Botanical Studies (2008) 49: 177-188.
*
Corresponding author: E-mail: chtsou@gate.sinica.edu.tw.
TECHNICAL REPORT
Technical report on the molecular phylogeny of Camellia
with nrITS: the need for high quality DNA and PCR
amplification with Pfu-DNA polymerase
Kunjupillai VIJAYAN and Chih-Hua TSOU*
Institute of Plant and Microbial Biology, Academia Sinica, Taipei, Taiwan 115, ROC
(Received May 17, 2007; Accepted December 6, 2008)
ABSTRACT.
Internal transcribed spacer (ITS) of nrDNA has been widely employed for reconstructing
phylogenetic relationships in plants, especially at the species level. Previous attempts to reconstruct the
molecular phylogeny of Camellia based on nrITS, however, have not succeeded due to technical difficulties.
In order to identify the major factors responsible for these difficulties and also to assess the efficacy of
nrITS in elucidating the interspecific relationships of Camellia, the present investigation was conducted with
seven closely or distantly related species. The purity of the DNA was found to be one of the major factors
affecting the success of PCR amplification and the errors in the sequences. Therefore, an efficient protocol has
been developed for extracting genomic DNA from dried leaf samples of Camellia. The purity of the DNA,
extracted using this method, was quite good as revealed by the A260/A280 ratio, which ranged from 1.84 to 1.89.
Further investigation on the effect of DNA polymerases on PCR induced variations revealed that the PCR
error rate was much higher in Taq-amplified sequences than Pfu-amplified sequences. The effect of the error
on phylogenetic analysis was evident from the wide dispersal of Ta q-amplified sequences across the gene tree
while the Pfu-amplified sequences from the same sample joined together to form a single clade. Our extensive
study of Camellia based on Pfu-amplified ITS sequences showed well-resolved interspecies relationships.
Since the results of the molecular phylogenetic investigation of Camellia needs to be reported in a series, due
to the technical and scientific complexity of the work, in this first report, we provide technical and scientific
insights into the major factors responsible for the failure of the PCR amplification, the occurrence of high
sequencing errors, and their effect on the phylogenetic interpretations. The results further stress the potential
of nrITS in deducing the phylogenetic relationships in Camellia.
Keywords: Camellia; DNA isolation; ITS; PCR error; Pfu; Ta q polymerase.
INTRODUCTION
Molecular phylogeny based on DNA sequences
has recently been used extensively to resolve intricate
problems at various taxonomic levels. The internal
transcribed spacer (ITS) of the nuclear ribosomal DNA
has been the most widely used molecular marker in
resolving phylogeny at the generic and specific levels
in the angiosperms (Baldwin et al., 1995; Alvarez and
Wendel, 2003). More than two-thirds of the related papers
published during 1998 to 2002 and in 2005 included
nrITS in the analyses (Alvarez and Wendel, 2003; Feliner
and Rossello, 2007). The advantages as well as the
disadvantages of nrITS in the phylogenetic application
have also been well explored (Buckler et al., 1997; Alvarez
and Wendel, 2003; Feliner and Rossello, 2007). Although
general concerns have been raised about its sequence
complexity and the existence of infra-species or even
infra-individual variations, which impact the accuracy of
the phylogeny being deduced, its advantages at a specific
level of phylogeny are unsurpassed by any other markers.
Recently, it has become customary that whenever nrITS
is used for phylogenetic analysis, a segment of plastid or
mitochondrial sequences is also incorporated to make a
comparison.
The genus Camellia, the largest genus in the family
Theaceae, is equally important for both the horticultural
and beverage industries. It is mainly distributed in East
Asia, with Southwest China as the center of distribution
(Ming, 2000). Taxonomically, the most popular
classification systems proposed by Chang (1981, 1998)
and Ming (2000) are very different. For example, Chang
recognized 284 species and treated them in 22 sections,
but Ming recognized 119 species and 14 sections only. The
difficulty in reaching a well-accepted classification system
resides in the large number of the species having natural
hybridizations (Ming, 2000) and in the great number of
pg_0002
178
Botanical Studies, Vol. 49, 2008
variations in floral characters. For example, the number of
bracteoles, sepals, and petals is not fixed in many species,
and the degree of filamentous and stylar fusion is variable
within species and difficult to quantify. In addition, floral
size often varies within a single species. In the past decade,
while molecular taxonomy has been overwhelmingly
applied to the ranks of all living organisms, some effort
has undoubtedly been devoted to the genus Camellia
as well. The earliest report on the molecular taxonomy
of Camellia, to our knowledge, was a presentation in a
conference by Thakor and Parks (1997). In this study, four
cpDNA regions (rbcL, trnT-trnL, trnL-trnF, atpB-rbcL)
were sequenced from 20 species which showed great
morphological diversity, but the sequence variability was
as low as 0% to 3%. Later on, in Thakor ¡¦s unpublished
thesis, one more cpDNA region (rpL16 intron) was used,
but it showed high intra-specific variation (Xiao, 2001).
Thakor¡¦s study based on morphology and DNA sequences
suggested that Camellia was paraphyletic (Xiao, 2001).
Xiao (2001) continued the study of molecular phylogeny
in Camellia, in the same laboratory, with two intron
sequences from the RNA polymerase (RPB2) gene. His
extensive survey, which included 149 Camellia species
and around 2000 nucleotides, resulted in a well resolved
tree in which four sections of the genus Camellia formed
distinct clades while the remaining, around three-fifths
of the species, formed a few clades that could be defined
by the color of the petals and the geographic distribution
range (Xiao and Parks, 2003). Nevertheless, this work
contributed many new insights and was different from
previous treatments done by Sealy (1958), Chang (1998),
and Ming (2000). Meanwhile, a few small-scale studies
also appeared like that of Tang and Zhong (2002), who
constructed a molecular phylogeny of 15 golden camellias
with nrITS using DNA from fresh leaf samples. Orel et
al. (2003) studied interrelationships of 32 species with
cpDNA trnL-trnF sequences and ISSR markers and found
that the former showed limited variations while the latter
appeared to be influenced by phylogenetic as well as non-
phylogenetic factors, rendering the tree topology difficult
to interpret. Yang et al. (2006) reported their surveys on
the utility of four DNA regions from 21 Camellia species.
They claimed that the cpDNA trnL-F and rpL16 were too
conserved to be used while nrDNA-ITS was difficult to
sequence technically. A waxy gene, however, appeared
to be potentially useful. Although Yang et al. (2006) was
the first to mention the difficulties of nrITS sequencing
in Camellia, our personal contact with many laboratories
revealed that, due to difficulties in sequencing, most of
the laboratories engaged in the molecular phylogeny of
Camellia with nrITS have abandoned their efforts. Since
our laboratory is involved in the phylogenetic analysis
of the family Theaceae, and nrITS appears to have
more potential than many other currently available gene
sequences in resolving the phylogeny of Camellia, w e
decided to undertake the present investigation to identify
and resolve the major bottlenecks in the phylogenetic
studies of Camellia with nrITS.
Our first major concern was the failure of PCR
amplification. Thus, we concentrated on the quality
of DNA as many contaminating agents¡Xlike
polysaccharides, phenolic compounds, tannins, resins,
and latex, present in the cell as secondary metabolites,
are present in the leaf samples¡Xusually coprecipitate
with DNA and interfere with the activity of the DNA
polymerase enzyme (Merlo and Kemp, 1976; Shioda and
Marakami-Musfushi, 1987; Fang et al., 1992; Panday
et al., 1996). In Camellia sinensis, i.e. tea, one of the
most common beverages, and most likely in many other
species of Camellia as well, a high content of secondary
metabolites is present in the leaves as 40% of the dry leaf
weight of Camellia sinensis comes from polyphenolic
compounds such as flavanols and their glycosides,
leucoanthocyanins, and phenolic acids, and more than
12% is polysaccharides (Graham, 1992). Therefore,
extraction of high quality genomic DNA from dried leaf
samples is challenging but a prerequisite for getting the
right sequences through PCR amplification. Initially we
tried a number of the existing protocols (Murray and
Thompson, 1980; Dellaporta et al., 1983; Rogers and
Bendich, 1988; Doyle and Doyle, 1990; Tai and Tanksley,
1990; Matsumoto et al., 1994; Takeuchi et al., 1994;
Wachira et al., 1995; Struwe et al., 1998), but none of
them was found efficient enough to isolate DNA suitable
for consistent PCR amplification. Therefore, we developed
a protocol suitable for extracting highly purified genomic
DNA, which is detailed in this report.
Our second concern was the possible occurrence of
intragenomic variations among the nrITS repeats, as our
attempts at direct sequencing of PCR product always
failed due to the large amount of noise (unreadable sites)
even though the DNA was of good quality. Thus, cloning
and sequencing of the PCR products became a necessity
to probe into the role of intragenomic heterogeneity
in it. Also proofreading DNA polymerase was tried to
examine the extent of the sequence variations caused by
the routinely used non-proofreading polymerase. Most
of the laboratories involved in phylogenetic studies with
large number of species and genera generally use the
thermostable polymerase from Thermus aquatics (Ta q )
for PCR amplifications due to its remarkable stability at
higher temperature, sensitivity, yield, and low economic
input. However, the PCR products amplified by Ta q
polymerase are subjected to nucleotide changes during
the amplification (Saiki et al., 1988; Tindall and Kunkel,
1988; Kwiatowski et al., 1991; Lundberg et al., 1991;
Kobayashi et al., 1999), and it was found that use of a high
fidelity DNA polymerase such as Pfu, Vent, DeepVent,
or UlTma could reduce the amplification error in PCR
products as compared to Ta q DNA polymerase (Flaman
et al., 1994; Cline et al., 1996). However, the higher cost
of these enzymes remains the major constraint on their
regular use. Considering this and the need of experimental
accuracy, it is prudent to carry out an initial polymerase
fidelity assay with the genomic DNA using non-
proofreading and proofreading polymerases. Therefore,
pg_0003
VIJAYAN and TSOU
¡X Technical report on molecular phylogeny of
Camellia
179
we made a comparison of the fidelity of Ta q polymerase
with that of Pfu polymerase, and the results showed
surprisingly high sequence variations in amplification
made by Ta q polymerases. During the last two years, we
have developed a protocol and succeeded in extending
the nrITS sequencing to about 100 species of Camellia.
The results obtained from the initial experiments were
insightful and useful in resolving intrageneric relationships
of Camellia. We intend to present our findings in a series
of reports to include both technical and scientific details of
the study. In this first report, we provide technical details
of the DNA extraction, methods of nrITS sequencing and
the major factors responsible for the failure of the PCR
amplification and the higher sequencing errors. We also
provide a snapshot of the potential of nrITS to resolve the
intrageneric relationships in Camellia.
MATERIALS AND METHODS
Plant material
Leaf samples of seven species belonging to four
sections, namely, C. assamica, C. euphlebia, C .
formosensis, C. microphylla, C. sasanqua, C. sinensis
var. sinensis, an d C. vietnanmensis were used for the
experiment (Table 1). The materials were kept in a sealable
bag containing silica gel immediately after detachment.
Voucher specimens were made and kept in the HAST
herbarium.
Solutions and reagents
¡E
Extraction buffer (Carlson et al., 1991). The extraction
buffer can be prepared by dissolving 2% (w/v)
cetyltrimethyl ammonium bromide (CTAB), 1.4
M sodium chloride, 20 mM EDTA, 100 mM Tris-
HCl, 1% PEG 8000. Just before use, add 1.5% (v/v)
£]-mercaptoethanol to the extraction buffer.
¡E
2% (w/v) Polyvinyl pyrrolidone (PVP; Sigma, St.
Louis, MO; MW 10,000)
¡E
Phenol: Chloroform: isoamyl alcohol (24:24:1)
¡E
Chloroform: isoamyl alcohol (24:1)
¡E
3 M Sodium acetate solution (pH 4.8)
¡E
Absolute alcohol
¡E
70% (v/v) ethanol
¡E
RNAse-A (10 mg/ml)
¡E
Tris-HCl (pH 8.0)
¡E
EDTA (pH 8.0)
¡E
GENECLEAN III-DNA purification kit.
DNA extraction protocol
One square cm of the dried leaf lamina was powdered
in a tube using a FASTPrep-FPI120 machine (BIO 101
Systems, New York). The extraction buffer along with
2% (w/v) PVP was added into the tube containing the leaf
powder, mixed thoroughly, and kept at 4¢XC for 5 days.
The slurry was subsequently incubated at 65¢XC for 1 h.
After incubation, an equal volume of chloroform: isoamyl
alcohol (24:1) was added, mixed well, and centrifuged at
14,000 rpm for two min. The supernatant was collected
in a fresh tube, and this step was repeated one more time.
To the supernatant, a 1/10 volume of 3 M-sodium acetate
and 2 volumes of pre-cooled ethyl alcohol were added
and mixed slowly by inverting the tube. The tubes were
kept at -20¢XC for 1h to precipitate the DNA before it was
spooled out into a fresh tube and air-dried at 37¢XC. The
air-dried DNA was dissolved in TE Buffer (10 mM Tris
HCl [pH 8.0] and 1 mM EDTA). RNA contamination
was removed by treating the DNA with bovine pancreatic
RNase-I at a final concentration of 40 £gg/ml at 37¢XC
for 30 min. The DNA, after RNAse treatment, was re-
extracted using a modified protocol of Struwe et al. (1998).
Table 1. List of samples, voucher information, and sequence accession numbers.
Species
Section
Collection locality
Voucher
(source origin) Accession number
Chang (1998) Ming (2000)
C. euphlebia
Chrysantha Archecamellia Xishanbana, Yunnan, China Wong Hong,
2004-3-4-a
Taq: EF544731-EF544734
Pfu: EF544735-EF544737
C. microphyla
Paracamellia Tuberculata Internatl. Camellia Gard.,
Zerjiang, China (cultivated)
Tsou-853T (J.
Y. gao)
Taq: EF544763-EF544767
Pfu: EF544768-EF544772
C. assamica
Thea
Thea Tea Improv. Res. Inst.
Nantou, Taiwan (cultivated)
Tsou-455S Taq: EF544721-EF544725
Pfu: EF544726-EF544730
C. formosensis
Thea
Thea Nantou, Taiwan
Tsou-318S Taq: EF544711-EF544715
Pfu: EF544716-EF544720
C. sinensis var. sinensis Thea
Thea West Tian-mu Mt., Zerjiang,
China
Tsou-894T Taq: EF544738-EF544741
Pfu: EF544742-EF544745
C. sasanqua
Oleifera Oleifera Yang-Ming Mt., Taipei,
Taiwan (cultivated)
Tsou-988T Taq: EF544755-EF544757
Pfu: EF544758-EF544762
C. vietnamensis
Oleifera Oleifera Internatl. Camellia Gard.,
Zerjiang, China (cultivated)
Tsou-856T (J.
Y. Gao)
Taq: EF544746-EF544749
Pfu: EF544750-EF544754
pg_0004
180
Botanical Studies, Vol. 49, 2008
An equal volume of Phenol-Chloroform-Isoamyl alcohol
(24:24:1) was added to the tube and mixed gently for 10
min. The upper aqueous layer, obtained after centrifuging
at 14,000 rpm for 2 min, was transferred into a new tube
and an equal volume of Chloroform-Isoamyl alcohol
(24:1) was added and mixed gently for 10 min. The tube
was subsequently centrifuged at 14,000 rpm for 2 min.
From the upper aqueous layer 300 £gl was transferred into
a new tube containing 900 £gl of sodium iodide solution
and 20 £gl of glass milk prepared from GENECLENE
R
kit (BIO101 Systems, New York), mixed gently for 20
min and centrifuged at 14,000 rpm for 30 s to pellet the
glass milk. The supernatant was poured off and 900 £gl of
NEW
TM
wash solution was added. The glass milk pellet
was resuspended by gently breaking the pellet with a
pipette tip and shaking the tube. The tube was centrifuged,
and the supernatant was poured off. The washing process
was repeated thrice with 900 £gl NEW
TM
wash solution,
and a fourth washing was done with 150 £gl NEW
TM
wash
solution. After centrifuging, the supernatant was removed
completely with a fine tipped pipette without disturbing
the pellet. The glass milk pellet was resuspended in TE
buffer and placed at 50¢XC for 10 min to elute the DNA
from the glass milk beads. The tube was centrifuged again
at 14,000 rpm for 2 min, and the supernatant was collected
into a new tube without disturbing the glass milk pellet.
DNA quantification
The quantity and quality of the DNA was estimated
with spectrophotometer (Hitachi U2000; Hitachi High
Technologies America, Inc. Life Sciences Division, San
Jose, CA, USA) as well as by electrophoresis on 1.0%
agarose gel in 1¡ÑTAE buffer.
PCR amplification with Taq and Pfu polymerases
The PCR amplification of the ITS1-5.8S-ITS2 regions
of the nrDNA was achieved with the primer pairs ITSleu1
(5¡¦-GTCCACTGAACCTTATCATTTAG-3¡¦ (Urbatsch et
al., 2000) and ITS4 (5¡¦-TCCTCCGCTTATTGATATGC-3¡¦)
(White et al., 1990). The PCR amplification reaction was
first carried out with Ta q polymerase (1U) in a reaction
mixture containing 5 mM Tris-HCl (pH 8.8 at 25¢XC), 10
mM NaCl, 0.01 mM EDTA, 0.1 mM DTT, 0.5% (v/v)
glycerol, 0.1% (v/v) Triton
R
X-100 (Promega corporation,
Madison, WI, USA), 2 mM MgCl
2
, 20 ng of Genomic
DNA, 100 £gM dNTPs, and 100 £gM primers. Then,
amplifications were carried out with Pfu polymerase
(Promega corporation, Madison, WI, USA) with the same
primers. The PCR cocktail contained Pfu 1U, 20 mM Tris-
HCl (pH 8.8 at 25¢XC), 10 mM KCl, 10 mM (NH
4
)
2
SO
4
,
2.0 mM MgSO
4
, 0.1% (v/v) Triton
R
X-100 and 0.1 mg/ml
nuclease-free BSA (Promega Corporation, Madison, WI,
USA), 20 ng of Genomic DNA, 100 £gM dNTPs, and 100
£gM primers. The PCR thermal cycler (GeneAmp PCR
system 2700, Applied Biosystem, Foster City, CA, USA)
was programmed as 94¢XC for 5 min for denaturing the
DNA followed by 30 cycles of 1 min at 94¢XC, 1 min at 50
¢XC, 1 min at 72¢XC, followed by a final extension of 7 min
at 72¢XC (Mast, 1998). PCR products were electrophoresed
on 1.0% agarose gel in TAE buffer, stained with ethidium
bromide (2 £gg/ml), and visualized under ultraviolet light.
Cloning and sequencing
PCR products were purified with QIAquick PCR
purification kit (QIAGEN Inc., Valencia, CA, USA). PCR
products of each sample were cloned to PCR II-TOPO
cloning vector and transformed into chemically competent
Ecoli- DH5£\
TM
-T1
R
cells (provided with the kit) following
the instructions of the manufacturer (Invitrogen Life
Technologies, Carlsbad, California, USA). Colonies were
cultured overnight at 37¢XC on LB (Luria-Bertani medium)
ampicillin/IPTG (isopropyl £]-D-1-thiogalactopyranoside)/
X-gal (5-bromo-4-chloro-3-idolyl beta-D-galactoside)
selective medium. Plasmids from white colonies were
isolated using Mini-M
TM
plasmid DNA extraction system
(VIOGENE, Sunnyvale, California, USA). The extracted
plasmids were digested with EcoR1 and tested on 1%
agarose gel for inserts. For each species, plasmids from
five colonies were sequenced using an ABI PRISM
dye terminator cycle sequencer, model 3700 (Applied
Biosystems, Foster City, CA, USA).
The full sequences were generated by aligning both
forward and reverse sequences of the same clone, and
any variation observed in the sequences was checked by
examining corresponding peaks in the chromatograms of
both forward and reverse sequences. Sequence alignments
were constructed with the help of the PILEUP program of
GCG, version 8.1 (Genetic Computer Group, 1994). Base
composition and length variability were estimated with
BioEdit, version 5.0.9 (Hall, 1999). Intragenomic variation
was calculated from the multiple clones as mean number of
nucleotide differences per site between pairs of sequences
according to Kimura¡¦s two-parameter model using MEGA
3.1 (Kumar et al., 2004). Standard errors of the sequence
divergence were estimated by applying 1000 bootstrap
replications. Total number of transitions and transversions
present within a set of sequences was estimated with
Tamura-Nei (gamma) algorithm (Tamura and Nei, 1993)
as implemented in MEGA 3.1. The PCR error rate was
calculated from the observed error frequency on the basis
of pair-wise comparison by taking into account the number
of doubling cycles using the formula (2 ¡Ñ observed
frequency) / (number of cycles) as described by first
Gelfand and White (1990) and later by Kwiatowski et al.
(1991). While calculating the PCR error, the variable sites
inherently present in the sequence were eliminated based
on the existence of patterns of Pfu-amplified sequences
present in each sample. For instance, in the aligned matrix
of the Pfu-amplified sequences of each species, if two or
more sequence types could be recognized, the variable
sites resulting from such differences in the sequence types,
as indicated in Figure 1, were identified and eliminated
from further analyses of errors.
The gene tree was generated with PAUP* 4.0b10
(Swofford, 2001) using a maximum likelihood algorithm.
pg_0005
VIJAYAN and TSOU
¡X Technical report on molecular phylogeny of
Camellia
181
Figure 1. Intragenomic variations inherently present in the sequences of Camellia assamica; portions highlighted with rectangles.
pg_0006
182
Botanical Studies, Vol. 49, 2008
To choose the nucleotide substitution model for the
maximum likelihood analysis, the computer program
Model Test 3.06 (Posada and Crandall, 1998) was used.
This program employs two statistical methods, a likelihood
ratio test (LRT) and an Akaike information criterion (AIC,
Akaike, 1974), to find out the best fitting model to be used
for the subsequent analysis with PAUP* 4.0b10. Based on
the results, the likelihood settings from the best-fit model
(GTR+G) were selected and implemented in a maximum
likelihood analysis with PAUP* 4.0b10 (Swofford,
2001). A heuristic search, containing 100 random taxon-
addition replicates, TBR branch swapping and MulTrees,
Collapse and Steepest Descent options were in effect, was
conducted with no upper limit imposed on the trees held
in memory. To ascertain the relative degree of support for
branches in the cladograms, jackknife (Farris et al., 1996)
support was estimated with 100 replicates using heuristic
searches and a random addition of sequences. Trees were
rooted using Pyreneria melanogaster as outgroup since
Pyrenaria (Tutcheria) is so far known as the sister genus
of Camellia (Prince and Parks, 2001).
RESULTS AND DISCUSSION
DNA quality
The DNA isolated from the dry leaf samples of
Camellia using our protocol was of high quality as is
evident from spectrophotometric and gel electrophoresis
analyses. The spectrophotometric analysis revealed that
the absorption ratio of the DNA at 260/280 nm was in the
range of 1.84 (C. vietnamensis) and 1.89 (C. formosensis).
A nearly 2.0 ratio indicates little contamination from
proteins, polyphenols, or polysaccharides in the DNA.
Electrophoresis of the DNA on agarose gel revealed
a conspicuous band nearly 22 kb in size with a little
shearing of the DNA. This may have resulted from the
purification with the GENECLEAN
R
III kit as shearing of
DNA with higher molecular weight (>10 kb) is a problem
often encountered with this kit (Application manual;
GENECLEAN
R
III kit, BIO101 Systems, New York).
This purification step was, however, essential as the DNA
obtained either from the CTAB extraction method alone
or from the method of Struwe et al. (1998) alone was not
amplified by PCR. Further, it should be noted that this
little shearing of the DNA has not interfered with the
subsequent PCR amplification of the DNA. The major
components we added into the extraction buffer were PVP,
CTAB, and PEG8000 to remove polyphenolic compounds,
polysaccharides, and proteins. PVP is reported to form
complex hydrogen bonds with phenolic compounds and
co-precipitates with cell debris upon cell lyses, and upon
centrifugation in the presence of chloroform the PVP
complexes accumulate at the interface between the organic
and the aqueous phases (Kim et al., 1997; Barnwell et
al., 1998). Similarly, CTAB, a cationic detergent which
solubilizes membranes, binds to fructans and other
polysaccharides to form complexes that are removed
during chloroform extraction. PEG-8000 also removes
proteins and polysacchriodes (Agudo et al., 1995).
Therefore, in our protocol most of the phenolics and
polysaccharides must have been removed in the first phase
of the extraction. In the second phase, the purification
steps with GENECLEAN
R
III kit removed the remaining
proteins, polyphenolics, and polysaccharides along with
tannins. This kit allows the DNA to bind onto the silica
particles present in the EZ-GLASSMILK under high salt
concentrations and to release from them under low salt
concentrations. The DNA, while adhering to the silica
particles, is washed many times to remove all contaminants
from it. Thus at the end, though the DNA was devoid of
most contaminants, the process left it slightly sheared.
Singh et al. (1999) also reported high degradation of DNA
upon its extraction from black tea, attributing it to the
extreme processing of the tea leaf and to the binding of
some of the phenolic compounds to the DNA upon cell
lyses as reported earlier (John, 1992). Nevertheless, in the
present study DNA shearing was insufficient to interfere
with PCR amplification.
PCR amplifications with Taq and Pfu
polymerases
The DNA extracted with our method was successfully
amplified with Ta q and Pfu DNA polymerases. Among
the 70 clones, i.e., five clones from each of the Ta q and
Pfu PCR amplifications for each of the seven samples, 63
clones were successfully sequenced (Table 1). However,
attempts to sequence the PCR products directly, without
cloning, were mostly unsuccessful. Sequences of the five
clones from each PCR amplification showed a certain
degree of intragenomic variability in the form of indels and
substitutions (Table 2). Those variations present among
the clones could be the main reason direct sequencing
of the PCR products failed. Further, it is obvious that
when sequences amplified with the Ta q polymerase
were compared with Pfu-amplified ones, the former
exhibited much higher intragenomic variability though
Taq polymerase is much commonly used in routine studies
than Pfu enzyme. The effect of the two DNA polymerases
on the sequence variability will be explicated in the next
paragraph.
The length of the ITS1-5.8S-ITS2 regions, of the seven
taxa studied, varied from 556 bp in Camellia sasanqua to
682 bp in C. microphylla when amplified with Ta q DNA
polymerase and varied from 634 bp in C. sinensis to 666
bp in C. microphylla when amplified with Pfu polymerase
(Table 1). The average length of sequences with Ta q was
643.52, and with Pfu 644.51. The average G+C content
of the sequences with Ta q was 66.12, and with Pfu 67.57.
Thus, both the length and G+C content of the sequences
amplified by Pfu polymerase were slightly higher.
Sequence variability
The substitution rate (cumulative transitions and
transversions) and indels were higher among Ta q -
amplified sequences than among Pfu-amplifed sequences
pg_0007
VIJAYAN and TSOU
¡X Technical report on molecular phylogeny of
Camellia
183
(Table 2). The average substitution+indel in sequences
amplified with Ta q polymerase was highest in C. euphebia
(15.8 per sequence) and was least in C. vietnamensis (3.5
per sequence) while the same in Pfu amplified sequences
was highest in C. sinensis (2.8 per sequence) and least
i n C. euphlebia (0.2 per sequence). When all the Taq -
and Pfu-amplified sequences of the same sample aligned
together and a consensus sequence was generated for each
species, the most common variations observed were C->T
and G->A in both Taq - and Pfu- amplified sequences.
Hofreiter et al. (2001) also found such biased substitutions
in the PCR products of DNA from ancient samples. They
found that a high frequency of deoxyadenosine residues
was incorporated to opposite positions where the template
carries deoxycytidine residues. One of the reasons for such
a high rate of G->A and C->T substitution is the tendency
of the Ta q polymerase to add deoxyadenosine residues
when it reaches the end of templates (Clark, 1988).
This has been shown to cause substitutions opposite to
deoxycytidine residues when degraded DNA or DNA with
high contaminants is used for PCR amplification (Kwok
et al., 1990; Paabo et al., 1990). Another reason for these
deamination-like substitutions in sequences amplified by
Ta q polymerase could be the preferential amplification
of templates with methylated residues. Methylation has
been implicated in nucleolar dominance phenomena, and
the resulting non-functional rDNA copies (pseudogenes)
therefore typically show patterns of deamination-like
substitutions (C->T and G->A) at methylation sensitive
sites relative to their active counterparts (Muri et al., 2001;
Marquez et al., 2003). Also noteworthy is that a higher
occurrence of C->T and G->A substitutions was noticed in
this study when G and C were consecutive, either in GC or
CG order, in both Ta q and Pfu-amplified sequences.
In order to prevent the inherently-present sequence
variations from influencing the rate of PCR/sequencing
errors, we identified and removed the variable sites
resulting from the intragenomic variations inherently
present in the nrDNA arrays. Out of the seven species
studied, three, namely C. assamica, C. microphylla, C.
sasanqua, showed definite patterns of variations. The
variations at the bases 12 to 15, 169 to173, 388, 390, 392,
473, and 620 (Figure 1) in C. assamica, at 9 to 12 in C.
microphylla, an d at 111 and 6 54 i n C. sasanqua were
considered variations inherently present in the nrITS
arrays, for they were consistently present in more than two
sequences amplified by Pfu-DNA polymerase. Therefore,
these variables sites were excluded from further analyses
of PCR/sequencing errors.
The mean pair-wise distance (kimura-2-parameters)
revealed that the highest distance was observed among
sequences amplified with Ta q polymerse (Table 2). The
number of indels present among the clones of each species
varied from 3.0 to 28.0 for Ta q and 0.0 to 5.0 for Pfu. This
high frequency of indels in Ta q-amplified sequences must
be due to PCR error. The PCR error rates also showed
considerable differences between Ta q - and Pfu-amplified
sequences. The error rate in the Ta q-amplified sequences
varied from 2.3 ¡Ñ 10
-4
/site/duplication in C. vietnamensis
to 1.4 ¡Ñ 10
-3
/site/duplication in C. euphlebia while in the
Pfu- amplified sequences it varied only from 0.0 ¡Ñ10
-5
/site/
duplication in C. euphlebia to 1.5 ¡Ñ 10
-4
/site/duplication
in C. sinensis var. sinensis (Table 3). This difference in
the error rates of Ta q and Pfu polymerases corroborates
the earlier reports that the error rate of Pfu polymerase
Table 2. Variations in length and G-C content of the sequences amplified with Taq and Pfu DNA polymerases.
Sample
PCR enzyme No. sequences
obtained
Range of
length (bp)
Range of G-C
content (%) Indels
Substitutions
ti* tv Total
C. euphlebia
Ta q
4
633-664 62.17-67.32
6 42 15 57
Pfu
3
647-649 66.72-66.77
1
0 0 0
C. microphylla
Ta q
5
650-682 66.86-68.46 28 14 9 23
Pfu
5
653-666 67.84-68.58
3
2 1 3
C. assamica
Ta q
5
640-658 62.97-68.13
6 17 4 21
Pfu
5
655-657 67.73-68.09
3
3 1 4
C. formosensis
Ta q
5
632-648 60.44-68.36
7 19 4 23
Pfu
5
646
67.96-68.27
0
1 0 1
C. sinensis var. sinensis Ta q
4
616-642 62.05-68.14 14 20 6 26
Pfu
4
634-639 67.87-68.45
5
4 2 6
C. sasanqua
Ta q
4
556-665 68.20-68.50 25 20 6 26
Pfu
5
657-661 67.37-68.08
3
3 0 3
C. vietnamensis
Ta q
4
635-664 66.67-67.62
5
7 2 9
Pfu
5
635
66.77-67.24
0
3 1 4
*ti: transition; tv: transversion.
pg_0008
184
Botanical Studies, Vol. 49, 2008
was 5 to 30 fold lower than the error rates of other proof
reading and non-proofreading enzymes (Cline et al., 1996;
Flaman et al., 1994). It has been noticed as well that the
Ta q polymerase has problems in amplifying regions with
high G+C contents (Innis et al., 1988). For instance, while
using Ta q polymerase to amplify a G+C-rich region of
the human 18S rRNA gene, Cariello et al. (1991) found
formation of deletions in the hairpin regions. This finding
supports an earlier report that using Ta q polymerase in
PCR amplification of sequences with secondary structures
may give rise to difficulties in the form of deletions and
amplification of unknown regions (Triglia et al., 1988;
Ochman et al., 1988). Since the ITS regions of Camellia
have an average of 66% GC content, which is towards
the high end of the range recorded for plants (Baldwin
et al., 1995), and do form secondary structures during
amplifications, the aforementioned problems in the PCR
amplification with Ta q polymerase can be anticipated. We
must note that the PCR error rates of the Pfu amplified
sequences in the seven Camellia samples (0.0 ¡Ñ 10
-5
, 6.0
¡Ñ 10
-5
, 8.1 ¡Ñ 10
-5
, 2.0 ¡Ñ 10
-4
, 1.5 ¡Ñ 10
-5
, 6.07 ¡Ñ 10
-5
, 8.4
¡Ñ 10
-5
) were higher than the commercially reported error
rate of 1.3 ¡Ñ 10
-6
for Pfu DNA polymerase (Slater et al.,
1998). This could be due to the high G+C content and
the tendency to form secondary structures in the nrITS
of Camellia, as the commercially reported error rate was
estimated from the E. coli lac I gene, which is devoid of
the above problems. Another problem with the Camellia
nrITS is the presence of mutational hot spots such as
mononucleotide stretches and microsatellites, which
are particularly prone to PCR induced errors (Lapegu et
al., 1999). An error rate of one in 1000 bp at each cycle
results in one error in 400 bp in the final product after
25 cycles (Saiki et al., 1988). Hence, Kobayashi et al.
(1999) opined that to obtain sequence data free from
PCR error, sequences for several clones are to be made
and compared. In the light of these factors, we stress that
the use of Pfu is essential to reducing PCR errors in GC-
rich regions like the nrITS of Camellia. Furthermore, the
findings of this study may very well explain the failure of
previous attempts to sequence nrITS in Camellia if they
were performed by routinely used Taq polymerase on low
quality DNA.
Phylogenetic applications
Phylogenetic analyses were made to examine the
usefulness of nrITS in Camellia and the impact of different
DNA polymerases on the analyses. The consequence
of not eliminating PCR-induced variability from the
phylogenetic analysis was clear from the gene trees
(Figures 2, 3). From the gene tree obtained from both Ta q
and Pfu DNA polymerases (Figure 2), it could be observed
that those clones amplified with Ta q polymerase showed
different patterns of grouping. Clones of the same species
either joined the sequences of other species or remained
as singletons. For instance, the clone C. formosensis-4 Ta q
joined with the clade formed by clones of C. euphlebia.
Similarly, the clones C. assamica-3 Ta q , C. formosensis-
5 Ta q, C. sasanqua-1 Taq , an d C. microphylla-5 Taq
remained as isolates showing no definite relationship
Table 3. Sequence divergence calculated using Kimura-2-parameter model the PCR error rates in the sequences amplified with Ta q
and Pfu DNA polymerases.
Species
DNA polymerase No. of clones Kimura-2-parameter distance Rate of PCR error*
C. euphlebia
Ta q
4
0.097¡Ó0.009
1.4¡Ñ10
-3
Pfu
3
0.000¡Ó0.000
0.0¡Ñ10
-5
C. microphylla
Ta q
5
0.035¡Ó0.005
4.4¡Ñ10
-4
Pfu
5
0.005¡Ó0.002
6.0¡Ñ10
-5
C. assamica
Ta q
5
0.033¡Ó0.005
4.2¡Ñ10
-4
Pfu
5
0.006¡Ó0.002
8.1¡Ñ10
-5
C. formosensis
Ta q
5
0.085¡Ó0.010
4.7¡Ñ10
-4
Pfu
5
0.002¡Ó0.001
2.0¡Ñ10
-5
C. sinensis var. sinensis
Ta q
4
0.052¡Ó0.007
6.8¡Ñ10
-4
Pfu
4
0.009¡Ó0.003
1.5¡Ñ10
-4
C. sasanqua
Ta q
4
0.044¡Ó0.007
6.8¡Ñ10
-4
Pfu
5
0.006¡Ó0.002
6.07¡Ñ10
-5
C. vietnamensis
Ta q
4
0.017¡Ó0.004
2.5¡Ñ10
-4
Pfu
5
0.006¡Ó0.002
8.4¡Ñ10
-5
*
Rate of nucleotide substitution/site/per duplication, estimated after eliminating the variable sites that are inherently present in the
sequences.
pg_0009
VIJAYAN and TSOU
¡X Technical report on molecular phylogeny of
Camellia
185
with any other groups. In contrast, all clones of the
same species amplified with Pfu polymerase grouped
together into one clade. Thus, it is clear that the Ta q DNA
polymerase induced PCR errors in the sequences and
greatly affected the topology of the cladogram. When
the Pfu-amplified sequences were analyzed separately,
all clones of the same taxon grouped together, except
i n C. sinensis (Figure 3). In this gene tree depicting
the interspecies relationships of the seven species, four
sections were well demonstrated. Species of C. assamica,
C. formosensis, an d C. sinensis belonging to Section
Thea are so closely related that they are often treated as
three varieties under C. sinensis (Su et al., 2007). The 14
sequences of these three species formed one clade, with
the five sequences of C. formosensis and four of the five
sequences of C. assamica forming their subgroups. The
clones of C. sinensis were not held together, and this
was not surprising as C. sinensis has a long history of
domestication, naturalization, and cultivation. The two
species, C. sasanqua and C. vietnamensis, belonging to
the same section viz., sect. Oleifera, are grouped together
in a clade with 100% support and sequences of each
species further joined together as subgroups. As for C.
euphebia and C. microphylla, they belong to the sections
Chrysantha and Paracamellia, respectively (Chang, 1998),
and their sequences form distinct clades. In summary, the
four sections (Chrysantha, Oleifera, Paracamellia, and
Thea) as well as the species within the sections can be
well separated in the gene tree generated by the Pfu-DNA
polymerase amplified sequences.
Figure 2. Maximum likelihood 50% majority rule consensus
tree derived from the sequences of seven species of Camellia
amplified with both Taq and Pfu polymerases.
Figure 3. Maximum likelihood 50% majority rule consensus
tree derived from the sequences of seven species of Camellia
amplified with only Pfu polymerases.
pg_0010
186
Botanical Studies, Vol. 49, 2008
In short, this investigation has demonstrated that
molecular phylogeny with nrITS has great potential to
resolve many vexing problems associated with both
intersectional and intrasectional relationships of the
Camellia. However, this is possible only if high quality
DNA can be extracted from the dried leaf samples and
proofreading enzyme like Pfu polymerase can be used for
PCR amplification to reduce the PCR/sequencing errors.
The DNA extraction protocol developed in this study
could be of much use to this end. Furthermore, owing
to the possibility of pseudogenes being present in the
genome, sequences from multiple clones have to be used
to generate a gene tree to distinguish the functional genes
from the pseudogenes and to minimize the interference of
pseudogenes in the phylogenetic analysis.
Acknowledgement. We would like to thank Mong-Huai,
Su, Hong Wong and Wen-Ju Zhang for assistance with
sample collection and Yaw-wen Yang for technical advice.
This research was supported by the National Science
Council, Taiwan, Republic of China (Project NSC95-
2621-B-001-009-MY2).
LITERATURE CITED
Agudo, L.C., P. Gavidia, P. Perez-Bermudez, and J. Segura.
19 95 . P E G pr ec i pi t at i on, a re q ui re d s te p fo r P C R
amplification of DNA from wild plants of Digitalis obscura
L. BioTechniques 18: 766-768.
Akaike, H. 1974. A new look at the statistical model
identification. IEEE Trans. Contr. 19: 716-723.
Alvarez, I. and J.F. Wendel. 2003. Ribosomal ITS sequences and
plant phylogenetic inference. Mol. Phylogenet.. Evol. 29:
417-434.
Baldwin, B.G., M.J. Sanderson, J.M. Porter, M.F.
Wojciechowski, C.S. Campbell, and M.J. Donoghue. 1995.
The ITS regions of nuclear ribosomal DNA: a valuable
s o urce of e vid enc e on a ngio sp erm phy log eny. Ann .
Missouri Bot. Gard. 82: 247-277.
Barnwell, P., A. N. Blanchard, J.A. Bryant, N. Smirnoff, and A.F.
Weir. 1998. Isolation of DNA from highly mucilaginous
succulent plant Sedum telephium. Plant. Mol. Biol. Reptr.
16: 133-138.
Buckler, E.S., A. Ippolito, and T.P. Holtsford. 1997. The
evolution of ribosomal DNA: divergent paralogues and
phylogenetic implications. Genetics 145: 821-832.
Cariello, N.F., W.G. Thilly, J.A. Swenberg, and R.S. Thamas.
1991. De letion m utagenes is duri ng pol ymeras e cha in
reaction: dependence on DNA polymerase. Gene 99:
105-108.
Carlson, J.E., L.K. Tulsieram, J.C. Glaubitz, V.W. K Luk, C.
Kauffeldt, and R. Rutledge. 1991. Segregation of random
amplified DNA markers in F1 progeny of conifers. Theor.
Appl. Genet. 83: 194-200.
Chang, H.T. 1981. A taxonomy of the genus Camellia. Acta
S c i ent a rum Na tur al ium . Univ ers i ta t is . S u nya ts e ni,
Monographic series 1: 1-180.
Chang, H.T. 1998. F lora of Reipublicae Popularis Sinicae,
Delectis F lorae Republicae Popularis S inicae, Agendae
Academiae S inicae Edita, Tomus, Beijing Science P ress,
China, Vol. 49, pp. 101-113.
Clark, J.M. 1988. Novel non-templated addition reactions cata-
lyzed by prokaryotic and eukaryotic DNA polymerases.
Nucleic Acids Res. 16: 9677-9686.
Cline, J., J.C. Barman, and H.H. Hogrefe. 1996. PCR fidelity of
Pfu polymerase and other thermostable DNA polymerases.
Nucleic Acids Res. 24: 3546-3551.
Dellaporta, S.L., J. Wood, and J.B. Hicks. 1983. A plant DNA
mini preparation a. Vers ion II. Plant Mol. Biol. Rep. 1:
19-23.
Doyle, J.J. and J.L. Doyle. 1990. Isolation of plant DNA from
fresh tissue. Focus 12: 13-15.
Fang, G., S. Hamm er, and R. Groumet. 1992. A quick and
inexpensive method for removing polysaccharides from
plant genomic DNA. Biofeedback 13: 52-54.
Farris, J.S., V.A. Albert, M. Kallersjo, D. Lipscomb, and A.G.
Kluge. 1996. Parsimony jackknifing outperforms neighbor-
joining. Cladistics 12: 99-124.
Feliner, G.N. and Rossello. 2007. Better the devil you know.
Guidelines for insightful utilization of nrITS in species-
level evolutionary studies in plants. Mol. Phylologenet.
Evol. 44: 911-919.
F lama n, J .M., T. Frebourg, V. More au, F. Charbonnie r, C.
Martin, C. Ishioka, S.H. Friend, and R. Iggo. 1994. A rapid
PCR fidelity assay. Nucleic Acids Res. 22: 3259-3260.
Gelfand, D.H. and T.J. White. 1990. Thermostable DNA
polymerases. In M.A. Innis, D.H. Gelfand, J.J. Sninsky and
T.J. White (eds.), PCR Protocols: a Guide to Methods and
Applications. Academic Press, London, pp. 129-141.
Graham, H.N. 1992. Green tea composition, consumption, and
polyphenols chemistry. Prev. Med. 21: 334-350.
Hall, T.A. 1999. BioEdit: a user friendly biological sequence
alignment editor and analysis program for windows 95/98/
NT. Nucleic. Acid. Sym. Ser. 41: 95-98.
Hofreiter, M., V. J aenicke, D. S erre, A. Haeseler Av. and S.
Paabo. 2001. DNA sequences from multiple amplifications
reveal artifacts induced by cytosine deamination in ancient
DNA. Nucleic Acids Res. 29:4793-4799.
Innis, M.A., K. B. Mayamboo, D.H. Gelfand, and M.A. Brow.
1988. DNA se que nc ing with Ther mus aquat icus DNA
polym eras e and direct s equencing of DNA polym eras e
chain reaction-amplified DNA. Proc. Natl. Acad. Sci. USA
85: 9436-9440.
John, M.E. 1992. An efficient method for isolation of RNA and
DNA from plants containing polyphenolics. Nucleic Acids.
Res. 20: 2381.
Kim, C.S., C.H. Lee, J.S . Shin, Y.S . Chung, and J.E. Hyung.
1997. A simple and rapid method of isolation of high quality
genomic DNA from fruit trees and conifers us ing PVP.
Nucleic Acids Res. 27: 66-58.
pg_0011
VIJAYAN and TSOU
¡X Technical report on molecular phylogeny of
Camellia
187
Kobayashi, N., K. Tamura, and T. Aotsuka.1999. PCR error
and molecular population genetics. Biochem. Genetics 3:
318-321.
Kumar, S., K. Tamura, and M. Nei. 2004. MEGA3: Integrated
software for molecular evolutionary genetics analysis and
sequence alignment. Briefings in Bioinformatics 5: 150-163.
Kwiatowski, J., D. Skarecky, S. Hernandez, D. Pham, F. Quijas,
and F.J. Ayala. 1991. High fidelity of polymerase chain
reaction. Mol. Biol. Evol. 8: 884-887.
Kwok, S ., D.E. Kellog, N. Mc Kinney, D. Spas ic, L . Goda,
C. Levenson, and J.J. S ninsky. 1990. Effects of primer-
template mismatches on polymerase chain reaction: human
immunodeficiency virus type I model studies. Nucleic Acids
Res. 18: 999-1005.
Lapegue, S.D., M.H. Pemonge, L. Gielly, P. Taberlt, and J. Petit.
1999. Amplification of oak DNA from ancient and moern
wood. Mol. Ecol. 8: 2137-2140.
Lun dberg, K.S., D.D. Sh oem aker, M.W.W. Adam s, J . M.
Short, J .A. S orge, and E.J . Marhur. 1991. High-fidelity
am pli fi cat ion us ing a the rmos tabl e DNA polyme ras e
isolated from Pyrococcus furiosus. Gene 108: 1-6.
Marquez, L.M., D.J. Miller, J.B. MacKenzie, and M.J.H. van
Oppen. 2003. Ps eudoge nes contribute to the e xtrem e
divers ity of nuclea r ribos omal DNA in the ha rd c ora l
Acropa. Mol. Bio. Evol. 20: 1077-1086.
Matsumoto, S., A. Takeuchi, M. Hayatsu, and S. Kondo. 1994.
Molecular cloning of phenyle alanine ammonia lyase cDNA
and classification of varieties and cultivars of tea lantsusing
the tea PAL cDNA as a probe. Theor. Appl. Genet. 89:
671-675.
Merlo, D.J. and J.D. Kemp. 1976. Effect of polysaccharides on
kinetics of DNA. Plant Physiol. 58: 1522-1526.
Ming, T. 2000. Monograph of the genus Camellia. Kunming
Institute of Botany, Chinese Academy of Sciences, Yunnan
Science and Technology Press, Kunming, P. R. China.
Muri, G., C.C. Fleming, and C. Schlotterer. 2001. Three
divergent rDNA clusters predate the species divergence in
Quercus petraea (Matt.) Liebl. and Quercus robur L. Mol.
Biol. Evol. 18: 112-119.
Murray, M.G. and W.F. Thompson. 1980. Rapid is olation of
high-molecular-weight plant DNA. Nucle Acids Res. 8:
4321-4325.
Ochm an, H., A.S . Ge rber, a nd D.L. Hart l. 1988. Genet ic
appl ica tion of an i nvers e polym eras e c hain re act ion.
Genetics 120: 621-623.
Orel, G., A. Marchent, and G. Richards. 2003. Evolutionary
relationships of yellow-flowered Camellia species from
southeast Asia. Inter. Camellia J. 35: 88-96.
Paabo, S., D.M. Irwin, and A.C. Wilson. 1990. DNA damage
prom otes jumping between templates during enzym atic
amplification. J. Biol. Chem. 265: 4718-4721.
Posada, D. and K.A. Crandall. 1998. MODELTEST: testing the
model of DNA substitution. Bioinformatics 14: 817-818.
Prince, L.M. and C.R. Parks. 2001. Phylogenetic relationships
of Theaceae inferred from chloroplast DNA sequences data.
Amer. J. Bot. 88: 2309-2320.
Rogers, S.U. and A.J. Bendich. 1988. Extraction of DNA from
m illigram am ounts of fresh, herbarium and mummified
plant tissues. Plant. Mol. Biol. 5: 69-76.
Saiki, R.K., D.H. Geifand, S. Stoffei, S.J. Scharf, R. Higuchi,
G.T. Horn, K.B. Mulfis, and H.A. Erlich. 1988. P rimer-
di r e ct e d e n z ym a t i c a m pl i f ic a t i on o f DNA wi th a
thermostable DNA polymerase. Science 239: 487-491.
S ealy, J.R. 1958. A Revis ion of genus Camellia. The Royal
Horticultural Society, pp. 168-191.
S hiod a, M. and K. Mara kam i-Muofus hi. 198 7. Se lec tiv e
inhibition of DNA polymerase by a polysaccharide purified
from slime of Physarum polycephalum. Biochem. Biophys.
Res. Commun. 146: 61-66.
Singh, M., Bandana, and P.S. Ahuja. 1999. Isolation and PCR
amplification of Genomic DNA from market samples of dry
tea. Plant. Mol. Biol. Reptr. 17: 171-178.
Slater, M., S . S elman, B.L. Mogilevsky, H. Ammons, and J.
Hartnett. 1998. P fu DNA polymera se: A high fideli ty
enzyme for nucleic acid amplification. Promega Notes 68:
7-12.
Struwe, L., M. Thiv, J.Kadereit, T. Motley, A.S .R. Pe pper,
J. Rova, K. Potgieter, P.W hi te, a nd V.A. Albert. 1998.
Saccifolium, an endemic of S ierra de La Neblina on the
Brazillian-Venezuelan frontier is related to a temperate
alpine lineage of Gentianaceae. Harvard Papers in Bot. 3:
199-214.
Su, M.H. 2007. Taxonom ic study of Camellia formosens is
(Mas amune et S uzuki). In M. H. Su, C. F. Hsieh et C.
H. Ts ou (The aceae) (eds .), P hD dis serta tion, Ins titute
of Ecology and Evolutionary Biology Colleage of L ife
Science, National Taiwan University. Taiwan.
Swofford, D.L. 2001. PAUP*: Phylogenetic Analys is Us ing
Parsimony (*and Other Methods). Version 4.0b10. Sinauer
Associates, Sunderland, Massachusetts.
Tai, T.H. and S.D. Tanks ley. 1990. A rapid and inexpensive
method for isolation of total DNA from dehydrated plant
tissue. Plant. Mol. Biol. Reptr. 8: 297-303.
Takeuchi, A., S. Matsumoto, and M. Hayatsu. 1994. Chalcone
synthase from C. sinensis: solation of cDNA and the organ
specific and sugar responsive expression of genes. Plant
Cell Physiol. 35: 1011-1018.
Tam ura, K. and M. Ne i. 1993. E st im ati on of the numbe r
o f nuc le ot ide s ubs t it uti ons i n t he con trol re gio n of
mitochondrial DNA in humans and chimpanzees. Mol. Biol.
Evol. 10: 512-526.
Tang, S. and Y. Zhong. 2002. A phylogenetic analysis of nrDNA
ITS sequences from S er. Chrysantha (Sect. Chrysantha,
Camellia, Theaceae). J. Genet. Mol. Biol. 103: 105-107.
Thakor, B.H. and C.R. Parks. 1997. Phylogenetic relationships
within the genus Camellia (Theace ae). S upplement to
Amer. J. Bot. 84: 237 (Abstract only).
Tindall, K.R. and T.A. Kunkel. 1988. Fidelity of DNA synthesis
pg_0012
188
Botanical Studies, Vol. 49, 2008
by the Thermus aquaticuc DNA polymerase. Biochemistry
27: 6008-6013.
Triglia, T., M.G. Peterson, and D.J. Kemp. 1988. A procedure
for in vitro amplification of DNA segments that lay out side
the boundaries of known sequences. Nucleic Acids Res. 16:
8186.
Urbat sch, L.E ., B.G. B aldwin, a nd M.J . Donoghue. 2000.
P hylogeny of cone flowers and relative s (Helianthease:
Ateraceae) based on nuclear DNA internal trans cribed
spacer (ITS) sequences and chloroplast DNA restriction site
data. Syst. Bot. 25: 1153-565.
Wachira, F.N., R. Waugh, C.A.Hackett, and W. Powell. 1995.
Detection of genetic diversity in tea using RAPD markers.
Genome 38: 201-210.
White, T.J., T. Bruns, S. Lee, and J. Taylor. 1990. Amplification
and direct sequencing of fungal ribosomal RNA genes for
phylogenetics. In M.A. Innis, D.H. Gelfand, J.J. Sninsky,
and T.J. White (eds.), PCR Protocols: a Guide to Methods
and Applications. Academic Press, London, pp. 315-322.
Xiao, T.J. 2001. Phylogenetic relationships of the genus
Camellia (Theaceae) bas ed on the RNA polyme ras e II
(RPB2) sequences. PhD dissertation, Univ. North Carolina,
Chapel Hill.
Xiao, T.J. and C.R. Parks. 2003. Molecular analysis of the genus
Camellia. Inter. Camellia J. 35: 57-65.
Yang, J.B., H.T. Li, S.X. Yang, D.Z. Li, and Y.Y. Yang. 2006.
T he a ppl ica ti on of fou r DNA se que nce s to s tudy ing
molecular phylogeny of Camellia (Theaceae). Acta Bot.
Yunn. 28: 108-114.