – 11309 65 Anderson, S. et al. (1981) Sequence and organization of the human
40 Adams, K.L. et al. (2001) Mitochondrial gene transfer in pieces: fission mitochondrial genome. Nature 290, 457– 465
of the ribosomal protein gene rpl2 and partial or complete gene 66 Lander, E.S. et al. (2001) Initial sequencing and analysis of the human
transfer to the nucleus. Mol. Biol. Evol. 18, 2289– 2297 genome. Nature 409, 860 – 921
41 Gray, M.W. et al. (1998) Genome structure and gene content in protist 67 Venter, J.C. et al. (2001) The sequence of the human genome. Science
mitochondrial DNAs. Nucleic Acids Res. 26, 865– 878 291, 1304– 1351
42 Seif, E.R. et al. (2003) Mitochondrial RNase P RNAs in ascomycete 68 Foury, F. et al. (1998) The complete sequence of the mitochondrial
fungi: lineage-specific variations in RNA secondary structure. RNA 9, genome of Saccharomyces cerevisiae. FEBS Lett. 440, 325 – 331
1073 – 1083 69 Giaever, G. et al. (2002) Functional profiling of the Saccharomyces
43 Bullerwell, C.E. et al. (2003) Evolution of monoblepharidalean fungi cerevisiae genome. Nature 418, 387 – 391
based on complete mitochondrial genome sequences. Nucleic Acids 70 Elmerot, C. et al. (2002) The mitochondrial genome of the pufferfish,
Res. 31, 1614– 1623 Fugu rubripes, and ordinal teleostean relationships. Gene 295,
44 Kennell, J.C. and Cohen, S.M. In Handbook of Fungal Biotechnology 163– 172
(Arora, D., ed.), Marcel Dekker (in press) 71 Aparicio, S. et al. (2002) Whole-genome shotgun assembly and analysis
45 Schardl, C.L. et al. (1985) Mitochondrial DNA rearrangements associ- of the genome of Fugu rubripes. Science 297, 1301– 1310
ated with fertile revertants of S-type male-sterile maize. Cell 43, 72 Thorsness, P.E. and Fox, T.D. (1990) Escape of DNA from mitochondria
361–368 to the nucleus in Saccharomyces cerevisiae. Nature 346, 376 – 379
46 Chesnick, J.M. et al. (2000) The mitochondrial genome of the 73 Woischnik, M. and Moraes, C.T. (2002) Pattern of organization of
stramenopile alga Chrysodidymus synuroideus. Complete sequence, human mitochondrial pseudogenes in the nuclear genome. Genome
gene content and genome organization. Nucleic Acids Res. 28, Res. 12, 885 – 893
2512–2518 74 Wolff, G. et al. (1994) Complete sequence of the mitochondrial DNA of
47 Forget, L. et al. (2002) Hyaloraphidium curvatum: a linear mitochon- the chlorophyte alga Prototheca wickerhamii. Gene content and
drial genome, tRNA editing, and an evolutionary link to lower fungi. genome organization. J. Mol. Biol. 237, 75 – 86
Mol. Biol. Evol. 19, 310 – 319 75 Kuck, U. et al. (2000) DNA sequence analysis of the complete
48 Bullerwell, C.E. et al. (2003) A comparison of three fission yeast mitochondrial genome of the green alga Scenedesmu obliquus:
mitochondrial genomes. Nucleic Acids Res. 31, 759 – 768 evidence for UAG being a leucine and UCA being a non-sense codon.
49 Marienfeld, J. et al. (1999) The mitochondrial genome of Arabidopsis is Gene 253, 13 – 18
composed of both native and immigrant information. Trends Plant Sci. 76 Gray, M.W. and Boer, P.H. (1988) Organization and expression of algal
4, 495 – 502 (Chlamydomonas reinhardtii) mitochondrial DNA. Philos. Trans.
50 Koulintchenko, M. et al. (2003) Plant mitochondria actively import DNA R. Soc. Lond. B Biol. Sci. 319, 135 – 147
via the permeability transition pore complex. EMBO J. 22, 1245–1254 77 Tessier, L.H. et al. (1997) The cox1 gene from Euglena gracilis: a protist
51 Adams, K.L. et al. (2000) Repeated, recent and diverse transfers of a mitochondrial gene without introns and genetic code modifications.
mitochondrial gene to the nucleus in flowering plants. Nature 408, Curr. Genet. 31, 208 – 213
354 – 357 78 Covello, P.S. and Gray, M.W. (1992) Silent mitochondrial and active
52 Adams, K.L. et al. (2001) Multiple losses and transfers to the nucleus of nuclear genes for subunit 2 of cytochrome c oxidase (cox2) in soybean:
two mitochondrial succinate dehydrogenase genes during angiosperm evidence for RNA-mediated gene transfer. EMBO J. 11, 3815 – 3820
evolution. Genetics 158, 1289 – 1300 79 Gabaldon, T. and Huynen, M.A. (2003) Reconstruction of the proto-
53 Adams, K.L. et al. (2002) Punctuated evolution of mitochondrial gene mitochondrial metabolism. Science 301, 609
content: high and variable rates of mitochondrial gene loss and 80 Rasmusson, A.G. et al. (1999) Homologues of yeast and bacterial
transfer to the nucleus during angiosperm evolution. Proc. Natl. Acad. rotenone-insensitive NADH dehydrogenases in higher eukaryotes:
Sci. U. S. A. 99, 9905 – 9912 two enzymes are present in potato mitochondria. Plant J. 20, 79 – 87
54 Chomyn, A. and Attardi, G. (2003) MtDNA mutations in aging and 81 Baldauf, S.L. (2003) The deep roots of eukaryotes. Science 300,
apoptosis. Biochem. Biophys. Res. Commun. 304, 519 – 529 1703– 1706
55 Wei, Y.H. and Lee, H.C. (2002) Oxidative stress, mitochondrial DNA 82 Cavalier-Smith, T. (1983) A 6-kingdom classification and a unified
mutation, and impairment of antioxidant enzymes in aging. Exp. Biol. phylogeny. In Endocytobiology II (Schwemmler, W. and Schenk,
Med. (Maywood) 227, 671 – 682 H.E.A., eds), pp. 1027– 1034, De Gruyter
56 Lonergan, K.M. and Gray, M.W. (1994) The ribosomal RNA gene region 83 Martin, W. and Müller, M. (1998) The hydrogen hypothesis for the first
in Acanthamoeba castellanii mitochondrial DNA. A case of evolution- eukaryote. Nature 392, 37 – 41
ary transfer of introns between mitochondria and plastids? J. Mol. 84 Cermakian, N. et al. (1997) On the evolution of the single-subunit RNA
Biol. 239, 476– 499 polymerases. J. Mol. Evol. 45, 671 – 681

http://tigs.trends.com

View publication stats

You might also like