SlideShare une entreprise Scribd logo
1  sur  13
1810
Mol. Biol. Evol. 18(9):1810–1822. 2001
᭧ 2001 by the Society for Molecular Biology and Evolution. ISSN: 0737-4038
Accelerated Evolution of Functional Plastid rRNA and Elongation Factor
Genes Due to Reduced Protein Synthetic Load After the Loss of
Photosynthesis in the Chlorophyte Alga Polytoma
Dawne Vernon,*1 Robin R. Gutell,† Jamie J. Cannone,† Robert W. Rumpf,*2
and C. William Birky Jr.*‡
*Department of Molecular Genetics, Ohio State University; †Institute of Cellular and Molecular Biology, University of Texas
at Austin; and ‡Department of Ecology and Evolutionary Biology and Graduate Interdisciplinary Program in Genetics,
University of Arizona
Polytoma obtusum and Polytoma uvella are members of a clade of nonphotosynthetic chlorophyte algae closely
related to Chlamydomonas humicola and other photosynthetic members of the Chlamydomonadaceae. Descended
from a nonphotosynthetic mutant, these obligate heterotrophs retain a plastid (leucoplast) with a functional protein
synthetic system, and a plastid genome (lpDNA) with functional genes encoding proteins required for transcription
and translation. Comparative studies of the evolution of genes in chloroplasts and leucoplasts can identify modes
of selection acting on the plastid genome. Two plastid genes—rrn16, encoding the plastid small-subunit rRNA, and
tufA, encoding elongation factor Tu—retain their functions in protein synthesis after the loss of photosynthesis in
two nonphotosynthetic Polytoma clades but show a substantially accelerated rate of base substitution in the P. uvella
clade. The accelerated evolution of tufA is due, at least partly, to relaxed codon bias favoring codons that can be
read without wobble, mainly in three amino acids. Selection for these codons may be relaxed because leucoplasts
are required to synthesize fewer protein molecules per unit time than are chloroplasts (reduced protein synthetic
load) and thus require a lower rate of synthesis of elongation factor Tu. Relaxed selection due to a lower protein
synthetic load is also a plausible explanation for the accelerated rate of evolution of rrn16, but the available data
are insufficient to test the hypothesis for this gene. The tufA and rrn16 genes in Polytoma oviforme, the sole member
of a second nonphotosynthetic clade, are also functional but show no sign of relaxed selection.
Introduction
Nonphotosynthetic land plants and algae serve as a
basis for interesting natural experiments on the evolution-
ary consequence of the loss of a significant cell function.
After losing the ability to do photosynthesis, nonphoto-
synthetic species use various alternative carbon sources,
with the plants becoming parasitic on other plants, while
the algae take up complex organic molecules from their
environment. Recognized by their lack of chlorophyll,
these nongreen organisms have unique plastid (‘‘leuco-
plast’’) genomes. The evolutionary consequences of the
loss of photosynthesis can be studied by comparing the
leucoplast genomes of nonphotosynthetic species with the
chloroplast genomes of their closest photosynthetic rela-
tives. The majority of the genes in the chloroplasts of pho-
tosynthetic green algae and land plants encode proteins
required for photosynthesis or gene expression (transcrip-
tion and translation [Gillham 1994]; for recent data from
complete chloroplast genome sequences, see Turmel, Otis,
and Lemieux [1999], Lemieux, Otis, and Turmel [2000],
and the NCBI chloroplast genome page at http://www.
ncbi.nlm.nih.gov:80/PMGifs/Genomes/plastids࿞tax.
html). These two functions account for 32 and 16 genes,
1 Present address: National Institute of Standards and Technology,
DNA Technologies Group, Gaithersburg, Maryland.
2 Present address: LabBook.com, Inc., Columbus, Ohio.
Key words: Polytoma, Chlamydomonas, chloroplast rRNA gene,
chloroplast elongation factor gene, substitution rate, codon bias.
Address for correspondence and reprints: C. William Birky Jr.,
Department of Ecology and Evolutionary Biology, Biological Sciences
West, University of Arizona, Tucson, Arizona 85721. E-mail:
birky@u.arizona.edu.
respectively, in Chlamydomonas reinhardtii, which also
encodes a minimal set of rRNA and tRNA genes (http:
//www.biology.duke.edu/chlamy࿞genome/chloro.html).
Genes encoding proteins with other functions, as well
as unidentified open reading frames, are found in some
taxa; C. reinhardtii has 15 of these. There is strong in-
direct evidence that at least one of these genes, yet to
be identified, codes for a protein that has an essential
nonphotosynthetic (ENP) function (Gillham 1994, pp.
83–86).
Investigations of leucoplast genome structure and
gene sequences of nonphotosynthetic organisms have
been limited to several parasitic angiosperm families
(Scrophulareae and Orobanchaceae; e.g., dePamphilis
and Palmer 1990; Wimpee, Morgan, and Wrobel 1992a,
1992b; Nickrent, Duff, and Konings 1997; Wolfe and
dePamphilis 1998; Young and dePamphilis 2000) and
the euglenoid alga Astasia (Siemeister, Buchholz, and
Hachtel 1990; Siemeister and Hachtel 1990a, 1990b).
The leucoplasts of these nonphotosynthetic species dif-
fer from the chloroplasts of their close green relatives
in numerous features. Morphologically, they are char-
acterized by reduction or elimination of the thylakoid
membranes. They all retain leucoplast ribosomes and
leucoplast DNA. However, their leucoplast genomes are
often reduced in size and complexity compared with
chloroplast genomes in photosynthetic relatives.
These leucoplast genomes allow investigation of
the effects of selection on rates of evolution. When the
ability to do photosynthesis is lost, the photosynthetic
and photorespiration genes lose their function and con-
sequently are no longer subject to selection; they are
expected to become pseudogenes and can be lost entire-
ly. In contrast, the leucoplast genes needed for transcrip-
Accelerated Evolution of Leucoplast Genes 1811
FIG. 1.—Cladogram showing the relationships of the taxa used in
this study based on Rrn18 sequences (Rumpf et al. 1996; unpublished
data).
tion and translation of the leucoplast genome probably
remain functional and subject to at least some degree of
selection because they are needed to transcribe and
translate the ENP genes. The leucoplast genomes of the
beech root parasite Epifagus virginiana (Orobancha-
ceae), the oak parasite Conopholis americana (Oroban-
chaceae), and the euglenoid Astasia longa display these
characteristics. Most expression genes are still present
in the leucoplast genomes and appear intact, and some
leucoplast RNA and protein products have been dem-
onstrated (Siemeister, Buchholz, and Hachtel 1990; Sie-
meister and Hachtel 1990a, 1990b; Wimpee, Morgan,
and Wrobel 1992; Wolfe, Morden, and Palmer 1992).
Conversely, many leucoplast genes that coded for pho-
tosynthetic proteins appear nonfunctional in key do-
mains, are grossly truncated, or are absent from these
leucoplast genomes. Although most expression genes
are selectively retained, the leucoplast genome in Epi-
fagus is missing more than a dozen expression genes
(tRNA genes, ribosomal protein genes, and all four
RNA polymerase subunit genes; Morden et al. 1991),
all four RNA polymerase genes are pseudogenes in
Lathraea (Lusson, Delavault, and Thalouarn 1998), and
the leucoplast genome in Conopholis is apparently miss-
ing several leucoplast tRNA genes (Wimpee, Morgan,
and Wrobel 1992b). Presumably, these tRNAs and pro-
teins are imported from the cytoplasm at a rate that may
be too low for detection but is sufficient for protein syn-
thesis in leucoplasts.
The tempo of evolution has also changed in the
leucoplast genomes of these nonphotosynthetic species.
Most, but not all, of the apparently still-functional genes
analyzed in leucoplasts show an increased rate of nu-
cleotide substitution compared with rates in their green
relatives. Most of the rate increases found in functional
leucoplast genes in Astasia and Epifagus are in the range
of 1.5-fold to 8-fold, while rrn16 genes in Epifagus and
Conopholis show 40-fold rate increases (references in
Results and Discussion). This suggests that selection on
these genes has been relaxed, even though it has not
been eliminated completely.
To test the generality of these results, we extended
the analysis of the sequence and evolution of leucoplast
translation genes to the nonphotosynthetic chlorophyte
algae, separated from the euglenoid plastid and from
land plants by over 400 Myr of evolution and large dif-
ferences in physiology and habitats. Phylogenetic anal-
yses of Rrn18 sequences show that the members of the
nonphotosynthetic genus Polytoma belong to two dif-
ferent lineages within the clade that includes all Chla-
mydomonas species as well as a number of other pho-
tosynthetic genera and another nonphotosynthetic clade,
Polytomella (Rumpf et al. 1996; unpublished data).
Many species of Chlamydomonas are facultative auxo-
trophs, capable of utilizing acetate as their sole carbon
and energy source, and nonphotosynthetic mutants are
readily isolated in C. reinhardtii (Harris 1989). It is as-
sumed that Polytoma species arose as nonphotosynthetic
mutants of facultative auxotrophs similar to the extant
Chlamydomonas. The single large cup-shaped leucoplast
in Polytoma does not have thylakoid membranes but still
contains ribosomes, DNA (lpDNA), rRNA, and stored
starch granules (Lang 1963; Scherbel, Behn, and Arnold
1974; Siu, Chiang, and Swift 1976; Vernon-Kipp, Kuhl,
and Birky 1989). Polytoma is sensitive to inhibitors of
chloroplast protein synthesis, which is additional evi-
dence that the leucoplast is synthesizing at least one
protein that is essential for auxotrophic growth and re-
production (Scherbel, Behn, and Arnold 1974).
Although these data strongly suggest that Polytoma
retains a functional leucoplast expression system, the
leucoplast genes involved have not been identified and
demonstrated to be functional. We sequenced the rrn16
gene and the tufA gene (encoding the plastid elongation
factor Tu) from two representatives of the Polytoma
uvella clade (P. uvella 964 and Polytoma obtusum
DH1), plus Polytoma oviforme, which is the sole mem-
ber of the second Polytoma lineage. For comparison,
these genes were also sequenced from two closely re-
lated photosynthetic relatives (Chlamydomonas humi-
cola SAG 11-9 and Chlamydomonas dysosmos UTEX
2399). The evolutionary relationships of these strains
and some others involved in the analysis are shown in
figure 1. This cladogram agrees with phylogenetic anal-
yses of the rrn16 and tufA genes (figs. 2 and 3). Rela-
tive-rate tests showed increased substitution rates in
rrn16 and tufA, compared with green relatives, in the
two P. uvella species. However, sequence analyses
showed that the genes were subject to selection and
therefore functional. The increase in substitution rate
was greater at sites subject to less stringent selection,
implicating a partial relaxation of selection. The tufA
gene of P. obtusum showed a large reduction in codon
preference, suggesting that the relaxed selection is due
at least partly to a reduced load of protein synthesis.
This was proposed earlier for the increased substitution
rates in Epifagus (Wolfe et al. 1992), but alternative ex-
planations were not ruled out. We observed no increase
in the substitution rate in the branch leading to P. ovi-
forme, suggesting that photosynthesis was lost more re-
cently in this lineage. This is the first analysis of the
molecular evolutionary consequences of the loss of pho-
tosynthesis in a chlorophyte alga.
Materials and Methods
Organisms
Polytoma uvella (UTEX 964) and P. oviforme
(SAG 62-27) were obtained from the University of Tex-
1812 Vernon et al.
FIG. 2.—Neighbor-joining tree of rrn16 sequences. Branch lengths in percentages of substitutions are shown above the branches; below
the branches are the lengths in the most parsimonious tree, which had an identical topology.
FIG. 3.—Neighbor-joining tree of tufA sequences. Branch lengths in percentages of substitutions are shown above the branches; below the
branches are the lengths in the most parsimonious tree and the maximum-likelihood tree, which had identical topologies.
as Culture Collection of Algae and from Sammlung von
Algenkulturen Gottingen, respectively. P. obtusum (des-
ignated strain DH1 by us) was obtained from David Her-
rin at the University of Texas at Austin; it originally
came from Luigi Provasoli’s collection at Yale. All cul-
tures were subcloned once or twice and grown in Po-
lytomella medium.
Chlamydomonas humicola UTEX 225 and C. dy-
sosmos UTEX 2399, from the University of Texas Cul-
ture Collection of Algae, were combined under the spe-
cies name Chlamydomonas applanata Pringsheim based
on morphology and autolysin cross-reactions (Ettl 1976)
and identity of the nuclear Rrn18 gene sequences (Gor-
don et al. 1995). Consistent with this, we found no sub-
stitution differences and only one insertion or deletion
difference in their chloroplast rrn16 sequences, while
the tufA sequences showed two synonymous differences
and one nonsynonymous differences and no insertions
or deletions. Consequently, we included only the C.
humicola sequences in the analyses described here.
DNA Preparation
The rrn16 gene of P. uvella was cloned. Whole-
cell DNA was isolated with a lysis method designed to
yield high-molecular-weight chloroplast DNA, modified
from Grant, Gillham, and Boynton (1980) as described
in Vernon (1996). This DNA was fractionated in a
CsClϩbisbenzimide equilibrium gradient. The top band
in the gradient was identified as leucoplast DNA by
Southern hybridization with an rrn16 probe and dot blot
hybridization with a tufA probe. The C. reinhardtii
cpDNA probes were provided by Elizabeth Harris (Duke
University) and Jeffrey Palmer (Indiana University). The
top band was used to prepare a HindIII library cloned
in pBluescript. DNA obtained from these clones by al-
kaline lysis plasmid minipreps (Sambrook, Fritsch, and
Maniatis 1989) was electrophoresed, and Southern blots
on GeneScreen Plus were hybridized with the cpDNA
probes. A clone containing the rrn16 gene in a 6.2-kb
insert was identified and purified for sequencing with a
GeneClean kit (Bio 101). All other new sequences used
in this study were of genes amplified from partially pu-
rified whole-cell DNA isolated from CTAB lysates of
1L algal cultures.
Polymerase Chain Reaction Amplifications
Primers located near the ends of the rrn16 and tufA
genes were used to obtain DNA templates for sequenc-
ing. The 5Ј and 3Ј primers for rrn16 were A-17 (5Ј-
GTTTGATCCTGGCTCAC-3Ј) and 5005-15 (3Ј-CA-
TGTGTGGCGGGCA-5Ј). The 5Ј and 3Ј degenerate
primers for all but one of the tufA genes were 1F (5Ј-
GGDCAYGTTGAYCAYGG-3Ј) and 5R (3Ј-TGA-
CANCCRCGRCCRCA-5Ј). Primer 5R did not amplify
tufA from P. obtusum, so the 3Ј ends of the tufA genes
from the other chlamydomonad species were inspected
for conserved areas, and an alternative 3Ј primer
(1130R: 3Ј-CCRATACGGDCCACTRGC-5Ј) was de-
signed and used, located 100 bases farther 5Ј of the orig-
inal 3Ј primer 5R. This amplified a tufA fragment from
P. obtusum that was approximately 100 bases shorter
than the other chlamydomonad sequences. The rrn16
Accelerated Evolution of Leucoplast Genes 1813
amplification products sequenced were about 1.3 kb
long, except for P. uvella, which was about 1.6 kb long;
the tufA amplification products were about 1.1 kb long,
except for P. obtusum, which was about 1.0 kb long.
Optimal amplification conditions were determined for
each gene empirically; multiple separate amplifications
were performed and pooled, then purified using
GeneClean.
Sequencing
Both strands of all genes were sequenced manually
using a modified dsDNA Cycle Sequencing kit (Life
Technologies). Most internal primers for sequencing
were obtained from Paul Fuerst for the rrn16 gene and
from Jeffrey Palmer for the tufA gene; additional inter-
nal primers in conserved regions were designed to fill
gaps in sequence coverage.
Alignment of rrn16 Sequences
The five new sequences for the study reported here
(P. uvella, P. obtusum, P. oviforme, C. humicola, and
C. dysosmos) were initially aligned using CLUSTAL W
in SeqApp (Gilbert 1992) to match the rrn16 primary
structure alignment in the Ribosomal Database Project
(Maidak et al. 1994). The alignment was further refined
by comparison with 70 publicly available plastid rrn16
sequences using a SUN Microsystems workstation with
the alignment editor AE2 (developed by T. Macke,
Scripps Research Institute, San Diego, Calif., and avail-
able at http://www.cme.msu.edu/RDP/html/index.html).
Sequences were initially aligned for maximum primary
structure similarity; then, all positions associated with
the comparatively inferred base pairs were checked to
assure that these base-paired positions were properly
aligned. The final alignment (with a complete list of
species and numerous chloroplast and Polytoma SSU
rRNA secondary-structure diagrams) is available in the
supplement (on the MBE web site) as GenBank files
(fig. 3c in the supplement); a subset of sequences used
for phylogenetic analysis is shown in less detail in se-
quential format (fig. 6 in the supplement) and in inter-
leaved Pretty Print format (fig. 7 in the supplement).
Alignment of tufA Sequences
To assist alignment of tufA sequences, C. Delwiche
and J. Palmer at Indiana University provided their align-
ment, with 18 eubacterial, 8 cyanobacterial, 26 algal,
and 4 land plant sequences (array described in Delwiche,
Kuhsel, and Palmer 1995). The Polytoma and Chlamy-
domonas sequences were aligned to various subsets of
this array using DNA sequences but were influenced by
the resulting amino acid alignment. One thousand fifty-
three base pairs of the tufA gene were aligned (85% of
the coding region), leaving out the first 72 5Ј positions
and the last 96 3Ј positions for lack of data in some or
all species. The complete alignment is available in the
supplement (fig. 5); a subset of sequences is shown in
less detail in sequential format (fig. 6 in the supplement)
and in interleaved Pretty Print format (fig. 8 in the
supplement).
Phylogenetic Analyses
Gene trees for were produced with PAUP* (Swof-
ford 1998) and PHYLIP, version 3.56 (Felsenstein
1993). Sequence differences were corrected for multiple
hits using the Jukes-Cantor one-parameter model (Jukes
and Cantor 1969); otherwise, all analyses used default
settings. Before a set of sequences was subjected to phy-
logenetic analysis or relative-rate tests, sites that were
missing in one or more species were removed from all
sequences.
Relative-Rate Tests
Relative-rate tests (Sarich and Wilson 1973; Wu
and Li 1985) were performed to detect differences be-
tween rates of nucleotide (or amino acid) substitution in
the three Polytoma species studied, compared with green
species. Each relative-rate test involved a Polytoma iso-
late (nongreen, N), its closest photosynthetic relative
(green, G), and a photosynthetic outgroup species (O).
The test parameters were KON and KOG, the estimated
numbers of base substitutions per site occurring along
the lineages leading from the outgroup to the nongreen
Polytoma and to the green ingroup, respectively. The
estimated numbers of substitutions per site (K) were ob-
tained by correcting the observed sequence differences
per site for multiple hits with the Jukes-Cantor model
implemented in the MEGA sequence analysis package,
PHYLIP, or PAUP*. Two other correction methods were
used for comparison: the Kimura (1980) two-parameter
method, which allows different rates of transition versus
transversion, and the Tamura (1992) method, which uses
information about GϩC content as well as separate tran-
sition and transversion rates, again using MEGA. All
three correction methods added approximately the same
number of unobserved substitutions (data not shown),
so the Jukes-Cantor method was used for the relative-
rate test because it had the smallest variance. KON and
KOG were related to the evolutionary rates EON and EOG
along the nongreen and green lineages by KON ϭ EONT
and KOG ϭ EOGT, where T is the time since divergence
of the two lineages and was, of course, the same for
both lineages. Any rate differences between green lin-
eages and the nongreen lineages can be expressed as the
difference between these two numbers of substitutions
(KON Ϫ KOG ϭ [EON Ϫ EOG]T). The significance of rate
differences was evaluated as in Muse and Weir (1992).
A second method was also used to separate ob-
served sequence differences into rates along different
green or nongreen lineages, employing phylogenetic
software. Gene trees in which the observed substitutions
were apportioned to the various branches of the tree by
phylogenetic algorithms provided the inferred substitu-
tions on each green or nongreen branch. The appor-
tioned substitutions from a nongreen Polytoma species
and from its nearest green relative (ingroup) to their
nearest ancestral node, KAN and KAG, respectively, were
used to calculate the ratio KAN/KAG or the difference KAN
1814 Vernon et al.
Table 1
Pairwise Numbers of Substitutions per Site Among rrn16 Genes
Chlamydomonas
reinhardtii
Chlamydomonas
moewusii
Polytoma
oviforme
Chlamydomonas
humicola
Polytoma
uvella
Polytoma
obtusum
Chlorella . . . . . . . . . . . . . . . . . . . . . .
Chlamydomonas reinhardtii. . . . . . .
Chlamydomonas moewusii . . . . . . . .
Polytoma oviforme . . . . . . . . . . . . . .
Chlamydomonas humicola . . . . . . . .
Polytoma uvella . . . . . . . . . . . . . . . .
0.162 0.185
0.138
0.171
0.123
0.106
0.170
0.119
0.125
0.095
0.227
0.212
0.213
0.194
0.152
0.220
0.193
0.200
0.166
0.142
0.054
NOTE.—Estimated number of substitutions per site (sequence divergence) ϭ sequence differences per site corrected for multiple hits by the Jukes-Cantor method.
Table 2
Relative-Rate Tests on rrn16 Based on Neighbor-Joining, Maximum-Parsimony, and Maximum-Likelihood Tree
Branch Lengths and on Pairwise Numbers of Substitutions
NONGREEN
SPECIES
GREEN
SPECIES
KAN/KAG
Neighbor
Joining
Maximum
Parsimony
Maximum
Likelihood
OUTGROUP
SPECIES KON ϪKOG
Polytoma uvella 964
Polytoma obtusum
Polytoma oviforme
Chlamydomonas humicola
C. humicola
Chlamydomonas moewusii
3.32
2.89
0.67
4.34
4.03
0.50
3.05
2.68
0.65
Chlamydomonas moewusii
Chlamydomonas reinhardtii
C. moewusii
C. reinhardtii
Chlamydomonas humicola
C. reinhardtii
0.088*
0.093**
0.075**
0.074**
Ϫ0.030
Ϫ0.015
* Significant at the 1% level.
** Significant at the 0.1% level.
Ϫ KAG. The difference divided by KAG, i.e., (KAN Ϫ
KAG)/KAG, can be used to compare the magnitudes of
the rate increases along two different nonphotosynthetic
lineages with different ingroups.
Codon usage and the amount of codon usage bias
in tufA were also investigated in P. obtusum, P. ovifor-
me, C. humicola, and C. reinhardtii. All gaps were re-
moved from the aligned sequences of these four species,
leaving 349 codons. DNA Strider was used to calculate
codon usage in these sequences. Relative synonymous
codon usage (RSCU) was calculated for each codon us-
ing MEGA. RSCU is the ratio of the observed frequency
of a particular codon to the expected frequency of that
codon calculated on the assumption that all codons are
used equally frequently; an RSCU value significantly
different from 1 is evidence of biased codon usage
(Sharp and Li 1987). A Pascal program provided by
Brian Morton was used to calculate the codon bias index
(CBI) and the codon adaptation index (CAI). The CBI
is an overall measure of codon bias for the entire gene
(Morton 1993); the CBI ranges from 0 (no codon bias
in the gene) to 1 (maximum codon bias). The CAI is
measure of bias in the use of a codon relative to its use
in a reference set of highly expressed genes (Sharp and
Li 1987).
Results and Discussion
The Evolutionary Rate of rrn16 is Accelerated in P.
uvella and P. obtusum but not in P. oviforme
Figure 2 shows the neighbor-joining tree of rrn16
sequences; the topology of the most parsimonious tree
from an exhaustive maximum parsimony search is iden-
tical. This tree is compatible with the trees of the nuclear
Rrn18 gene (fig. 1). Above each line in figure 2 is the
length of the branch in the Neighbor-Joining tree in per-
centage of substitutions; below each line is the length
of the same branch in the parsimony tree. The tree
shows a strong acceleration of substitution rate along
the branches leading to the nonphotosynthetic P. uvella
lineage.
We used the distances on the tree in figure 2 to
calculate the ratio of substitution rates on nonphotosyn-
thetic and photosynthetic lineages, as well as the differ-
ence between nonphotosynthetic and photosynthetic
rates. We also used the method of Wu and Li (1985) for
relative-rate tests based on corrected frequencies of pair-
wise substitutions (table 1). Table 2 shows the results of
these relative-rate calculations. The tests for P. uvella
and P. obtusum used C. humicola as their closest relative
and Chlamydomonas moewusii or C. reinhardtii as the
outgroup; the test for Polytoma oviforme used C. moe-
wusii as the closest green relative and C. humicola or
C. reinhardtii as the outgroup. All relative-rates tests
showed significantly increased substitution rates in the
branch leading to P. obtusum versus the branch leading
to the ingroup (C. humicola), and an even greater rate
increase was seen in P. uvella. Figure 2 shows that the
branch leading from the common ancestor of the P.
uvella clade and C. humicola to the common ancestor
of P. uvella and P. obtusum is longer, i.e., has more
substitutions, than the branch leading to C. humicola.
This shows that the acceleration began in the common
ancestor of the P. uvella clade, as expected. As a con-
trol, we performed a relative-rate test on the nuclear
Rrn18 gene of P. uvella (not shown). The test showed
a small increase in this nongreen species, but it was not
Accelerated Evolution of Leucoplast Genes 1815
Table 3
Pairwise Numbers of Substitutions per Site Among tufA Genes
ORGANISMS
SEQUENCE DIVERGENCE
All Sites 3rd Position
1st ϩ 2nd
Positions
Polytoma obtusum–Chlamydomonas reinhardtii. . . .
Chlamydomonas humicola–C. reinhardtii . . . . . . . . .
P. obtusum–C. humicola . . . . . . . . . . . . . . . . . . . . . . .
Polytoma oviforme–C. humicola. . . . . . . . . . . . . . . . .
P. oviforme–C. reinhardtii . . . . . . . . . . . . . . . . . . . . .
0.25334
0.13349
0.22619
0.15501
0.15988
0.74805
0.31078
0.62276
0.43202
0.40108
0.09112
0.05843
0.08612
0.04732
0.06324
NOTE.—Estimated number of substitutions per site (sequence divergence) ϭ sequence differences per site corrected for multiple hits by the Jukes-Cantor method.
Table 4
Relative-Rate Tests on tufA Based on Tree Branch Lengths and Pairwise Distance Matrices
NONGREEN
SPECIES
GREEN
SPECIES
KAN/KAG
Parsimony
Neighbor
Joining
Maximum
Likelihood OUTGROUP
KON Ϫ KOG
All Sites
3rd
Position
1st ϩ 2nd
Positions
Polytoma obtusum
Polytoma oviforme
Chlamydomonas humicola
Chlamydomonas reinhardtii
3.23
1.26
2.7
1.18
4.39
1.34
C. reinhardtii
C. humicola
0.1998**
0.0215
0.4408*
0.1167
0.0278*
Ϫ0.00951
* Significant at the 1% level.
** Significant at the 0.1% level.
statistically significant. No rate increase was seen in the
branch leading to P. oviforme.
The Evolutionary Rate of tufA is Accelerated in P.
obtusum but not in P. oviforme
Sequences of tufA are available for C. humicola, P.
obtusum, C. reinhardtii, P. oviforme, and a number of
green algae outside of the Chlamydomonadaceae. Of
these, Codium is the closest relative, but when we used
it as an outgroup, all three tree-making algorithms
grouped Codium with P. obtusum, presumably due to
long-branch attraction. We therefore used only the four
Chlamydomonadaceae, with C. reinhardtii serving as
the outgroup for C. humicola and P. obtusum, and C.
humicola serving as the outgroup for C. reinhardtii and
P. oviforme. The tufA sequences of these four species
have 1,014 sites in common. The topology of the neigh-
bor-joining tree of these genes (fig. 2) is consistent with
the Rrn18 tree (fig. 1). However, long-branch attraction
was still a problem with the parsimony and maximum-
likelihood algorithms, which favored the tree that placed
P. obtusum with C. reinhardtii. The correct parsimony
tree (the one with the same topology as the Neighbor-
Joining tree and all trees involving rrn16 or Rrn18) was
the least parsimonious and had the lowest likelihood
scores, although not by much. The branch lengths for
the correct trees from all three algorithms are shown in
figure 2. In every case, the branch leading to P. obtusum
is much longer than that leading to C. humicola, while
P. oviforme shows no acceleration. Table 3 shows the
estimated pairwise numbers of substitutions among
these species; the branch leading to P. obtusum is ac-
celerated in all three trees (parsimony, Neighbor-Join-
ing, and maximum likelihood).
We performed relative-rate tests of the evolution of
tufA in P. obtusum and P. oviforme, using the pairwise
differences with the Jukes-Cantor correction for multiple
hits (table 3). In addition to calculating relative rates of
nucleotide substitutions for all aligned sites, we com-
pared first ϩ second codon positions with third codon
positions. The results are shown in table 4; all tests
showed a significantly higher substitution rate in the
branch leading to the nonphotosynthetic P. obtusum than
in the branch leading to the photosynthetic ingroup, C.
humicola. The increase was greater in the third codon
positions than in the first and second positions. No sig-
nificant difference was found between the branches lead-
ing to P. oviforme and C. reinhardtii, in agreement with
the data from rrn16.
The Plastid tufA Genes in Polytoma Remain
Functional After the Loss of Photosynthesis
One possible explanation for the accelerated evo-
lution of rrn16 and tufA is that the genes became non-
functional in the nonphotosynthetic lineages. This is un-
likely, given the evidence that they remain functional in
nonphotosynthetic land plants. We found additional ev-
idence that tufA remained subject to selection, and hence
functional, in the Polytoma lineages:
1. There are no premature stop codons in the entire
gene. This could be because no stop mutations oc-
curred since the loss of photosynthesis or because
they were eliminated by selection. For the P. uvella
clade, we estimated the probability of no stop mu-
tations occurring as follows: First, we assumed that
a truncation would not inactivate the protein if it oc-
curred between the carboxyl terminal of the protein
and the 14th amino acid, since the first 13 amino
acids are not involved in intermolecular bonding in
Escherichia coli (Kawashima et al. 1996). In the re-
mainder of the protein, we found 108 codons that
were one substitution away from being stop codons.
As described above, we know that more synonymous
1816 Vernon et al.
substitutions occurred along the branch leading to P.
obtusum than along the branch leading to C. humi-
cola from their common ancestor. There were 0.2235
extra substitutions per site in third codon positions,
which must have occurred after photosynthesis was
lost; this is also an estimate of the number of muta-
tions per site. We found 81 sense codons in the tufA
gene of P. obtusum that could have become a stop
codon as a result of one kind of substitution (e.g.,
UCG to UAG); the expected number of such substi-
tutions in the absence of selection was 81 ϫ 0.2235
ϫ 1/3 ϭ 6.034. We found 27 sense codons that could
have become stop codons as a result of either of two
kinds of substitutions (e.g., UAC to UAA or UAG);
the expected number of such substitutions was 27 ϫ
0.2235 ϫ 2/3 ϭ 4.023. Consequently, the expected
number of premature stop codons in the absence of
selection was 10.057, and from the Poisson distri-
bution the probability of finding no premature stop
codons was eϪ10.057 ϭ 4.3 ϫ 10Ϫ5. We conclude that
the tufA gene of P. obtusum must have been under
selection that eliminated genes with premature stop
codons most or all of the time since the loss of
photosynthesis.
Additional evidence was obtained using a tufA se-
quence obtained from P. uvella by Nedelcu (2001)
using the UTEX stock without subcloning. We
aligned 999 bp, or 333 complete codons, of P. uvella
and P. obtusum. The sequences of these species dif-
fered by 0.07892 synonymous substitutions per site,
all of which must have occurred since they diverged
from a common ancestor, after photosynthesis was
lost. We used parsimony to reconstruct 321 codons
of the sequence of their most recent common ances-
tor. This sequence contained 72 codons which could
have become stop codons if they had incurred single
specific mutations. The expected number of such
substitutions in the absence of selection was 72 ϫ
0.07892 ϫ 1/3 ϭ 1.894. The ancestral sequence also
contained 28 codons that could have become stop
codons as a result of either of two kinds of substi-
tutions; the expected number of such substitutions
was 28 ϫ 0.07892 ϫ 2/3 ϭ 1.473. Consequently, the
expected number of premature stop codons in the ab-
sence of selection is 3.367, and from the Poisson dis-
tribution the probability of finding no premature stop
codons was eϪ3.367 ϭ 0.0345.
2. All of the amino acid substitutions that occurred (out-
side of the hypervariable region discussed below) in
P. oviforme must be compatible with the normal
function of tufA, because each of the substituted ami-
no acids can be found at a comparable position in at
least one functional algal, cyanobacterial, or nonpho-
tosynthetic bacterial gene in the alignment array of
Delwiche, Kuhsel, and Palmer (1995). The same is
true of all but six amino acid substitutions seen in P.
obtusum. Both Polytoma sequences contain only con-
servative amino acid substitutions, except for some
nonconservative substitutions on the surface of the
EF-Tu protein of P. obtusum. None of these amino
acid substitutions are likely to change the folding of
the EF-Tu protein.
3. Nucleotide substitution rates in the tufA sequences at
first and second codon positions are much lower than
the rates at third positions (table 3), a difference that
can only be due to selection.
4. Relative to all other species, the tufA sequences from
Polytoma and Chlamydomonas contain numerous
amino acid substitutions, insertions, and deletions in
the hypervariable region. Despite the variability in
this region, we believe that it is compatible with
functionality of the protein in both Polytoma species
for the following reasons. First, the hypervariable re-
gion of P. oviforme is identical in length to that of
C. reinhardtii, and nearly identical in sequence, with
only two conserved amino acid differences between
the two species. The hypervariable region in P. ob-
tusum differs from that in C. reinhardtii in 13 sub-
stitutions and 3 gaps. However, this region is also
hypervariable and unusually long in functional EF-
Tu proteins from C. reinhardtii (Baldauf and Palmer
1990), C. humicola, and C. dysosmos, which are pho-
tosynthetic and therefore have functional EF-Tu pro-
teins. Second, the hypervariable region is on the out-
side surface of the protein in functional domain 3,
where amino acid changes or extra amino acids
would probably not affect the conformational chang-
es that occur during catalysis (Berthtold et al. 1993),
especially since the amino acid composition of the
hypervariable region is even more hydrophilic in P.
obtusum than in C. humicola and C. reinhardtii.
Moreover, the face of domain 3 that interacts with
other EF-Tu molecules (Kawashima et al. 1996) and
with the acceptor stem or T stem of the tRNA (Nis-
sen et al. 1995) is opposite the hypervariable region.
The Plastid rrn16 Genes in Polytoma Remain
Functional After the Loss of Photosynthesis
The growth of P. uvella and of another member of
the same clade, P. uvella 62-3 ϭ P. mirum, is inhibited
by the antibiotics erythromycin, streptomycin, and spec-
tinomycin at 800, 400, and 50 mg/ml, respectively (data
not shown). Scherbel, Behn, and Arnold (1974) previ-
ously found that growth of P. mirum is inhibited by
streptomycin. This antibiotic is known to inhibit chlo-
roplast protein synthesis in C. reinhardtii at similar or
lower (erythromycin) concentrations (Harris 1989).
Spectinomycin sensitivity is especially interesting: it has
no known side effects, and Chlamydomonas mutants re-
sistant to high concentrations have mutations only in the
rrn16 gene. These data suggest that Polytoma, like
Chlamydomonas, synthesizes at least one essential pro-
tein on plastid ribosomes which contain functional 16S
rRNA molecules.
Consistent with the antibiotic studies, an analysis of
the primary and secondary structures of the 16S rRNA
molecules, inferred from the rn16 sequences, strongly sup-
ports functionality of the molecules. Here we present only
a summary; the complete analysis is included with the
secondary-structure figures and sequence alignments in the
Accelerated Evolution of Leucoplast Genes 1817
FIG. 4.—Relative synonymous codon usage values of codons in
Polytoma obtusum (open bars) compared with Chlamydomonas hum-
icola (filled bars). Separate graphs are shown for amino acids that are
sixfold-, fourfold-, and twofold-degenerate. Stars represent codons for
which the complementary anticodons are found in tRNAs encoded in
plant chloroplast genomes.
supplement and at the web site http://www.rna.icmb.
utexas.edu/PUBLICATIONS/BIRKY/.
1. The primary and secondary structures for the three
Polytoma rRNA sequences contain all of the struc-
tural elements present in the chloroplast rRNAs that
are functional, and, apart from a few insertions dis-
cussed below, all of the nucleotide positions in the
Polytoma sequences correspond to all of the positions
present in these 70 functional 16S rRNA chloroplast
sequences (figs. 1, 2, 3a, and 3d–f in the supplement
and at our web site mentioned above; see also Gutell
1994).
2. Differences between the Polytoma sequences occur
at positions that also vary in the functional SSU
rRNAs in the nuclear genes of the Eucarya, Bacteria,
and Archaea and in chloroplast and mitochondrial
genes (http://www.rna.icmb.utexas.edu/RDBMS/,
http://www.rna.icmb.utexas.edu/CSI/BPFREQ/
16S-MODEL-BP/, and figs. 2, 3a, and 3c in the sup-
plement and at our web site mentioned above). Con-
versely, positions that are conserved in the three phy-
logenetic domains plus chloroplasts and mitochon-
dria are also conserved in the Polytoma sequences.
3. Base pairs were predicted with comparative sequence
analysis (Gutell 1996; http://www.rna.icmb.utexas.
edu/METHODS/); two aligned positions that change
coordinately are considered possible base pairs.
These base pairs are highly conserved in the Poly-
toma SSU rRNA and are consistent with function-
ality. There are 70 base-paired positions at which the
Polytoma sequences differ from the chloroplast con-
sensus: of these, 24 are compensatory changes; 10
involve an A·U or G·C interchange to a G·U base
pair; and only 12 change an A·U, G·C, or G·U pair
to a noncanonical pair (figs. 2 and 3 at our web site
mentioned above). The number of noncanonical base
pairs in the Polytoma SSU rRNA is approximately
the same as in other rRNAs that are known to be
active in protein synthesis (unpublished data).
4. A comparison of the sequences of P. obtusum and P.
uvella reveals functional and structural constraints
acting on these 16S rRNA sequences since their com-
mon ancestor; because this ancestor was nonphoto-
synthetic, all of the differences between the two Po-
lytoma isolates arose in nonphotosynthetic lineages.
Among the 55 variable paired positions, 30 are in-
volved in a covariation at base pairs predicted with
comparative sequence analysis (fig. 4 in the supple-
ment and at our web site mentioned above). Of these
positions, 12 involve an exchange between Watson-
Crick base pairs, while three involve a Watson-Crick
base pair with something else. Of the 25 paired po-
sitions with a single change, 14 exchanged between
AU or GC and GU. Since our covariation algorithms
search for simultaneous changes regardless of the
base pair types, the fact that AU and GC are the most
frequent pairs identified at positions that covary in-
dicates that these genes are still undergoing positive
selection. In addition, several of the noncanonical co-
variation base pairs have been experimentally veri-
fied (reviewed in Gutell 1996), suggesting that all of
these covarying bases could be base-paired in the
secondary structure.
5. Substitution rates were higher at paired sites than at
unpaired sites, presumably because single substitu-
tions at paired sites are detrimental but pairs of com-
pensating changes are neutral (data not shown).
There are three large uncharacteristic insertions in
the Polytoma sequences. The P. oviforme gene has an
incompletely sequenced insertion larger than 438 bp be-
tween positions 427 and 428 (E. coli numbering). Po-
lytoma obtusum and P. uvella have insertions of 223 bp
and Ͼ596 bp, respectively, between positions 746 and
747. Finally, P. uvella has a second insertion of 85 nt
between (E. coli) positions 138 and 140. All of these
insertions occur in regions of the 16S rRNA that are
variable in sequence and not considered to be function-
ally important. The first, between positions 138 and 140
(E. coli numbering) contains 85 extra nucleotides in P.
1818 Vernon et al.
uvella; 75 of these are A or U. We believe that this
insertion may be an internal transcribed spacer that is
excised after transcription for two reasons. First, this
extra sequence has a very high AU content, a charac-
teristic of internal transcribed spacer (ITS) sequences.
Second, an ITS has been found in the bacterium Cam-
pylobacter sputorum 16S rRNA at position 220 (E. coli
numbering; VanCamp et al. 1993; accession number
X67775), which is directly across the helix from our
presumed ITS. The two other insertions occur at the
ends of helices and may be either introns or ITSs.
Relaxed Selection Is at Least Partly Responsible for
the Accelerated Evolution
The preceding sections show that the rrn16 and
tufA genes in Polytoma remained subject to selection,
and hence functional, long after the loss of photosyn-
thesis. Although we cannot exclude the possibility that
these genes lost their function recently, they were prob-
ably still functional when we tested their antibiotic sen-
sitivity. Consequently the increased rate of base pair
substitution in the tufA and rrn16 genes in the P. uvella
lineage is not due to complete loss of function. We
therefore considered four other possible explanations for
their accelerated evolution:
1. An increased mutation rate would increase the rate
of substitution, provided, of course, that the fixation
probability was not reduced by the same factor. How-
ever, an increase in mutation rate cannot explain our
data because it would affect different functional re-
gions of a gene to the same extent; in contrast, we
found that the substitution rate in the rrn16 gene was
greater in regions coding for the stems of the rRNA
than in regions coding for the loops, and in tufA there
was a much greater increase in the synonymous rate
than in the nonsynonymous rate. These observations
show that an increased mutation rate is not the sole
cause of the accelerated evolution in rrn16 and tufA
but do not rule it out as a contributing factor.
2. There could be an increased level of adaptive or pos-
itive selection by fixation of mutations that were for-
merly neutral or detrimental but are now advanta-
geous. This explanation is unlikely because of the
large number of such mutations that would be re-
quired to explain the data. For example, parsimony
analysis (data not shown) suggests that about 152
base pair substitutions occurred in the rrn16 gene
along the lineage leading to P. uvella from the most
recent common ancestor of P. uvella and C. humi-
cola, while only 35 occurred in the C. humicola lin-
eage. The nonphotosynthetic lineage thus incurred
152 Ϫ 35 ϭ 117 more substitutions that would have
to be adaptive; this is 77% of all of the substitutions
after photosynthesis was lost. Moreover, it is proba-
bly not a complete explanation because it is not likely
to increase the synonymous substitution rate. How-
ever, it is possible that positive selection contributed
in part to the increase in substitution rate.
3. A reduced effective population size (Ne) would make
natural selection against detrimental mutations less
effective. By making some mutations that were for-
merly detrimental effectively neutral, it would in-
crease the overall fixation probability of new muta-
tions. A decrease in the frequency of recombination
of plastid genes would reduce Ne as a result of the
Hill-Robertson effect (Birky and Walsh 1988), but
there is no reason to suspect this in nonphotosyn-
thetic lineages. A more plausible reason for a de-
crease in Ne is that an obligate heterotroph has fewer
niches available to it, which might reduce Ne by de-
creasing the total population size or increasing the
variance in offspring number. In any event, a reduced
Ne cannot explain our data by itself, because the ratio
KAN/KAG was greater for kinds of substitutions sub-
ject to weak selection (third positions in tufA; stems
in rrn16) than for those subject to relatively strong
selection (first and second positions in tufA; loops in
rrn16). A theoretical analysis of the combined effects
of directional selection and drift, using Kimura’s
(1957) equation for the fixation probability of a mu-
tation, showed that a decrease in effective population
size by itself would cause KAN/KAG to be larger, not
smaller, for strongly selected detrimental mutations
than for relatively weakly selected mutations (unpub-
lished data). In addition, a reduced Ne would affect
nuclear genes as well as organelle genes, whereas we
observed no significant increase in substitution rate
in the nuclear Rrn18 genes of P. obtusum and P.
uvella. We cannot rule out the possibility that some
of the observed acceleration in evolutionary rate is
due to a reduction in effective population size too
small to be detected in the nuclear gene, but this
cannot be a complete explanation.
4. Relaxed selection against detrimental mutations is a
plausible explanation for the increase in substitution
rate in nonphotosynthetic lineages. This would re-
quire that some mutations that were detrimental in a
chloroplast are neutral or effectively neutral in a leu-
coplast and thus more likely to be fixed. The rrn16
and tufA genes are involved in translation of plastid
proteins, and mutations in these genes are likely to
be detrimental if they directly or indirectly reduce the
rate or accuracy of protein synthesis. Selection would
be relaxed if heterotrophic algae could tolerate either
a lower rate of plastid protein synthesis or a higher
proportion of amino acid substitutions in plastid pro-
teins due to mistranslation. At least the first of these
conditions is very likely to be met. Most of the chlo-
roplast protein-coding genes code for proteins in-
volved in photosynthesis, and all mutations affecting
these genes are selectively neutral after photosynthe-
sis is lost in the ancestor of a nonphotosynthetic
clade; as a result, they will accumulate mutations that
block transcription or translation or will be deleted
entirely. Leucoplasts thus have to translate only about
two-thirds as many proteins. Among the genes that
do not have to be expressed in heterotrophic algae or
plants is rbcL, which encodes one of the most abun-
dant proteins in photosynthetic organisms, ribulose
bisphosphate carboxylase. A number of photosyn-
Accelerated Evolution of Leucoplast Genes 1819
Table 5
Codon Adaptation Index (CAI) and Codon Bias Index
(CBI) for tufA in Chlamydomonas and Polytoma Species
Organism CAI(Cre) CAI(Cum) CBI
Chlamydomonas reinhardtii. . . .
Polytoma oviforme . . . . . . . . . . .
Chlamydomonas humicola . . . . .
Polytoma obtusum. . . . . . . . . . . .
0.487
0.359
0.432
0.22
0.77
0.688
0.687
0.37
0.775
0.774
0.722
0.455
NOTE.—CAI(Cre) is calculated relative to the codon usage in C. reinhardtii
psbA gene; CAI(Cum) is calculated relative to the codon usage in the psbA genes
of C. reinhardtii, Porphyra purpurea, Odontella sinensis, and Cyanophora par-
adoxa.
thetic genes show high codon bias characteristic of
highly expressed genes.
We hypothesized that the loss of photosynthesis in
Polytoma permitted a lower rate of synthesis of elon-
gation factor Tu because these heterotrophic organisms
can tolerate a lower overall rate of protein synthesis than
their photosynthetic relatives. This hypothesis predicts
that the tufA gene will show less codon bias in P. ob-
tusum than in C. humicola, and probably less codon bias
than in other Chlamydomonas species. Codon bias is
generally greater in highly expressed genes, probably
because selection favors codons that are read by more
abundant tRNAs (e.g., Morton 1993, 1996; Sharp et al.
1995). We calculated two measures of codon bias for
the tufA genes of the Chlamydomonadaceae. The CBI
measures the overall level of codon bias in a gene and
ranges from 0 to 1; the CAI (Sharp and Li 1987) mea-
sures codon bias for a gene relative to one or more high-
ly expressed genes that show high codon bias. Two ref-
erence databases of codon use in a highly expressed
plastid gene were provided by Brian Morton: the psbA
gene of C. reinhardtii and the combined psbA genes of
C. reinhardtii, Porphyra purpurea, Cyanophora para-
doxa, and Odontella sinensis. The CBI and both CAI
values were lower for P. obtusum than for the two Chla-
mydomonas species and P. oviforme (table 5).
These data show that codon bias in tufA is reduced
in P. obtusum compared with its closest photosynthetic
relative, C. humicola. The change in codon bias might
be due to a change in relative mutation rates, e.g., to
favor AT pairs over GC. To test this possibility, we cal-
culated expected numbers of the four codons in all four-
fold-degenerate amino acids and in the fourfold-degen-
erate codon set of sixfold-degenerate amino acids, as-
suming that the expected proportions of A, T, G, and C
were the averages of those found in the third positions
of all codons. The deviation of observed numbers from
these expected numbers can be summarized by the chi-
square value, which was much greater for C. humicola
(64.3) than for P. obtusum (46.2). We also found that
the percentages of AϩT were not significantly different
in the first ϩ second codon positions in C. humicola and
P. obtusum (53.0% vs. 52.6% respectively; P k 0.05
by Fisher’s exact test) but were significantly different in
the third codon positions (60.5% vs. 75.9%, respective-
ly; P K 0.001). Since mutation rates do not differ sys-
tematically between codon positions, we conclude that
the third-position base composition difference between
Chlamydomonas and Polytoma is due to selection on
synonymous codons rather than to mutation pressure.
The reduced codon bias in Polytoma is presumably due
to relaxed selection against mutations that substitute
less-used codons for those that are most used in the C.
humicola lineage.
A more detailed comparison of codon bias in P.
obtusum and C. humicola suggests that the bias that is
relaxed in the nonphotosynthetic species is largely a
preference for codons that can be read without wobble
(a mismatched base pair between tRNA and the mRNA
third codon position). Several plant genomes have been
completely sequenced and found to have a set of 29
tRNAs. Gene sequences of 15 of these have been
mapped on the chloroplast genome of C. humicola (E.
Boudreau, personal communication), and no tRNA
genes have been identified in any green alga other than
those known from plant chloroplasts, so it is likely that
these algae have the same set of chloroplast tRNAs as
do plants. Many codons require wobble to be read by
this set of tRNAs, and several require ‘‘superwobble,’’
in which one tRNA reads A, G, C, or U in the third
codon position. We calculated the RSCU for all degen-
erate amino acids (those using more than one codon) in
P. obtusum and C. humicola. Figure 4 shows that much
of the difference in codon bias is in the sixfold-degen-
erate amino acids leucine, serine, and arginine, which
preferentially use codons that can be read without wob-
ble. This preference is much greater for C. humicola
than for P. obtusum (chi-square test; 0.01 Ͼ P Ͼ 0.001).
Similar analyses for fourfold- and twofold-degenerate
amino acids show no significant difference between
these species; both species show preference for no wob-
ble over wobble in the twofold-degenerate amino acids
(P ഠ 0.9) and for no wobble or normal wobble over
superwobble in those fourfold-degenerate amino acids
that require superwobble (0.8 Ͼ P Ͼ 0.7).
In the case of rrn16, one can imagine that the re-
duced load of protein synthesis in the leucoplast has
resulted in less intense selection for translation rate or
fidelity. Alternatively, there may be a greater tolerance
for mutations that affect the rate of transcription of the
gene, of posttranscriptional processing, or of assembly
of the ribosome. Relaxed selection for translational rate
or fidelity might result in decreased stability of the sec-
ondary structure of the small-subunit rRNA. Compared
with C. humicola, both P. uvella and P. obtusum had
fewer GC pairs and more AU pairs, with GC/AU ratios
of 1.80, 1.40, and 1.34, respectively. These numbers are
compatible with the hypothesis that the loss of photo-
synthesis has been accompanied by a relaxation of se-
lection for helix stability in the leucoplast small-subunit
rRNA molecule. However, it is also compatible with a
decrease in the GϩC content of the genome as a whole.
A more definitive test of the hypothesis will require
comparisons of secondary-structure stability in a larger
sample of nonphotosynthetic species and their close
photosynthetic relatives, using the nuclear small-subunit
rRNA molecule as a control, together with data on the
GϩC contents of the genomes.
1820 Vernon et al.
The bulk of the evidence suggests that the in-
creased rate of base pair substitution in the tufA and
rrn16 genes of Polytoma is due at least in part to relaxed
selection, such that mutations that reduce the rate or
fidelity of translation are less likely to be eliminated.
Rigorous tests of this hypothesis will require more de-
tailed analyses of a larger number of sequences of these
and other plastid expression genes.
Polytoma oviforme May Have Lost Photosynthesis
More Recently
The tufA gene of P. oviforme shows codon bias
similar to that of several close green relatives, and nei-
ther the rrn16 gene nor the tufA gene shows an increased
substitution rate in the lineage leading to P. oviforme
relative to green lineages. In contrast, the expression
gene sequences that we sampled from the P. uvella clade
all show significant rate increases compared with green
relatives, and tufA from P. obtusum (the only protein-
coding gene we obtained from the P. uvella clade)
shows minimal codon bias. These data are consistent
with the hypothesis that the P. uvella clade may have
resulted from a more ancient loss of photosynthesis than
the P. oviforme lineage, and only the P. uvella clade has
been evolving without photosynthesis long enough for
the consequences of relaxed selection to be evident.
Accelerated Evolution of Leucoplast Expression Genes
in Other Organisms
Smaller increases in evolutionary rates have been
demonstrated in the leucoplast expression genes rrn16,
rrn23, and tufA, as well as in the rbcL gene, of the
heterotrophic euglenoid alga Astasia longa (Siemeister,
Buchholz, and Hachtel 1990; Siemeister and Hachtel
1990a, 1990b). An increase in the substitution rate of
the rrn16 genes of some nonphotosynthetic angiosperms
was reported by Nickrent, Duff, and Konings (1997),
although it was not verified by relative-rate tests. The
increase was accompanied by an increase in AϩT con-
tent in the genes and consequently by an increase in
destabilizing AU pairs in the rRNA, such as we ob-
served. The authors did not compare the rates in stems
and loops. Nickrent and Starr (1994) also reported an
increased substitution rate in the nuclear Rrn18 genes of
a different set of species of holoparasitic angiosperms
and presented evidence that the increase could not be
explained by a decrease in generation time or effective
population size. In contrast, we observed no increase in
the evolutionary rate of Rrn18 in Polytoma (Rumpf et
al. 1996) or Polytomella (unpublished data). The holo-
parasitic angiosperm Epifagus virginiana was subjected
to an intensive analysis of evolutionary rates in leuco-
plast genes; rate increases were detected in rrn16 and
rrn23 and the pooled data for 17 tRNAs and 15 ribo-
somal proteins (Wolfe et al. 1992). In the ribosomal pro-
teins, accelerated evolution was seen in both nonsynon-
ymous and synonymous substitutions; in contrast to our
results, the increase was greater for nonsynonymous
substitutions. Epifagus also differed from Polytoma in
showing no change in codon bias (Morden et al. 1991).
Wolfe et al. (1992) proposed that the increase in non-
synonymous substitutions was probably due to relaxed
selection for both the rate and fidelity of protein syn-
thesis, while the increase in synonymous substitutions
probably resulted from an increased mutation rate. How-
ever, alternative explanations were not ruled out. In par-
ticular, both the synonymous and the nonsynonymous
rate increase might be entirely due to a decreased effec-
tive population size, which under some circumstances
can increase the rate of detrimental substitutions more
when they are under stronger selection (unpublished
data). dePamphilis, Young, and Wolfe (1997) found that
the substitution rate of the ribosomal protein gene rps2
was significantly accelerated in some, but not all, hol-
oparasitic (nonphotosynthetic) angiosperm lineages in
the Orobanchaceae and Sdrophulariaceae. In some cas-
es, there was a significant increase in synonymous but
not nonsynonymous substitutions, and in some other
cases, the reverse held. dePamphilis, Young, and Wolfe
(1997) proposed that increases in nonsynonymous sub-
stitution rates are probably due to reduced functional
constraint when abundant photosynthetic proteins do not
have to be synthesized. The increases in synonymous
substitutions were attributed to increased mutation rates.
However, no formal analyses of the data were given in
support of these conclusions.
The independent losses of photosynthesis in a num-
ber of different plants and algae provide biologists with
a series of natural experiments in which selection has
been reduced or eliminated to different extents in genes
with different functions. Future comparative studies of
leucoplast expression and photosynthetic gene sequenc-
es from Polytoma and Polytomella will help to unravel
the roles of mutation and selection in the molecular evo-
lution of plastid genes. The data to date suggest that
accelerated evolution of plastid expression genes is a
general consequence of the loss of photosynthesis and
that the rate of evolution of expression genes is strongly
influenced by the rate of protein synthesis that they must
sustain.
Supplementary Material
The DNA sequences have been deposited in
GenBank under accession numbers AF352839 (P. ob-
tusum tufA), AF352840 (P. oviforme tufA), AF352838
(C. humicola tufA), AF397587 (P. obtusum rrn16),
AF397589 (P. uvella rrn16), AF397588 (P. oviforme
rrn16), AF397590 (C. reinhardtii rrn16), and
AF397586 (C. humicola rrn16). Additional supplemen-
tary materials on the Molecular Biology and Evolution
web site include tufA and rrn16 sequence alignments
in sequential GenBank format and in interleaved Pretty
Print format, and the complete argument and additional
data showing that the rrn16 gene is functional in
Polytoma.
Acknowledgments
We are grateful to Brian Morton for supplying his
Macintosh Pascal program and reference databases for
calculating measures of codon bias, and to Aurora Ned-
Accelerated Evolution of Leucoplast Genes 1821
elcu for allowing us to analyze her tufA sequence from
P. uvella before it was published. Sally Otto analyzed
Kimura’s (1957) equation to verify that reducing Ne af-
fects strongly selected sites more than weakly selected
sites. Primers were kindly provided by Paul Fuerst and
Jeffrey Palmer. The large tufA database was kindly pro-
vided by Charles Delwiche. Other members of the Birky
lab at Ohio State University, notably Pamela Mackowski
and Stacy Seibert, assisted in numerous ways. We are
grateful to Jennifer Wernergreen, Aurora Nedelcu, and
two anonymous reviewers for helpful comments on ear-
lier manuscripts. D.V. submitted the data and a prelim-
inary analysis in a thesis in partial fulfillment of the
requirements for the Ph.D. degree at the Ohio State Uni-
versity. This work was supported in part by research
grants from NIH (GM34094) and NSF (BSR-9107069)
to C.W.B. and from NIH (GM48207) to R.R.G., and by
funds from the Ohio State University, the University of
Arizona, and the University of Texas.
LITERATURE CITED
BALDAUF, S. L., and J. D. PALMER. 1990. Evolutionary transfer
of the chloroplast tufA gene to the nucleus. Nature 344:
262–265.
BERTHTOLD, H., L. RESHETNIKOVA, C. O. A. REISER, N. K.
SCHIRMER, M. SLPRINZL, and R. HILGENFELD. 1993. Crystal
structure of active elongation factor Tu reveals major do-
main rearrangements. Nature 365:126–132.
BIRKY, C. W. JR., and J. B. WALSH. 1988. Effects of linkage
on rates of molecular evolution. Proc. Natl. Acad. Sci. USA
85:6414–6418.
DELWICHE, C., M. KUHSEL, and J. D. PALMER. 1995. Phylo-
genetic analysis of tufA sequences indicates a cyanobacter-
ial origin of all plastids. Mol. Phylogenet. Evol. 4:110–128.
DEPAMPHILIS, C. W., and J. D. PALMER. 1990. Loss of photo-
synthetic and chlororespiratory genes from the plastid ge-
nome of a parasitic flowering plant. Nature 348:337–339.
DEPAMPHILIS, C. W., N. D. YOUNG, and A. D. WOLFE. 1997.
Evolution of plastid gene rps2 in a lineage of hemiparasitic
and holoparasitic plants: many losses of photosynthesis and
complex patterns of rate variation. Proc. Natl. Acad. Sci.
USA 94:7367–7372.
ETTL, H. 1976. Die Gattung Chlamydomonas Ehrenberg. Beih.
Nova Hedwigia 49:1–1122.
FELSENSTEIN, J. 1993. PHYLIP (phylogeny inference package).
Version 3.56. Distributed by the author, Department of Ge-
netics, University of Washington, Seattle.
GILBERT, D. G. 1992. SeqApp, a biological sequence editor
and analysis program for Macintosh computers. Available
via gopher or anonymous ftp to ftp.biol.indiana.edu.
GILLHAM, N. W. 1994. Organelle genes and genomes. Oxford
University Press, New York.
GORDON, J., R. RUMPF, S. L. SHANK, D. VERNON, and C. W.
BIRKY JR. 1995. Sequences of the rrn18 genes of Chla-
mydomonas humicola and C. dysosmos are identical, in
agreement with their combination in the species C. applan-
ata (Chlorophyta). J. Phycol. 31:312–313.
GRANT, D. M., N. W. GILLHAM, and J. E. BOYNTON. 1980.
Inheritance of chloroplast DNA in Chlamydomonas rein-
hardtii. Proc. Natl. Acad. Sci. USA 77:6067–6071.
GUTELL, R. R. 1994. Collection of small subunit (16S- and
16S-like) ribosomal RNA structures: 1994. Nucleic Acids
Res. 22:3502–3507.
———. 1996. Comparative sequence analysis and the structure
of 16S and 23S rRNA. Pp. 111–128 in R. A. ZIMMERMAN
and A. E. DAHLBERG, eds. Ribosomal RNA. Structure, evo-
lution, processing, and function in protein biosynthesis.
CRC Press, New York.
HARRIS, E. H. 1989. The Chlamydomonas sourcebook. Aca-
demic Press, New York.
JUKES, T. H., and C. R. CANTOR. 1969. Evolution of protein
molecules. Pp. 21–123 in H. N. MUNRO, eds. Mammalian
protein metabolism. Academic Press, New York.
KAWASHIMA, T., C. BERTHET-COLOMINAS, M. WULFF, S. CU-
SACK, and R. LEBERMAN. 1996. The structure of the E. coli
EF-Tu/EF-Ts complex at 2.5 A resolution. Nature 379:511–
518.
KIMURA, M. 1957. Some problems of stochastic processes in
genetics. Ann. Math. Stat. 28:882–901.
———. 1980. A simple method for estimating evolutionary
rates of base substitutions through comparative studies of
nucleotide sequences. J. Mol. Evol. 16:111–120.
LANG, N. J. 1963. Electron-microscopic demonstration of plas-
tids in Polytoma. J. Protozool. 10:333–339.
LEMIEUX, C., C. OTIS, and M. TURMEL. 2000. Ancestral chlo-
roplast genome in Mesostigma viride reveals an early
branch of green plant evolution. Nature 249:649–652.
LUSSON, N. A., P. M. DELAVAULT, and P. A. THALOUARN.
1998. The rbcL gene from the non-photosynthetic parasite
Lathraea clandestina is not transcribed by a plastid-encoded
RNA polymerase. Curr. Genet. 34:212–215.
MAIDAK, B. L., N. LARSEN, M. J. MCCAUGHEY, R. OVERBEEK,
G. J. OLSEN, K. FOGEL, J. BLANDY, and C. R. WOESE. 1994.
The Ribosomal Database Project. Nucleic Acids Res. 22:
3485–3487.
MORDEN, C. W., K. H. WOLFE, C. W. DEPAMPHILIS, and J. W.
PALMER. 1991. Plastid translation and transcription genes
in a non-photosynthetic plant: intact, missing and pseudo
genes. EMBO J. 10:3281–3288.
MORTON, B. R. 1993. Chloroplast DNA codon use: evidence
for selection at the psbA locus based on tRNA availability.
J. Mol. Evol. 37:273–280.
———. 1996. Selection on the codon bias of Chlamydomonas
reinhardtii chloroplast genes and the plant psbA gene. J.
Mol. Evol. 43:28–31.
MUSE, S. V., and B. S. WEIR. 1992. Testing for equality of
evolutionary rates. Genetics 132:269–276.
NEDELCU, A. M. 2001. Complex patterns of plastid 16S rRNA
gene evolution in nonphotosynthetic green algae. J. Mol.
Evol. (in press).
NICKRENT, D. L., R. J. DUFF, and D. A. M. KONINGS. 1997.
Structural analyses of plastid-derived 16S rRNAs in holo-
parasitic angiosperms. Plant Mol. Biol 34:731–743.
NICKRENT, D. L., and E. M. STARR. 1994. High rates of nu-
cleotide substitution in nuclear small-subunit (18S) rDNA
from holoparasitic flowering plants. J. Mol. Evol. 39:62–
70.
NISSEN, P., M. KJELDGAARD, S. THIRUP, G. POLEKHINA, L.
RESHETNIKOVA, B. F. C. CLARK, and J. NYBORG. 1995.
Crystal structure of the ternary complex of Phe-tRNA, EF-
Tu, and a GTP analog. Science 270:1464–1472.
RUMPF, R., D. VERNON, D. SCHREIBER, and C. W. BIRKY JR.
1996. Evolutionary consequences of the loss of photosyn-
thesis in Chlamydomonadaceae: phylogenetic analysis of
Rrn18 (18S rDNA) in 13 Polytoma strains (Chlorophyta).
J. Phycol. 32:119–126.
SAMBROOK, J., E. F. FRITSCH, and T. MANIATIS. 1989. Molec-
ular cloning: a laboratory manual. Cold Spring Harbor Lab-
oratory Press, Cold Spring Harbor, N.Y.
1822 Vernon et al.
SARICH, V. M., and A. C. WILSON. 1973. Generation time and
genomic evolution in primates. Science 179:1144–1147.
SCHERBEL, G., W. BEHN, and C. G. ARNOLD. 1974. Untersu-
chungen zur genetischen Funktion des farblosen Plastiden
von Polytoma mirum. Arch. Microbiol. 96:205–222.
SHARP, P. M., M. AVEROF, A. T. LLOYD, G. MATASSI, and J.
F. PEDEN. 1995. DNA sequence evolution: the sounds of
silence. Philos. Trans. R. Soc. Lond. B Biol. Sci. 349:241–
247.
SHARP, P. M., and W.-H. LI. 1987. The codon adaptation in-
dex—a measure of directional synonymous codon usage
bias, and its potential applications. Nucleic Acids Res. 15:
1281–1295.
SIEMEISTER, G., C. BUCHHOLZ, and W. HACHTEL. 1990. Genes
for the plastid elongation factor Tu and ribosomal protein
S7 and six tRNA genes on the 73 kb DNA from Astasia
longa that resembles the chloroplast DNA of Euglena. Mol.
Gen. Genet. 220:425–432.
SIEMEISTER, G., and W. HACHTEL. 1990a. Organization and
nucleotide sequence of ribosomal RNA genes on a circular
73 kbp DNA from the colourless flagellate Astasia longa.
Curr. Genet. 17:433–438.
———. 1990b. Structure and expression of a gene encoding
the large subunit of ribulose-1,5-bisphosphate carboxylase
(rbcL) in the colourless euglenoid flagellate Astasia longa.
Plant Mol. Biol. 14:825–833.
SIU, C.-H., K.-S. CHIANG, and H. SWIFT. 1976. Characteriza-
tion of cytoplasmic and nuclear genomes in the colorless
alga Polytoma. III. Ribosomal RNA cistrons of the nucleus
and leucoplast. J. Cell Biol. 69:383–392.
SWOFFORD, D. L. 1998. PAUP*: phylogenetic analysis using
parsimony (*and many other methods). Version 4. Sinauer,
Sunderland, Mass.
TAMURA, K. 1992. Estimation of the number of nucleotide sub-
stitutions when there are strong transition-transversion and
GϩC content biases. Mol. Biol. Evol. 9:678–687.
TURMEL, M., C. OTIS, and C. LEMIEUX. 1999. The complete
chloroplast DNA sequence of the green alga Nephroselmis
olivacea: insights into the architecture of ancestral chloro-
plast genomes. Proc. Natl. Acad. Sci. USA 96:10248–
10253.
VANCAMP, G., Y. VAN DE PEER, J. M. NEEFS, P. VANDAMME,
and R. DEWACHTER. 1993. Presence of internal transcribed
spacers in the 16S and 23S ribosomal RNA genes of Cam-
pylobacter. Syst. Appl. Microbiol. 16:361–368.
VERNON, D. 1996. Evolutionary consequences of the loss of
photosynthesis in the nonphotosynthetic chlorophyte alga
Polytoma. Dissertation, Ohio State University, Columbus.
VERNON-KIPP, D., S. A. KUHL, and C. W. J. BIRKY. 1989.
Molecular evolution of Polytoma, a non-green chlorophyte.
Pp. 284–286 in C. T. BOYER, J. C. SHANNON, and R. C.
HARDISON, eds. Physiology, biochemistry, and genetics of
nongreen plastids. American Society of Plant Physiologists,
Rockville, Md.
WIMPEE, C. F., R. MORGAN, and R. L. WROBEL. 1992a. An
aberrant plastid ribosomal RNA gene cluster in the root
parasite Conopholis americana. Plant Mol. Biol. 18:275–
285.
———. 1992b. Loss of transfer RNA genes from the plastid
16S-23S ribosomal RNA gene spacer in a parasitic plant.
Curr. Genet. 21:417–422.
WOLFE, A. D., and C. W. DEPAMPHILIS. 1998. The effect of
relaxed functional constraints on the photosynthetic gene
rbcL in photosynthetic and nonphotosynthetic parasitic
plants. Mol. Biol. Evol. 15:1243–1258.
WOLFE, K. H., C. W. MORDEN, S. C. EMS, and J. D. PALMER.
1992. Rapid evolution of the plastid translational apparatus
in a nonphotosynthetic plant: loss or accelerated sequence
evolution of tRNA and ribosomal protein genes. J. Mol.
Evol. 35:304–317.
WOLFE, K. H., C. W. MORDEN, and J. D. PALMER. 1992. Func-
tion and evolution of a minimal plastid genome from a non-
photosynthetic parasitic plant. Proc. Natl. Acad. Sci. USA
89:10648–10652.
WU, C.-I., and W.-H. LI. 1985. Evidence for higher rates of
nucleotide substitution in rodents than in man. Proc. Natl.
Acad. Sci. USA 82:1741–1745.
YOUNG, D. Y., and C. W. DEPAMPHILIS. 2000. Purifying selec-
tion detected in the plastid gene matK and flanking ribo-
zyme regions within a group II intron of nonphotosynthetic
plants. Mol. Biol. Evol. 17:1933–1941.
DAVID RAND, reviewing editor
Accepted May 30, 2001

Contenu connexe

Tendances

Mendel's principle of inheritance
Mendel's principle of inheritanceMendel's principle of inheritance
Mendel's principle of inheritance
DepEd
 
Reeta yadav. roll no. 02. transposable element in eukaryotes.
Reeta yadav. roll no. 02. transposable element in eukaryotes.Reeta yadav. roll no. 02. transposable element in eukaryotes.
Reeta yadav. roll no. 02. transposable element in eukaryotes.
Manisha Jangra
 
AP Biology Genetic Diversity and Operons in bacteria
AP Biology Genetic Diversity and Operons in bacteriaAP Biology Genetic Diversity and Operons in bacteria
AP Biology Genetic Diversity and Operons in bacteria
Stephanie Beck
 

Tendances (20)

Transposons ppt
Transposons pptTransposons ppt
Transposons ppt
 
Properties and effects of transposable elements
Properties and effects of transposable elementsProperties and effects of transposable elements
Properties and effects of transposable elements
 
2014 intro-genetics
2014 intro-genetics2014 intro-genetics
2014 intro-genetics
 
Mendel's principle of inheritance
Mendel's principle of inheritanceMendel's principle of inheritance
Mendel's principle of inheritance
 
Reeta yadav. roll no. 02. transposable element in eukaryotes.
Reeta yadav. roll no. 02. transposable element in eukaryotes.Reeta yadav. roll no. 02. transposable element in eukaryotes.
Reeta yadav. roll no. 02. transposable element in eukaryotes.
 
Lesson 14.3
Lesson 14.3Lesson 14.3
Lesson 14.3
 
Mutation & its detection
Mutation & its detectionMutation & its detection
Mutation & its detection
 
Terminology related to molecular biology and genetics
Terminology related to molecular biology and geneticsTerminology related to molecular biology and genetics
Terminology related to molecular biology and genetics
 
2014 genetics-concepts
2014 genetics-concepts2014 genetics-concepts
2014 genetics-concepts
 
Biotechnology and recombinant dna technique(UOG)
Biotechnology and recombinant dna technique(UOG)Biotechnology and recombinant dna technique(UOG)
Biotechnology and recombinant dna technique(UOG)
 
Transposons in drosophila - P element
Transposons in drosophila - P elementTransposons in drosophila - P element
Transposons in drosophila - P element
 
TRANSPOSABLE ELEMENTS
TRANSPOSABLE ELEMENTSTRANSPOSABLE ELEMENTS
TRANSPOSABLE ELEMENTS
 
How do gene work1
How do gene work1How do gene work1
How do gene work1
 
Transposons ask
Transposons askTransposons ask
Transposons ask
 
Microbial genetics lectures 10, 11, and 12
Microbial genetics lectures 10, 11, and 12 Microbial genetics lectures 10, 11, and 12
Microbial genetics lectures 10, 11, and 12
 
AP Biology Genetic Diversity and Operons in bacteria
AP Biology Genetic Diversity and Operons in bacteriaAP Biology Genetic Diversity and Operons in bacteria
AP Biology Genetic Diversity and Operons in bacteria
 
type of genes - jumping genes
type of genes - jumping genestype of genes - jumping genes
type of genes - jumping genes
 
TRANSPOSABLE ELEMENTS
TRANSPOSABLE   ELEMENTSTRANSPOSABLE   ELEMENTS
TRANSPOSABLE ELEMENTS
 
Transposons (2) (3)
Transposons (2) (3)Transposons (2) (3)
Transposons (2) (3)
 
Human multi gene families
Human multi gene familiesHuman multi gene families
Human multi gene families
 

En vedette

OPA, T Liu (Sept2014)
OPA, T Liu (Sept2014)OPA, T Liu (Sept2014)
OPA, T Liu (Sept2014)
Tiffany Liu
 
Liu.Tiffany.WineGrapes
Liu.Tiffany.WineGrapesLiu.Tiffany.WineGrapes
Liu.Tiffany.WineGrapes
Tiffany Liu
 
Getting Scientfic Results in the Lab
Getting Scientfic Results in the LabGetting Scientfic Results in the Lab
Getting Scientfic Results in the Lab
oliviagreen1
 
The Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
The Business Strategy that Makes Competition Irrelevant, Blue Ocean StrategyThe Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
The Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
Flevy.com Best Practices
 
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
Dpaps Saputra
 
Africa- Social Entrepreneuship and Technology
Africa- Social Entrepreneuship and TechnologyAfrica- Social Entrepreneuship and Technology
Africa- Social Entrepreneuship and Technology
Tiffany Liu
 

En vedette (14)

Tips for plumbing emergencies
Tips for plumbing emergenciesTips for plumbing emergencies
Tips for plumbing emergencies
 
All cirtificates مهندس استشارى حسن فرج ENG. HASSAN FARAG EL-SAYED PMP INSTRUC...
All cirtificates مهندس استشارى حسن فرج ENG. HASSAN FARAG EL-SAYED PMP INSTRUC...All cirtificates مهندس استشارى حسن فرج ENG. HASSAN FARAG EL-SAYED PMP INSTRUC...
All cirtificates مهندس استشارى حسن فرج ENG. HASSAN FARAG EL-SAYED PMP INSTRUC...
 
OPA, T Liu (Sept2014)
OPA, T Liu (Sept2014)OPA, T Liu (Sept2014)
OPA, T Liu (Sept2014)
 
Liu.Tiffany.WineGrapes
Liu.Tiffany.WineGrapesLiu.Tiffany.WineGrapes
Liu.Tiffany.WineGrapes
 
无锁编程
无锁编程无锁编程
无锁编程
 
Leipzig june-2017
Leipzig june-2017Leipzig june-2017
Leipzig june-2017
 
Culture shock DevOps meetup Wellington 27 Sep 2016
Culture shock DevOps meetup Wellington 27 Sep 2016Culture shock DevOps meetup Wellington 27 Sep 2016
Culture shock DevOps meetup Wellington 27 Sep 2016
 
Demystifying people aspect of DevOps Complexity Cube-Story Line
Demystifying people aspect of DevOps Complexity Cube-Story LineDemystifying people aspect of DevOps Complexity Cube-Story Line
Demystifying people aspect of DevOps Complexity Cube-Story Line
 
Getting Scientfic Results in the Lab
Getting Scientfic Results in the LabGetting Scientfic Results in the Lab
Getting Scientfic Results in the Lab
 
The Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
The Business Strategy that Makes Competition Irrelevant, Blue Ocean StrategyThe Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
The Business Strategy that Makes Competition Irrelevant, Blue Ocean Strategy
 
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
020 srt permoh_penyemprotan_nyamuk_ke_puskesmas_mei_2013_arsip
 
Day 1 cta dakar 0915 nepad_caadp_malabo_i_jallow
Day 1 cta dakar 0915 nepad_caadp_malabo_i_jallowDay 1 cta dakar 0915 nepad_caadp_malabo_i_jallow
Day 1 cta dakar 0915 nepad_caadp_malabo_i_jallow
 
Day 1 cta dakar 0915 mali_agriculture-nutrition-nexus_dagnoko
Day 1 cta dakar 0915 mali_agriculture-nutrition-nexus_dagnokoDay 1 cta dakar 0915 mali_agriculture-nutrition-nexus_dagnoko
Day 1 cta dakar 0915 mali_agriculture-nutrition-nexus_dagnoko
 
Africa- Social Entrepreneuship and Technology
Africa- Social Entrepreneuship and TechnologyAfrica- Social Entrepreneuship and Technology
Africa- Social Entrepreneuship and Technology
 

Similaire à Gutell 078.mbe.2001.18.1810

genomics final paper 3 after peer
genomics final paper 3 after peergenomics final paper 3 after peer
genomics final paper 3 after peer
Roshan Kumar
 
The Roles Of HMGA Proteins Essay
The Roles Of HMGA Proteins EssayThe Roles Of HMGA Proteins Essay
The Roles Of HMGA Proteins Essay
Jennifer Wood
 
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
Pat (JS) Heslop-Harrison
 
Gutell 055.rna.1996.02.0134
Gutell 055.rna.1996.02.0134Gutell 055.rna.1996.02.0134
Gutell 055.rna.1996.02.0134
Robin Gutell
 
DNA And RNA As The Basic Unit Of The Living System
DNA And RNA As The Basic Unit Of The Living SystemDNA And RNA As The Basic Unit Of The Living System
DNA And RNA As The Basic Unit Of The Living System
Alison Reed
 

Similaire à Gutell 078.mbe.2001.18.1810 (20)

Chloroplast dna
Chloroplast dnaChloroplast dna
Chloroplast dna
 
Arabidopsis Climate Change
Arabidopsis Climate ChangeArabidopsis Climate Change
Arabidopsis Climate Change
 
genomics final paper 3 after peer
genomics final paper 3 after peergenomics final paper 3 after peer
genomics final paper 3 after peer
 
The Roles Of HMGA Proteins Essay
The Roles Of HMGA Proteins EssayThe Roles Of HMGA Proteins Essay
The Roles Of HMGA Proteins Essay
 
assignment on inheritance and expressio of organeller dna 1
 assignment on inheritance and expressio of organeller dna 1 assignment on inheritance and expressio of organeller dna 1
assignment on inheritance and expressio of organeller dna 1
 
Molecular systematics.pdf
Molecular systematics.pdfMolecular systematics.pdf
Molecular systematics.pdf
 
Ayyappan et al., 2016
Ayyappan et al., 2016Ayyappan et al., 2016
Ayyappan et al., 2016
 
Identification and expression analysis of LEA gene family members in cucumber...
Identification and expression analysis of LEA gene family members in cucumber...Identification and expression analysis of LEA gene family members in cucumber...
Identification and expression analysis of LEA gene family members in cucumber...
 
Drosophila Melanogaster Genome And its developmental process
Drosophila Melanogaster  Genome And its developmental processDrosophila Melanogaster  Genome And its developmental process
Drosophila Melanogaster Genome And its developmental process
 
Taxonomy a synthetic science
Taxonomy a synthetic scienceTaxonomy a synthetic science
Taxonomy a synthetic science
 
3.wall.transcription
3.wall.transcription3.wall.transcription
3.wall.transcription
 
Sandlund
SandlundSandlund
Sandlund
 
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
Chromosomes and molecular cytogenetics of oil palm: impact for breeding and g...
 
Gutell 120.plos_one_2012_7_e38320_supplemental_data
Gutell 120.plos_one_2012_7_e38320_supplemental_dataGutell 120.plos_one_2012_7_e38320_supplemental_data
Gutell 120.plos_one_2012_7_e38320_supplemental_data
 
Gutell 055.rna.1996.02.0134
Gutell 055.rna.1996.02.0134Gutell 055.rna.1996.02.0134
Gutell 055.rna.1996.02.0134
 
DNA And RNA As The Basic Unit Of The Living System
DNA And RNA As The Basic Unit Of The Living SystemDNA And RNA As The Basic Unit Of The Living System
DNA And RNA As The Basic Unit Of The Living System
 
Gene expression in eukaryotes
Gene expression in eukaryotesGene expression in eukaryotes
Gene expression in eukaryotes
 
GJB6
GJB6GJB6
GJB6
 
Gutell 100.imb.2006.15.533
Gutell 100.imb.2006.15.533Gutell 100.imb.2006.15.533
Gutell 100.imb.2006.15.533
 
Essay On Arabidopsis Thaliaa
Essay On Arabidopsis ThaliaaEssay On Arabidopsis Thaliaa
Essay On Arabidopsis Thaliaa
 

Plus de Robin Gutell

Plus de Robin Gutell (20)

Gutell 124.rna 2013-woese-19-vii-xi
Gutell 124.rna 2013-woese-19-vii-xiGutell 124.rna 2013-woese-19-vii-xi
Gutell 124.rna 2013-woese-19-vii-xi
 
Gutell 123.app environ micro_2013_79_1803
Gutell 123.app environ micro_2013_79_1803Gutell 123.app environ micro_2013_79_1803
Gutell 123.app environ micro_2013_79_1803
 
Gutell 122.chapter comparative analy_russell_2013
Gutell 122.chapter comparative analy_russell_2013Gutell 122.chapter comparative analy_russell_2013
Gutell 122.chapter comparative analy_russell_2013
 
Gutell 121.bibm12 alignment 06392676
Gutell 121.bibm12 alignment 06392676Gutell 121.bibm12 alignment 06392676
Gutell 121.bibm12 alignment 06392676
 
Gutell 119.plos_one_2017_7_e39383
Gutell 119.plos_one_2017_7_e39383Gutell 119.plos_one_2017_7_e39383
Gutell 119.plos_one_2017_7_e39383
 
Gutell 118.plos_one_2012.7_e38203.supplementalfig
Gutell 118.plos_one_2012.7_e38203.supplementalfigGutell 118.plos_one_2012.7_e38203.supplementalfig
Gutell 118.plos_one_2012.7_e38203.supplementalfig
 
Gutell 114.jmb.2011.413.0473
Gutell 114.jmb.2011.413.0473Gutell 114.jmb.2011.413.0473
Gutell 114.jmb.2011.413.0473
 
Gutell 117.rcad_e_science_stockholm_pp15-22
Gutell 117.rcad_e_science_stockholm_pp15-22Gutell 117.rcad_e_science_stockholm_pp15-22
Gutell 117.rcad_e_science_stockholm_pp15-22
 
Gutell 116.rpass.bibm11.pp618-622.2011
Gutell 116.rpass.bibm11.pp618-622.2011Gutell 116.rpass.bibm11.pp618-622.2011
Gutell 116.rpass.bibm11.pp618-622.2011
 
Gutell 115.rna2dmap.bibm11.pp613-617.2011
Gutell 115.rna2dmap.bibm11.pp613-617.2011Gutell 115.rna2dmap.bibm11.pp613-617.2011
Gutell 115.rna2dmap.bibm11.pp613-617.2011
 
Gutell 113.ploso.2011.06.e18768
Gutell 113.ploso.2011.06.e18768Gutell 113.ploso.2011.06.e18768
Gutell 113.ploso.2011.06.e18768
 
Gutell 112.j.phys.chem.b.2010.114.13497
Gutell 112.j.phys.chem.b.2010.114.13497Gutell 112.j.phys.chem.b.2010.114.13497
Gutell 112.j.phys.chem.b.2010.114.13497
 
Gutell 111.bmc.genomics.2010.11.485
Gutell 111.bmc.genomics.2010.11.485Gutell 111.bmc.genomics.2010.11.485
Gutell 111.bmc.genomics.2010.11.485
 
Gutell 110.ant.v.leeuwenhoek.2010.98.195
Gutell 110.ant.v.leeuwenhoek.2010.98.195Gutell 110.ant.v.leeuwenhoek.2010.98.195
Gutell 110.ant.v.leeuwenhoek.2010.98.195
 
Gutell 109.ejp.2009.44.277
Gutell 109.ejp.2009.44.277Gutell 109.ejp.2009.44.277
Gutell 109.ejp.2009.44.277
 
Gutell 108.jmb.2009.391.769
Gutell 108.jmb.2009.391.769Gutell 108.jmb.2009.391.769
Gutell 108.jmb.2009.391.769
 
Gutell 107.ssdbm.2009.200
Gutell 107.ssdbm.2009.200Gutell 107.ssdbm.2009.200
Gutell 107.ssdbm.2009.200
 
Gutell 106.j.euk.microbio.2009.56.0142.2
Gutell 106.j.euk.microbio.2009.56.0142.2Gutell 106.j.euk.microbio.2009.56.0142.2
Gutell 106.j.euk.microbio.2009.56.0142.2
 
Gutell 105.zoologica.scripta.2009.38.0043
Gutell 105.zoologica.scripta.2009.38.0043Gutell 105.zoologica.scripta.2009.38.0043
Gutell 105.zoologica.scripta.2009.38.0043
 
Gutell 104.biology.direct.2008.03.016
Gutell 104.biology.direct.2008.03.016Gutell 104.biology.direct.2008.03.016
Gutell 104.biology.direct.2008.03.016
 

Dernier

Cloud Frontiers: A Deep Dive into Serverless Spatial Data and FME
Cloud Frontiers:  A Deep Dive into Serverless Spatial Data and FMECloud Frontiers:  A Deep Dive into Serverless Spatial Data and FME
Cloud Frontiers: A Deep Dive into Serverless Spatial Data and FME
Safe Software
 

Dernier (20)

MINDCTI Revenue Release Quarter One 2024
MINDCTI Revenue Release Quarter One 2024MINDCTI Revenue Release Quarter One 2024
MINDCTI Revenue Release Quarter One 2024
 
DEV meet-up UiPath Document Understanding May 7 2024 Amsterdam
DEV meet-up UiPath Document Understanding May 7 2024 AmsterdamDEV meet-up UiPath Document Understanding May 7 2024 Amsterdam
DEV meet-up UiPath Document Understanding May 7 2024 Amsterdam
 
Navigating the Deluge_ Dubai Floods and the Resilience of Dubai International...
Navigating the Deluge_ Dubai Floods and the Resilience of Dubai International...Navigating the Deluge_ Dubai Floods and the Resilience of Dubai International...
Navigating the Deluge_ Dubai Floods and the Resilience of Dubai International...
 
Mcleodganj Call Girls 🥰 8617370543 Service Offer VIP Hot Model
Mcleodganj Call Girls 🥰 8617370543 Service Offer VIP Hot ModelMcleodganj Call Girls 🥰 8617370543 Service Offer VIP Hot Model
Mcleodganj Call Girls 🥰 8617370543 Service Offer VIP Hot Model
 
Strategize a Smooth Tenant-to-tenant Migration and Copilot Takeoff
Strategize a Smooth Tenant-to-tenant Migration and Copilot TakeoffStrategize a Smooth Tenant-to-tenant Migration and Copilot Takeoff
Strategize a Smooth Tenant-to-tenant Migration and Copilot Takeoff
 
Repurposing LNG terminals for Hydrogen Ammonia: Feasibility and Cost Saving
Repurposing LNG terminals for Hydrogen Ammonia: Feasibility and Cost SavingRepurposing LNG terminals for Hydrogen Ammonia: Feasibility and Cost Saving
Repurposing LNG terminals for Hydrogen Ammonia: Feasibility and Cost Saving
 
How to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected WorkerHow to Troubleshoot Apps for the Modern Connected Worker
How to Troubleshoot Apps for the Modern Connected Worker
 
Artificial Intelligence Chap.5 : Uncertainty
Artificial Intelligence Chap.5 : UncertaintyArtificial Intelligence Chap.5 : Uncertainty
Artificial Intelligence Chap.5 : Uncertainty
 
Connector Corner: Accelerate revenue generation using UiPath API-centric busi...
Connector Corner: Accelerate revenue generation using UiPath API-centric busi...Connector Corner: Accelerate revenue generation using UiPath API-centric busi...
Connector Corner: Accelerate revenue generation using UiPath API-centric busi...
 
Apidays New York 2024 - The Good, the Bad and the Governed by David O'Neill, ...
Apidays New York 2024 - The Good, the Bad and the Governed by David O'Neill, ...Apidays New York 2024 - The Good, the Bad and the Governed by David O'Neill, ...
Apidays New York 2024 - The Good, the Bad and the Governed by David O'Neill, ...
 
Cloud Frontiers: A Deep Dive into Serverless Spatial Data and FME
Cloud Frontiers:  A Deep Dive into Serverless Spatial Data and FMECloud Frontiers:  A Deep Dive into Serverless Spatial Data and FME
Cloud Frontiers: A Deep Dive into Serverless Spatial Data and FME
 
Platformless Horizons for Digital Adaptability
Platformless Horizons for Digital AdaptabilityPlatformless Horizons for Digital Adaptability
Platformless Horizons for Digital Adaptability
 
Exploring Multimodal Embeddings with Milvus
Exploring Multimodal Embeddings with MilvusExploring Multimodal Embeddings with Milvus
Exploring Multimodal Embeddings with Milvus
 
DBX First Quarter 2024 Investor Presentation
DBX First Quarter 2024 Investor PresentationDBX First Quarter 2024 Investor Presentation
DBX First Quarter 2024 Investor Presentation
 
Apidays New York 2024 - Scaling API-first by Ian Reasor and Radu Cotescu, Adobe
Apidays New York 2024 - Scaling API-first by Ian Reasor and Radu Cotescu, AdobeApidays New York 2024 - Scaling API-first by Ian Reasor and Radu Cotescu, Adobe
Apidays New York 2024 - Scaling API-first by Ian Reasor and Radu Cotescu, Adobe
 
Strategies for Landing an Oracle DBA Job as a Fresher
Strategies for Landing an Oracle DBA Job as a FresherStrategies for Landing an Oracle DBA Job as a Fresher
Strategies for Landing an Oracle DBA Job as a Fresher
 
Apidays New York 2024 - Passkeys: Developing APIs to enable passwordless auth...
Apidays New York 2024 - Passkeys: Developing APIs to enable passwordless auth...Apidays New York 2024 - Passkeys: Developing APIs to enable passwordless auth...
Apidays New York 2024 - Passkeys: Developing APIs to enable passwordless auth...
 
EMPOWERMENT TECHNOLOGY GRADE 11 QUARTER 2 REVIEWER
EMPOWERMENT TECHNOLOGY GRADE 11 QUARTER 2 REVIEWEREMPOWERMENT TECHNOLOGY GRADE 11 QUARTER 2 REVIEWER
EMPOWERMENT TECHNOLOGY GRADE 11 QUARTER 2 REVIEWER
 
Polkadot JAM Slides - Token2049 - By Dr. Gavin Wood
Polkadot JAM Slides - Token2049 - By Dr. Gavin WoodPolkadot JAM Slides - Token2049 - By Dr. Gavin Wood
Polkadot JAM Slides - Token2049 - By Dr. Gavin Wood
 
"I see eyes in my soup": How Delivery Hero implemented the safety system for ...
"I see eyes in my soup": How Delivery Hero implemented the safety system for ..."I see eyes in my soup": How Delivery Hero implemented the safety system for ...
"I see eyes in my soup": How Delivery Hero implemented the safety system for ...
 

Gutell 078.mbe.2001.18.1810

  • 1. 1810 Mol. Biol. Evol. 18(9):1810–1822. 2001 ᭧ 2001 by the Society for Molecular Biology and Evolution. ISSN: 0737-4038 Accelerated Evolution of Functional Plastid rRNA and Elongation Factor Genes Due to Reduced Protein Synthetic Load After the Loss of Photosynthesis in the Chlorophyte Alga Polytoma Dawne Vernon,*1 Robin R. Gutell,† Jamie J. Cannone,† Robert W. Rumpf,*2 and C. William Birky Jr.*‡ *Department of Molecular Genetics, Ohio State University; †Institute of Cellular and Molecular Biology, University of Texas at Austin; and ‡Department of Ecology and Evolutionary Biology and Graduate Interdisciplinary Program in Genetics, University of Arizona Polytoma obtusum and Polytoma uvella are members of a clade of nonphotosynthetic chlorophyte algae closely related to Chlamydomonas humicola and other photosynthetic members of the Chlamydomonadaceae. Descended from a nonphotosynthetic mutant, these obligate heterotrophs retain a plastid (leucoplast) with a functional protein synthetic system, and a plastid genome (lpDNA) with functional genes encoding proteins required for transcription and translation. Comparative studies of the evolution of genes in chloroplasts and leucoplasts can identify modes of selection acting on the plastid genome. Two plastid genes—rrn16, encoding the plastid small-subunit rRNA, and tufA, encoding elongation factor Tu—retain their functions in protein synthesis after the loss of photosynthesis in two nonphotosynthetic Polytoma clades but show a substantially accelerated rate of base substitution in the P. uvella clade. The accelerated evolution of tufA is due, at least partly, to relaxed codon bias favoring codons that can be read without wobble, mainly in three amino acids. Selection for these codons may be relaxed because leucoplasts are required to synthesize fewer protein molecules per unit time than are chloroplasts (reduced protein synthetic load) and thus require a lower rate of synthesis of elongation factor Tu. Relaxed selection due to a lower protein synthetic load is also a plausible explanation for the accelerated rate of evolution of rrn16, but the available data are insufficient to test the hypothesis for this gene. The tufA and rrn16 genes in Polytoma oviforme, the sole member of a second nonphotosynthetic clade, are also functional but show no sign of relaxed selection. Introduction Nonphotosynthetic land plants and algae serve as a basis for interesting natural experiments on the evolution- ary consequence of the loss of a significant cell function. After losing the ability to do photosynthesis, nonphoto- synthetic species use various alternative carbon sources, with the plants becoming parasitic on other plants, while the algae take up complex organic molecules from their environment. Recognized by their lack of chlorophyll, these nongreen organisms have unique plastid (‘‘leuco- plast’’) genomes. The evolutionary consequences of the loss of photosynthesis can be studied by comparing the leucoplast genomes of nonphotosynthetic species with the chloroplast genomes of their closest photosynthetic rela- tives. The majority of the genes in the chloroplasts of pho- tosynthetic green algae and land plants encode proteins required for photosynthesis or gene expression (transcrip- tion and translation [Gillham 1994]; for recent data from complete chloroplast genome sequences, see Turmel, Otis, and Lemieux [1999], Lemieux, Otis, and Turmel [2000], and the NCBI chloroplast genome page at http://www. ncbi.nlm.nih.gov:80/PMGifs/Genomes/plastids࿞tax. html). These two functions account for 32 and 16 genes, 1 Present address: National Institute of Standards and Technology, DNA Technologies Group, Gaithersburg, Maryland. 2 Present address: LabBook.com, Inc., Columbus, Ohio. Key words: Polytoma, Chlamydomonas, chloroplast rRNA gene, chloroplast elongation factor gene, substitution rate, codon bias. Address for correspondence and reprints: C. William Birky Jr., Department of Ecology and Evolutionary Biology, Biological Sciences West, University of Arizona, Tucson, Arizona 85721. E-mail: birky@u.arizona.edu. respectively, in Chlamydomonas reinhardtii, which also encodes a minimal set of rRNA and tRNA genes (http: //www.biology.duke.edu/chlamy࿞genome/chloro.html). Genes encoding proteins with other functions, as well as unidentified open reading frames, are found in some taxa; C. reinhardtii has 15 of these. There is strong in- direct evidence that at least one of these genes, yet to be identified, codes for a protein that has an essential nonphotosynthetic (ENP) function (Gillham 1994, pp. 83–86). Investigations of leucoplast genome structure and gene sequences of nonphotosynthetic organisms have been limited to several parasitic angiosperm families (Scrophulareae and Orobanchaceae; e.g., dePamphilis and Palmer 1990; Wimpee, Morgan, and Wrobel 1992a, 1992b; Nickrent, Duff, and Konings 1997; Wolfe and dePamphilis 1998; Young and dePamphilis 2000) and the euglenoid alga Astasia (Siemeister, Buchholz, and Hachtel 1990; Siemeister and Hachtel 1990a, 1990b). The leucoplasts of these nonphotosynthetic species dif- fer from the chloroplasts of their close green relatives in numerous features. Morphologically, they are char- acterized by reduction or elimination of the thylakoid membranes. They all retain leucoplast ribosomes and leucoplast DNA. However, their leucoplast genomes are often reduced in size and complexity compared with chloroplast genomes in photosynthetic relatives. These leucoplast genomes allow investigation of the effects of selection on rates of evolution. When the ability to do photosynthesis is lost, the photosynthetic and photorespiration genes lose their function and con- sequently are no longer subject to selection; they are expected to become pseudogenes and can be lost entire- ly. In contrast, the leucoplast genes needed for transcrip-
  • 2. Accelerated Evolution of Leucoplast Genes 1811 FIG. 1.—Cladogram showing the relationships of the taxa used in this study based on Rrn18 sequences (Rumpf et al. 1996; unpublished data). tion and translation of the leucoplast genome probably remain functional and subject to at least some degree of selection because they are needed to transcribe and translate the ENP genes. The leucoplast genomes of the beech root parasite Epifagus virginiana (Orobancha- ceae), the oak parasite Conopholis americana (Oroban- chaceae), and the euglenoid Astasia longa display these characteristics. Most expression genes are still present in the leucoplast genomes and appear intact, and some leucoplast RNA and protein products have been dem- onstrated (Siemeister, Buchholz, and Hachtel 1990; Sie- meister and Hachtel 1990a, 1990b; Wimpee, Morgan, and Wrobel 1992; Wolfe, Morden, and Palmer 1992). Conversely, many leucoplast genes that coded for pho- tosynthetic proteins appear nonfunctional in key do- mains, are grossly truncated, or are absent from these leucoplast genomes. Although most expression genes are selectively retained, the leucoplast genome in Epi- fagus is missing more than a dozen expression genes (tRNA genes, ribosomal protein genes, and all four RNA polymerase subunit genes; Morden et al. 1991), all four RNA polymerase genes are pseudogenes in Lathraea (Lusson, Delavault, and Thalouarn 1998), and the leucoplast genome in Conopholis is apparently miss- ing several leucoplast tRNA genes (Wimpee, Morgan, and Wrobel 1992b). Presumably, these tRNAs and pro- teins are imported from the cytoplasm at a rate that may be too low for detection but is sufficient for protein syn- thesis in leucoplasts. The tempo of evolution has also changed in the leucoplast genomes of these nonphotosynthetic species. Most, but not all, of the apparently still-functional genes analyzed in leucoplasts show an increased rate of nu- cleotide substitution compared with rates in their green relatives. Most of the rate increases found in functional leucoplast genes in Astasia and Epifagus are in the range of 1.5-fold to 8-fold, while rrn16 genes in Epifagus and Conopholis show 40-fold rate increases (references in Results and Discussion). This suggests that selection on these genes has been relaxed, even though it has not been eliminated completely. To test the generality of these results, we extended the analysis of the sequence and evolution of leucoplast translation genes to the nonphotosynthetic chlorophyte algae, separated from the euglenoid plastid and from land plants by over 400 Myr of evolution and large dif- ferences in physiology and habitats. Phylogenetic anal- yses of Rrn18 sequences show that the members of the nonphotosynthetic genus Polytoma belong to two dif- ferent lineages within the clade that includes all Chla- mydomonas species as well as a number of other pho- tosynthetic genera and another nonphotosynthetic clade, Polytomella (Rumpf et al. 1996; unpublished data). Many species of Chlamydomonas are facultative auxo- trophs, capable of utilizing acetate as their sole carbon and energy source, and nonphotosynthetic mutants are readily isolated in C. reinhardtii (Harris 1989). It is as- sumed that Polytoma species arose as nonphotosynthetic mutants of facultative auxotrophs similar to the extant Chlamydomonas. The single large cup-shaped leucoplast in Polytoma does not have thylakoid membranes but still contains ribosomes, DNA (lpDNA), rRNA, and stored starch granules (Lang 1963; Scherbel, Behn, and Arnold 1974; Siu, Chiang, and Swift 1976; Vernon-Kipp, Kuhl, and Birky 1989). Polytoma is sensitive to inhibitors of chloroplast protein synthesis, which is additional evi- dence that the leucoplast is synthesizing at least one protein that is essential for auxotrophic growth and re- production (Scherbel, Behn, and Arnold 1974). Although these data strongly suggest that Polytoma retains a functional leucoplast expression system, the leucoplast genes involved have not been identified and demonstrated to be functional. We sequenced the rrn16 gene and the tufA gene (encoding the plastid elongation factor Tu) from two representatives of the Polytoma uvella clade (P. uvella 964 and Polytoma obtusum DH1), plus Polytoma oviforme, which is the sole mem- ber of the second Polytoma lineage. For comparison, these genes were also sequenced from two closely re- lated photosynthetic relatives (Chlamydomonas humi- cola SAG 11-9 and Chlamydomonas dysosmos UTEX 2399). The evolutionary relationships of these strains and some others involved in the analysis are shown in figure 1. This cladogram agrees with phylogenetic anal- yses of the rrn16 and tufA genes (figs. 2 and 3). Rela- tive-rate tests showed increased substitution rates in rrn16 and tufA, compared with green relatives, in the two P. uvella species. However, sequence analyses showed that the genes were subject to selection and therefore functional. The increase in substitution rate was greater at sites subject to less stringent selection, implicating a partial relaxation of selection. The tufA gene of P. obtusum showed a large reduction in codon preference, suggesting that the relaxed selection is due at least partly to a reduced load of protein synthesis. This was proposed earlier for the increased substitution rates in Epifagus (Wolfe et al. 1992), but alternative ex- planations were not ruled out. We observed no increase in the substitution rate in the branch leading to P. ovi- forme, suggesting that photosynthesis was lost more re- cently in this lineage. This is the first analysis of the molecular evolutionary consequences of the loss of pho- tosynthesis in a chlorophyte alga. Materials and Methods Organisms Polytoma uvella (UTEX 964) and P. oviforme (SAG 62-27) were obtained from the University of Tex-
  • 3. 1812 Vernon et al. FIG. 2.—Neighbor-joining tree of rrn16 sequences. Branch lengths in percentages of substitutions are shown above the branches; below the branches are the lengths in the most parsimonious tree, which had an identical topology. FIG. 3.—Neighbor-joining tree of tufA sequences. Branch lengths in percentages of substitutions are shown above the branches; below the branches are the lengths in the most parsimonious tree and the maximum-likelihood tree, which had identical topologies. as Culture Collection of Algae and from Sammlung von Algenkulturen Gottingen, respectively. P. obtusum (des- ignated strain DH1 by us) was obtained from David Her- rin at the University of Texas at Austin; it originally came from Luigi Provasoli’s collection at Yale. All cul- tures were subcloned once or twice and grown in Po- lytomella medium. Chlamydomonas humicola UTEX 225 and C. dy- sosmos UTEX 2399, from the University of Texas Cul- ture Collection of Algae, were combined under the spe- cies name Chlamydomonas applanata Pringsheim based on morphology and autolysin cross-reactions (Ettl 1976) and identity of the nuclear Rrn18 gene sequences (Gor- don et al. 1995). Consistent with this, we found no sub- stitution differences and only one insertion or deletion difference in their chloroplast rrn16 sequences, while the tufA sequences showed two synonymous differences and one nonsynonymous differences and no insertions or deletions. Consequently, we included only the C. humicola sequences in the analyses described here. DNA Preparation The rrn16 gene of P. uvella was cloned. Whole- cell DNA was isolated with a lysis method designed to yield high-molecular-weight chloroplast DNA, modified from Grant, Gillham, and Boynton (1980) as described in Vernon (1996). This DNA was fractionated in a CsClϩbisbenzimide equilibrium gradient. The top band in the gradient was identified as leucoplast DNA by Southern hybridization with an rrn16 probe and dot blot hybridization with a tufA probe. The C. reinhardtii cpDNA probes were provided by Elizabeth Harris (Duke University) and Jeffrey Palmer (Indiana University). The top band was used to prepare a HindIII library cloned in pBluescript. DNA obtained from these clones by al- kaline lysis plasmid minipreps (Sambrook, Fritsch, and Maniatis 1989) was electrophoresed, and Southern blots on GeneScreen Plus were hybridized with the cpDNA probes. A clone containing the rrn16 gene in a 6.2-kb insert was identified and purified for sequencing with a GeneClean kit (Bio 101). All other new sequences used in this study were of genes amplified from partially pu- rified whole-cell DNA isolated from CTAB lysates of 1L algal cultures. Polymerase Chain Reaction Amplifications Primers located near the ends of the rrn16 and tufA genes were used to obtain DNA templates for sequenc- ing. The 5Ј and 3Ј primers for rrn16 were A-17 (5Ј- GTTTGATCCTGGCTCAC-3Ј) and 5005-15 (3Ј-CA- TGTGTGGCGGGCA-5Ј). The 5Ј and 3Ј degenerate primers for all but one of the tufA genes were 1F (5Ј- GGDCAYGTTGAYCAYGG-3Ј) and 5R (3Ј-TGA- CANCCRCGRCCRCA-5Ј). Primer 5R did not amplify tufA from P. obtusum, so the 3Ј ends of the tufA genes from the other chlamydomonad species were inspected for conserved areas, and an alternative 3Ј primer (1130R: 3Ј-CCRATACGGDCCACTRGC-5Ј) was de- signed and used, located 100 bases farther 5Ј of the orig- inal 3Ј primer 5R. This amplified a tufA fragment from P. obtusum that was approximately 100 bases shorter than the other chlamydomonad sequences. The rrn16
  • 4. Accelerated Evolution of Leucoplast Genes 1813 amplification products sequenced were about 1.3 kb long, except for P. uvella, which was about 1.6 kb long; the tufA amplification products were about 1.1 kb long, except for P. obtusum, which was about 1.0 kb long. Optimal amplification conditions were determined for each gene empirically; multiple separate amplifications were performed and pooled, then purified using GeneClean. Sequencing Both strands of all genes were sequenced manually using a modified dsDNA Cycle Sequencing kit (Life Technologies). Most internal primers for sequencing were obtained from Paul Fuerst for the rrn16 gene and from Jeffrey Palmer for the tufA gene; additional inter- nal primers in conserved regions were designed to fill gaps in sequence coverage. Alignment of rrn16 Sequences The five new sequences for the study reported here (P. uvella, P. obtusum, P. oviforme, C. humicola, and C. dysosmos) were initially aligned using CLUSTAL W in SeqApp (Gilbert 1992) to match the rrn16 primary structure alignment in the Ribosomal Database Project (Maidak et al. 1994). The alignment was further refined by comparison with 70 publicly available plastid rrn16 sequences using a SUN Microsystems workstation with the alignment editor AE2 (developed by T. Macke, Scripps Research Institute, San Diego, Calif., and avail- able at http://www.cme.msu.edu/RDP/html/index.html). Sequences were initially aligned for maximum primary structure similarity; then, all positions associated with the comparatively inferred base pairs were checked to assure that these base-paired positions were properly aligned. The final alignment (with a complete list of species and numerous chloroplast and Polytoma SSU rRNA secondary-structure diagrams) is available in the supplement (on the MBE web site) as GenBank files (fig. 3c in the supplement); a subset of sequences used for phylogenetic analysis is shown in less detail in se- quential format (fig. 6 in the supplement) and in inter- leaved Pretty Print format (fig. 7 in the supplement). Alignment of tufA Sequences To assist alignment of tufA sequences, C. Delwiche and J. Palmer at Indiana University provided their align- ment, with 18 eubacterial, 8 cyanobacterial, 26 algal, and 4 land plant sequences (array described in Delwiche, Kuhsel, and Palmer 1995). The Polytoma and Chlamy- domonas sequences were aligned to various subsets of this array using DNA sequences but were influenced by the resulting amino acid alignment. One thousand fifty- three base pairs of the tufA gene were aligned (85% of the coding region), leaving out the first 72 5Ј positions and the last 96 3Ј positions for lack of data in some or all species. The complete alignment is available in the supplement (fig. 5); a subset of sequences is shown in less detail in sequential format (fig. 6 in the supplement) and in interleaved Pretty Print format (fig. 8 in the supplement). Phylogenetic Analyses Gene trees for were produced with PAUP* (Swof- ford 1998) and PHYLIP, version 3.56 (Felsenstein 1993). Sequence differences were corrected for multiple hits using the Jukes-Cantor one-parameter model (Jukes and Cantor 1969); otherwise, all analyses used default settings. Before a set of sequences was subjected to phy- logenetic analysis or relative-rate tests, sites that were missing in one or more species were removed from all sequences. Relative-Rate Tests Relative-rate tests (Sarich and Wilson 1973; Wu and Li 1985) were performed to detect differences be- tween rates of nucleotide (or amino acid) substitution in the three Polytoma species studied, compared with green species. Each relative-rate test involved a Polytoma iso- late (nongreen, N), its closest photosynthetic relative (green, G), and a photosynthetic outgroup species (O). The test parameters were KON and KOG, the estimated numbers of base substitutions per site occurring along the lineages leading from the outgroup to the nongreen Polytoma and to the green ingroup, respectively. The estimated numbers of substitutions per site (K) were ob- tained by correcting the observed sequence differences per site for multiple hits with the Jukes-Cantor model implemented in the MEGA sequence analysis package, PHYLIP, or PAUP*. Two other correction methods were used for comparison: the Kimura (1980) two-parameter method, which allows different rates of transition versus transversion, and the Tamura (1992) method, which uses information about GϩC content as well as separate tran- sition and transversion rates, again using MEGA. All three correction methods added approximately the same number of unobserved substitutions (data not shown), so the Jukes-Cantor method was used for the relative- rate test because it had the smallest variance. KON and KOG were related to the evolutionary rates EON and EOG along the nongreen and green lineages by KON ϭ EONT and KOG ϭ EOGT, where T is the time since divergence of the two lineages and was, of course, the same for both lineages. Any rate differences between green lin- eages and the nongreen lineages can be expressed as the difference between these two numbers of substitutions (KON Ϫ KOG ϭ [EON Ϫ EOG]T). The significance of rate differences was evaluated as in Muse and Weir (1992). A second method was also used to separate ob- served sequence differences into rates along different green or nongreen lineages, employing phylogenetic software. Gene trees in which the observed substitutions were apportioned to the various branches of the tree by phylogenetic algorithms provided the inferred substitu- tions on each green or nongreen branch. The appor- tioned substitutions from a nongreen Polytoma species and from its nearest green relative (ingroup) to their nearest ancestral node, KAN and KAG, respectively, were used to calculate the ratio KAN/KAG or the difference KAN
  • 5. 1814 Vernon et al. Table 1 Pairwise Numbers of Substitutions per Site Among rrn16 Genes Chlamydomonas reinhardtii Chlamydomonas moewusii Polytoma oviforme Chlamydomonas humicola Polytoma uvella Polytoma obtusum Chlorella . . . . . . . . . . . . . . . . . . . . . . Chlamydomonas reinhardtii. . . . . . . Chlamydomonas moewusii . . . . . . . . Polytoma oviforme . . . . . . . . . . . . . . Chlamydomonas humicola . . . . . . . . Polytoma uvella . . . . . . . . . . . . . . . . 0.162 0.185 0.138 0.171 0.123 0.106 0.170 0.119 0.125 0.095 0.227 0.212 0.213 0.194 0.152 0.220 0.193 0.200 0.166 0.142 0.054 NOTE.—Estimated number of substitutions per site (sequence divergence) ϭ sequence differences per site corrected for multiple hits by the Jukes-Cantor method. Table 2 Relative-Rate Tests on rrn16 Based on Neighbor-Joining, Maximum-Parsimony, and Maximum-Likelihood Tree Branch Lengths and on Pairwise Numbers of Substitutions NONGREEN SPECIES GREEN SPECIES KAN/KAG Neighbor Joining Maximum Parsimony Maximum Likelihood OUTGROUP SPECIES KON ϪKOG Polytoma uvella 964 Polytoma obtusum Polytoma oviforme Chlamydomonas humicola C. humicola Chlamydomonas moewusii 3.32 2.89 0.67 4.34 4.03 0.50 3.05 2.68 0.65 Chlamydomonas moewusii Chlamydomonas reinhardtii C. moewusii C. reinhardtii Chlamydomonas humicola C. reinhardtii 0.088* 0.093** 0.075** 0.074** Ϫ0.030 Ϫ0.015 * Significant at the 1% level. ** Significant at the 0.1% level. Ϫ KAG. The difference divided by KAG, i.e., (KAN Ϫ KAG)/KAG, can be used to compare the magnitudes of the rate increases along two different nonphotosynthetic lineages with different ingroups. Codon usage and the amount of codon usage bias in tufA were also investigated in P. obtusum, P. ovifor- me, C. humicola, and C. reinhardtii. All gaps were re- moved from the aligned sequences of these four species, leaving 349 codons. DNA Strider was used to calculate codon usage in these sequences. Relative synonymous codon usage (RSCU) was calculated for each codon us- ing MEGA. RSCU is the ratio of the observed frequency of a particular codon to the expected frequency of that codon calculated on the assumption that all codons are used equally frequently; an RSCU value significantly different from 1 is evidence of biased codon usage (Sharp and Li 1987). A Pascal program provided by Brian Morton was used to calculate the codon bias index (CBI) and the codon adaptation index (CAI). The CBI is an overall measure of codon bias for the entire gene (Morton 1993); the CBI ranges from 0 (no codon bias in the gene) to 1 (maximum codon bias). The CAI is measure of bias in the use of a codon relative to its use in a reference set of highly expressed genes (Sharp and Li 1987). Results and Discussion The Evolutionary Rate of rrn16 is Accelerated in P. uvella and P. obtusum but not in P. oviforme Figure 2 shows the neighbor-joining tree of rrn16 sequences; the topology of the most parsimonious tree from an exhaustive maximum parsimony search is iden- tical. This tree is compatible with the trees of the nuclear Rrn18 gene (fig. 1). Above each line in figure 2 is the length of the branch in the Neighbor-Joining tree in per- centage of substitutions; below each line is the length of the same branch in the parsimony tree. The tree shows a strong acceleration of substitution rate along the branches leading to the nonphotosynthetic P. uvella lineage. We used the distances on the tree in figure 2 to calculate the ratio of substitution rates on nonphotosyn- thetic and photosynthetic lineages, as well as the differ- ence between nonphotosynthetic and photosynthetic rates. We also used the method of Wu and Li (1985) for relative-rate tests based on corrected frequencies of pair- wise substitutions (table 1). Table 2 shows the results of these relative-rate calculations. The tests for P. uvella and P. obtusum used C. humicola as their closest relative and Chlamydomonas moewusii or C. reinhardtii as the outgroup; the test for Polytoma oviforme used C. moe- wusii as the closest green relative and C. humicola or C. reinhardtii as the outgroup. All relative-rates tests showed significantly increased substitution rates in the branch leading to P. obtusum versus the branch leading to the ingroup (C. humicola), and an even greater rate increase was seen in P. uvella. Figure 2 shows that the branch leading from the common ancestor of the P. uvella clade and C. humicola to the common ancestor of P. uvella and P. obtusum is longer, i.e., has more substitutions, than the branch leading to C. humicola. This shows that the acceleration began in the common ancestor of the P. uvella clade, as expected. As a con- trol, we performed a relative-rate test on the nuclear Rrn18 gene of P. uvella (not shown). The test showed a small increase in this nongreen species, but it was not
  • 6. Accelerated Evolution of Leucoplast Genes 1815 Table 3 Pairwise Numbers of Substitutions per Site Among tufA Genes ORGANISMS SEQUENCE DIVERGENCE All Sites 3rd Position 1st ϩ 2nd Positions Polytoma obtusum–Chlamydomonas reinhardtii. . . . Chlamydomonas humicola–C. reinhardtii . . . . . . . . . P. obtusum–C. humicola . . . . . . . . . . . . . . . . . . . . . . . Polytoma oviforme–C. humicola. . . . . . . . . . . . . . . . . P. oviforme–C. reinhardtii . . . . . . . . . . . . . . . . . . . . . 0.25334 0.13349 0.22619 0.15501 0.15988 0.74805 0.31078 0.62276 0.43202 0.40108 0.09112 0.05843 0.08612 0.04732 0.06324 NOTE.—Estimated number of substitutions per site (sequence divergence) ϭ sequence differences per site corrected for multiple hits by the Jukes-Cantor method. Table 4 Relative-Rate Tests on tufA Based on Tree Branch Lengths and Pairwise Distance Matrices NONGREEN SPECIES GREEN SPECIES KAN/KAG Parsimony Neighbor Joining Maximum Likelihood OUTGROUP KON Ϫ KOG All Sites 3rd Position 1st ϩ 2nd Positions Polytoma obtusum Polytoma oviforme Chlamydomonas humicola Chlamydomonas reinhardtii 3.23 1.26 2.7 1.18 4.39 1.34 C. reinhardtii C. humicola 0.1998** 0.0215 0.4408* 0.1167 0.0278* Ϫ0.00951 * Significant at the 1% level. ** Significant at the 0.1% level. statistically significant. No rate increase was seen in the branch leading to P. oviforme. The Evolutionary Rate of tufA is Accelerated in P. obtusum but not in P. oviforme Sequences of tufA are available for C. humicola, P. obtusum, C. reinhardtii, P. oviforme, and a number of green algae outside of the Chlamydomonadaceae. Of these, Codium is the closest relative, but when we used it as an outgroup, all three tree-making algorithms grouped Codium with P. obtusum, presumably due to long-branch attraction. We therefore used only the four Chlamydomonadaceae, with C. reinhardtii serving as the outgroup for C. humicola and P. obtusum, and C. humicola serving as the outgroup for C. reinhardtii and P. oviforme. The tufA sequences of these four species have 1,014 sites in common. The topology of the neigh- bor-joining tree of these genes (fig. 2) is consistent with the Rrn18 tree (fig. 1). However, long-branch attraction was still a problem with the parsimony and maximum- likelihood algorithms, which favored the tree that placed P. obtusum with C. reinhardtii. The correct parsimony tree (the one with the same topology as the Neighbor- Joining tree and all trees involving rrn16 or Rrn18) was the least parsimonious and had the lowest likelihood scores, although not by much. The branch lengths for the correct trees from all three algorithms are shown in figure 2. In every case, the branch leading to P. obtusum is much longer than that leading to C. humicola, while P. oviforme shows no acceleration. Table 3 shows the estimated pairwise numbers of substitutions among these species; the branch leading to P. obtusum is ac- celerated in all three trees (parsimony, Neighbor-Join- ing, and maximum likelihood). We performed relative-rate tests of the evolution of tufA in P. obtusum and P. oviforme, using the pairwise differences with the Jukes-Cantor correction for multiple hits (table 3). In addition to calculating relative rates of nucleotide substitutions for all aligned sites, we com- pared first ϩ second codon positions with third codon positions. The results are shown in table 4; all tests showed a significantly higher substitution rate in the branch leading to the nonphotosynthetic P. obtusum than in the branch leading to the photosynthetic ingroup, C. humicola. The increase was greater in the third codon positions than in the first and second positions. No sig- nificant difference was found between the branches lead- ing to P. oviforme and C. reinhardtii, in agreement with the data from rrn16. The Plastid tufA Genes in Polytoma Remain Functional After the Loss of Photosynthesis One possible explanation for the accelerated evo- lution of rrn16 and tufA is that the genes became non- functional in the nonphotosynthetic lineages. This is un- likely, given the evidence that they remain functional in nonphotosynthetic land plants. We found additional ev- idence that tufA remained subject to selection, and hence functional, in the Polytoma lineages: 1. There are no premature stop codons in the entire gene. This could be because no stop mutations oc- curred since the loss of photosynthesis or because they were eliminated by selection. For the P. uvella clade, we estimated the probability of no stop mu- tations occurring as follows: First, we assumed that a truncation would not inactivate the protein if it oc- curred between the carboxyl terminal of the protein and the 14th amino acid, since the first 13 amino acids are not involved in intermolecular bonding in Escherichia coli (Kawashima et al. 1996). In the re- mainder of the protein, we found 108 codons that were one substitution away from being stop codons. As described above, we know that more synonymous
  • 7. 1816 Vernon et al. substitutions occurred along the branch leading to P. obtusum than along the branch leading to C. humi- cola from their common ancestor. There were 0.2235 extra substitutions per site in third codon positions, which must have occurred after photosynthesis was lost; this is also an estimate of the number of muta- tions per site. We found 81 sense codons in the tufA gene of P. obtusum that could have become a stop codon as a result of one kind of substitution (e.g., UCG to UAG); the expected number of such substi- tutions in the absence of selection was 81 ϫ 0.2235 ϫ 1/3 ϭ 6.034. We found 27 sense codons that could have become stop codons as a result of either of two kinds of substitutions (e.g., UAC to UAA or UAG); the expected number of such substitutions was 27 ϫ 0.2235 ϫ 2/3 ϭ 4.023. Consequently, the expected number of premature stop codons in the absence of selection was 10.057, and from the Poisson distri- bution the probability of finding no premature stop codons was eϪ10.057 ϭ 4.3 ϫ 10Ϫ5. We conclude that the tufA gene of P. obtusum must have been under selection that eliminated genes with premature stop codons most or all of the time since the loss of photosynthesis. Additional evidence was obtained using a tufA se- quence obtained from P. uvella by Nedelcu (2001) using the UTEX stock without subcloning. We aligned 999 bp, or 333 complete codons, of P. uvella and P. obtusum. The sequences of these species dif- fered by 0.07892 synonymous substitutions per site, all of which must have occurred since they diverged from a common ancestor, after photosynthesis was lost. We used parsimony to reconstruct 321 codons of the sequence of their most recent common ances- tor. This sequence contained 72 codons which could have become stop codons if they had incurred single specific mutations. The expected number of such substitutions in the absence of selection was 72 ϫ 0.07892 ϫ 1/3 ϭ 1.894. The ancestral sequence also contained 28 codons that could have become stop codons as a result of either of two kinds of substi- tutions; the expected number of such substitutions was 28 ϫ 0.07892 ϫ 2/3 ϭ 1.473. Consequently, the expected number of premature stop codons in the ab- sence of selection is 3.367, and from the Poisson dis- tribution the probability of finding no premature stop codons was eϪ3.367 ϭ 0.0345. 2. All of the amino acid substitutions that occurred (out- side of the hypervariable region discussed below) in P. oviforme must be compatible with the normal function of tufA, because each of the substituted ami- no acids can be found at a comparable position in at least one functional algal, cyanobacterial, or nonpho- tosynthetic bacterial gene in the alignment array of Delwiche, Kuhsel, and Palmer (1995). The same is true of all but six amino acid substitutions seen in P. obtusum. Both Polytoma sequences contain only con- servative amino acid substitutions, except for some nonconservative substitutions on the surface of the EF-Tu protein of P. obtusum. None of these amino acid substitutions are likely to change the folding of the EF-Tu protein. 3. Nucleotide substitution rates in the tufA sequences at first and second codon positions are much lower than the rates at third positions (table 3), a difference that can only be due to selection. 4. Relative to all other species, the tufA sequences from Polytoma and Chlamydomonas contain numerous amino acid substitutions, insertions, and deletions in the hypervariable region. Despite the variability in this region, we believe that it is compatible with functionality of the protein in both Polytoma species for the following reasons. First, the hypervariable re- gion of P. oviforme is identical in length to that of C. reinhardtii, and nearly identical in sequence, with only two conserved amino acid differences between the two species. The hypervariable region in P. ob- tusum differs from that in C. reinhardtii in 13 sub- stitutions and 3 gaps. However, this region is also hypervariable and unusually long in functional EF- Tu proteins from C. reinhardtii (Baldauf and Palmer 1990), C. humicola, and C. dysosmos, which are pho- tosynthetic and therefore have functional EF-Tu pro- teins. Second, the hypervariable region is on the out- side surface of the protein in functional domain 3, where amino acid changes or extra amino acids would probably not affect the conformational chang- es that occur during catalysis (Berthtold et al. 1993), especially since the amino acid composition of the hypervariable region is even more hydrophilic in P. obtusum than in C. humicola and C. reinhardtii. Moreover, the face of domain 3 that interacts with other EF-Tu molecules (Kawashima et al. 1996) and with the acceptor stem or T stem of the tRNA (Nis- sen et al. 1995) is opposite the hypervariable region. The Plastid rrn16 Genes in Polytoma Remain Functional After the Loss of Photosynthesis The growth of P. uvella and of another member of the same clade, P. uvella 62-3 ϭ P. mirum, is inhibited by the antibiotics erythromycin, streptomycin, and spec- tinomycin at 800, 400, and 50 mg/ml, respectively (data not shown). Scherbel, Behn, and Arnold (1974) previ- ously found that growth of P. mirum is inhibited by streptomycin. This antibiotic is known to inhibit chlo- roplast protein synthesis in C. reinhardtii at similar or lower (erythromycin) concentrations (Harris 1989). Spectinomycin sensitivity is especially interesting: it has no known side effects, and Chlamydomonas mutants re- sistant to high concentrations have mutations only in the rrn16 gene. These data suggest that Polytoma, like Chlamydomonas, synthesizes at least one essential pro- tein on plastid ribosomes which contain functional 16S rRNA molecules. Consistent with the antibiotic studies, an analysis of the primary and secondary structures of the 16S rRNA molecules, inferred from the rn16 sequences, strongly sup- ports functionality of the molecules. Here we present only a summary; the complete analysis is included with the secondary-structure figures and sequence alignments in the
  • 8. Accelerated Evolution of Leucoplast Genes 1817 FIG. 4.—Relative synonymous codon usage values of codons in Polytoma obtusum (open bars) compared with Chlamydomonas hum- icola (filled bars). Separate graphs are shown for amino acids that are sixfold-, fourfold-, and twofold-degenerate. Stars represent codons for which the complementary anticodons are found in tRNAs encoded in plant chloroplast genomes. supplement and at the web site http://www.rna.icmb. utexas.edu/PUBLICATIONS/BIRKY/. 1. The primary and secondary structures for the three Polytoma rRNA sequences contain all of the struc- tural elements present in the chloroplast rRNAs that are functional, and, apart from a few insertions dis- cussed below, all of the nucleotide positions in the Polytoma sequences correspond to all of the positions present in these 70 functional 16S rRNA chloroplast sequences (figs. 1, 2, 3a, and 3d–f in the supplement and at our web site mentioned above; see also Gutell 1994). 2. Differences between the Polytoma sequences occur at positions that also vary in the functional SSU rRNAs in the nuclear genes of the Eucarya, Bacteria, and Archaea and in chloroplast and mitochondrial genes (http://www.rna.icmb.utexas.edu/RDBMS/, http://www.rna.icmb.utexas.edu/CSI/BPFREQ/ 16S-MODEL-BP/, and figs. 2, 3a, and 3c in the sup- plement and at our web site mentioned above). Con- versely, positions that are conserved in the three phy- logenetic domains plus chloroplasts and mitochon- dria are also conserved in the Polytoma sequences. 3. Base pairs were predicted with comparative sequence analysis (Gutell 1996; http://www.rna.icmb.utexas. edu/METHODS/); two aligned positions that change coordinately are considered possible base pairs. These base pairs are highly conserved in the Poly- toma SSU rRNA and are consistent with function- ality. There are 70 base-paired positions at which the Polytoma sequences differ from the chloroplast con- sensus: of these, 24 are compensatory changes; 10 involve an A·U or G·C interchange to a G·U base pair; and only 12 change an A·U, G·C, or G·U pair to a noncanonical pair (figs. 2 and 3 at our web site mentioned above). The number of noncanonical base pairs in the Polytoma SSU rRNA is approximately the same as in other rRNAs that are known to be active in protein synthesis (unpublished data). 4. A comparison of the sequences of P. obtusum and P. uvella reveals functional and structural constraints acting on these 16S rRNA sequences since their com- mon ancestor; because this ancestor was nonphoto- synthetic, all of the differences between the two Po- lytoma isolates arose in nonphotosynthetic lineages. Among the 55 variable paired positions, 30 are in- volved in a covariation at base pairs predicted with comparative sequence analysis (fig. 4 in the supple- ment and at our web site mentioned above). Of these positions, 12 involve an exchange between Watson- Crick base pairs, while three involve a Watson-Crick base pair with something else. Of the 25 paired po- sitions with a single change, 14 exchanged between AU or GC and GU. Since our covariation algorithms search for simultaneous changes regardless of the base pair types, the fact that AU and GC are the most frequent pairs identified at positions that covary in- dicates that these genes are still undergoing positive selection. In addition, several of the noncanonical co- variation base pairs have been experimentally veri- fied (reviewed in Gutell 1996), suggesting that all of these covarying bases could be base-paired in the secondary structure. 5. Substitution rates were higher at paired sites than at unpaired sites, presumably because single substitu- tions at paired sites are detrimental but pairs of com- pensating changes are neutral (data not shown). There are three large uncharacteristic insertions in the Polytoma sequences. The P. oviforme gene has an incompletely sequenced insertion larger than 438 bp be- tween positions 427 and 428 (E. coli numbering). Po- lytoma obtusum and P. uvella have insertions of 223 bp and Ͼ596 bp, respectively, between positions 746 and 747. Finally, P. uvella has a second insertion of 85 nt between (E. coli) positions 138 and 140. All of these insertions occur in regions of the 16S rRNA that are variable in sequence and not considered to be function- ally important. The first, between positions 138 and 140 (E. coli numbering) contains 85 extra nucleotides in P.
  • 9. 1818 Vernon et al. uvella; 75 of these are A or U. We believe that this insertion may be an internal transcribed spacer that is excised after transcription for two reasons. First, this extra sequence has a very high AU content, a charac- teristic of internal transcribed spacer (ITS) sequences. Second, an ITS has been found in the bacterium Cam- pylobacter sputorum 16S rRNA at position 220 (E. coli numbering; VanCamp et al. 1993; accession number X67775), which is directly across the helix from our presumed ITS. The two other insertions occur at the ends of helices and may be either introns or ITSs. Relaxed Selection Is at Least Partly Responsible for the Accelerated Evolution The preceding sections show that the rrn16 and tufA genes in Polytoma remained subject to selection, and hence functional, long after the loss of photosyn- thesis. Although we cannot exclude the possibility that these genes lost their function recently, they were prob- ably still functional when we tested their antibiotic sen- sitivity. Consequently the increased rate of base pair substitution in the tufA and rrn16 genes in the P. uvella lineage is not due to complete loss of function. We therefore considered four other possible explanations for their accelerated evolution: 1. An increased mutation rate would increase the rate of substitution, provided, of course, that the fixation probability was not reduced by the same factor. How- ever, an increase in mutation rate cannot explain our data because it would affect different functional re- gions of a gene to the same extent; in contrast, we found that the substitution rate in the rrn16 gene was greater in regions coding for the stems of the rRNA than in regions coding for the loops, and in tufA there was a much greater increase in the synonymous rate than in the nonsynonymous rate. These observations show that an increased mutation rate is not the sole cause of the accelerated evolution in rrn16 and tufA but do not rule it out as a contributing factor. 2. There could be an increased level of adaptive or pos- itive selection by fixation of mutations that were for- merly neutral or detrimental but are now advanta- geous. This explanation is unlikely because of the large number of such mutations that would be re- quired to explain the data. For example, parsimony analysis (data not shown) suggests that about 152 base pair substitutions occurred in the rrn16 gene along the lineage leading to P. uvella from the most recent common ancestor of P. uvella and C. humi- cola, while only 35 occurred in the C. humicola lin- eage. The nonphotosynthetic lineage thus incurred 152 Ϫ 35 ϭ 117 more substitutions that would have to be adaptive; this is 77% of all of the substitutions after photosynthesis was lost. Moreover, it is proba- bly not a complete explanation because it is not likely to increase the synonymous substitution rate. How- ever, it is possible that positive selection contributed in part to the increase in substitution rate. 3. A reduced effective population size (Ne) would make natural selection against detrimental mutations less effective. By making some mutations that were for- merly detrimental effectively neutral, it would in- crease the overall fixation probability of new muta- tions. A decrease in the frequency of recombination of plastid genes would reduce Ne as a result of the Hill-Robertson effect (Birky and Walsh 1988), but there is no reason to suspect this in nonphotosyn- thetic lineages. A more plausible reason for a de- crease in Ne is that an obligate heterotroph has fewer niches available to it, which might reduce Ne by de- creasing the total population size or increasing the variance in offspring number. In any event, a reduced Ne cannot explain our data by itself, because the ratio KAN/KAG was greater for kinds of substitutions sub- ject to weak selection (third positions in tufA; stems in rrn16) than for those subject to relatively strong selection (first and second positions in tufA; loops in rrn16). A theoretical analysis of the combined effects of directional selection and drift, using Kimura’s (1957) equation for the fixation probability of a mu- tation, showed that a decrease in effective population size by itself would cause KAN/KAG to be larger, not smaller, for strongly selected detrimental mutations than for relatively weakly selected mutations (unpub- lished data). In addition, a reduced Ne would affect nuclear genes as well as organelle genes, whereas we observed no significant increase in substitution rate in the nuclear Rrn18 genes of P. obtusum and P. uvella. We cannot rule out the possibility that some of the observed acceleration in evolutionary rate is due to a reduction in effective population size too small to be detected in the nuclear gene, but this cannot be a complete explanation. 4. Relaxed selection against detrimental mutations is a plausible explanation for the increase in substitution rate in nonphotosynthetic lineages. This would re- quire that some mutations that were detrimental in a chloroplast are neutral or effectively neutral in a leu- coplast and thus more likely to be fixed. The rrn16 and tufA genes are involved in translation of plastid proteins, and mutations in these genes are likely to be detrimental if they directly or indirectly reduce the rate or accuracy of protein synthesis. Selection would be relaxed if heterotrophic algae could tolerate either a lower rate of plastid protein synthesis or a higher proportion of amino acid substitutions in plastid pro- teins due to mistranslation. At least the first of these conditions is very likely to be met. Most of the chlo- roplast protein-coding genes code for proteins in- volved in photosynthesis, and all mutations affecting these genes are selectively neutral after photosynthe- sis is lost in the ancestor of a nonphotosynthetic clade; as a result, they will accumulate mutations that block transcription or translation or will be deleted entirely. Leucoplasts thus have to translate only about two-thirds as many proteins. Among the genes that do not have to be expressed in heterotrophic algae or plants is rbcL, which encodes one of the most abun- dant proteins in photosynthetic organisms, ribulose bisphosphate carboxylase. A number of photosyn-
  • 10. Accelerated Evolution of Leucoplast Genes 1819 Table 5 Codon Adaptation Index (CAI) and Codon Bias Index (CBI) for tufA in Chlamydomonas and Polytoma Species Organism CAI(Cre) CAI(Cum) CBI Chlamydomonas reinhardtii. . . . Polytoma oviforme . . . . . . . . . . . Chlamydomonas humicola . . . . . Polytoma obtusum. . . . . . . . . . . . 0.487 0.359 0.432 0.22 0.77 0.688 0.687 0.37 0.775 0.774 0.722 0.455 NOTE.—CAI(Cre) is calculated relative to the codon usage in C. reinhardtii psbA gene; CAI(Cum) is calculated relative to the codon usage in the psbA genes of C. reinhardtii, Porphyra purpurea, Odontella sinensis, and Cyanophora par- adoxa. thetic genes show high codon bias characteristic of highly expressed genes. We hypothesized that the loss of photosynthesis in Polytoma permitted a lower rate of synthesis of elon- gation factor Tu because these heterotrophic organisms can tolerate a lower overall rate of protein synthesis than their photosynthetic relatives. This hypothesis predicts that the tufA gene will show less codon bias in P. ob- tusum than in C. humicola, and probably less codon bias than in other Chlamydomonas species. Codon bias is generally greater in highly expressed genes, probably because selection favors codons that are read by more abundant tRNAs (e.g., Morton 1993, 1996; Sharp et al. 1995). We calculated two measures of codon bias for the tufA genes of the Chlamydomonadaceae. The CBI measures the overall level of codon bias in a gene and ranges from 0 to 1; the CAI (Sharp and Li 1987) mea- sures codon bias for a gene relative to one or more high- ly expressed genes that show high codon bias. Two ref- erence databases of codon use in a highly expressed plastid gene were provided by Brian Morton: the psbA gene of C. reinhardtii and the combined psbA genes of C. reinhardtii, Porphyra purpurea, Cyanophora para- doxa, and Odontella sinensis. The CBI and both CAI values were lower for P. obtusum than for the two Chla- mydomonas species and P. oviforme (table 5). These data show that codon bias in tufA is reduced in P. obtusum compared with its closest photosynthetic relative, C. humicola. The change in codon bias might be due to a change in relative mutation rates, e.g., to favor AT pairs over GC. To test this possibility, we cal- culated expected numbers of the four codons in all four- fold-degenerate amino acids and in the fourfold-degen- erate codon set of sixfold-degenerate amino acids, as- suming that the expected proportions of A, T, G, and C were the averages of those found in the third positions of all codons. The deviation of observed numbers from these expected numbers can be summarized by the chi- square value, which was much greater for C. humicola (64.3) than for P. obtusum (46.2). We also found that the percentages of AϩT were not significantly different in the first ϩ second codon positions in C. humicola and P. obtusum (53.0% vs. 52.6% respectively; P k 0.05 by Fisher’s exact test) but were significantly different in the third codon positions (60.5% vs. 75.9%, respective- ly; P K 0.001). Since mutation rates do not differ sys- tematically between codon positions, we conclude that the third-position base composition difference between Chlamydomonas and Polytoma is due to selection on synonymous codons rather than to mutation pressure. The reduced codon bias in Polytoma is presumably due to relaxed selection against mutations that substitute less-used codons for those that are most used in the C. humicola lineage. A more detailed comparison of codon bias in P. obtusum and C. humicola suggests that the bias that is relaxed in the nonphotosynthetic species is largely a preference for codons that can be read without wobble (a mismatched base pair between tRNA and the mRNA third codon position). Several plant genomes have been completely sequenced and found to have a set of 29 tRNAs. Gene sequences of 15 of these have been mapped on the chloroplast genome of C. humicola (E. Boudreau, personal communication), and no tRNA genes have been identified in any green alga other than those known from plant chloroplasts, so it is likely that these algae have the same set of chloroplast tRNAs as do plants. Many codons require wobble to be read by this set of tRNAs, and several require ‘‘superwobble,’’ in which one tRNA reads A, G, C, or U in the third codon position. We calculated the RSCU for all degen- erate amino acids (those using more than one codon) in P. obtusum and C. humicola. Figure 4 shows that much of the difference in codon bias is in the sixfold-degen- erate amino acids leucine, serine, and arginine, which preferentially use codons that can be read without wob- ble. This preference is much greater for C. humicola than for P. obtusum (chi-square test; 0.01 Ͼ P Ͼ 0.001). Similar analyses for fourfold- and twofold-degenerate amino acids show no significant difference between these species; both species show preference for no wob- ble over wobble in the twofold-degenerate amino acids (P ഠ 0.9) and for no wobble or normal wobble over superwobble in those fourfold-degenerate amino acids that require superwobble (0.8 Ͼ P Ͼ 0.7). In the case of rrn16, one can imagine that the re- duced load of protein synthesis in the leucoplast has resulted in less intense selection for translation rate or fidelity. Alternatively, there may be a greater tolerance for mutations that affect the rate of transcription of the gene, of posttranscriptional processing, or of assembly of the ribosome. Relaxed selection for translational rate or fidelity might result in decreased stability of the sec- ondary structure of the small-subunit rRNA. Compared with C. humicola, both P. uvella and P. obtusum had fewer GC pairs and more AU pairs, with GC/AU ratios of 1.80, 1.40, and 1.34, respectively. These numbers are compatible with the hypothesis that the loss of photo- synthesis has been accompanied by a relaxation of se- lection for helix stability in the leucoplast small-subunit rRNA molecule. However, it is also compatible with a decrease in the GϩC content of the genome as a whole. A more definitive test of the hypothesis will require comparisons of secondary-structure stability in a larger sample of nonphotosynthetic species and their close photosynthetic relatives, using the nuclear small-subunit rRNA molecule as a control, together with data on the GϩC contents of the genomes.
  • 11. 1820 Vernon et al. The bulk of the evidence suggests that the in- creased rate of base pair substitution in the tufA and rrn16 genes of Polytoma is due at least in part to relaxed selection, such that mutations that reduce the rate or fidelity of translation are less likely to be eliminated. Rigorous tests of this hypothesis will require more de- tailed analyses of a larger number of sequences of these and other plastid expression genes. Polytoma oviforme May Have Lost Photosynthesis More Recently The tufA gene of P. oviforme shows codon bias similar to that of several close green relatives, and nei- ther the rrn16 gene nor the tufA gene shows an increased substitution rate in the lineage leading to P. oviforme relative to green lineages. In contrast, the expression gene sequences that we sampled from the P. uvella clade all show significant rate increases compared with green relatives, and tufA from P. obtusum (the only protein- coding gene we obtained from the P. uvella clade) shows minimal codon bias. These data are consistent with the hypothesis that the P. uvella clade may have resulted from a more ancient loss of photosynthesis than the P. oviforme lineage, and only the P. uvella clade has been evolving without photosynthesis long enough for the consequences of relaxed selection to be evident. Accelerated Evolution of Leucoplast Expression Genes in Other Organisms Smaller increases in evolutionary rates have been demonstrated in the leucoplast expression genes rrn16, rrn23, and tufA, as well as in the rbcL gene, of the heterotrophic euglenoid alga Astasia longa (Siemeister, Buchholz, and Hachtel 1990; Siemeister and Hachtel 1990a, 1990b). An increase in the substitution rate of the rrn16 genes of some nonphotosynthetic angiosperms was reported by Nickrent, Duff, and Konings (1997), although it was not verified by relative-rate tests. The increase was accompanied by an increase in AϩT con- tent in the genes and consequently by an increase in destabilizing AU pairs in the rRNA, such as we ob- served. The authors did not compare the rates in stems and loops. Nickrent and Starr (1994) also reported an increased substitution rate in the nuclear Rrn18 genes of a different set of species of holoparasitic angiosperms and presented evidence that the increase could not be explained by a decrease in generation time or effective population size. In contrast, we observed no increase in the evolutionary rate of Rrn18 in Polytoma (Rumpf et al. 1996) or Polytomella (unpublished data). The holo- parasitic angiosperm Epifagus virginiana was subjected to an intensive analysis of evolutionary rates in leuco- plast genes; rate increases were detected in rrn16 and rrn23 and the pooled data for 17 tRNAs and 15 ribo- somal proteins (Wolfe et al. 1992). In the ribosomal pro- teins, accelerated evolution was seen in both nonsynon- ymous and synonymous substitutions; in contrast to our results, the increase was greater for nonsynonymous substitutions. Epifagus also differed from Polytoma in showing no change in codon bias (Morden et al. 1991). Wolfe et al. (1992) proposed that the increase in non- synonymous substitutions was probably due to relaxed selection for both the rate and fidelity of protein syn- thesis, while the increase in synonymous substitutions probably resulted from an increased mutation rate. How- ever, alternative explanations were not ruled out. In par- ticular, both the synonymous and the nonsynonymous rate increase might be entirely due to a decreased effec- tive population size, which under some circumstances can increase the rate of detrimental substitutions more when they are under stronger selection (unpublished data). dePamphilis, Young, and Wolfe (1997) found that the substitution rate of the ribosomal protein gene rps2 was significantly accelerated in some, but not all, hol- oparasitic (nonphotosynthetic) angiosperm lineages in the Orobanchaceae and Sdrophulariaceae. In some cas- es, there was a significant increase in synonymous but not nonsynonymous substitutions, and in some other cases, the reverse held. dePamphilis, Young, and Wolfe (1997) proposed that increases in nonsynonymous sub- stitution rates are probably due to reduced functional constraint when abundant photosynthetic proteins do not have to be synthesized. The increases in synonymous substitutions were attributed to increased mutation rates. However, no formal analyses of the data were given in support of these conclusions. The independent losses of photosynthesis in a num- ber of different plants and algae provide biologists with a series of natural experiments in which selection has been reduced or eliminated to different extents in genes with different functions. Future comparative studies of leucoplast expression and photosynthetic gene sequenc- es from Polytoma and Polytomella will help to unravel the roles of mutation and selection in the molecular evo- lution of plastid genes. The data to date suggest that accelerated evolution of plastid expression genes is a general consequence of the loss of photosynthesis and that the rate of evolution of expression genes is strongly influenced by the rate of protein synthesis that they must sustain. Supplementary Material The DNA sequences have been deposited in GenBank under accession numbers AF352839 (P. ob- tusum tufA), AF352840 (P. oviforme tufA), AF352838 (C. humicola tufA), AF397587 (P. obtusum rrn16), AF397589 (P. uvella rrn16), AF397588 (P. oviforme rrn16), AF397590 (C. reinhardtii rrn16), and AF397586 (C. humicola rrn16). Additional supplemen- tary materials on the Molecular Biology and Evolution web site include tufA and rrn16 sequence alignments in sequential GenBank format and in interleaved Pretty Print format, and the complete argument and additional data showing that the rrn16 gene is functional in Polytoma. Acknowledgments We are grateful to Brian Morton for supplying his Macintosh Pascal program and reference databases for calculating measures of codon bias, and to Aurora Ned-
  • 12. Accelerated Evolution of Leucoplast Genes 1821 elcu for allowing us to analyze her tufA sequence from P. uvella before it was published. Sally Otto analyzed Kimura’s (1957) equation to verify that reducing Ne af- fects strongly selected sites more than weakly selected sites. Primers were kindly provided by Paul Fuerst and Jeffrey Palmer. The large tufA database was kindly pro- vided by Charles Delwiche. Other members of the Birky lab at Ohio State University, notably Pamela Mackowski and Stacy Seibert, assisted in numerous ways. We are grateful to Jennifer Wernergreen, Aurora Nedelcu, and two anonymous reviewers for helpful comments on ear- lier manuscripts. D.V. submitted the data and a prelim- inary analysis in a thesis in partial fulfillment of the requirements for the Ph.D. degree at the Ohio State Uni- versity. This work was supported in part by research grants from NIH (GM34094) and NSF (BSR-9107069) to C.W.B. and from NIH (GM48207) to R.R.G., and by funds from the Ohio State University, the University of Arizona, and the University of Texas. LITERATURE CITED BALDAUF, S. L., and J. D. PALMER. 1990. Evolutionary transfer of the chloroplast tufA gene to the nucleus. Nature 344: 262–265. BERTHTOLD, H., L. RESHETNIKOVA, C. O. A. REISER, N. K. SCHIRMER, M. SLPRINZL, and R. HILGENFELD. 1993. Crystal structure of active elongation factor Tu reveals major do- main rearrangements. Nature 365:126–132. BIRKY, C. W. JR., and J. B. WALSH. 1988. Effects of linkage on rates of molecular evolution. Proc. Natl. Acad. Sci. USA 85:6414–6418. DELWICHE, C., M. KUHSEL, and J. D. PALMER. 1995. Phylo- genetic analysis of tufA sequences indicates a cyanobacter- ial origin of all plastids. Mol. Phylogenet. Evol. 4:110–128. DEPAMPHILIS, C. W., and J. D. PALMER. 1990. Loss of photo- synthetic and chlororespiratory genes from the plastid ge- nome of a parasitic flowering plant. Nature 348:337–339. DEPAMPHILIS, C. W., N. D. YOUNG, and A. D. WOLFE. 1997. Evolution of plastid gene rps2 in a lineage of hemiparasitic and holoparasitic plants: many losses of photosynthesis and complex patterns of rate variation. Proc. Natl. Acad. Sci. USA 94:7367–7372. ETTL, H. 1976. Die Gattung Chlamydomonas Ehrenberg. Beih. Nova Hedwigia 49:1–1122. FELSENSTEIN, J. 1993. PHYLIP (phylogeny inference package). Version 3.56. Distributed by the author, Department of Ge- netics, University of Washington, Seattle. GILBERT, D. G. 1992. SeqApp, a biological sequence editor and analysis program for Macintosh computers. Available via gopher or anonymous ftp to ftp.biol.indiana.edu. GILLHAM, N. W. 1994. Organelle genes and genomes. Oxford University Press, New York. GORDON, J., R. RUMPF, S. L. SHANK, D. VERNON, and C. W. BIRKY JR. 1995. Sequences of the rrn18 genes of Chla- mydomonas humicola and C. dysosmos are identical, in agreement with their combination in the species C. applan- ata (Chlorophyta). J. Phycol. 31:312–313. GRANT, D. M., N. W. GILLHAM, and J. E. BOYNTON. 1980. Inheritance of chloroplast DNA in Chlamydomonas rein- hardtii. Proc. Natl. Acad. Sci. USA 77:6067–6071. GUTELL, R. R. 1994. Collection of small subunit (16S- and 16S-like) ribosomal RNA structures: 1994. Nucleic Acids Res. 22:3502–3507. ———. 1996. Comparative sequence analysis and the structure of 16S and 23S rRNA. Pp. 111–128 in R. A. ZIMMERMAN and A. E. DAHLBERG, eds. Ribosomal RNA. Structure, evo- lution, processing, and function in protein biosynthesis. CRC Press, New York. HARRIS, E. H. 1989. The Chlamydomonas sourcebook. Aca- demic Press, New York. JUKES, T. H., and C. R. CANTOR. 1969. Evolution of protein molecules. Pp. 21–123 in H. N. MUNRO, eds. Mammalian protein metabolism. Academic Press, New York. KAWASHIMA, T., C. BERTHET-COLOMINAS, M. WULFF, S. CU- SACK, and R. LEBERMAN. 1996. The structure of the E. coli EF-Tu/EF-Ts complex at 2.5 A resolution. Nature 379:511– 518. KIMURA, M. 1957. Some problems of stochastic processes in genetics. Ann. Math. Stat. 28:882–901. ———. 1980. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 16:111–120. LANG, N. J. 1963. Electron-microscopic demonstration of plas- tids in Polytoma. J. Protozool. 10:333–339. LEMIEUX, C., C. OTIS, and M. TURMEL. 2000. Ancestral chlo- roplast genome in Mesostigma viride reveals an early branch of green plant evolution. Nature 249:649–652. LUSSON, N. A., P. M. DELAVAULT, and P. A. THALOUARN. 1998. The rbcL gene from the non-photosynthetic parasite Lathraea clandestina is not transcribed by a plastid-encoded RNA polymerase. Curr. Genet. 34:212–215. MAIDAK, B. L., N. LARSEN, M. J. MCCAUGHEY, R. OVERBEEK, G. J. OLSEN, K. FOGEL, J. BLANDY, and C. R. WOESE. 1994. The Ribosomal Database Project. Nucleic Acids Res. 22: 3485–3487. MORDEN, C. W., K. H. WOLFE, C. W. DEPAMPHILIS, and J. W. PALMER. 1991. Plastid translation and transcription genes in a non-photosynthetic plant: intact, missing and pseudo genes. EMBO J. 10:3281–3288. MORTON, B. R. 1993. Chloroplast DNA codon use: evidence for selection at the psbA locus based on tRNA availability. J. Mol. Evol. 37:273–280. ———. 1996. Selection on the codon bias of Chlamydomonas reinhardtii chloroplast genes and the plant psbA gene. J. Mol. Evol. 43:28–31. MUSE, S. V., and B. S. WEIR. 1992. Testing for equality of evolutionary rates. Genetics 132:269–276. NEDELCU, A. M. 2001. Complex patterns of plastid 16S rRNA gene evolution in nonphotosynthetic green algae. J. Mol. Evol. (in press). NICKRENT, D. L., R. J. DUFF, and D. A. M. KONINGS. 1997. Structural analyses of plastid-derived 16S rRNAs in holo- parasitic angiosperms. Plant Mol. Biol 34:731–743. NICKRENT, D. L., and E. M. STARR. 1994. High rates of nu- cleotide substitution in nuclear small-subunit (18S) rDNA from holoparasitic flowering plants. J. Mol. Evol. 39:62– 70. NISSEN, P., M. KJELDGAARD, S. THIRUP, G. POLEKHINA, L. RESHETNIKOVA, B. F. C. CLARK, and J. NYBORG. 1995. Crystal structure of the ternary complex of Phe-tRNA, EF- Tu, and a GTP analog. Science 270:1464–1472. RUMPF, R., D. VERNON, D. SCHREIBER, and C. W. BIRKY JR. 1996. Evolutionary consequences of the loss of photosyn- thesis in Chlamydomonadaceae: phylogenetic analysis of Rrn18 (18S rDNA) in 13 Polytoma strains (Chlorophyta). J. Phycol. 32:119–126. SAMBROOK, J., E. F. FRITSCH, and T. MANIATIS. 1989. Molec- ular cloning: a laboratory manual. Cold Spring Harbor Lab- oratory Press, Cold Spring Harbor, N.Y.
  • 13. 1822 Vernon et al. SARICH, V. M., and A. C. WILSON. 1973. Generation time and genomic evolution in primates. Science 179:1144–1147. SCHERBEL, G., W. BEHN, and C. G. ARNOLD. 1974. Untersu- chungen zur genetischen Funktion des farblosen Plastiden von Polytoma mirum. Arch. Microbiol. 96:205–222. SHARP, P. M., M. AVEROF, A. T. LLOYD, G. MATASSI, and J. F. PEDEN. 1995. DNA sequence evolution: the sounds of silence. Philos. Trans. R. Soc. Lond. B Biol. Sci. 349:241– 247. SHARP, P. M., and W.-H. LI. 1987. The codon adaptation in- dex—a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 15: 1281–1295. SIEMEISTER, G., C. BUCHHOLZ, and W. HACHTEL. 1990. Genes for the plastid elongation factor Tu and ribosomal protein S7 and six tRNA genes on the 73 kb DNA from Astasia longa that resembles the chloroplast DNA of Euglena. Mol. Gen. Genet. 220:425–432. SIEMEISTER, G., and W. HACHTEL. 1990a. Organization and nucleotide sequence of ribosomal RNA genes on a circular 73 kbp DNA from the colourless flagellate Astasia longa. Curr. Genet. 17:433–438. ———. 1990b. Structure and expression of a gene encoding the large subunit of ribulose-1,5-bisphosphate carboxylase (rbcL) in the colourless euglenoid flagellate Astasia longa. Plant Mol. Biol. 14:825–833. SIU, C.-H., K.-S. CHIANG, and H. SWIFT. 1976. Characteriza- tion of cytoplasmic and nuclear genomes in the colorless alga Polytoma. III. Ribosomal RNA cistrons of the nucleus and leucoplast. J. Cell Biol. 69:383–392. SWOFFORD, D. L. 1998. PAUP*: phylogenetic analysis using parsimony (*and many other methods). Version 4. Sinauer, Sunderland, Mass. TAMURA, K. 1992. Estimation of the number of nucleotide sub- stitutions when there are strong transition-transversion and GϩC content biases. Mol. Biol. Evol. 9:678–687. TURMEL, M., C. OTIS, and C. LEMIEUX. 1999. The complete chloroplast DNA sequence of the green alga Nephroselmis olivacea: insights into the architecture of ancestral chloro- plast genomes. Proc. Natl. Acad. Sci. USA 96:10248– 10253. VANCAMP, G., Y. VAN DE PEER, J. M. NEEFS, P. VANDAMME, and R. DEWACHTER. 1993. Presence of internal transcribed spacers in the 16S and 23S ribosomal RNA genes of Cam- pylobacter. Syst. Appl. Microbiol. 16:361–368. VERNON, D. 1996. Evolutionary consequences of the loss of photosynthesis in the nonphotosynthetic chlorophyte alga Polytoma. Dissertation, Ohio State University, Columbus. VERNON-KIPP, D., S. A. KUHL, and C. W. J. BIRKY. 1989. Molecular evolution of Polytoma, a non-green chlorophyte. Pp. 284–286 in C. T. BOYER, J. C. SHANNON, and R. C. HARDISON, eds. Physiology, biochemistry, and genetics of nongreen plastids. American Society of Plant Physiologists, Rockville, Md. WIMPEE, C. F., R. MORGAN, and R. L. WROBEL. 1992a. An aberrant plastid ribosomal RNA gene cluster in the root parasite Conopholis americana. Plant Mol. Biol. 18:275– 285. ———. 1992b. Loss of transfer RNA genes from the plastid 16S-23S ribosomal RNA gene spacer in a parasitic plant. Curr. Genet. 21:417–422. WOLFE, A. D., and C. W. DEPAMPHILIS. 1998. The effect of relaxed functional constraints on the photosynthetic gene rbcL in photosynthetic and nonphotosynthetic parasitic plants. Mol. Biol. Evol. 15:1243–1258. WOLFE, K. H., C. W. MORDEN, S. C. EMS, and J. D. PALMER. 1992. Rapid evolution of the plastid translational apparatus in a nonphotosynthetic plant: loss or accelerated sequence evolution of tRNA and ribosomal protein genes. J. Mol. Evol. 35:304–317. WOLFE, K. H., C. W. MORDEN, and J. D. PALMER. 1992. Func- tion and evolution of a minimal plastid genome from a non- photosynthetic parasitic plant. Proc. Natl. Acad. Sci. USA 89:10648–10652. WU, C.-I., and W.-H. LI. 1985. Evidence for higher rates of nucleotide substitution in rodents than in man. Proc. Natl. Acad. Sci. USA 82:1741–1745. YOUNG, D. Y., and C. W. DEPAMPHILIS. 2000. Purifying selec- tion detected in the plastid gene matK and flanking ribo- zyme regions within a group II intron of nonphotosynthetic plants. Mol. Biol. Evol. 17:1933–1941. DAVID RAND, reviewing editor Accepted May 30, 2001