ArticlePDF Available

Starch Gel Electrophoresis of Ferns: A Compilation of Grinding Buffers, Gel and Electrode Buffers, and Staining Schedules

Authors:
American Fern Society
is collaborating with JSTOR to digitize, preserve and extend access to
American Fern Journal.
http://www.jstor.org
American Fern Society
Starch Gel Electrophoresis of Ferns: A Compilation of Grinding Buffers, Gel and Electrode
Buffers, and Staining Schedules
Author(s): Douglas E. Soltis, Christopher H. Haufler, David C. Darrow and Gerald J. Gastony
Source:
American Fern Journal,
Vol. 73, No. 1 (Jan. - Mar., 1983), pp. 9-27
Published by: American Fern Society
Stable URL: http://www.jstor.org/stable/1546611
Accessed: 05-06-2015 16:02 UTC
REFERENCES
Linked references are available on JSTOR for this article:
http://www.jstor.org/stable/1546611?seq=1&cid=pdf-reference#references_tab_contents
You may need to log in to JSTOR to access the linked references.
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at http://www.jstor.org/page/
info/about/policies/terms.jsp
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content
in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship.
For more information about JSTOR, please contact support@jstor.org.
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN FERN JOURNAL:
VOLUME
73 NUMBER
1 (1983)
Starch Gel Electrophoresis of Ferns:
A Compilation of Grinding Buffers,
Gel and Electrode Buffers, and Staining Schedules
DOUGLAS E. SOLTIS*, CHRISTOPHER H. HAUFLER**,
DAVID C. DARROW***, and GERALD J. GASTONY***
The homosporous pteridophytes have been largely uninvestigated by electrophore-
sis, despite the fact that they offer many exciting research possibilities (Soltis et al.,
1980). The paucity of electrophoretic studies of ferns and fern allies may be due in
large part to the high concentrations of condensed tannins that many species contain
(Cooper-Driver, 1976 and pers. comm.). These compounds render enzymes inactive
by binding with them following cellular disruption, thereby frustrating researchers
who have attempted electrophoretic analysis utilizing standard methods of sample
preparation.
The method of sample preparation developed by Kelley and Adams (1977a, b) in
their analysis of enzyme variation in Juniperus was an important procedural
breakthrough in overcoming the difficulties that result from the liberation of large
amounts of phenolic compounds during tissue preparation. Recently, a simplified
version of that method was applied by Soltis et al. (1980) to fern leaf tissue,
facilitating rapid preparation
of active enzyme samples and thereby making electro-
phoretic analyses of large numbers of individuals more feasible.
In an attempt to improve methods of analysis of fern enzymes in starch gel
electrophoresis, we have experimented with modifications of the method of sample
preparation
outlined by Soltis et al. (1980). We also have examined several different
methods of sample preparation such as those of Gottlieb (1981a), Mitton et al.
(1979), and Werth et al. (1982), and have evaluated the relative merits of each with
fern tissue. Finally, during the course of our electrophoretic investigations of ferns
we found that standard gel and electrode buffers and staining schedules, such as
those of Brewer (1970) and Shaw and Prasad (1970), often provided unsatisfactory
results when applied to ferns. We have determined gel and electrode buffers, as well
as staining schedules, that provide clear starch gel enzyme banding for 22 enzyme
systems in ferns. Requests for advice resulting from the recent surge of interest in
fern enzyme electrophoresis have prompted
us to compile our procedural
data so that
other researchers can take advantage
of our experimentation. We hope that these data
will stimulate more extensive electrophoretic investigation
of pteridophytes
and other
electrophoretically difficult taxa.
Gottlieb (1981b) recently reviewed aspects of enzyme electrophoresis primarily in
gymnosperms and angiosperms. His discussion is equally relevant to understanding
the potential applications and limitations of electrophoretic evidence in pterido-
phytes. Since homosporous pteridophytes have high chromosome numbers, it is
tempting to invoke polyploidy in interpreting
their enzyme band patterns. It is well
*Department of Biology, University of North Carolina, Greensboro, NC 27412.
**Department of Botany, University of Kansas, Lawrence, KS 66045.
***Department of Biology, Indiana University, Bloomington, IN 47405.
9
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
TABLE 1. ELECTRODE AND GEL BUFFER RECIPES USED SUCCESSFULLY IN ELECTROPHORETIC ANALYSIS OF FERNS (Gram
amounts are given for one liter final volume of buffer except where noted).
Electrode Buffer
1. 0.400 M Citric Acid, trisodium salt;
1-17.64 g Citric Acid, trisodium salt
dihydrate, 1.0 M HCI to pH 7.0
2. 0.135 M Tris, 0.032 M Citric Acid;
16.35 g Trisa, 6.10 g Citric Acidb
3. 0.135 M Tris, 0.017 M Citric Acid;
16.35 g Trisa, 3.35 g Citric Acidb
4. 0.223 M Tris, 0.086 M Citric Acid;
27.00 g Trisa, 16.52 g Citric Acidb,
NaOH to pH 7.5
5. 0.223 M Tris, 0.069 M Citric Acid;
27.00 g Trisa, 13.33 g Citric Acidb
6. 0.100 M NaOH, 0.300 M Boric Acid;
4.00 g NaOH, 18.55 g Boric Acidb
7. 0.038 M LiOH, 0.188 M Boric Acid;
1.60 g LiOH-H20, 11.60 g Boric Acidb
Adjust to pH 8.3 with dry components
pH
(a 22?C
7.0
8.0
8.5
7.5
7.2
8.6
8.3
Gel Buffer
0.020 M Histidine-HCl; 4.19 g L-Histidine-HCl
monohydrate, 1.0 M NaOH to pH 7.0
0.009 M Tris, 0.002 M Citric Acid;
dilute 67 ml of electrode buffer to I liter
0.009 M Tris, 0.001 M Citric Acid;
dilute 67 ml of electrode buffer to 1 liter
0.008 M Tris, 0.003 M Citric Acid;
dilute 35 ml of electrode buffer to I liter
0.008 M Tris, 0.002 M Citric Acid;
dilute 35 ml of electrode buffer to I liter
0.015 M Tris, 0.004 M Citric Acid;
1.84 g Trisa, 0.69 g Citric Acidb
0.045 M Tris, 0.007 M Citric Acid,
0.004 M LiOH, 0.019 M Boric Acid;
Tris-citrate buffer (5.45 g Trisa,
1.28 g Citric Acidb, bring volume to
900 ml), add 100 ml electrode buffer
to give 9:1 ratio, 1.0 M NaOH to pH 8.3
pH
(a 22?C
7.0
8.0C
8.5c
7.5'
m
7.2' g
m
7.8 o
2C
.
8.3 ;
r-
ci
m
Oo
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
8. 0.039 M LiOH, 0.263 M Boric Acid;
1.64 g LiOH-H20, 16.23 g Boric Acidb
9. 0.065 M L-Histidine free base, ca.
0.015 to 0.016 M Citric Acid;
10.09 g L-Histidine free base, Citric Acidb to
pH 5.7 (= ca. 2.9 to 3.1 g)
10. 0.180 M Tris, 0.004 M EDTA, 0.100 M
Boric Acid; 21.80 g Trisa, 1.52 g
EDTA tetrasodium salt dihydrate,
6.18 g Boric Acidb, Boric Acid to
pH 8.6
11. 0.400 M Citric Acid, trisodium salt;
117.64 g Citric Acid, trisodium salt
dihydrate, 1.0 M HC1 to pH 7.0
8.0
5.7
8.6
7.0
0.042 M Tris, 0.007 M Citric Acid;
0.004 M LiOH, 0.025 M Boric acid;
5.04 g Trisa, 1.25 g Citric Acidb,
0.16 g LiOHH2O, 1.56 g Boric Acidb,
1.0 M HC1 to pH 7.6
0.009 M L-Histidine, 0.002 M Citric Acid;
dilute 140 ml of electrode buffer to 1 liter
0.045 M Tris, 0.001 M EDTA, 0.025 M Boric Acid;
dilute 250 ml of electrode buffer to 1 liter
0.005 M Histidine-HCl; 1.05 g L-Histidine-HCl
monohydrate, 1.0 M NaOH to pH 7.0
aTrizma-base (Sigma T1503; Sigma T1378 works as well and is much less expensive).
bAmounts of Citric Acid and Boric Acid are given in grams of anhydrous free acid per liter.
CCheck
pH of solution after mixing; pH of Tris buffers will change with dilution.
Buffer systems 1 and 11 are from Gottlieb (198 la).
Buffer systems 2 and 3 are modifications of buffer system I of Shaw and Prasad (1970).
Buffer systems 4 and 5 are modifications of buffer system XII of Shaw and Prasad (1970).
Buffer system 6 is from Mitton et al. (1977).
Buffer system 7 is a modification of the buffer system of Gottlieb (1981a).
Buffer system 8 is a modification of the buffer system of Adams and Joly (1980).
Buffer systems 9 and 10 are from Gottlieb (pers. comm.).
7.6 a
C,,
m
7.6 m
-
r-
I
m
C)
m
C)
7.0
C/,
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN FERN
JOURNAL:
VOLUME
73 (1983)
known, however, that multiple forms of a given enzyme may be coded by different
alleles at a single locus (allozymes) or by genes at more than one locus (isozymes).
Furthermore,
many standardly assayed plant enzymes have isozymes located in two
or more subcellular compartments (e.g., the cytosolic and chloroplastic isozymes of
PGI and PGM discussed by Gottlieb, 1981b, 1982-for interpretation
of enzyme
symbols see Table 2). When found in pteridophytes, multiple, subcellularly com-
partmentalized
isozymes should not be misinterpreted
as products of duplicated loci
resulting from polyploidy.
According to Chapman et al. (1979), attempts to determine the amount of
heterozygosity per locus in homosporous pteridophytes
are complicated by recombi-
nation between homoeologous duplicated loci. An obvious prerequisite
to recombina-
tion of this kind is the actual presence of such loci. Gastony and Gottlieb (1982)
developed a means of demonstrating whether duplicated loci are, in fact, present in
pteridophytes. By electrophoretically analyzing sporophytes taken from nature and
individual gametophytes grown from spores of these sporophytes, segregational
analysis of sporophytic enzyme banding patterns can be conducted. This permits
genetic interpretation
of parental sporophytic enzyme phenotypes without the time-
consuming crossing programs required by total reliance on sporophytic tissue. Use
of this methodology enables the investigator to determine whether apparent
heterozygosity of sporophytes is coded by alleles at a single locus or by genes at
duplicated (homoeologous) loci. Our application of this methodology to several fern
genera (e.g., Athyrium, Bommeria, and Pellaea) has demonstrated that the genetic
variability observed in these taxa results from allelic diversity and segregation at
single, not duplicated, loci.
MATERIALS
Living sporophytes of Athyrium filix-femina, Bommeria ehrenbergiana,
B. hispida,
B. pedata, B. subpaleacea, Botrychium virginianum, Ceratopteris thalictroides,
Cystopteris bulbifera, C. dickieana, C. fragilis, C. laurentiana, C. protrusa, C.
reevesiana, C. tennesseensis, C. tenuis, Isoetes butleri, I. engelmannii, Lvgodium
japonicum, Nephrolepis exaltata, Ophioglossum engelmannii, Pellaea andromedi-
folia, P. atropurpurea, P. glabella, Polypodium polypodioides, P. virginianum,
Polystichum acrostichoides, Pteridium aquilinum, Woodsia obtusa, and W. oregana
were maintained in greenhouse culture and utilized in this investigation. Living
gametophytes representing the species of Athyrium,
Bommeria, Cystopteris, Pellaea
and Pteridium listed above were cultured on nutrient
agar (Gastony & Haufler, 1976)
and also provided material for this study.
GRINDING BUFFER SOLUTIONS
We routinely utilize modifications of either the phosphate grinding buffer-poly-
vinylpyrrolidone (PVP) solution employed by Mitton et al. (1979), the Tris-maleate
grinding buffer-PVP solution of Soltis et al. (1980), or the Tris-HCl grinding
buffer-PVP solution of Gottlieb (1981a) with PVP substituted for polyvinylpoly-
pyrrolidone (PVPP), as described below. Recipes for preparing
these grinding buffer
solutions are provided below; molarity or percent volume values are provided and
12
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D.
E.
SOLTIS
ET
AL.: STARCH
GEL ELECTROPHORESIS
OF FERNS
gram
or milliliter amounts
required
to prepare
25 ml of buffer
solution are given in
parentheses.
Phosphate grinding buffer-PVP solution.-0.029 M (0.28 g) sodium tetraborate,
0.017 M (0.08 g)
sodium metabisulfite, 0.20 M (1.0 g) L-ascorbic acid sodium salt, 0.016 M (0.07 g) diethyldithiocar-
bamic acid sodium salt, 4% w/v (1.0 g) PVP average molecular weight 40,000 (Sigma PVP 40T), or
36-40% w/v (9.0-10.0 g) PVP average molecular weight 10,000. Dissolve gram amounts in 25 ml of
0.10 M phosphate buffer pH 7.5 (to make 100 ml of phosphate buffer dissolve 1.36 g KH,PO4 in H,O.
add 9.0 ml 1M NaOH, and bring volume to 100 ml with H20) and then add 0.25 ml (1%)
2-mercaptoethanol.
Tris-maleate grinding buffer-PVP solution.-0.20 M (1.91 g) sodium tetraborate, 0.02 M (0.095
g) sodium metabisulfite, 0.25 M (1.24 g) L-ascorbic acid sodium salt. 0.026 M (0.113 g)
diethyldithiocarbamic acid sodium salt, 0.10 M (0.29 g) maleic acid, 0.10 M (0.30 g) Tris, 4% w/v (1.0
g) PVP average molecular weight 40,000 or 32-40% w/v (8.0-10.0 g) PVP average molecular weight
10,000. For 25 ml of buffer, dissolve amounts indicated in 19 ml distilled water; mix thoroughly and
crush out lumps; adjust to pH 7.5 with 1.0 M HCI; add 0.025 ml (0.1%) 2-mercaptoethanol;
add H2O to
25 ml. Originally (when we utilized PVP 10,000) it was necessary to allow the PVP to hydrate
overnight
before using the grinding buffer solution. When employing PVP 40,000, it is possible to prepare the
grinding buffer solution immediately prior to sample preparation
by stirring the PVP into solution.
Tris-HCI grinding buffer-PVP solution.-0.1% v/v (0.025 ml) 2-mercaptoethanol, 0.001 M
(0.010 g) EDTA (tetrasodium salt), 0.010 M (0.019 g) potassium chloride. 0.010 M (0.050 g)
magnesium chloride hexahydrate, 4 or 20% w/v (1 or 5 g) PVP 40,000, 25 ml 0.10 M Tris-HCl buffer,
pH 7.5. Stir the PVP into solution or allow it to hydrate in the buffer overnight.
TABLE 2. GEL AND ELECTRODE BUFFER SYSTEMS THAT WE HAVE FOUND TO YIELD THE
BEST BANDING IN FERNS WE HAVE ASSAYED.
Enzyme Symbol Gel and Electrode Buffer System
(from Table 1)
Acid phosphatase APH 6, 7
Aconitase ACN 1, 5, 9
Aldolase ALD 1-5, 7, 10, 11
Aspartate aminotransferase AAT (or GOT) 6, 7, 8
Catalase CAT 4, 7, 8, 10
Esterase (Colorimetric) EST 6, 7
Esterase (Fluorescent) FE 8
Fructose-
1,6-diphosphatase F1,6DP 1, 11
Glucose-6-phosphate dehydrogenase G6PDH 4, 5, 6, 7
Glutamate dehydrogenase GDH 3, 5, 7
Glyceraldehyde-3-phosphate
dehydrogenase G3PDH 1, 11
Hexokinase HK 2, 3, 5, 11
Isocitrate dehydrogenase IDH 1, 2, 3, 4, 5
Leucine aminopeptidase LAP 3, 7, 8
Malate dehydrogenase MDH 1, 4, 5, 9
Malic enzyme ME 7, 10
Peroxidase PER 2, 7
Phosphoglucoisomerase PGI 5, 6, 7
Phosphoglucomutase PGM 3, 5, 6, 7, 8
6-Phosphogluconate dehydrogenase 6-PGD 1, 2, 4, 5
Shikimate dehydrogenase SkDH 1, 2, 4, 5, 11
Triosephosphate isomerase TPI 2, 5, 6, 7, 8
13
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN FERN
JOURNAL:
VOLUME 73 (1983)
Comparison of these three grinding buffer solutions indicates that for most
enzymes the staining results are highly comparable. For some enzymes, however,
the Tris-HCl-PVP solution seems to improve enzyme band clarity (e.g., PGI,
PGM, LAP), while for other enzymes the reverse is true. In Pellaea andromedifblia,
for example, PGI banding was sharp with the Tris-HCI-PVP solution but was
inhibited by the presence of ascorbic acid and sodium tetraborate in the
Tris-maleate-PVP buffer. In Athyrium filix-femina, however, the Tris-maleate-PVP
solution gives superior banding for PGI and one more observable EST locus when
compared to the Tris-HCl-PVP solution. The phosphate-PVP solution often result-
ed in reduced APH activity when compared to the other two grinding buffer
solutions.
In working with species of Asplenium, Werth et al. (1982) used a method of
sample preparation
that uses caffeine, but not PVP. Following this technique, equal
weights of tissue and caffeine were ground in a 0.1 M HEPES pH 7.0 buffer with
0.2% 2-mercaptoethanol
and 0.5% sodium metabisulfite to produce
a slurry.
We found
that for virtually all enzyme systems investigated, the method of Werth et al.
provided results roughly comparable to those obtained with grinding buffers contain-
ing PVP. The presence of 0.5% sodium metabisulfite in the grinding buffer appears
to be of great importance. When this ingredient is omitted from the buffer, activity
is noticeably reduced or almost totally lost for many enzymes. Significantly, the full
complement of PGI enzyme bands obtained with PVP was not expressed in most of
the fern taxa investigated when sodium metabifsulfite was omitted. These results are
in agreement with similar observations of Werth
et al. (1982).
It should be emphasized that to obtain the best results, the appropriate
amount of
PVP depends upon the taxon under investigation and in part upon the molecular
weight of the PVP employed. Soltis et al. (1980) utilized a grade of PVP with an
average molecular weight of 10,000 (although this is not stated in their report) and
incorporated
40% w/v PVP (4 g PVP in 10 ml of Tris-maleate grinding buffer). At
present, we routinely employ a grade of PVP with an average molecular weight of
40,000. Both molecular weights of PVP are suitable, but less PVP 40,000 is
required than PVP 10,000. The amount of PVP used is an important
consideration
because, as reported by Soltis et al. (1980), use of excessive amounts of PVP in the
preparation of grinding buffer-PVP solutions frequently results in a decrease or
complete loss of enzyme activity. Although PVPP is effective with a wide range of
angiosperms, its substitution for PVP failed to produce banding for PGI in Pellaea
andromedifblia whether the PVPP was hydrated in the grinding buffer or added
directly to the leaf tissue during grinding.
This discussion stresses the importance of selecting optimal grinding buffer
components and procedures for the taxon under investigation. As noted above,
certain components of these buffers may inhibit band expression, whereas in other
cases the full complement of components may be required to obtain clear banding.
We have found that LAP activity is greater if sodium ascorbate, sodium tetraborate,
and sodium metabisulfite are eliminated from the grinding buffer. On the other
hand, activity for some enzymes, (e.g., SkDH, G6PDH, MDH, CAT, 6PGD) may
be very noticeably reduced when a buffer solution incorporating only PVP is
14
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D.
E.
SOLTIS ET
AL.:
STARCH
GEL ELECTROPHORESIS
OF FERNS
utilized. A further complication in some taxa is that when a grinding buffer lacking
2-mercaptoethanol is employed, artifactual ("ghost") bands proliferate for some
enzymes and can hinder accurate interpretation
of the band patterns. Given these
examples, we strongly encourage comparative experimentation with grinding buffer
solutions when initiating electrophoretic studies.
SAMPLE PREPARATION AND ELECTROPHORESIS
Leaf samples to be analyzed electrophoretically were taken from mature sporo-
phytes. The methods discussed herein, however, are applicable to gametophytic as
well as sporophytic tissue. In our original electrophoretic analyses of ferns, leaf
samples were prepared according to the method of Soltis et al. (1980) in which
small amounts of leaf tissue are ground under liquid nitrogen with a porcelain
mortar and pestle until a fine dry powder is obtained. The powder is quickly mixed
with grinding buffer to form a thick slurry. This slurry is absorbed directly into thick
paper wicks (made from Whatmann 3 MM chromatography paper or some other
suitable wick paper).
It should be emphasized that the use of the grinding buffer-PVP solutions alone
(outlined above), without the use of liquid nitrogen in the preparation
of samples,
provide clear enzyme banding in all of the fern taxa that we have examined so far.
Therefore, at present we utilize a simplified method of sample preparation
in which
liquid nitrogen has been eliminated. In standardizing our grinding procedure by
consistently homogenizing 100 mg of leaf material in 0.5 ml of grinding buffer, we
obtained roughly equivalent staining intensity for all samples, which facilitates
comparing individuals and scoring gels. If small amounts of tissue are being ground,
it may be possible to substitute 43 mm2 plastic weigh boats for mortars and glass or
plexiglass rods for pestles. As discussed by Gastony and Gottlieb (1982), it is
possible to obtain enzyme banding from single gametophytes. Their technique
involves smearing individual gametophytes onto wicks pre-moistened with extraction
buffer-PVP solution. With all of the described grinding/extraction procedures, the
amount of grinding buffer and the type and size of wicks should be tailored to the
taxon and tissues under investigation. Once the plant material has been absorbed
into wicks, the wicks are inserted into a vertical slit in the starch gel and subjected
to horizontal electrophoresis at 4?C. Band definition may be improved
by removing
wicks from the gel after the first 10-20 minutes of electrophoresis. It is important
to
maintain electrical contact across the slit in the gel. We routinely place some sort of
spacer (e.g., strips of Whatmann paper, a plastic straw, a thin piece of plexiglass) at
the cathodal end of the gel to compress the gel slightly and to insure that the slit
does not open.
To permit comparison of banding patterns between gels, it is important to
maintain a constant starch concentration throughout a study. We have experimented
with a variety of starch concentrations throughout
the practical range of 11.5 to 15%
and have found concentrations of 12 to 13.2% most suitable, although in some
instances alternative concentrations may improve band resolution. Connaught, Fish-
er, and Sigma starch have been used and all provide clear enzyme banding, provided
appropriate gel and electrode buffer systems and staining schedules are utilized.
15
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN FERN JOURNAL:
VOLUME
73 (1983)
In order to facilitate determining the location of the anodal front, it is useful
to insert a wick bearing the marker dye Bromphenol Blue (Sulfone Form, Sigma
B0126; 0.04% w/v in 95% ethanol). Depending on the gel and electrode buffer
system employed, as well as the taxon under investigation, the enzyme bands will
migrate varying distances behind the Bromphenol Blue marker. Optimal running
times for each enzyme system will therefore have to be determined empirically.
GEL AND ELECTRODE BUFFER RECIPES
We have found that standard gel and electrode buffer systems, such as those of
Shaw and Prasad (1970), and Brewer (1970), often yield unsatisfactory
results with
fern tissue. We have determined gel and electrode buffer systems that provide clear
banding for 22 enzymes in ferns. Recipes for the gel and electrode buffers we most
commonly employ are provided in Table 1.
Several of the gel and electrode buffer systems that we employ are modifications
of those of Shaw and Prasad (1970). It should be noted, therefore, that errors are
present in several of the recipes they provided. For example, in Shaw and Prasad
buffer system I (see page 299, Table 1 of their report), utilization of 16.35 g of Tris
yields an electrode buffer with a molarity of 0.135 (see present report, Table 1,
electrode buffers 2 and 3), rather than 0.155 as reported by Shaw and Prasad.
Similarly, use of 27.0 g of Tris in the preparation
of the electrode buffer of Shaw
and Prasad system XII yields a molarity of not 0.233 but 0.223 (see present report,
Table 1, electrode buffers 4 and 5). Furthermore,
use of the electrode buffer recipe
of Shaw and Prasad system XII yields a buffer of approximately pH 5.5, which
differs significantly from the estimate of pH 7.0 listed by Shaw and Prasad. Use of
the amounts of Tris and citric acid given for system 4 in Table I of the present report
will also yield an electrode buffer of pH 5.5; sodium hydroxide is then added to
adjust the buffer to pH 7.5.
Of the 11 gel and electrode buffer systems for which recipes are provided in Table
1, some clearly are preferable to others for certain of the 22 enzymes for which we
have obtained sharp banding. For example, best enzyme banding for MDH is
obtained when system 1, 4, 5, or 9 is used. Which system works best is dependent,
in part, on the taxon under investigation. We have found, for example, that the best
systems for Athyrium and Cystopteris differ somewhat from those that supply clear
enzyme banding for Bommeria. The systems provided here, however, should provide
an excellent framework for conducting electrophoretic analyses of most ferns. The
gel and electrode buffer systems that we have found to yield the clearest banding for
each of the 22 enzymes with which we have experimented are listed in Table 2.
STANDARD STAINING SCHEDULES
During the course of our electrophoretic investigations of ferns, we found that
many published staining schedules, such as those of Shaw and Prasad (1970),
yielded unsatisfactory results. Enzyme banding often can be improved dramatically
when these standard
recipes are modified. One very important
modification that we
employ involves increasing the pH of the buffer in staining schedules for ALD,
GDH, G6PDH, HK, IDH, MDH, ME, PGI, PGM, and 6PGD. Tris-HCl staining
buffers of pH 7.0 or 7.1 typically are used in staining for these enzymes (Shaw &
16
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D.
E.
SOLTIS ET AL.:
STARCH GEL
ELECTROPHORESIS OF FERNS
Prasad, 1970). We have found, however, that in ferns the pH optima for these
enzymes are in the 8.0-8.5 range, and we therefore use a Tris-HCl staining buffer
of pH 8.0-8.5 for them. For some enzymes, this simple modification improves
staining dramatically. For example, standard
MDH staining recipes use a buffer of
pH 7.0. Application of such a protocol to ferns, however, often results in very little
or inconsistent staining. Well stained bands are obtained consistently for MDH when
a Tris-HCl staining buffer of pH 8.0-8.5 is used (provided an appropriate
gel and
electrode buffer system is employed; see Table 2).
It also should be noted that we have consistently been unable to obtain observable
activity with fern leaf tissue for some enzymes, despite considerable experimenta-
tion. These include Alcohol dehydrogenase, Alkaline phosphatase, o-Glycerophos-
phate dehydrogenase and Lactate dehydrogenase. There are several enzymes for
which we have obtained observable activity with ferns, but have not yet obtained
clear banding, such as Diaphorase and Peptidase, and these are therefore not
included in this report. In addition, there are a number of enzyme staining schedules
that we have not yet attempted. We hope that the staining schedules provided below
will stimulate additional experimentation in this regard.
The staining schedules that we utilize for the 22 enzymes for which we obtain
clear banding are provided below in alphabetical order. The final volume for all stain
recipes is 100 ml. Depending on gel size and staining container capacity, it may be
possible to assay two gel slices simultaneously with each of the following 22
solutions. Alternatively, when assaying single gel slices, the recipes may be halved
to 50 ml final volume. Readers are encouraged to consult Enzyme Nomenclature
1978 (International Union of Biochemistry, Nomenclature Committee, 1979 Ihere-
after referred to as I.U.B., 1979]) for details of enzyme specificity and Gottlieb
(1981b) for a recent review of the electrophoretic technique and its application to
plant populations. It should be noted that enzymes such as peroxidases, esterases,
and acid phosphatases operate on a large number of substrates in vitro; therefore,
standard staining protocols for these enzymes result in the staining of various
numbers of isozymes whose homologies are not apparent. Although the utility of
these isozymes in between-species comparisons is therefore limited, their banding
patterns can be useful in assessing variability within populations.
Several different organisms serve as the source for commercial preparations of
Glucose-6-phosphate dehydrogenase. This enzyme typically is specific for NADP,
but some forms are capable of utilizing either NADP or NAD. It should be
emphasized that NADP is much more expensive than NAD, hence selecting a
Glucose-6-phosphate dehydrogenase type that can use NAD will save considerable
money. Therefore, in those recipes that call for added Glucose-6-phosphate dehydro-
genase, the notation NAD(P)* reflects this option.
Acid phosphatase (EC 3.1.3.2)
0.05 M sodium acetate buffer, pH 5.0 100 ml
1.0 M MgC12 0.5 ml
o-naphthyl acid phosphate, sodium salt 100 mg
Fast garnet GBC salt 80 mg
We have obtained results with buffer molarities ranging from 0.05 M to 0.2 M
and within a pH range of 5.0 to 6.0. Stain at room temperature;
a modification of
17
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN FERN
JOURNAL:
VOLUME 73 (1983)
Scandalios (1969). Gottlieb (1981b) noted that plants may have a dozen or more
acid phosphatases. We consistently have observed one very intense zone of acid
phosphatase activity, with several fainter bands occasionally evident.
Aconitase (EC 4.2.1.3.; = Aconitate hydratase
in I.U.B., 1979)
1.0 M Tris-HC1
buffer, pH 8.5 10 ml
H0O 90 ml
cis-aconitic acid 70 mg
Isocitrate ( = Isocitric) dehydrogenase 7 units
1.0 M MgCI2 1 ml
NADP 10 mg
MTT 5 mg
PMS 1 mg
Stain in the dark at 30?C; a modification of Shaw and Prasad (1970).
Aldolase (EC 4.1.2.13; = Fructose-biphosphate
aldolase in I.U.B., 1979)
1.0 M Tris-HCl buffer, pH 8.5 10 ml
H,O 90 ml
Fructose-
1,6-diphosphate,
tetra(cyclohexylammonium)
salt 500 mg
or
trisodium salt 200 mg
1.0 M arsenic acid, sodium salt 1 ml
Glyceraldehyde-3-phosphate dehydrogenase 200 units
NAD 20 mg
MTT 20 mg
PMS 5 mg
Stain at room temperature;
a modification of Shaw and Prasad (1970).
Aspartate aminotransferase (EC 2.6.1.1.; = Glutamate oxaloacetate trans-
aminase)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H20 90 ml
L-aspartic acid 100 mg
cx-ketoglutaric
acid 100 mg
Adjust pH to 8.0 with 1.0 M NaOH as necessary, then add:
Pyridoxal-5'-phosphate 5 mg
Fast
blue BB salt 100 mg
Stain in the dark at room temperature; a modification of Gottlieb (1973a) and
Selander et al. (1971, appendix). Our results show that this enzyme is very sensitive
to the pH of the staining buffer; very little staining is observed at pH 7.0 or at pH
9.0. The number of loci coding for proteins that are capable of this aminotransferase
reaction is known to vary in plants. Although we typically encounter a single locus,
Gottlieb (1981b) has reported the likelihood of specific subcellular localizations for
the several dimeric isozymes.
Catalase (EC 1.11.1.6)
3% H202 5 ml
0.1 M phosphate buffer, pH 7.0 10 ml
0.06 M Na2S203-5H2O 7 ml
H20 78 ml
18
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
19
D.
E.
SOLTIS
ET AL.:
STARCH GEL ELECTROPHORESIS OF FERNS
Incubate in this solution at room temperature for 1-2(30) minutes; pour off the
solution, rinse several times with distilled water, then add:
0.09 M KI 50 ml
H2O 50 ml
Catalase activity will appear as white bands on a dark blue background; a
modification of Shaw and Prasad (1970). For some of the buffer systems in Table 1,
it may be necessary to add several drops (approximately 2 ml) of glacial acetic acid
in order to induce staining. An alternative method is to substitute 50 ml 0.05 M
sodium acetate buffer (pH 5.0) for the 50 ml H20. The gel may turn completely blue
in a very short time. Be prepared either to photograph the gel or score it while
staining. One locus has been observed in ferns. The protein is reported to be
tetrameric (Scandalios, 1969).
Esterase (Colorimetric; EC 3.1.1.-)
a-naphthyl acetate 40 mg
B-naphthyl
acetate 40 mg
dissolved in acetone 2 ml
1.0 M phosphate buffer, pH 6.0 10 ml
H20 90 ml
Fast blue RR salt 100 mg
Stain at room temperature;
a modification of Gottlieb (1974).
Esterase (Fluorescent; EC 3.1.1.-)
4-methylumbelliferyl acetate 42 mg
dissolved
in acetone 25 ml
1.0 M sodium acetate
buffer,
pH 5.0 18 ml
H20 57 ml
Stain in the dark at room temperature and observe under long-wave ultraviolet
light; staining schedule of Mitton et al. (1979). Observe immediately after staining
(bands fade quickly). Caution should be exercised in scoring gels because flavonoid
"bands" may also be visualized under UV light. To determine whether flavonoid
"bands" are present, the gel slice should be observed under UV light prior to
staining.
Fructose-1,6-diphosphatase (EC 3.1.3.11; = Fructose-biphosphatase in I.U.B.,
1979)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H20 90 ml
Fructose-
1,6-diphosphate,
tetra(cyclohexylammonium) salt 250 mg
or
trisodium salt 100 mg
1.0 M MgC12 1 ml
Phosphoglucoisomerase 50 units
Glucose-6-phosphate dehydrogenase 50 units
NAD(P)* 10 mg
MTT 5 mg
PMS 1 mg
Stain in the dark at 30? C; a modification of Shaw and Prasad (1970).
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN
FERN
JOURNAL:
VOLUME
73
(1983)
Glucose-6-phosphate
dehydrogenase
(EC 1.1.1.49)
1.0 M Tris-HCl buffer, pH 8.0 or 8.5 10 ml
H,O 90 ml
Glucose-6-phosphate, disodium salt 100 mg
NADP 20 mg
MTT (or NBT) 10 mg
PMS 2 mg
Stain in the dark at 37? C; a modification of Shaw and Prasad (1970).
Glutamate
dehydrogenase
(EC 1.4.1.2)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H,O 70 ml
1.0 M L-glutamic acid, pH 8.0 (use free acid
or monosodium salt and add NaOH to pH 8.0) 20 ml
NAD 20 mg
MTT (or NBT) 10 mg
PMS 2 mg
Stain in the dark at room temperature;
a modification of Gottlieb (1973b) and of
Shaw and Prasad (1970).
Glyceraldehyde-3-phosphate
dehydrogenase
(EC 1.2.1.12; = Glyceraldehyde-
phosphate dehydrogenase
in I.U.B., 1979)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H,O 90 ml
Fructose-1,6-diphosphate, trisodium salt 100 mg
Aldolase 10 units
Incubate above mixture approx. 30 min. at 30-37? C, then add:
1.0 M arsenic acid, sodium salt 1 ml
NAD 10 mg
MTT 5 mg
PMS 1 mg
Stain in the dark at 30? C; a modification of Shaw and Prasad (1970).
Hexokinase
(EC 2.7.1.1)
1.0 M Tris-HCl buffer, pH 8.5 10 ml
H,O 90 ml
Glucose 90 mg
1.0 M MgC1, 5 ml
EDTA, tetrasodium salt, dihydrate 40 mg
NAD(P)* 10 mg
MTT (or NBT) 15 mg
Glucose-6-phosphate dehydrogenase 40 units
PMS 2 mg
ATP 25 mg
Stain in the dark at room temperature;
a modification of Shaw and Prasad (1970).
Isocitrate dehydrogenase
[NADP+] (EC 1.1.1.42)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H,O 85 ml
Isocitric acid, trisodium salt 100 mg
1.0 M MgCl, 5 ml
NADP 10 mg
20
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D. E.
SOLTIS
ET
AL.: STARCH GEL ELECTROPHORESIS
OF FERNS
MTT (or NBT) 15 mg
PMS 2 mg
Depending on the taxon being analyzed, results have been obtained by using
buffers ranging from pH 7.2 to pH 8.5. Stain in the dark at room temperature;
a
modification of Shaw and Prasad (1970).
Leucine aminopeptidase
(EC 3.4.11.-)
L-leucine-p-naphthylamide (free base or
acid salt) 20 mg
dissolved in dimethyl
formamide 5 ml
1.0 M phosphate
buffer,
pH 6.0 10 ml
H,O 90 ml
Black
K salt or fast black K salt 50 mg
Stain in the dark at room temperature;
a modification of Gottlieb (1973c). This
enzyme has broad activity; it can cleave a number of N-terminal amino acids.
Although commonly referred to as Leucine aminopeptidase ("LAP"), it is more
precisely referred to as Aminopeptidase. There are a number of aminopeptidases,
some of which are specific for certain N-terminal amino acids. Aminopeptidase
(cytosol; EC 3.4.11.1) is activated by heavy metals, whereas Aminopeptidase
(microsomal; EC 3.4.11.2) is not activated by heavy metals (I.U.B., 1979). It may
be necessary to modify the pH of the stain in order to increase staining intensity.
One isozyme is usually observed and a second is occasionally apparent; both are
monomeric.
Malate dehydrogenase
(EC 1.1.1.37)
1.0 M Tris-HCl
buffer,
pH 8.0 or 8.5 10 ml
2.0 M DL-malic
acid, pH 8.0 (add
NaOH to pH 8.0) 10 ml
H20 80 ml
NAD 10 mg
MTT
(or NBT) 10 mg
PMS 2 mg
Stain in the dark at room temperature;
a modification of Shaw and Prasad (1970).
Some investigators add 38 mg EDTA (tetrasodium salt, dihydrate)
to this recipe. We
use 2.0 M DL-malic acid rather than 1.0 M L-malic acid as most standard
staining
schedules require because DL-malic acid is much less expensive than purified
L-malic acid, which is the actual substrate. There are at least four malate NAD+
dehydrogenases reported (I.U.B., 1979). The one most frequently reported in
routine electrophoresis is L-Malate: NAD + oxidoreductase (EC 1.1.1.37), of which
at least two putative isozymes have been observed. The protein is dimeric; subcellu-
lar compartmentalization
of the isozymes is discussed by Gottlieb (1981b).
Malic enzyme (EC 1.1.1.40; = Malate dehydrogenase [Oxaloacetate-decar-
boxylating, NADP+] in I.U.B., 1979)
1.0 M Tris-HC1 buffer, pH 8.0 or 8.5 10 ml
H2O 80 ml
2.0 M DL-malic acid, pH 8.0 (prepare as in Malate dehydrogenase) 10 ml
1.0 M MgC12 2 ml
NADP 20 mg
MTT (or NBT) 20 mg
PMS 2 mg
21
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN
FERN
JOURNAL:
VOLUME
73 (1983)
Stain in the dark at room temperature; a modification of Richmond (1972 and
pers. comm.).
Peroxidase
(EC 1.11.1.7)
3-amino-9-ethyl carbazole 65 mg
dissolved in dimethyl formamide 5 ml
0.05 M sodium acetate buffer, pH 5.0 95 ml
0.1 MCaC12 2 ml
3% H,O, 2 ml
Incubate the gel in a refrigerator
until the bands appear
(30-60 min); a modification
of Shaw and Prasad (1970) and of Gottlieb (1973b).
Phosphoglucoisomerase
(EC 5.3.1.9; = Glucosephosphate
isomerase in I.U.B.,
1979)
1.0 M Tris-HC1 buffer, pH 8.0 10 ml
H20 90 ml
1.0 M MgCl2 1 ml
Fructose-6-phosphate, disodium salt 30 mg
Glucose-6-phosphate dehydrogenase 40 units
NAD(P)* 10 mg
MTT (or NBT) 20 mg
PMS 2 mg
Stain in the dark at 37?C or at room temperature;
a modification of Shaw and
Prasad (1970). This dimeric enzyme has numerous synonyms (I.U.B., 1979)
including: Phosphohexose isomerase, Phosphohexomutase, Oxoisomerase, Hexose
phosphate isomerase, Glucose-6-phosphate isomerase, Phosphosaccharomutase,
and
Phosphohexoisomerase. Cytosolic and chloroplastic isozymes should be expected
(Gottlieb, 1981b, 1982).
Phosphoglucomutase
(EC 2.7.5.1)
1.0 M Tris-HC1 buffer, pH 8.0 or 8.5 10 ml
H,O 90 ml
1.0 M MgCl2 2 ml
Glucose-l-phosphate, disodium (Sigma G7000)
or dipotassium (Sigma G6875) salt 100 mg
1.7 x 10-4 M ca-D-glucose-l,6-diphosphate,
tetra(cyclohexylammonium) salt 5 ml
Glucose-6-phosphate dehydrogenase 40 units
NAD(P)* 10 mg
MTT (or NBT) 20 mg
PMS 2 mg
A modification of Shaw and Prasad (1970). An equally suitable alternate recipe
which eliminates the need to purchase a-D-glucose-l,6-diphosphate is:
1.0 M Tris-HC1 buffer, pH 8.0 10 ml
H20 90 ml
1.0 M MgCl2 2 ml
Glucose-l-phosphate, disodium salt
(Sigma G1259) 50 mg
Glucose-6-phosphate dehydrogenase 40 units
NAD(P)* 10 mg
MTT 10 mg
PMS 2 mg
Stain in the dark at 37?C or at room temperature.
22
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D. E.
SOLTIS
ET AL.: STARCH GEL ELECTROPHORESIS
OF
FERNS
6-Phosphogluconate dehydrogenase (EC 1.1.1.44)
1.0 M Tris-HC1
buffer,
pH 8.0 10 ml
H20 90 ml
6-phosphogluconic
acid, barium
salt 40 mg
1.0 M MgC12 2 ml
NADP 10 mg
MTT 10 mg
PMS 2 mg
Stain in the dark at room temperature or at 37?C; a modification of Shaw and
Prasad (1970).
Shikimate dehydrogenase (EC 1.1.1.25)
1.0 M Tris-HC1
buffer,
pH 8.5 10 ml
H20 90 ml
Shikimic acid 100 mg
NADP 10 mg
MTT
(or NBT) 20 mg
PMS 2 mg
Stain in the dark at room temperature
or at 37?C.
Triosephosphate isomerase (EC 5.3.1.1)
1.0 M Tris-HCl
buffer,
pH 8.0 10 ml
H20 90 ml
Dihydroxyacetone phosphate,
lithium salt 10 mg
EDTA,
tetrasodium
salt, dihydrate 38 mg
NAD 30 mg
MTT 10 mg
PMS 2 mg
Arsenic
acid, sodium salt 460 mg
Glyceraldehyde-3-phosphate
dehydrogenase 300 units
Stain in the dark at 37?C; Gottlieb (pers. comm.). An alternative and much less
expensive method of staining for TPI follows (also see agarose recipe):
1.0 Tris-HCl buffer,
pH 8.0 10 ml
H20 90 ml
DL-a-glycerophosphate 200 mg
Pyruvic
acid, sodium salt 100 mg
a-Glycerophosphate dehydrogenase 100 units
Lactate
dehydrogenase 100 units
NAD 20 mg
Incubate the above solution for 2 hours at 30-37?C. At the end of the incubation
period, inactivate the enzymes by adjusting the solution to pH 2.0 with 1.0 M HC1
(taking care not to go below pH 2.0) and then re-adjust to pH 8.0 with 1.0 M
NaOH. Add:
1.0 M arsenic
acid, sodium salt 1 ml
Glyceraldehyde-3-phosphate dehydrogenase 100 units
NAD 20 mg
MTT 5 mg
PMS 1 mg
Stain in the dark at 30?C. The bands will be dark blue on a light blue background.
Readers who follow this protocol must first show the lack of cx-Glycerophosphate
dehydrogenase activity in their taxa. This can be done simply by staining a gel slice
with:
23
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN
FERN JOURNAL:
VOLUME
73 (1983)
1.0 M Tris-HCl buffer, pH 8.0 10 ml
H,0 90 ml
DL-a-glycerophosphate 200 mg
NAD 20 mg
MTT 5 mg
PMS I mg
AGAROSE STAINING SCHEDULES
Several of the reagents commonly required in enzyme electrophoresis, such as
NADP, are expensive. Costs may be reduced by using staining schedules employing
agarose (modified from Mitton et al., 1979 and Gaines, pers. comm.), which
require much smaller quantities of most ingredients than do standard staining
schedules but yield comparable results.
Agarose staining schedules are provided below for ALD, G6PDH, GDH, IDH,
MDH, PGI, PGM, 6-PGD, SkDH, and TPI. The following general procedure is
employed to stain one starch gel slice using the agarose technique: dissolve all
ingredients required for enzyme staining (see below) in 6 ml of Tris-HC1 buffer; in
a second flask mix 0.06 g agarose (Agarose type II, Sigma A6877) with 6 ml
distilled water (a 1% agarose solution) and bring this solution to a boil while stirring
constantly; combine the two solutions and very quickly apply to surface of one gel
slice (it may be necessary to increase ml amounts of buffer and 1% agarose
depending on the size of the gel slice); stain all slices in the dark at room
temperature.
Aldolase (EC 4.1.2.13)
PMS 1 mg
MTT 5 mg
NAD 10 mg
Fructose-1,6-diphosphate, trisodium salt
or tetra(cyclohexylammonium) salt 75 mg
Arsenic acid 25 mg
Glyceraldehyde-3-phosphate dehydrogenase 40 units
0.1 M Tris-HC1 buffer, pH 8.5 6 ml
1% agarose 6 ml
Glucose-6-phosphate dehydrogenase
(EC 1.1.1.49)
PMS 1 mg
MTT 4 mg
NADP 4 mg
EDTA, tetrasodium salt, dihydrate 6 mg
Glucose-6-phosphate, disodium salt 5 mg
0.1 M Tris-HCl buffer, pH 8.0 6 ml
1% agarose 6 ml
Glutamate
dehydrogenase
(EC 1.4.1.2)
PMS 1 mg
MTT 4 mg
NAD 3 mg
L-glutamic acid, monosodium salt 100 mg
0.1 M Tris-HCl buffer, pH 8.0 6 ml
1% agarose 6 ml
24
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D.
E.
SOLTIS
ET AL.:
STARCH GEL ELECTROPHORESIS
OF
FERNS
Isocitrate dehydrogenase
[NADP+] (EC 1.1.1.42)
PMS 1 mg
MTT 4 mg
NADP 4 mg
MgC 12 40 mg
Isocitric acid, trisodium salt 10 mg
0.1 M Tris-HC1 buffer, pH 8.0 6 ml
1% agarose 6 ml
Malate dehydrogenase
(EC 1.1.1.37)
PMS 2 mg
MTT 7 mg
NAD 7 mg
L-malic acid 16 mg
0.25 M Tris-HC1 buffer, pH 8.6 6 ml
1% agarose 6 ml
Keep PMS and L-malic acid powders separate until adding the buffer.
Phosphoglucoisomerase
(EC 5.3.1.9)
PMS I mg
MTT 4 mg
NAD(P)* 1 mg
0.1 M Tris-HCl buffer, pH 8.0 4 ml
10% MgC12 1 ml
0.018 M fructose-6-phosphate, disodium salt I ml
Glucose-6-phosphate dehydrogenase 10 units
1% agarose 6 ml
Phosphoglucomutase
(EC 2.7.5.1)
PMS 1 mg
MTT 5 mg
NAD(P)* 4 mg
0.1 M Tris-HCl buffer, pH 8.0 2 ml
0.05 M glucose-l-phosphate, disodium salt (Sigma G7000) 2.5 ml
Glucose-6-phosphate dehydrogenase 10 units
0.001 M fructose- ,6-diphosphate, trisodium salt 1 ml
10% MgC12 1 ml
1% agarose 6 ml
6-Phosphogluconate
dehydrogenase
(EC 1.1.1.4.4)
PMS 1 mg
MTT 4 mg
NADP 4 mg
6-phosphogluconic acid, barium salt 10 mg
10% MgC12 0.5 ml
0.1 M Tris-HCl buffer, pH 8.5 5 ml
1% agarose 6 ml
Dissolve the 6-phosphogluconic acid barium salt in 1.0 ml of 0.1 M Tris-HC1
buffer (pH 8.5) approximately 15 minutes before mixing in other ingredients.
25
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
AMERICAN
FERN
JOURNAL:
VOLUME
73
(1983)
Shikimate dehydrogenase (EC 1.1.1.25)
PMS 1 mg
MTT 4 mg
NADP 4 mg
Shikimic acid 10 mg
0.1 M Tris-HCl buffer, pH 8.0 6 ml
1% agarose 6 ml
Triosephosphate isomerase (EC 5.3.1.1)
PMS 1 mg
MTT 4 mg
NAD 10 mg
EDTA, tetrasodium salt, dihydrate 10 mg
Arsenic acid 150 mg
Dihydroxyacetone phosphate 1.5 mg
Glyceraldehyde-3-phosphate dehydrogenase 75 units
0.1 M Tris-HCl buffer, pH 8.0 6 ml
1% agarose 6 ml
GEL FIXATION
AND DOCUMENTATION
AAT and PER gel slices are fixed in 50% glycerol; all other gel slices are fixed in
50% ethanol.
In order to keep a permanent record of results, we routinely photograph
gel slices
using Kodak Technical Pan Film 2415 following the exposure index and developing
procedure recommended for high contrast.
ACKNOWLEDGEMENTS
D.E.S. and C.H.H. thank Jeff Atwood, Harriet Blanton, Leslie Gottlieb, James
Hamrick, Terry Lastovicka, and Andrew Torres for laboratory
assistance and,helpful
comments and Charles Werth for providing details of his sample preparation
methodology. G.J.G. is grateful to Leslie Gottlieb for the use of laboratory
facilities
and advice during work with Pellaea andromedifolia. Grants-in-aid of research to
D.E.S. from the Research Council of the University of North Carolina at Greens-
boro, to C.H.H. from General Research Fund Grants 3005 and 3588 of the
University of Kansas, and to D.C.D. from the Department of Biology of Indiana
University, NSF Doctoral Dissertation Improvement
Grant DEB-8117239, and the
Society of the Sigma Xi are gratefully acknowledged.
LITERATURE
CITED
ADAMS, W. T. and R. J. JOLY. 1980. Genetics of allozyme variants in loblolly pine. J. Hered.
71:33-40.
BREWER, G. J. 1970. An Introduction to Isozyme Techniques. Academic Press, New York.
CHAPMAN, R. H., E. J. KLEKOWSKI, Jr., and R. K. SELANDER. 1979. Homoeologous
heterozygosity and recombination in the fern Pteridium aquilinum. Science 204:1207-1209.
COOPER-DRIVER, G. 1976. Chemotaxonomy and phytochemical ecology of bracken. J. Linn. Soc..
Bot. 73:35-46.
GASTONY, G. J. and L. D. GOTTLIEB. 1982. Evidence for genetic heterozygosity in a homosporous
fern. Amer. J. Bot. 69:634-637.
, and C. H. HAUFLER. 1976. Chromosomes and apomixis in the fern genus Bommeria
(Gymnogrammaceae). Biotropica 8:1-11.
26
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
D.
E.
SOLTIS ET AL.:
STARCH
GEL ELECTROPHORESIS OF FERNS
GOTTLIEB, L. D. 1973a. Genetic control of Glutamate oxaloacetate transaminase isozymes in the
diploid plant Stephanomeria
exigua and its allotetraploid
derivative. Biochem. Genet. 9:97-107.
1973b. Genetic differentiation, sympatric speciation, and the origin of a diploid species of
Stephanomeria. Amer. J. Bot. 60:545-553.
. 1973c. Enzyme differentiation and phylogeny in Clarkia franciscana, C. rubicunda and C.
amoena. Evolution 27:205-214.
1974. Genetic confirmation of the origin of Clarkia lingulata. Evolution 28:244-250.
1981a. Gene numbers in species of Astereae that have different chromosome numbers. Proc.
Natl. Acad. Sci. USA 78:3726-3729.
1981b. Electrophoretic evidence and plant populations. Progress in Phytochemistry 7:1-45.
1982. Conservation and duplication of isozymes in plants. Science 216:373--380.
INTERNATIONAL UNION OF BIOCHEMISTRY, NOMENCLATURE COMMITTEE. 1979. Enzyme
Nomenclature 1978. Academic Press, New York.
KELLEY, W. A. and R. P. ADAMS. 1977a. Preparation of extracts from Juniper leaves for
electrophoresis. Phytochemistry 16:513-516.
and R. P. ADAMS. 1977b. Seasonal variation of isozymes in Juniperus scopulorum:
systematic significance. Amer. J. Bot. 64:1092-1096.
MITTON, J. B., Y. B. LINHART, J. L. HAMRICK, and J. S. BECKMAN. 1977. Observations on the
genetic structure and mating system of ponderosa pine in the Colorado front range. Theor.
Appl. Genet. 51:5-13.
, Y. B. LINHART, K. B. STURGEON, and J. L. HAMRICK. 1979. Allozyme polymorphism
detected in mature needle tissue of ponderosa pine. J. Hered. 70:86-89.
RICHMOND, R. C. 1972. Enzyme variability in the Drosophila willistoni group. II. Amounts of
variability in the superspecies, D. paulistorum. Genetics 70:87-112.
SCANDALIOS, J. G. 1969. Genetic control of multiple molecular forms of enzymes in plants: a
review. Biochem. Genet. 3:37-79.
SELANDER, R. K., M. H. SMITH, S. Y. YANG, W. E. JOHNSON, and J. B. GENTRY. 1971. IV.
Biochemical polymorphism and systematics in the genus Peromyscus. I. Variation in the
old-field mouse (Peromyscus polionotus). Studies in genetics VI. Univ. Texas Publ.
7103:49-90.
SHAW, C. R. and R. PRASAD. 1970. Starch gel electrophoresis of enzymes-a compilation of recipes.
Biochem. Genet. 4:297-320.
SOLTIS, D. E., C. H. HAUFLER, and G. J. GASTONY. 1980. Detecting enzyme variation in the fern
genus Bommeria: an analysis of methodology. Syst. Bot. 5:30-38.
WERTH, C. R., A. A. KARLIN, and S. I. GUTTMAN. 1982. Enzyme extraction from phenolic-
containing plant tissues using caffeine. Isozyme Bull. 15:139.
27
This content downloaded from 128.227.169.130 on Fri, 05 Jun 2015 16:02:09 UTC
All use subject to JSTOR Terms and Conditions
... Each gel slice was individually placed into a staining tray. Staining followed the recipes of Brewer (1970), Shaw and Prasad (1970), Harris and Hopkinson (1976), and Soltis et al. (1983). ...
... CC-BY-NC 4.0 International license made available under a (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is Brewer (1970), Shaw and Prasad (1970), Harris and Hopkinson (1976), and Soltis et al. (1983). ...
Preprint
Full-text available
We used common garden growth experiments to study genetic variation among geographic isolates (Greenland, Massachusetts, and Connecticut, USA) of the filamentous brown seaweed Pilayella littoralis, including the free-living form unique to Nahant Bay, Massachusetts. Ecotypic variation for temperature growth maximum was demonstrated for a west Greenland isolate (10° C versus 15° C for other attached isolates) and between Nahant Bay attached (narrower phenotypic plasticity) and free-living forms (broader phenotypic plasticity) of the species. Morphological and reproductive characteristics of attached and free-living isolates remained distinctive under identical culture conditions after four years. The attached forms were characteristically cabled, twisted, and clumped; unilocular reproductive cells were common and plurilocular reproductive cells were present. The free-living form was characteristically loosely branched and ball-like; only vegetative reproduction occurred, with a few unilocular reproductive cells observed in one experiment. Free-living and attached isolates cultured using no water movement and turbulent conditions to mimic surf and surge conditions did not develop forms that resembled each other after eight months. We additionally used starch gel electrophoresis to study genetic variability in attached and free-living forms of P. littoralis from Nahant Bay. Free-living and attached populations were not different at the isozyme level because a limited number of isozymes were resolved (six out of 39 enzymes tested). One isozyme (PGI) was polymorphic, with two alleles present. The two alleles shared in the attached and free-living populations suggest that the free-living form is not one large identical clone. For attached and free-living P. littoralis, both transplant and growth studies in the laboratory provide convincing evidence of ecotypic differentiation.
... They were stored at -70 o C until electrophoresed. Ten enzymes were resolved on 11% starch gels with two buffer systems (Soltis et al., 1983). System I had an electrode buffer of 0.065 M L-histidine free base titrated to pH 6.5 with 0.007 M citric acid monohydrate and a gel buffer of a 1:3 dilution of the electrode buffer. ...
... System II was employed to resolve phosphoglucose isomerase (PGI), alcohol dehydrogenase (ADH), and malic enzyme (ME). Enzyme activity staining and agarose overlays mostly followed the protocols of Soltis et al. (1983). Loci and alleles were numbered sequentially and lettered alphabetically, beginning with the most anodal form. ...
Article
To elucidate the ancestry of the allopolyploids E. stevenii and E. boöphthona , I examined eleven isozyme loci and 24 morphological characters from 28 populations representing five related Euphorbia species from Australia. According to an analysis of genetic and morphological data, three diploid species differentiated recently, but two independent polyploid species are estimated to have differentiated a relatively long time ago. Fixed heterozygosity for most isozymes in E. stevenii and E. boöphthona strongly suggests that these two species are allopolyploids rather than autopolyploids. The isozyme profiles of E. stevenii indicate that it is an allopolyploid that evolved from interspecific hybridization between the diploid E. tannensis and unidentified or extinct tetraploid species. In addition, isozyme patterns strongly suggest that E. stevenii was one of the ancestors of E. boöphthona . However, E. boöphthona showed a large number of fixed alleles that were not detected in any other Australian Eremophyton species. The most likely hypothesis for the origin of E. boöphthona is that it was formed by hybridization and chromosomal doubling between an extinct diploid species and the hexaploid E. stevenii .
... Corolla colour, corolla shape, colour of corolla spots, anther colour, style colour and mature fruit colour were thought to distinguish Forty plants from thebackcross family were used for studies of morphological markers and isozyme segregation. For isozyme analysis, very young leaves were macerated in an extraction medium and the extracts were subjected to horizontal starch gel electrophoresis (Soltis et al., 1983). The formulation of 100 ml of stock solution (pH 7.8) for preparing extraction medium was 1.211 g Tris, 1.761 g ascorbic acid, 0.074 g KCl, 0.002 g Na2 EDTA, and 0.037 g MgCL2.6H2O. ...
... After electrophoresis, the gel was cut into several slices (normally 5 or 6 slices, each 1.5 mm thick) and each slice was stained for a different enzyme system. The slices were in the appropriate staining solution for the enzyme to be visualized in an oven at approximately 37 o C in the dark (Soltis et al., 1983). The staining solution was poured off and the gel was rinsed a couple of times with distilled water. ...
... For each individual analysed, approximately 1 cm 2 of leaf was ground in 80 μl of a Tris-HCl extraction buffer (Soltis et al. 1983), modified by replacing ß-mercaptoethanol with dithiothreitol. Extracts were absorbed onto paper wicks and proteins were separated on horizontal starch gels at 60 mV for 30 mins until removal of the wicks, after which current was increased to 70 mV for a further 3 -4 h. ...
Article
Full-text available
Between 1959 and 1988, three populations of purple-flowered terrestrial orchids attributable to Dacty-lorhiza subgenus Dactylorhiza were discovered in Canada. The populations at Timmins, Ontario, and St John's, New-foundland were strongly marked on both flowers and leaves, in contrast with the anthocyanin-deficient population at Tilt Cove, Newfoundland. All three populations have since experienced a wide range of taxonomic assignments; debates are also ongoing regarding their origin and most appropriate conservation status. Here, we address these questions by combining detailed in situ morphometric analyses based on 52 characters with allozyme profiles and data from nrITS, 15 plastid microsatellites and seven nuclear microsatellites. The allozyme data alone are sufficient to both confirm allopolyploidy and categorically refute past assignments of these populations to D. incarnata, D. maculata, D. fuchsii, D. majalis or D. purpurella. Several morphometric characters, nuclear microsatellites and nrITS all reliably distinguish each of the three study populations, whereas the two sampled subpopulations from St John's proved near-identical morphologically. In contrast, morphological variation within each of the three populations is strikingly low, particularly in characters other than those influenced by plant vigour. Similarly, compared with 14 European populations, the three Canadian populations proved genetically impoverished (two were near-invariant) and likely experienced recent, extreme genetic bottlenecks during establishment. The three populations differ substantially, both morphologically and molecularly, therefore probably representing independent immigration events. Although clearly attributable to D. praetermissa, all three populations deviate significantly in morphology and DNA data from comparable populations sampled across Europe, preventing identification of their precise geographic origins. Any attempt to determine their mode or origin-through natural long-distance transport, or accidental or deliberate introduction by humans-is challenged to explain why three lineages of a single Euro-pean Marsh-orchid species, each in different ways atypical of that species, arrived independently in North America whereas no other European dactylorchid species has become established there.
... For each individual analysed, approximately 1 cm 2 of leaf was ground in 80 μl of a Tris-HCl extraction buffer (Soltis et al. 1983), modified by replacing ß-mercaptoethanol with dithiothreitol. Extracts were absorbed onto paper wicks and proteins were separated on horizontal starch gels at 60 mV for 30 mins until removal of the wicks, after which current was increased to 70 mV for a further 3 -4 h. ...
Article
Full-text available
The intensively studied Eurasian orchid genus Dactylorhiza has become a model system for exploring allopolyploid evolution, yet determining the optimal circumscriptions of, and most appropriate ranks for, its constituent taxa remain highly controversial topics. Here, novel allozyme data and detailed morphometric data for 16 Scottish marsh-orchid populations are interpreted in the context of recent DNA sequencing studies. Despite being derived from the same pair of parental species, the two allopolyploid species that currently occur in Scotland can reliably be distinguished using allozymes, haplotypes, ribotypes or sequences of nuclear genes. A modest range of diverse morphological characters are shown to distinguish the two molecularly-circumscribed species, but they have in the past been obscured by equivalent levels of infraspecific variation in characters rooted in anthocyanin pigments; these characters are better employed for distinguishing infraspecific taxa. Dactylorhiza francis-drucei (formerly D. traunsteinerioides ) is confirmed as being distinct from the continental D. traunsteineri/lapponica , probably originating through allopatric isolation once the continental lineage reached Britain. All Scottish populations are attributed to the comparatively small-flowered, anthocyanin-rich subsp. francis-drucei , which includes as a variety the former D. 'ebudensis' ; the less anthocyanin-rich subsp. traunsteinerioides is confined to Ireland, North Wales and northern England. In contrast with D. francis-drucei , only a minority of Scottish populations of D. purpurella are attributed to the anthocyanin-rich race, var. cambrensis . This species most likely originated through an allopolyploidy event that occurred comparatively recently within the British Isles, as it contains allozyme alleles distinctive of British rather than continental D. incarnata (its diploid pollen-parent). In contrast, the rare Scottish population of D. incarnata subsp. cruenta shares with its Irish counterparts a continental genotype, and is most likely a recent arrival in Scotland through long-distance dispersal. Among all European allotetraploid dactylorchids, D. purpurella is the species that most closely resembles D. incarnata , both molecularly and morphologically.
... Samples were centrifuged at 16,000 rpm for 20 min at 4 °C, and supernatants were kept at − 20 °C until used. Native polyacrylamide gel electrophoresis at 10% (w/v) was utilized for separation of isozymes (α-and β-esterase), and the gels were dyed substrates for the desired enzyme according to Soltis et al. [49]. ...
Article
Full-text available
Background In the present study, Allium test bioassay was utilized to evaluate the effects of mixed wastewater of agricultural and sewage effluents at Kitchener pool, Gharbia governate, Middle Delta region, North Egypt. Germination indices, mitotic index and aberrations, α, β-esterase isozymes and inter-simple sequence repeat (ISSR) fingerprinting were tested by different concentrations of the wastewater (tap water as control, 25%, 50% and 100% wastewater). Results Water analysis recorded high levels of electrical conductivity, cations and anions compared to control, but were in the permitted limits according to FAO (Food and Agricultural Organization) except Mg ²⁺ and K ¹⁺ were above the limits. P, N and heavy metals as Pb, Mn and Ni were also higher than the control. Germination indices showed reduction for all parameters studied (root and shoot lengths, root and shoot fresh and dry weights, and tolerance index). Mitotic index decreased, and the percentage of mitotic aberrations increased as the concentration of treatments increased and the time prolonged. Different types of aberrations were recorded in all treatments and its percentage is time and dose independent. Goat cells are the most common type recorded after different times in all treatments. The expression of α, β-esterase enzymes showed variation in different treatments compared to control and ISSR profiles showed considerable polymorphism. Concentration of 25% mixed water induced different profiles for expression of both α- and β-esterase from other treatments, and the cluster analysis based on polymorphism in ISSR fingerprinting revealed the distinction of plants treated with this concentration and the control plants from those treated with high concentrations. Conclusion It was suggested that concentration of 25% mixed water may be suitable for growth and act as fertilizer. Mixed water from this pool may be genotoxic for Allium cepa plants at early growth if it is used for irrigation in its present form and usage of this wastewater for agricultural purposes may be harmful and must be partially treated and biologically tested before use.
Preprint
Full-text available
The relative allocation within a plant species of asexual vs. sexual reproduction is predicted to vary according to changing environmental conditions. In bryophytes, sex expression is predicted to be reduced in unfavourable environments because sporophyte maturation is resource‒limited. The allocation between reproductive strategies in the bryophyte, Polytrichum juniperinum was examined in the low‒arctic Torngat Mountains (Labrador, Canada) in 2008/2009 along both an elevation and moisture gradient. Using genetic markers, no pattern was observed in the allocation between sexual and asexual reproductive strategies among the different environmental conditions. Although all populations showed evidence of sexual reproduction, effective sexual reproduction was depressed along both gradients. All populations showed high rates of fertilization among closely related ramets that resulted in inbreeding. This might limit the capacity to maintain high genetic diversity in local populations, which would be favoured in an environment experiencing warming.
Article
Premise: Previously published evidence suggests that Draba maguirei, a mustard endemic to a few localities in the Bear River, Wellsville, and Wasatch Mountains of northern Utah, may represent a cryptic species complex rather than a single species. Conservation concerns prompted an in-depth systematic study of this taxon and its putative relatives. Methods: Sampling most known populations of D. maguirei s.l. (D. maguirei var. maguirei and D. maguirei var. burkei), we integrate data from geography, ecology, morphology, cytogenetics and pollen, enzyme electrophoresis, and the phylogenetic analysis of nuclear ITS sequences to explore potential taxonomic diversity in the species complex. Results: Draba maguirei var. burkei is shown here to be a distinct species (D. burkei) most closely related to D. globosa, rather than to D. maguirei. Within D. maguirei s.s., the northern (high elevation) and southern (low elevation) population clusters are genetically isolated and morphologically distinguishable, leading to the recognition here of the southern taxon as D. maguirei subsp. stonei. Conclusions: Our study reveals that plants traditionally assigned to D. maguirei comprise three genetically divergent lineages (D. burkei and two newly recognized subspecies of D. maguirei), each exhibiting a different chromosome number and occupying a discrete portion of the geographic range. Although previously overlooked and underappreciated taxonomically, the three taxa are morphologically recognizable based on the distribution and types of trichomes present on the leaves, stems, and fruit. Our clarification of the diversity and distribution of these taxa provides an improved framework for conservation efforts. This article is protected by copyright. All rights reserved.
Preprint
Full-text available
Tissue culture technique was carried out in order to assess the possible effects of calcium nanoparticles either Ca(OH) 2 or CaO at concentration of 100 and 50 µM, respectively on peroxidase and catalase enzymes, total soluble carbohydrates content, starch content and isozymes of peroxidase and amylase of Lupinus termis callus at 30 days of growth. The experimental plant tissues were grown in culture using B 5 medium amended with either different concentrations of CdCl 2 (20 and 60 mg/l) or calcium nanoparticles either Ca(OH) 2 or CaO at concentrations of 100 and 50 µM, respectively. Ca nanoparticles decreased the harmful effects of Cadmium as the activities of peroxidase (POD) and catalase (CAT) were reduced by applying nanoparticles either Ca(OH) 2 or CaO to media having different concentrations of Cd. The content of Starch Tissue culture technique was carried out on Lupinus termis plant.was increased only by addition of Ca nanoparticles either Ca(OH) 2 or CaO and their mixture to 60 mg/l Cd. Isozyme profile of peroxidase and amylase in callus showed 2 isoforms of POD and three isoforms of amylase varied in their intensity and thickening in response to different treatments.
Article
Full-text available
A method for extraction of enzymes from mature leaves of a diversity of trees is described. The methods of tissue preparation allow samples to be collected at any time of year regardless of the reproductive state of the plant. We have used this method to detect six enzyme polymorphisms in ponderosa pine and we present evidence to show that the genes coding for these proteins are simple Mendelian loci.
Article
Full-text available
The semispecies composing the superspecies, Drosophila paulistorum, have been analyzed for genetic variation at 17 enzyme loci. On the average a population of D. paulistorum is polymorphic for 55-67% of its loci and an average individual is heterozygous at 21% of its loci. The pattern of genetic variability found supports the hypothesis that allozyme variation is maintained in natural populations by some form of balancing selection. Evidence is presented which supports the hypothesis that glucose-metabolizing enzymes are less genetically variable than non-glucose-metabolizing enzymes. The known genetic relationships between the semispecies of D. paulistorum are discussed in the light of the frequencies of alleles at allozyme loci.
Article
Although numerous flowering plant species have been subjected to electrophoretic enzyme analysis, very few such studies have involved pteridophytes. This is largely because standard methodology typically results in little or no electrophoretically detectable enzymatic activity in pteridophytes. The presence of large amounts of phenolic compounds in leaves of most pteridophytes suggests that the electrophoretic difficulties heretofore encountered in this group result from complexing of enzymes by these compounds following cellular disruption. Use of a grinding buffer containing dimethylsulfoxide and several compounds inhibitory to phenolic complexing, in combination with a grinding procedure utilizing polyvinyl-pyrrolidone and liquid nitrogen, was found to overcome the difficulties of tissue preparation that result from high concentrations of phenolic compounds. Modifications of this grinding buffer and grinding procedure often resulted in reduced clarity of enzyme bands, or even a total loss of enzyme activity. In Bommeria, preparation of leaf tissue in liquid nitrogen and use of sodium tetraborate, sodium ascorbate, sodium meta-bisulfite, and sodium diethyldithiocarbamate were found to be of less importance than use of polyvinylpyrrolidone in obtaining active enzyme samples. We hope that discussion of our procedural data and provision of gel and electrode buffer recipes and staining schedules will stimulate further electrophoretic investigation of pteridophytes.
Article
Variation of peroxidase enzymes is analyzed both in mature needle tissue and in open-pollinated seedling families of ponderosa pine, Pinus ponderosa, and is identified as being controlled by a single Mendelian locus. Variation at this locus, analyzed in 1, 386 individuals, is used in the analysis of population differentiation and the mating system. Significant variation of gene frequencies is detected over distances of several hundred meters, and is found to be associated with slopes of different aspects. Ponderosa pine is wind-pollinated, and an analysis of the mating system indicates that the level of outcrossing is greater than 90 %. Selection specific for different environments is evidently strong enough to overcome the homogenizing force of migration and produce population fissuring in ponderosa pine.
Article
Variations in flavonoids extracted from fronds suggest that Pteridium aquilinum (L.) Kuhn is a monospecific genus. A wide range of secondary plant substances also occur in bracken including sesquiterpenoids, echydones, cyanogenic glycosides, tannins and phenolic acids.
Article
The multiple forms of glutamate oxaloacetate transaminase in the annual diploid plant Stephanomeria exigua (Compositae) are controlled by three unlinked gene loci with two, four, and five alleles, respectively. All alleles are codominant, and heterozygotes for any pair of them produce a more darkly staining enzyme with intermediate mobility, suggesting that the enzymes have a dimeric subunit structure. In natural populations, the same allele is predominant or fixed at each locus. Stephanomeria elata, the allotetraploid derivative of S. exigua and the closely related S. virgata, produces multiple enzyme variants coded by one pair of its duplicated loci which are identical in electrophoretic mobility to those of diploid individuals heterozygous at this locus. The formation of multiple enzyme variants in all individuals of the tetraploid may provide a degree of biochemical versatility that contributes to its ability to colonize disturbed habitats.