LONG-TERM FITNESS EFFECTS OF ABIOTIC STRESS TOLERANCE TRANSGENES IN
ARABIDOPSIS THALIANA POPULATIONS UNDER COMPETITIVE CONDITIONS
By
Patrick James Bigelow
A DISSERTATION
Submitted to
Michigan State University
in partial fulfillment of the requirements
for the degree of
Plant Breeding, Genetics and Biotechnology - Horticulture - Doctor of Philosophy
2013
ABSTRACT
LONG-TERM FITNESS EFFECTS OF ABIOTIC STRESS TOLERANCE TRANSGENES
IN ARABIDOPSIS THALIANA POPULATIONS UNDER COMPETITIVE CONDITIONS
By
Patrick James Bigelow
The demands on agricultural lands from a growing world population will rise at the same
time that climate change is predicted to increase the abiotic stresses (e.g. drought, heat-waves,
frosts and salinity) that decrease crop yield today. Efforts to increase crop abiotic stress tolerance
are ongoing, including via transgenic approaches. However, unlike past transgenes, abiotic
stress tolerance genes function through indirect alterations to regulatory, signaling and metabolic
pathways, increasing the possibility of complex secondary effects. Concerns have been raised
that such genes could alter crop persistence and ferality and, through gene flow, the invasiveness
and ecological range of recipient interfertile wild or weedy relatives. These possible ecological
risks are influenced by fitness effects conferred by the transgene. Despite the link between
competitive fitness and long-term environmental risks, competitive fitness has been rarely
empirically determined in ecological risk assessment. In this study three transgenes, which
increase salinity tolerance in Arabidopsis thaliana in growth chamber studies, were examined for
impacts on plant fitness: (I) the abiotic stress response transcription factor C-repeat binding
factor 3/drought responsive element binding factor 1a (CBF3/DREB1a), (II) the plasma
membrane Na+/H+ antiporter Salt Overly-Sensitive 1 (SOS1), and (III) the mannitol biosynthetic
enzyme mannose-6-phosphate reductase (M6PR). Transgene fitness impacts were examined
across six field seasons and in the presence and absence of competition with the wild-type
2
parental genotype at planting densities (2600/m ) chosen to replicate conditions observed in wild
populations. Fourteen replicate competitive populations, initially 1:1 transgenic:wild-type
mixes, of each transgenic line (2-3 lines/transgene) were maintained separately for six
generations to allow transgene frequencies to fluctuate according to genetic drift and field
selective pressures. Transgene frequencies were monitored each generation via phenotypic
screening of progeny seed for presence of the co-integrated kanamycin resistance trait; low
frequency populations were verified by qPCR analysis to rule out artifacts due to possible gene
silencing. The fitness effects observed in competition with WT differed from relative fitness in
pure populations. In pure populations, CBF3 lines showed moderately negative fitness relative to
wild-type, but decreased to near extinction in direct competition. SOS1 lines performed
equivalently to wild-type in pure populations but decreased in frequency by 50% in competition.
The fitness of both M6PR lines was enhanced relative to wild-type in field pure populations, but
in competition one line exhibited a competitive advantage while the other was selectively neutral
and exhibited random drift. Selection and drift modeling, incorporating short-term noncompetitive and competitive transgene fitness measurements, determined that only models which
utilized competitive fitness values yielded long-term transgene frequency patterns comparable to
trends observed in the field. Significant relative fitness gains were observed from all three
transgenes under salt stress in the growth chamber, but from only SOS1 and M6PR lines in the
greenhouse. In competition with wild-type plants no advantage was observed, indicating that like
the field, competition reduced observed transgene fitness. The implications of these findings,
together with prior transcriptomic analysis of the three transgenes, were examined in the context
of environmental risk assessment (ERA) practices. Together these results indicate the important
role competition has on the success or failure of a transgene to establish and support the use of
competitive field assessments in estimating the risk of transgene establishment.
ACKNOWLEDGEMENTS
I would like to thank all my committee members Drs. Wayne Loescher, Jim Hancock,
Paul Thompson, and Carolyn Malmstrom, with special thanks to my advisor Dr. Rebecca
Grumet for the opportunity to design and conduct the experiments of this dissertation and for her
guidance and support through the writing process.
I would also like to thank all current and past members of the Grumet lab, with special
thanks to Zhulong Chan for the project’s growth chamber ground work, Kaori for her advice as I
was beginning research, Jessica for the humor needed to get through countless hours of seed
screening, and Sue for help with said seed screening. I would also like to thank the
undergraduate students who assisted, whether they wanted to or not, in the seed cleaning process
and especially Jean for her many years of work in the field and lab with this project.
Thanks go out to all ‘joint lab members’ who have provided so much input and advice
over the years, especially Veronica and Ann for being sounding boards for resolving strange
issues.
Most importantly I would like to thank my wonderful family for their eternal patience
with my long career as a student. To my mom Deb, dad Leon and brother Chris, I have finally
made it. I know you are proud of me and I am so proud of all of you.
To my wife Erin, you have been there supporting me throughout this whole adventure
with love, humor and wit. Thank you!
iv
TABLE OF CONTENTS
LIST OF TABLES………………………………………………………………………….…....vii
LIST OF FIGURES……………………………………………………………………..………viii
Chapter 1: Literature Review………………………………………………………………….…..1
Introduction………………………………………………………………………………..1
Salinity Stress …………………………………………………………………………….1
Genetic engineering for abiotic stress tolerance…………………………………………..2
Ecological risk assessment of abiotic stress tolerance enhanced crops………….………..5
Feral Crops………………………………………………………………………..7
Crop x Wild hybridization………………………………………………………..8
Wild x GM crop hybridization…………………………………………………..10
Transgene Fitness effects ………………………………………………………………..11
Measuring impacts on fitness…………………………………………………….12
Transgenes assessed……………………………………………………………………...13
CBF3/DREB1a…………………………………………………………………..14
SOS1…………………………………………………………………..…………16
M6PR…………………………………………………………………………….17
Objectives of dissertation………………………………………………………………..18
LITERATURE CITED..…………………………………………………………………20
Chapter 2: Multigenerational study of the establishment of abiotic stress tolerance transgenes in
Arabidopsis thaliana populations under competitive field conditions…………………………..32
Introduction…….………………………………………………………………………..32
Materials and Methods…………………………………………………………………..36
Arabidopsis lines for field experiment……………………….………………….36
Experimental design for field experiment……………………………………….36
Genotyping field grown progeny via kanamycin screening…………………….42
qPCR verification of transgene frequency calculated from kanamycin
screening………………………………………………………………………...43
Determination of selective pressure under competitive conditions……………..44
Modeling transgene frequency…………………………………………………..44
Statistical Analysis………………………………………………………………45
Results………………………………………………………………………………........45
Effects of seasonal differences on plant development and productivity…………45
Effects of transgene expression on plant development………………………….46
Effects of transgene expression on non-competitive plant productivity…………54
Productivity, fitness and transgene frequency in competition with wild-type
plants…………………………………………………………………………….54
Relative fitness in non-competitive verses competitive conditions……………..62
Modeling transgene frequency…………………………………………………..62
Discussion………………………………………………………………………………..67
v
LITERATURE CITED..…………………………………………………………………73
Chapter 3: The fitness effects of three abiotic stress tolerance transgenes in Arabidopsis thaliana
in the presence of salinity stress and competition………………………………………………..81
Introduction…….….……………………………………………………………………..81
Materials and Methods…………………………………………………………………...83
Arabidopsis lines…………………………………………………………………83
Experimental design and salt treatment………………………………………….84
Genotyping progeny from competitive populations via kanamycin screening….87
Determination of selective pressure under competitive conditions……….……..88
Statistical Analysis……………………………………………………………….88
Results………………………………………………………………………………........89
Effects of salinity stress on plant development and productivity………………..89
Effects of transgene expression on plant development and productivity in the
absence of salinity stress…………………………………………………………90
Effects of transgene expression on plant development and productivity in the
presence of salinity stress……………………………………………………….102
Effects of salinity and transgene expression on competitive ability……………106
Discussion………………………………………………………………………………110
LITERATURE CITED..……..…………………………………………………………115
Chapter 4: Crop improvement utilizing abiotic stress tolerance enhancing transgenes and the
implications for ecological risk assessments…………………………………………………...119
Introduction…….……………………………………………………………………….119
Analysis of implications for ecological risk assessments..……………………………..120
The current status of abiotic stress tolerant crops………………………………120
Challenges in the development and implementation of abiotic stress tolerant
Crops…………………………………………………………………………...123
Ecological risks of abiotic stress tolerant crops………………………………..124
Implications of abiotic stress tolerant GE crops for current ecological risk
assessment methodologies……………………………………………………..125
European Food Safety Association guidance report on the ERA of transgenic
plants..………………………………………………………………………….127
Persistence, invasiveness and plant-to-plant gene flow………………..128
Interactions with target and non-target species…………………………132
Cultivation and management changes………………………………….133
Assessing risk of abiotic stress tolerance enhancing transgenes by the
measurement of secondary fitness effects………………………………………134
Assessing risk of abiotic stress tolerance enhancing transgenes using ‘omic’
approaches………………………………………………….…………………..137
Discussion………………………………………………………………………………140
LITERATURE CITED…………………………………………………………………144
Conclusions and future work…………………………………………………………………...155
APPENDIX………………………………………………………...…………………………...160
vi
LIST OF TABLES
Table 2.1. Transgenic Arabidopsis thaliana lines engineered with abiotic stress tolerance genes
and the number of replicate populations tested over six generations in the field.……..39
Table 2.2. Seasonal effect on development and productivity averaged across all lines and
populations. ……………………………………………………………………………...47
Table 2.3. Analysis of variance showing the effects of transgene, transformation event,
environment and their interactions on seed yield from non-competitive pure populations
grown for six seasons in the field ……………..………………………………………..48
Table 2.4. Transgene effect on productivity as observed in non-competitive populations averaged
across all field seasons. …………………………………………………………………55
Table 2.5. Comparison of transgene frequencies estimated by phenotypic selectable marker
screening (nptII) and qPCR analysis. …………………………………………………...59
Table 2.6. . Measured single generation fitness values, from non-competitive growth chamber
(GC), non-competitive field (NCF), and competitive field (CF) populations, used in
2
transgene frequency modeling with PopGene.S modeling program as well as predicted
and observed sixth generation competitive fitness values. ………………………..……64
Table 3.1. Transgenic lines of Arabidopsis thaliana and the number of replicate populations
tested in three greenhouse experiments under control and salinity stressed conditions
(75mM NaCl)…………………………………………………………………...……….86
Table 3.2. Salinity treatment effect on development and productivity averaged across all lines,
populations and experiments……………………………………………………………..93
Table 3.3. Transgene effect on development and productivity averaged across lines and
populations…..…………………………………………………………….…………...101
vii
LIST OF FIGURES
Figure 1.1. The yearly number of submissions to the USDA for field trial of abiotic stress related
traits from 1990-2013. Submission data was collected from the Information Systems for
Biotechnology (http://www.isb.vt.edu/search-release-data.aspx) on January 23, 2013…..3
Figure 2.1. Verification of transgenic Arabidopsis thaliana lines via Southern (A) and northern
analysis (B). SOS1 and CBF3 lines show the respective transgene and the endogenous
gene, while M6PR lines show only the transgene since the gene is not endogenous to
Arabidopsis thaliana. Verification was performed by Zhulong Chan…………………...37
Figure 2.2. The tray-in-flat field planting method showing the trickle hose for irrigation and the
high density of planting for interplant competition (A) and the screening of progeny seed
on ½ MS 1% agar 100mg/L kanamycin containing media (B). Germination rates and
transgene frequencies can be calculated from the number of: bleaching of wild-type
seedlings (blue arrows), healthy transgenic seedlings (green arrows) and non-germinating
seeds (red arrows). For interpretation of the references to color in this and all other
figures, the reader is referred to the electronic version of this dissertation……..……….41
Figure 2.3. Seasonal differences in maximum and minimum air temperature (A) and daily total
solar flux density (B) across the six field generations. Each value is the mean ± SE across
all days populations were in the field for that growing season. Bars with the same letter
were not significantly different from each other at P<0.05 (Duncan’s)………..………..49
Figure 2.4. Mean genotype yield in each season in relation to the environmental yield capacity of
that season. The environmental yield capacity was the mean yield for all genotypes
grown in pure line plots (n=50) (dark line). The CBF3, SOS1 and M6PR values are the
mean of 15, 10 and 10 populations, respectively (three, two and two transgenic lines per
transgene, with five populations per line)…..……………………………………………51
Figure 2.5. Mean days to reach various lifecycle stages for pure populations of wild-type (solid)
and transgenic plants (dashed) CBF3 overexpression lines (A40, A30) and their
background ecotype WS (A), M6PR lines (M2-1, M5-1) and their wild-type Col (B), and
SOS1 lines S1-1 and S7-6 with their wild-type Col(gl) (C). Values are the mean of five
field generations with five replicate populations per genotype. The stages are days to:
first germination (1), 75% of the population with two true leaves (2), 75% of the
population with five to six true leaves forming a rosette (3), first bolting (4), first
flowering (5), 75% reaching bolting (6) , 75% flowering and 75% mature (75% of
siliques drying down). The transition from vegetative to reproductive growth is indicated
(black arrow), highlighting the effect of CBF3 overexpression…………………………52
Figure 2.6 Transgene frequency within mixed populations of wild-type WS and three CBF3
overexpression lines, A40 (A), A30 (B) and A28 (C), as determined by selectable marker
screening. All populations began at 50% starting frequency (FG0) and were maintained
separately in subsequent generations. Changes in transgene frequency from one
viii
generation to the next in each of the fourteen replicate mixed populations are shown by
the black lines, with the mean of all fourteen replicates indicated by the gray line. Dashed
lines indicate 95% confidence intervals predicted for mean transgene frequency
undergoing solely genetic drift…...………………………………..…………………….57
Figure 2.7. Transgene frequency within mixed populations of wild-type Col(gl) and two SOS1
overexpression lines, S1-1(A), S7-6 (B), as determined by selectable marker screening.
All populations began at 50% starting frequency (FG0) and were maintained separately in
subsequent generations. Changes in transgene frequency from one generation to the next
in each of the fourteen replicate mixed populations are shown by the black lines, with the
mean of all fourteen replicates indicated by the gray line. Dashed lines indicate 95%
confidence intervals predicted for mean transgene frequency undergoing solely genetic
drift……………………………………….………………………………….…………...60
Figure 2.8. Transgene frequency within mixed populations of wild-type Col and M6PR
overexpression lines, M2-1(A) andM5-1(B), as determined by selectable marker
screening. All populations began at 50% starting frequency (FG0) and were maintained
separately in subsequent generations. Changes in transgene frequency from one
generation to the next in each of the fourteen replicate mixed populations are shown by
the black lines, with the mean of all fourteen replicates indicated by the gray line. Dashed
lines indicate 95% confidence intervals predicted for mean transgene frequency
undergoing solely genetic drift………………..…………………....……………………61
Figure 2.9. Comparison of non-competitive fitness estimates to competitive fitness estimates for
the three abiotic stress tolerance enhancing transgenes. A fitness value of 100% (dashed
line) indicates fitness equal to wild-type. Each non-competitive fitness value is the mean
± SE transgene fitness, calculated from seed yields relative to WT, of five replicate pure
populations averaged across all six field seasons. Each competitive fitness value is the
mean ± SE transgene fitness, calculated by selectable marker screening of fourteen
replicate, but separately maintained, populations which began as 1:1 wild-type:transgenic
mixes. Lines where non-competitive and competitive fitness values significantly differ at
P<0.05 are marked (*)…………………………………………………………………...63
Figure 2.10. Examples of predicted transgene frequencies for 100 modeled mixed populations
(black lines) of wild-type and transgenic lines. Negatively selected CBF3 line A40 (top
row, a-c), positively selected M6PR line M2-1 (middle row, d-f) and neutral M6PR line
M5-1 (bottom row, g-i). Also indicated are the mean predicted transgene frequency (solid
gray line) and the mean field observed frequency (dashed gray line). The model
incorporated both genetic drift and selection based on growth chamber derived fitness
values (left column, a,d,g), non-competitive fitness estimates from field generation 1
(middle column, b,e,h) and competitive fitness estimates from field generation 1 (right
column, c,f,i)……………………………………………………………………………..65
Figure 3.1. Viable seed yield of pure populations of transgenic and wild type plants in relation to
yield capacity of the environment for six growing conditions, three seasons ±salt (closed
and open symbols respectively). The environmental yield capacity was calculated as the
ix
mean viable seed yield of all genotypes wild-type and transgenic (n=50, gray solid line).
The yield capacity for specific genotypes was the mean viable seed yield for each
transgenic line (n=5, black dashed lines) and wild-type background (n=5, black solid
line). The three CBF3 lines and their WS wild-type background (A), the two SOS1 lines
and their wild-type Col(gl), and the two M6PR lines and their wild-type Col (C)..…….91
Figure 3.2. Mean productivity measures for greenhouse-grown pure line populations under
control and salinity stress (75mM NaCl); aboveground dry weight (A), viable seed yield
based on germination rates of progeny seed (B), partitioning to seed (C) and progeny
seed germination rate (D). Salt treatment reduced dry weight, seed yield, partitioning to
seed and progeny seed viability across all genotypes (P<0.01). Each value is the mean
from five replicate populations per genotype per treatment per experiment. Data are
pooled from three replicate experiments; equivalent trends were observed in each
experiment. Mean values with the same letter were not significantly different from each
other at P<0.05. (Duncan’s)....…………………………………………………………...94
Figure 3.3. Mean productivity measures for growth chamber grown pure line populations under
control and salinity stress (100mM NaCl); aboveground dry weight (A), viable seed yield
based on germination rates of progeny seed (B), partitioning to seed (C) and progeny
seed germination rate (D). Each value is the mean relative fitness from three replicate
populations per genotype per treatment . Mean values with the same letter were not
significantly different from each other at P<0.05. (Duncan’s). The dry weight data for Col
and M6PR lines were previously included in Chan et al. 2011 and were gathered by
Zhulong Chan…………..………………………………………………………………...96
Figure 3.4. Mean days to reach various lifecycle stages under control (black line) and salt treated
(75mM NaCl, grey lines) conditions for pure populations of wild-type (solid) and
transgenic plants (dashed). The development of CBF3 overexpression lines (A40, A30,
A28) and their background ecotype WS (A), M6PR lines (M2-1, M5-1, M7-6) and their
wild-type Col (B), and SOS1 lines (S1-1, S7-6) with their wild-type Col(gl) (C). Values
are the mean of three experiments with five replicate populations per genotype. The
stages are days to: first germination (1), 75% of the population with two true leaves (2),
five to six true leaves forming a rosette (3), first bolting (4), first flowering (5), 75%
reaching bolting , flowering and maturation(6, 7 & 8 respectively)...…………………...98
Figure 3.5. Mean relative fitness for pure populations of CBF3 (lines A40, A30 and A28 ), SOS1
(lines 1-1 and 7-6) , and M6PR (lines M2-1 and M5-1) transgenic plants under control
and salt treated (75mM NaCl) conditions in the greenhouse (A) and control and salt
treated condions in the growth chamber (100 mM NaCl) (B). Relative fitness was
calculated as the proportion of viable transgenic seed yield to viable wild-type seed yield
under the same treatment. The relative fitness of SOS1 and M6PR lines under salinity
stress in the growth chamber exceeded 20 and were cut off for figure clarity. Each value
is the mean relative fitness from replicate populations, with n=15 in the greenhouse and
n=3 in the growth chamber. Mean fitness values which differ from wild-type at P<0.1 or
P<0.05 are marked by ( .) or (*) respectively…………………....……..………………103
x
Figure 3.6. Mean competitive fitness of transgenic lines, CBF3 lines (A40, A30 and A28), SOS1
(lines 1-1 and 7-6) and M6PR (lines M2-1, M5-1), in mixed populations under control
and salt treated (75mM NaCl) conditions. All populations were planted as 1:1 wildtype:transgenic mixes, with each transgenic line planted with its respective wild-type
background: WS with CBF3 lines, Col(gl) with SOS1 lines and Col with M6PR lines.
Competitive fitness was calculated from selectable marker screening of progeny seed.
Each value is the mean fitness from fourteen replicate populations per treatment in three
repeated greenhouse experiments. Mean competitive fitness values outside the 95%
confidence intervals, calculated based on theoretical drift distributions, indicate negative
selective pressures (marked by *)……….…………...…………………………………107
Figure 3.7. Seed yield of competitive mixed populations of transgenic plants and their wild type
backgrounds in relation to yield capacity of the environment for six growing conditions, 3
seasons ±salt (closed and open symbols respectively). The three CBF3 lines in
competition with their WS wild-type background (A), the two SOS1 lines in competition
with wild-type Col(gl), and the two M6PR in competition with wild-type Col (C). All
values are the mean of 14 replicate populations, each planted with an initial transgene
frequency of 50%. Overall yield capacity was calculated as mean seed yield of all
genotypes for each of the three experiments multiplied by 50% (the starting transgene
frequency). Yield capacity of each transgenic line was calculated as the mean seed yield
multiplied by the mean transgene frequency measured by selectable marker phenotyping
of progeny seed………………………………………………….……………………...108
Figure 4.1. The tiered hazard characterization system used by EFSA to address ecological
concerns related to persistence, ferality, and invasiveness of transgenic plants, whether
the genetically engineered crop itself or recipient wild plants due to gene flow.
Reproduced from the EFSA 2010 report, Guidance on the environmental risk assessment
of genetically modified plants…………………………………………………...….…..129
Figure S.1. The proportion of plants within ten natural Arabidopsis thaliana populations at three
broad developmental and reproductive stages. These categories were: pre-bolting, which
included all plants from the cotyledon to rosette stages of vegetative development; bolted,
which included plants which had bolted but not yet flowered and flowering plants
without developed siliques; and mature, which included flowering plants with well
developed siliques and plants undergoing senescence………………………………….162
Figure S.2. The density in plants per square meter of ten natural Arabidopsis thaliana
populations near East Lansing, MI. All ten populations were in disturbed ground along
the margins of a broccoli field at the ‘Sand Hill site’ near the Michigan State University
Tree Research Facility. Populations were distributed over several acres. Population
density was determined by counting all Arabidopsis thaliana plants within a 26x26 cm
quadrant…………………………..……………………………………………………..163
Figure S.3. Photographs of the summer/fall 2007 preliminary field trial. The field layout with the
tray-in-tray design allowing subsoil watering by trickle hose (A) and the protective lids
temporarily deployed prior to inclement weather which could endanger the plot (B)…168
xi
Figure S.4. Fitness of transgenic plants in competition with their wild-type background ecotype
in a preliminary field study fall 2007. Fitness was calculated based on selectable marker
screening of progeny seed from mixed populations of CBF3 lines A28 and wild-type WS
and line A40 and WS (A and B respectively) and M6PR line M2-1 and wild-type Col
(C). A fitness of 1 would indicate seed production equal to wild-type (red line). Both
CBF3 lines showed significantly reduced fitness in competition (P<0.05) while M6PR
line M2-1 showed increased fitness (P<0.05)………………..………………………...169
Chapter 1
Literature Review
xii
Introduction
The Green Revolution was driven in part by selection for a handful of genes for
dwarfism, increased yield and shorter growing seasons (NRC 2002). While the new varieties
carrying these genes fed the rising human population, the environmental, social, and cultural
repercussions of the revolution remain to this day. Today plant breeders possess a much larger
toolbox of genes, and the potential to move genes seemingly at will without species barriers.
Given the sweeping and often unexpected changes that resulted from the use of a handful of
genes (NRC 2002), what changes could result from today’s ever expanding toolkit? The first
generation of genetically modified (GM) crops, mostly herbicide- and pest-resistant varieties,
targeted a specific problem with a single transgene (Chua and Tingey 2006). Herbicide tolerance
was usually conferred by an altered enzyme and insect pest resistance by the production of
proteins that while toxic to insects are merely inert within the plant cells (Warwick et al. 2009).
Although there are now commercialized varieties with up to eight stacked transgenes to resist a
variety of insect pests as well as herbicides, the transgenes involved remain simplistic in their
function. The next generation of GM crops to be commercialized will target larger and more
complex issues and arguably one of the most important of these will be increasing crop tolerance
to abiotic stresses, including drought, heat waves, frost, and soil salinity that cause an estimated
50% global yield reduction (Wang et al. 2003).
Salinity Stress
Salinity stress is one of the fastest increasing abiotic stresses on arable land across the
globe. Salinity currently affects at least 20% of crop land, but more than 50% of arable land is
predicted to be under salinity stress by 2050 (Wang et al. 2003, Chinnusamy et al. 2005).
Irrigation has added 110 million ha of agricultural land and played a vital role in increasing food
resources globally, but this increase is not without side effects. Increased irrigation usage in arid
and semi-arid environments has resulted in increasing salt stress levels. Arid and semi-arid areas
1
comprise 40% of the world’s arable land, but do not receive enough annual rainfall to provide
the necessary leaching needed to remove excess salts from the soil (Smedema and Shiati 2002).
Crop irrigation can affect the salinity levels of soils by several means. The use of
irrigation water pumped from deep fossil aquifers, which tend to be more saline than typical
surface or ground waters, directly applies additional salt loads to the soil surface. Overuse of
irrigation can waterlog soils, which brings soil borne salts closer to the surface where they can
accumulate due to evaporation, an issue currently impacting ~25% of irrigated land in semi-arid
regions worldwide (Smedema and Shiati 2002). Lastly by altering water tables and affecting
groundwater movement, irrigation can unlock fossil salt deposits, left behind from past marine
events (Smedema and Shiati 2002). As market pressure to produce more food, feed, fiber and
fuel rises, the high salinity issues facing farmers today will continue to increase as irrigation is
adopted in previously non-irrigated regions and more salt is deposited in already irrigated lands.
This will mean that salinity-tolerant varieties will become more important to farmers trying to
meet the demands of a growing population.
Genetic engineering for abiotic stress tolerance
Submissions to the USDA for field trials of abiotic stress related traits have increased
significantly in the last decade, indicating a concerted effort on the part of plant biotechnologists
and breeders to produce crops with enhanced abiotic stress tolerance (Figure 1.1). Abiotic
stresses can result from a broad range of environmental factors such as heat, drought, cold and
salinity and numerous candidate genes have been described to potentially increase tolerance to
one or more types of abiotic stress (Zhang et al. 2004, Sreenivasulu et al. 2007, BhatnagarMathur et al. 2008, Warwick et al. 2009). The mechanisms of resistance supplied by these
candidate genes vary and include such functions as membrane protection, stress signaling,
protein and RNA stabilization, transcriptional activation, and the detoxification of toxic
molecules and free radicals (Wang et al. 2003).
2
Figure 1.1. The yearly number of submissions to the USDA for field trial of abiotic stress related
traits from 1990-2013. Submission data was collected from the Information Systems for
Biotechnology (http://www.isb.vt.edu/search-release-data.aspx) on January 23, 2013.
3
Although the causes of abiotic stress differ, many have similar effects upon plant tissues.
Drought, salinity and freezing all create osmotic stress resulting from the loss of water available
for homeostasis and chemical reactions (Wang et al. 2003, Warwick et al. 2009). Oxidative
stress results from reactive oxygen species or free radicals, which can be produced by plant cells
under stress from high temperatures, drought or salinity. Therefore the introduction of one gene
could potentially alter resistance against multiple stresses. In addition, multiple stresses, such as
drought and heat, can combine to create greater levels of stress (Mittler 2006). These
combinations could occur in many ways, any of which might influence the effectiveness of a
given abiotic stress resistance gene.
At the time of writing, the first and only commercialized genetically engineered crop with
enhanced abiotic stress tolerance is the corn line MON 87460 (APHIS 2011a,b), a parent to
Monsanto’s DroughtGard hybrid corn lines. These corn hybrids express a cold shock protein
from Bacillus subtilis named CspB (Reeves 2010). This protein belongs to a class of RNA
chaperones which prevent RNA misfolding, allowing cells to continue to produce vital proteins
while under environmental stress (reviewed by Horn et al. 2007). Under water deficit stress,
hybrids constitutively expressing CspB had increased growth rates, maintained higher
chlorophyll content and photosynthetic rates and significantly out-yielded the untransformed
conventional hybrids (Castiglioni et al. 2008). No cost of resistance was observed in high
yielding environments. CspB was also shown to increase abiotic stress tolerance (cold, heat and
drought) in Arabidopsis thaliana and rice (Castiglioni et al. 2008). With no reported cost of
resistance and conferring increased tolerance to multiple abiotic stresses in both monocot and
dicot species, cold shock proteins show potential to increase stress tolerance in many crops.
Another abiotic stress tolerant crop in the regulatory process is freezing tolerant hybrid
Eucalyptus (Eucalyptus grandis x E. urophylla) (Nehra & Pearson 2011). Interest in Eucalyptus
as a short rotation woody crop for fiber and biofuel production has been hindered by the tropical
plant’s low tolerance to cold temperatures. Previous research had shown that over-expression of
Eucalyptus CBF1a and CBF1b homologs increased freezing tolerance; however, the transformed
4
plants showed significant negative phenotypic effects that rendered them unsuitable for
commercialization (Navarro et al. 2011). The genetically engineered Eucalyptus lines submitted
for deregulation express the abiotic stress response transcription factor C-repeat binding factor 2
(CBF2) gene from Arabidopsis thaliana under the cold inducible promoter rd29A. The
transformation cassette also includes a barnase gene linked to an anther-specific promoter
PrMC2 to confer pollen sterility. Transgenic lines showed significantly less winter die-back than
the untransformed background: at multiple sites after 5 years the transgenic lines were 40-50 feet
tall, while the background line had died back to less than 1 ft over the winter (Nehra & Pearson
2011). Under freeze-free conditions, the transgenic lines had small but significant reductions in
growth compared to background. Flowers on the transgenic lines produced no viable pollen. In
February 2013, the USDA released a notification of intent to perform an Environmental Impact
Statement on the transgenic lines submitted for deregulation (APHIS 2013). Thus freeze-tolerant
Eucalyptus could become the next abiotic stress tolerant crop to be released to the market. The
addition of the pollen sterility trait could aid that process by reducing the ecological concerns
related to feral crop plants and gene flow to compatible relatives that are frequently raised about
crops with abiotic stress tolerance traits.
Ecological risk assessment of abiotic stress tolerance enhanced crops
Although many studies of the ecological risks of genetically modified crops are focused
on the possibility of gene flow, gene flow in and of itself does not constitute an environmental
harm (EFSA 2012). While exact definitions of ecological harm vary, most consider a reduction
in biodiversity to be harmful (Sanvido et al. 2012). Thus, an ecological harm could result from
gene flow if it resulted in a negative impact on biodiversity. A reduction in biodiversity could
occur at the species level within an ecosystem or at the genetic level for an individual species
(Hancock 2011). The effects of abiotic stress tolerance transgenes that could alter biodiversity
include altered fitness, increased weediness or invasiveness, and range expansion (Tiedje et al.
1989, Hancock et al. 1996, Ellstrand 2003, Weaver & Morris 2005, Hails & Morley 2005).
Unlike most first generation transgenes used for crop improvement, where the gene product
5
directly conferred the desired phenotype, e.g. herbicide resistance or Bt-mediated pest resistance,
abiotic stress tolerance traits have the potential to affect a wide range of plant growth and
developmental functions (Chua and Tingey 2006, Ellstrand 2003). Thus risk assessment of
abiotic stress tolerant crops will require careful consideration of possible secondary and fitness
effects due to transgene expression on recipient plants, for both the transformed crop and wild
compatible relatives which could receive the transgene through pollen-mediated gene flow.
The transgene could confer significant competitive advantages allowing the transformed
crop to become feral, growing along field margins and in ruderal areas, and possibly invasive in
natural environments. Transgene movement into a compatible wild relative could give recipient
plants a selective advantage under conditions of abiotic stress (Hancock et al. 1996, Ellstrand
2003). This advantage could result in the transgene establishing within wild germplasm. For a
transgene that conferred a strong selective advantage, establishment could further alter the wild
gene pool as other crop genes are also introgressed through linkage drag (Ellstrand 2003). The
transgene, and possibly other crop genes, could allow the recipient plants to become weeds in
agricultural systems and invasive in natural ecosystems. If the transgene instead conferred a
selective disadvantage in natural ecosystems, high rates of gene flow from nearby transgenic
crop fields could result in demographic swamping and in cases of transgenes with severely
negative fitness effects, localized extinction of compatible relatives (Hancock 2011). However,
even with a strongly negatively selected transgene, this effect would likely only occur to weedy
or wild relatives that grow along field margins due to the inability of the transgene to persist
outside of agricultural settings. Thus, local biodiversity could be impacted by the fitness effect of
the transgene on recipient plants, whether those effects are positive or negative.
In breeding for abiotic stress tolerance, plant breeders will be potentially removing a
factor that limits an individual plant’s survival and fecundity, and a species range and
distribution. Salt tolerance has been found to be a key attribute in the invasiveness of a number
of plant species. Together with the increasing levels of salinity in rivers due to irrigation, and the
use of salt as a de-icing agent, salt tolerance has enabled some species to invade habitats that
6
would otherwise be resistant to invasion (Thompson 1991, Bauer and Geber 2002, Glenn et al.
1998, Glenn and Nalger 2005). Proper risk assessment of these second generation crops
expressing abiotic stress tolerance traits will be vital to their successful introduction while
meeting ecological concerns over increased invasiveness.
Feral crops
The addition of transgenic traits to domesticated crops, has led to concerns that
volunteers could become feral and become weedy or invasive in agricultural and natural
ecosystems. Although feral or weedy versions of domesticated crops have been widely reported,
the process by which domesticated crops become feral has not been extensively researched
(Gressel 2005). This process of de-domestication has been classified into two categories
dependent upon the source of the genetic changes needed to revert to a weedy phenotype.
Endoferality occurs when a domesticated species is able to become feral on its own due to
selection on existing or new genetic variation leading to the loss of domestication traits. While
this has been reported and studied widely in animals, feral pigs for example, it has been less
studied in plants (Gressel 2005) as most cases of crop ferality have been linked to the influx of
wild genes and are thus the result of exoferality (Ellstrand 2003).
The best characterized example of endoferality occurred with wheat (Triticum aestivum)
in the Tibetan highlands of China (Ghimire et al. 2006). This semi-wild wheat is found as an
agricultural weed in barley and wheat fields and has been deemed a separate subspecies Triticum
aestivum ssp. tibetanum (Shao et al. 1983). This subspecies is believed to have evolved its feral
traits on its own without input from wild relative as the region is geographically isolated with no
compatible relatives (Chen et al. 1991, Sun et al. 1998). Also the subspecies is hexaploid and
there are no wild hexaploid progenitor species with from which it could have received genetic
material (Ayal and Levy 2005).
The evolution of endoferality was examined in the cultivated radish (Raphanus sativus)
(Campbell & Snow 2009). Experimental cultivated radish plots were established >1km from the
nearest wild radish (R . raphanistrum ) populations and were allowed to self seed for four
7
generations (Campbell & Snow 2009). Phenotypic assessments for traits associated with feral
crops (early flowering, reduced root biomass, and high seed production) were conducted at each
generation. Three populations went extinct, while two were observed to contain early flowering
individuals by the third generation. However, these individuals were shown to be inadvertent
wild-crop hybrids. To examine the endoferality potential of cultivated radish, greenhouse
populations were grown under strong selective pressure for early flowering for two generations,
with all but the first 10% of plants to flower culled. Two of three replicate populations had
significantly earlier flowering and increased seed yield in common garden studies relative to
control populations that had undergone random 90% culling. However, further gains in feral
traits were not observed, leading the researchers to consider that the genetic variability of the
cultivate radish was insufficient for rapid development of endoferality. The authors postulated
that the domestication process appears to reduce genetic diversity and eliminate weediness traits
(such as seed shattering) resulting in the rare occurrence of natural endoferality in plants and the
difficulty in artificially inducing it. This indicates that hybridization and introgression with
compatible wild relatives are the primary concerns for possible ecological impacts from the
introduction of abiotic stress tolerance enhanced transgenic crops.
Crop x Wild hybridization
When outside genetic sources, i.e. wild relatives, supply the genes needed to dedomesticate, the process is termed exoferality. Hybridization between cultivated and wild
relatives has been documented in a wide number of crop species. As early as 1965, crop-wild
hybrids were found in wheat, sorghum and barley (Harlan 1965). From the 1970s through the
1990s, hybrids were found in almost every category of crop plants from grains to root crops (Chu
and Oka 1970, Zohary 1971, Ladizinsky and Zohary 1971, Brunken et al. 1977a, Brunken et al.
1977b, Klinger et al. 1991). Hybridization between closely related crop and wild species has
resulted in feral hybrid populations growing along the edges of fields or within the field from
shed seed. Thus the risk of transgene flow, despite the species barriers between cultivated and
wild plants potentially causing loss of fitness and partial sterility, exists where hybrid
8
populations can be maintained by high gene flow from crop sources (Wright 1969, Slatkin 1985,
Hails and Morley 2005, Ellstrand 2003). Once hybrids form, introgression into the wild species
can occur through repeated backcrossings. The conditions required for hybrids to form exist at
any large farm where wild interfertile relatives persist as weeds within the field or along its
margins.
One of the most well studied examples of wild-crop hybridization and its effects on plant
phenotype and fitness is with sugar beet (Beta vulgaris ssp. vulgaris) and its wild relative sea
beet (B. vulgaris ssp. maritima). Insufficient isolation distances in sugar beet seed production
areas from natural sea beet populations resulted in an outbreak of weedy sugar beets in fields
across Europe in the 1980s and 1990s (Hornsey & Arnold 1979, Longden 1989, Ford-Llyod
1995). Where sugar beet grows as a biennial, the hybrids bolted, flowered and set seed in one
year (Boudry et al. 1994, Mucher et al. 2000). These weedy beets decreased farm yields and
caused difficulties for processing facilities resulting in millions of dollars worth of economic
losses. Control of these weedy beets proved difficult as seed banks rapidly built up in soils and
remained for several years (Longden 1993). Later gene flow studies determined that the isolation
distance for sugar beet seed production at the time (1km), fell far short the distances viable sea
beet pollen was able to travel (up to nearly 10km) (Fenart et al. 2007). Genetic marker analysis
not only confirmed the sea beet paternal origins of the weedy beets, but also found widespread
multidirectional gene flow (wild↔weedy ↔cultivated) (Desplanque et al. 1999, Bartsch et al.
1999, Viard et al. 2002, Viard et al. 2004, Andersen et al. 2005, Fenart et al. 2008). This level of
gene flow and the difficulties found in preventing it, raised ecological concerns related to the
future release of transgenic sugar beets in Europe (Desplanque et al. 2002).
The evolutionary and ecological consequences of hybridization have been studied in
sunflower (Helianthus annuus) and its compatible wild relatives. Sunflower is an annual native
to the Great Plains and Southwest regions of North America which occurs in cultivated, weedy
and wild types (Rieseberg & Seiler 1990). Hybridization between cultivated sunflower and wild
H. annuus can occur at rates of 27-42% along field margins, with gene flow detectable out to
9
1km, and significant persistence of crop alleles in wild populations for at least five generations
after initial hybridization (Arias & Rieseberg 1994, Whitton et al. 1997). Hybridization between
related wild and weedy annual and perennial Helianthus species has been studied for over 60
years (Heiser 1947). In co-occurring populations of H. annuus and the wild annual prairie
sunflower (H. petiolaris), hybrid swarms have been frequently observed despite differences in
initial flowering times and high levels of pollen and seed sterility among hybrids. Backcrossed
and fully introgressed individuals are also observed and genetically confirmed in hybrid zones,
indicating bidirectional gene flow among both parental species and the hybrids (Heiser 1947,
reviewed in Rieseberg et al. 2007). The impacts of such hybridization can include heterosis,
increased genetic diversity, and reduced levels of negative mutations (Rieseberg et al. 2007).
These hybridization effects could increase the adaptability of the hybrids relative to the
parent species, possibly resulting in speciation, while introgression of locally adapted germplasm
has been proposed as a means of parental range expansion (Heiser 1947). Three species, H.
anomalus, H. deserticola, and H. paradoxus, are the result of selection on transgressive hybrid
(H. annuus x H. petiolaris ) individuals, and are now ecologically isolated from the parent
species inhabiting deserts, sand dunes and saline wetlands respectively (Rieseberg et al. 1990b,
1991a, 1991b, reviewed by Rieseberg et al. 2007). Expansion of H. annuus into Texas was
facilitated by introgression of H. debilis ssp. cucumerifolius germplasm, resulting in the weedy
subspecies H. annuus ssp. Texanus (Rieseberg et al. 1990a, 2007). These introgressed regions
conferred indeterminate branching and resistance to damage from seed midges (Neolasioptera
helianthis) which can destroy 90% of seed in a field of H. annuus in central Texas. Studies with
sunflower have shown widespread hybridization between cultivated, weedy, and wild relatives
resulting in evolutionary and ecological impacts including increased fitness, range expansion,
and speciation.
Wild x GM crop hybridization
Examples of natural hybridization between genetically modified crops and compatible
wild relatives are limited for two reasons. First, a relatively low number of crop species have
10
deregulated transgenic lines on the market (James 2012). Second, most transgenic crops are
currently grown away from their origins of domestication and so are in regions low in
compatible wild-relatives (e.g. the majority of GM maize and soybean acreage, Hancock 2012,
James 2012). However, this trend is unlikely to continue as the acreage of genetically modified
crops continues to increase and a more diverse range of transgenic crops become
commercialized.
The first detection of transgene introgression into a wild relative was observed in 2005 in
Canada between transgenic glyphosate-resistant canola, Brassica napus, and its weedy relative,
Brassica rapa (Warwick et al. 2008). The first detected hybridization between GM canola and
the wild species was observed in 2001 (Beckie et al. 2003). A follow-up study monitored
populations of volunteer canola and wild B. rapa growing along commercial field margins from
initial planting in 2000 to 2005, with glyphosate herbicide applied only once after 2000. Despite
the lack of fresh seed input and low level of selective pressure, there was successful introgression
of the herbicide resistance gene into the wild gene pool, as evidenced by occurrence of a plant
isolated in 2005 that was morphologically and genetically, by AFLP analysis, B. rapa, but
contained the gene for glyphosate resistance (Warwick et al. 2008).
While this detection occurred once, volunteer canola is widespread, with normal in-field
seed loss estimated from 3,000 to 10,000 seeds per meter square (Harker et al. 2006). The
planting of herbicide resistant canola in Canada began in 1995, and since then, double- and even
triple-herbicide resistant canola volunteers have been found as farmers switched from one
herbicide resistant line to another over time to control both volunteers, both in field and along
margins, and developing resistance in other weeds (Simard et al. 2005, Knispel et al. 2008). The
production of these multi-resistant volunteers and the successful introgression of the herbicide
tolerance gene into a wild relative, indicates a high level of gene flow between compatible plants
both within field and between the field and margins.
Transgene Fitness effects
11
While most risk assessments of first generation transgenic crops focused on the potential
for transgene movement via pollen and seed, the rate of transgene establishment outside of
cultivation is governed not by the stringency of containment procedures, but by the selective
advantage or disadvantage resulting from transgene expression (Hancock et al. 1996, Ellstrand
and Hoffman 1990, Haygood et al. 2004). Current GM crops include pathogen resistance and
insect resistance, useful traits to any plant under attack by biotic agents (Warwick et al. 2009).
Herbicide resistance, while unlikely to be of selective advantage in an environment without
human presence, could be a great advantage in a weed trying to survive in an edge habitat in
agricultural lands. The abiotic resistance genes currently under development for the second
generation of GM crops pose potentially even greater advantages, as abiotic stress is often the
greatest limiter of a plant species potential range (Davis and Shaw 2001, Ellstrand 2003, Thuiller
et al. 2005, Warwick et al. 2009). Given the high impacts abiotic stress events like heat waves,
droughts and freezes can have on plant reproduction, the selective pressures of such stresses
could give a crop-wild hybrid possessing a transgene for abiotic stress tolerance a significant
fitness advantage and thus hasten introgression of the transgene into the wild germplasm.
Measuring impacts on fitness
Fitness refers to the average contribution of an allele to succeeding generations. Changes
in fitness have frequently been measured in one of two manners (Bourguet et al. 2004). First, a
study can be performed comparing components of fitness (traits that directly influence fecundity)
between wild-type lines and lines containing the gene of interest. This method was been widely
used in the assessment of first generation GM crops to establish substantial equivalence for
regulatory purposes (NRC 2002). Biomass, branching pattern, fruit set, seed production, timing
of developmental stages and many other factors were reported in these papers and used to assess
fitness.
The second approach calculates the contribution of an allele to future generations directly
determining the fitness costs or benefits within a given population (Bourguet et al. 2004). Using
this direct method requires a population of plants to be carried through several generations and
12
necessitates that the population possess a known number of homozygous positive and
homozygous negative parent plants in the first generation (Gilliland et al. 1998, Roux et al.
2005). The fitness costs can then be calculated as percentage of total seed production of the
population for each group. While this method could be problematic for studies involving plant
species with long generation times, Arabidopsis is well suited to this method. With short
generation times, high seed production and the ability to grow in high density stands, it combines
all the traits needed for multigenerational risk assessment studies. These traits make Arabidopsis
thaliana an excellent candidate to perform a primary assessment of the fitness costs of transgenes
likely to be used in second generation abiotic stress resistant crops.
The duration of the experiment can be an important factor in the accuracy of a study’s
results. Most early GM trials designed to determine the selective advantage of a transgene were
performed for only one or two generations (reviewed by Bergelson and Purrington 1996). Longer
studies using a variety of mutants found fitness effects which would have gone undetected in
shorter studies (Gilliland et al. 1998, Asmussen et al. 1998, Roux et al. 2005). In analyzing the
functional role of three Arabidopsis thaliana actin genes, the fitness costs of their knockouts did
not appear until the second and third generation (Gilliland et al. 1998, Asmussen et al. 1998).
Another study used six herbicide resistant EMS mutant lines that had been studied previously,
and tested them in competitive environments over seven generations (Roux et al. 2005). Fitness
costs of resistance were found to change over time and across densities in some lines. One
mutation found previously to cause negative fitness (Bergelson and Purrington 2002) was
actually neutral over multiple generations (Roux et al. 2005). Another line exhibited densitydependent fitness costs. Its fitness cost increased from 38% to 64% as the prevalence of plants
containing it decreased in the population. These studies show that when attempting to determine
fitness costs it is best to use multiple generations, large populations, and a high number of
replicates (Roux et al. 2005, Whitlock 2000). These considerations were used in the design of
the field and greenhouse experiments of this dissertation.
Transgenes assessed
13
As discussed earlier, a variety of genes that function through differing modes of action
are in development to improve crop abiotic stress tolerance. For the risk assessment studies
described herein, three diverse transgenes shown to increase salinity tolerance in the growth
chamber were selected: (i) the transcription factor [C-repeat binding factor 3/drought responsive
element binding factor 1a (CBF3/DREB1a)], (ii) the plasma membrane Na+/H+ antiporter [Salt
Overly-Sensitive 1 (SOS1)], and (iii) the metabolic enzyme [mannose-6-phosphate reductase
(M6PR)]. Salinity stress related transgenes were chosen due to the large and increasing impact of
salt stress on crop yield worldwide. While cold, heat and drought stress are generally transient
stresses in farmer fields, occurring for generally a matter of days, in irrigated arid and semi-arid
regions salt stress occurs continually and increases due to insufficient rainfall to leach excess salt
from the soil (Smedema and Shiati 2002). The three transgenes were assessed for secondary
effects of transgene expression that could impact plant fitness and transgene establishment in
natural populations.
CBF3/DREB1a. Previous research has shown that upon exposure to cold or dehydration,
a set of genes termed COR (cold-regulated) or DR (dehydration-responsive) genes is rapidly
induced. Sequence analysis of the promoters of these COR/DR genes in Arabidopsis thaliana
revealed the presence of a conserved motif, TACCGACAT, which was simultaneously
discovered by two groups and named both the C-repeat element and the dehydration-responsive
element (DRE) (Baker et al. 1994, Yamaguchi-Shinozaki et al. 1994). Yeast one-hybrid
screening revealed this sequence to be the target of a family of abiotic stress-responsive
transcription factors, named C-repeat binding factors (CBF) / drought responsive element
binding factors (DREB) (Stockinger et al. 1997, Liu et al. 1998, Yang et al. 2005). Among the
genes discovered in the family are CBF1/DREB1b, CBF2/DREB1c, and CFB3/DREB1a
(Kasuga et al. 1999, Gilmour et al. 2000). All three transcription factors have been found to be
functionally redundant, with the expression of each resulting in the induction of COR gene
expression, phenotypic changes in growth and development, and increased plant tolerance to
drought, cold, and salinity (Gilmour et al. 2004).
14
The CBF3/DREB1a transcription factor was found by northern and microarray analysis,
to activate not only known genes in the cold regulated (COR) regulon controlled by the CBF
family of transcription factors, but also a number of other genes closely linked to drought,
salinity and osmotic stress tolerance (Jaglo-Ottosen et al. 1998, Kasuga et al. 1999, Gilmour et
al. 2000, Seki et al. 2001, Fowler and Thomashow 2002, Maruyama et al. 2004, Chan et al.
2012). CBF3/DREB1a over-expression resulted in the constitutive expression of COR genes as
well as other abiotic stress tolerance genes such as dehydrins and metabolic enzymes for the
production of compatible solutes (Gilmour et al. 2000, Fowler and Thomashow 2002, Chan et al.
2012). Given that the CBF regulon includes other transcription factors, such as members of the
ZAT and RAP gene families, over-expression of CBF3 results in a cascade of gene expression
activations and repressions that influence over 1300 genes (Seki et al. 2001, Maruyama et al.
2004, Vogel et al. 2005, Chan et al. 2012).
Over-expression of CBF3/DREB1a increases abiotic stress resistance of transgenic
Arabidopsis through this activation of stress tolerance genes and pathways. In abiotic stress
experiments, increased freezing, drought and salinity tolerance were observed (Jaglo-Ottosen et
al. 1998, Kasuga et al. 1999). Transgenic lines developed levels of soluble sugars and proline
three and five times higher than the controls, compounds which have been demonstrated to
increase tolerance to dehydration and salinity stress (Gilmour et al. 2000, Cook et al. 2004).
CBF3 over-expression was also associated with significant secondary effects resulting in
a clear CBF3 over-expression phenotype for Arabidopsis thaliana. Varying degrees of dwarfism,
delayed flowering, and reduced seed production, are seen in the majority of CBF3 overexpression lines (Kasuga et al. 1999, Gilmour et al. 2000, Chan et al. 2012). In addition,
transgenic lines exhibited changes in plant architecture including decreased axillary shoot
formation, shorter petiole lengths, and leaves that remain close to the ground (Gilmour et al.
2000, Chan et al. 2012). These phenotypic and transcriptomic effects could have impacts on
plant fitness and the ability to establish under field conditions.
15
SOS1. Mutational screening of A. thaliana revealed a series of loci, Salt Overly Sensitive
1 through 3 (SOS1-SOS3) that showed a highly salt-sensitive phenotype when mutated (Wu et al.
1996, Zhu et al. 1998). Crossing of the mutant lines revealed that the loci act together to regulate
cellular sodium ion levels. Positional cloning, sequence, and protein analyses revealed SOS1 to
encode a putative plasma membrane Na+/H+ antiporter (Shi et al. 2000, Shi et al. 2002). SOS3
was determined to encode a calcium binding protein, while SOS2 encodes a protein kinase
(Halfter et al. 2000). The SOS pathway begins with high Na+ concentration inducing the release
2+
of a Ca
signal which binds to SOS3 causing it to complex with SOS2 (Halfter et al. 2000). The
SOS2-SOS3 complex then activates SOS1 via phosphorylation causing the antiporter to begin
the extrusion of Na+ ions from the cytosol to the extracellular space (Shi et al. 2000, Qiu et al.
2002, Guo et al. 2004). Recent work has also found that SOS1 can be activated by RCD1, an
oxidative-stress regulator, bringing together Arabidopsis thaliana’s response to salinity stress
and oxidative stress (Katiyar-Agarwal et al. 2006). SOS1 has also been found to be constitutively
expressed in salt cress (Thellungiella halophila), a relative of Arabidopsis which grows in salt
marshes, and may be one of the reasons why plants of the species are able to survive 500 mM
NaCl concentrations (Taji et al. 2004).
Transformed Arabidopsis plants constitutively expressing the SOS1 gene showed higher
levels of SOS1 transcript than wild-type plants with or without stress (Shi et al. 2003). SOS1
transcript levels in over-expression and wild-type lines increased in the presence of salt,
indicating post-transcriptional regulation of SOS1 mRNA levels via increased mRNA stability in
the presence of reactive oxygen species produced under salinity stress (Shi et al. 2003, Chung et
al. 2008). Transcriptomic analysis of SOS1 over-expression lines detected roughly 600 genes
with altered expression profiles in the absence of salinity stress, including down-regulation of a
large number of oxidative and redox related genes (Chan et al. 2012). These results corroborated
prior studies indicating interactions between the SOS response pathway and oxidative and redox
signaling pathways (Katiyar-Agarwal et al. 2006, Verslues et al. 2007). Minimal overlap
16
between CBF3 and SOS1 transcriptomic changes indicate that the SOS signaling pathway is
independent of CBF, ABA and MYB regulated pathways (Kamei et al. 2005, Chan et al. 2012).
SOS1 over-expression plants had reduced levels of sodium ions in both cellular tissues
and xylem sap after salt exposure than wild-type plants under the same stress (Shi et al. 2003).
The over-expression lines showed greater ability to germinate, grow and set seed under salinity
stress than wild-type plants (Shi et al. 2003, Chan et al. 2012). Without the presence of salt
stress, no phenotypic differences were observed between the transgenic and WT plants. In the
presence of salinity stress, SOS1 over-expression lines were better able to maintain cellular
functions than wild-type plants, with reduced impacts to root growth, total protein and
chlorophyll levels, and photosynthetic efficiency (Shi et al. 2003). By maintaining cellular
homeostasis under abiotic stress, SOS1 over-expression could alter plant fitness and
establishment in the field.
M6PR. The mannose-6-phosphate reductase (M6PR) gene was cloned from celery,
Apium graveolens L., a salt-tolerant plant whose tolerance has been linked to its production of
mannitol as a primary product of photosynthesis (Everard et al. 1997). By catalyzing the first
irreversible reaction in the mannitol biosynthetic pathway, M6PR was shown to be the key
enzyme in the production of mannitol (Zhifang and Loescher 2003). Mannitol has been found to
act as a compatible solute, counterbalancing the osmotic effects of sodium and chloride ions, as
well as an osmoprotectant, reducing the damage caused by free radicals produced under salinity
stress (Shen et al. 1997, Sickler et al. 2007, Zhifang and Loescher 2003). Additionally,
transcriptomic analysis of M6PR over-expression lines in the absence of salt found over 1700
genes with altered expression levels relative to wild-type (Chan et al. 2011). Approximately 50%
of these overlapped with those altered by CBF3 over-expression under the same conditions
(Chan et al. 2012). Strong overlap was also detected between transcripts affected by M6PR and
those induced in wild-type plants in response to salinity stress. Also observed was significant upregulation of biotic stress responsive genes by M6PR over-expression (Chan et al. 2011).
Mannitol is not commonly produced by plants but is produced by a number of pathogenic fungal
17
species and thus could be acting as an elicitor of biotic stress response. Together these indicate
multiple levels of protection against the damaging effects of abiotic stress and suggest possible
effects on the response to biotic stress as well.
Constitutive expression of M6PR in A. thaliana resulted in the accumulation of mannitol
and increased salinity tolerance without visible phenotypic effects in the absence of salinity
stress. Mannitol levels ranged from 0.5 to nearly 6 umol per gram fresh weight depending on the
transgenic line, with the metabolite present in all tissues but highest in floral tissues and seeds
(Zhifang and Loescher 2003). Dark room experiments showed no ability to catabolize mannitol
by the transgenic Arabidopsis thaliana plants. Plants expressing M6PR were able to grow and set
seed in salt concentrations up to 300mM NaCl, levels which prevent wild-type seed formation
(Zhifang and Loescher 2003). Transgenic lines were also able to maintain their photosystem II
efficiency at wild-type stress-free levels in the presense of salinity stress and showed reduced salt
stress impacts on CO2 assimilation (Sickler et al. 2007). In the absence of salt stress, the
transgenic 35S-M6PR lines showed no observable differences in phenotype to the wild-type
Columbia ecotype background (Sickler et al. 2007, Chan et al. 2011). These experiments
demonstrated that M6PR confers primary and secondary effects which could significantly alter
plant fitness and confer advantages that might increase the likelihood of transgene establishment
in the natural environment.
Objectives of dissertation
The fitness effects conferred by a transgene play a significant role in the possible
environment risks from a transgenic crop cultivar or recipient wild plants, and thus are closely
evaluated during the environmental risk assessment conducted before any commercialization.
My first objective was to examine whether transgenes involved in different abiotic stress
tolerance mechanisms conferred differing secondary fitness effects and whether those transgene
fitness impacts differed among growing environments. I performed a series of experiments in the
field and greenhouse to determine the secondary fitness effects of three transgenes; CBF3, SOS1
and M6PR, whose expression in Arabidopsis thaliana confers increased tolerance to salinity
18
stress. Each gene represents a different cellular mechanism for tolerance and a possible route for
the production of salt tolerant genetically engineered crops. Given the potentially broad-reaching
effects of altering abiotic stress tolerance via their effects on gene regulation, cellular ion
balance, and carbon metabolism, we expected to find secondary effects of transgene expression
including fitness impacts, either positive or negative. Transgene fitness was assessed under a
variety of field growing environments ranging across spring, summer and fall (described in
chapter 2). Additionally fitness effects were assessed in the greenhouse, in the presence and
absence of salinity stress, and compared to previously performed growth chamber studies
(described in chapter 3).
My second objective was to assess whether transgenic plant performance in competition
with other genotypes, as would occur in the wild, differs from non-competitive pure line
performance, as occurs in agriculture. Plant productivity and fitness in competition with wildtype plants was observed across six generations in the field (detailed in chapter 2) and in three
repeated single-generation greenhouse competition experiments in the presence and absence of
salinity stress (detailed in chapter 3).
The last objective was to examine the implications of the data and methodology,
including previous transcriptomic analysis of the three transgenes, for ecological risk
assessments of abiotic stress tolerance enhancing transgenes (described in chapter 4). This
project measured transgene fitness effects not predicted from prior growth chamber assessments,
determined that competition with wild-type significantly affected the observed fitness of
transgenic plants and developed methodology for the analysis of secondary fitness effects of
future abiotic stress resistance transgenes using the model species Arabidopsis thaliana.
19
LITERATURE CITED
20
LITERATURE CITED
Andersen, N. S., Siegismund, H. R., Meyer, V., & Jørgensen, R. B. (2005). Low level of gene
flow from cultivated beets (Beta vulgaris L. ssp. vulgaris) into Danish populations of sea
beet (Beta vulgaris L. ssp. maritima).Molecular Ecology, 14(5), 1391-1405.
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2011a). Monsanto Company Petition (07-CR-191U) for determination of
non-regulated status of event MON 87460. Draft Environmental Assessment
http://www.aphis.usda.gov/brs/aphisdocs/09_05501p_dea.pdf (accessed March 2013).
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2011b). Determination of Nonregulated Status for MON 87460 Corn (Zea
mays L). http://www.aphis.usda.gov/brs/aphisdocs/09_05501p_det.pdf (accessed March
2013).
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2013). Notice of intent to prepare an environmental impact statement for
determination of nonregulated status of freeze tolerant Eucalyptus lines and notice of
virtual public meetings. Federal Register, 78 (39) 13309- 13312.
http://www.aphis.usda.gov/brs/fedregister/11_01901p_fr_noi.pdf (accessed April 2013).
Arias, D. M., & Rieseberg, L. H. (1994). Gene flow between cultivated and wild
sunflowers. Theoretical and Applied Genetics, 89(6), 655-660.
Asmussen, M. A., Gilliland, L. U., & Meagher, R. B. (1998). Detection of deleterious genotypes
in multigenerational studies. II. Theoretical and experimental dynamics with selfing and
selection. Genetics, 149(2), 727-737.
Ayal, S., Levy, A. A., & Gressel, J. (2005). Wheat Domestication and Dedomestication—What
Are the Odds?. In Crop ferality and volunteerism. CRC Press.
Baker, S. S., Wilhelm, K. S., & Thomashow, M. F. (1994). The 5′-region of Arabidopsis
thaliana cor15a has cis-acting elements that confer cold-, drought-and ABA-regulated
gene expression. Plant molecular biology, 24(5), 701-713.
Bartsch, D., Lehnen, M., Clegg, J., Pohl‐Orf, M., Schuphan, I., & Ellstrand, N. C. (1999). Impact
of gene flow from cultivated beet on genetic diversity of wild sea beet
populations. Molecular Ecology, 8(10), 1733-1741.
Beckie, H. J., Warwick, S. I., Nair, H., & Séguin-Swartz, G. (2003). Gene flow in commercial
fields of herbicide-resistant canola (Brassica napus). Ecological applications, 13(5),
1276-1294.
21
Bergelson, J., Purrington, C. B., Palm, C. J., & LOpez-Gutierrez, J. C. (1996). Costs of
resistance: a test using transgenic Arabidopsis thaliana. Proceedings of the Royal Society
of London. Series B: Biological Sciences, 263(1377), 1659-1663.
Bergelson, J., & Purrington, C. B. (2002). Factors affecting the spread of resistant Arabidopsis
thaliana populations. In Genetically engineered organisms: assessing environmental and
human health effects. CRC Press.
Brauer, J., & Geber, M. A. (2002). Population differentiation in the range expansion of a native
maritime plant, Solidago sempervirens L. International Journal of Plant Sciences, 163(1),
141-150.
Bhatnagar-Mathur, P., Vadez, V., & Sharma, K. K. (2008). Transgenic approaches for abiotic
stress tolerance in plants: retrospect and prospects. Plant Cell Reports, 27(3), 411-424.
Boudry, P., Wieber, R., Saumitou-Laprade, P., Pillen, K., Van Dijk, H., & Jung, C. (1994).
Identification of RFLP markers closely linked to the bolting gene B and their significance
for the study of the annual habit in beets (Beta vulgaris L.). Theoretical and Applied
Genetics, 88(6-7), 852-858.
Bourguet, D., Guillemaud, T., Chevillon, C., & Raymond, M. (2004). Fitness costs of insecticide
resistance in natural breeding sites of the mosquito Culex pipiens. Evolution, 58(1), 128135.
Brunken, J., de Wet, J. M., & Harlan, J. R. (1977a). The morphology and domestication of pearl
millet. Economic Botany, 31(2), 163-174.
Brunken, J. N. (1977b). A systematic study of Pennisetum sect. Pennisetum
(Gramineae). American Journal of Botany, 161-176.
Campbell, L. G., & Snow, A. A. (2009). Can feral weeds evolve from cultivated radish
(Raphanus sativus, Brassicaceae)?. American Journal of Botany, 96(2), 498-506.
Chan, Z., Grumet, R., & Loescher, W. (2011). Global gene expression analysis of transgenic,
mannitol-producing, and salt-tolerant Arabidopsis thaliana indicates widespread changes
in abiotic and biotic stress-related genes. Journal of Experimental Botany, 62(14), 47874803.
Chan, Z., Bigelow, P. J., Loescher, W., & Grumet, R. (2012). Comparison of salt stress
resistance genes in transgenic Arabidopsis thaliana indicates that extent of transcriptomic
change may not predict secondary phenotypic or fitness effects. Plant Biotechnology
Journal, 10(3), 284-300.
Castiglioni, P., Warner, D., Bensen, R. J., Anstrom, D. C., Harrison, J., Stoecker, M., ... &
Heard, J. E. (2008). Bacterial RNA chaperones confer abiotic stress tolerance in plants
22
and improved grain yield in maize under water-limited conditions. Plant
Physiology, 147(2), 446-455.
Chen, P.D., D.J. Liu, G.Z. Pei, L.L. Qi & L. Huang, 1991. The chromosome constitution of three
endemic hexaploid wheats in western China. Proc 7th International Wheat Genetics
Symposium, pp. 75–80.
Chinnusamy, V., Jagendorf, A., & Zhu, J. K. (2005). Understanding and improving salt tolerance
in plants. Crop Science, 45(2), 437-448.
Chu, Y. E., & Oka, H. I. (1970). Introgression across isolating barriers in wild and cultivated
Oryza species. Evolution, 24(2), 344-355.
Chua, N. H., & Tingey, S. V. (2006). Plant biotechnology: Looking forward to the next ten
years. Current Opinion in Biotechnology, 17(2), 103-104.
Chung, J. S., Zhu, J. K., Bressan, R. A., Hasegawa, P. M., & Shi, H. (2008). Reactive oxygen
species mediate Na+‐induced SOS1 mRNA stability in Arabidopsis. The Plant
Journal, 53(3), 554-565.
Cook, D., Fowler, S., Fiehn, O., & Thomashow, M. F. (2004). A prominent role for the CBF cold
response pathway in configuring the low-temperature metabolome of
Arabidopsis. Proceedings of the National Academy of Sciences of the United States of
America, 101(42), 15243-15248.
Davis, M. B., & Shaw, R. G. (2001). Range shifts and adaptive responses to Quaternary climate
change. Science, 292(5517), 673-679.
Desplanque, B., Boudry, P., Broomberg, K., Saumitou-Laprade, P., Cuguen, J., & Van Dijk, H.
(1999). Genetic diversity and gene flow between wild, cultivated and weedy forms of
Beta vulgaris L.(Chenopodiaceae), assessed by RFLP and microsatellite
markers. Theoretical and Applied Genetics, 98(8), 1194-1201.
Desplanque, B., Hautekèete, N., & Van Dijk, H. (2002). Transgenic weed beets: possible,
probable, avoidable?. Journal of Applied Ecology, 39(4), 561-571.
EFSA (2012) Scientific Opinion on a request from the European Commission related to the
prolongation of prohibition of the placing on the market of genetically modified oilseed
rape events Ms8, Rf3 and Ms8 x Rf3 for import, processing and feed uses in Austria.
EFSA J 2878:1-33, http://www.efsa.europa.eu/en/efsajournal/doc/2878.pdf (accessed
April 2013).
Ellstrand, N. C., & Hoffman, C. A. (1990). Hybridization as an avenue of escape for engineered
genes. BioScience, 40(6), 438-442.
Ellstrand, N. C. (2003). Dangerous Liaisons?. Johns Hopkins University Press.
23
Everard, J. D., Cantini, C., Grumet, R., Plummer, J., & Loescher, W. H. (1997). Molecular
cloning of mannose-6-phosphate reductase and its developmental expression in
celery. Plant physiology, 113(4), 1427-1435.
Fenart, S., Austerlitz, F., Cuguen, J., & ARNAUD, J. F. (2007). Long distance pollen‐mediated
gene flow at a landscape level: the weed beet as a case study. Molecular Ecology, 16(18),
3801-3813.
Fénart, S., Arnaud, J. F., De Cauwer, I., & Cuguen, J. (2008). Nuclear and cytoplasmic genetic
diversity in weed beet and sugar beet accessions compared to wild relatives: new insights
into the genetic relationships within the Beta vulgaris complex species. Theoretical and
Applied Genetics, 116(8), 1063-1077.
Ford-Lloyd, B. V. (1995). Sugarbeet and other cultivated beets. In Evolution of crop plants.
Longman Sinagpore, 2nd edition.
Fowler, S., & Thomashow, M. F. (2002). Arabidopsis transcriptome profiling indicates that
multiple regulatory pathways are activated during cold acclimation in addition to the CBF
cold response pathway. The Plant Cell, 14(8), 1675-1690.
Gilliland, L. U., McKinney, E. C., Asmussen, M. A., & Meagher, R. B. (1998). Detection of
deleterious genotypes in multigenerational studies. I. Disruptions in individual
Arabidopsis actin genes. Genetics, 149(2), 717-725.
Gilmour, S. J., Sebolt, A. M., Salazar, M. P., Everard, J. D., & Thomashow, M. F. (2000).
Overexpression of the Arabidopsis CBF3 transcriptional activator mimics multiple
biochemical changes associated with cold acclimation. Plant Physiology, 124(4), 18541865.
Gilmour, S. J., Fowler, S. G., & Thomashow, M. F. (2004). Arabidopsis transcriptional
activators CBF1, CBF2, and CBF3 have matching functional activities. Plant molecular
biology, 54(5), 767-781.
Glenn, E., Tanner, R., Mendez, S., Kehret, T., Moore, D., Garcia, J., & Valdes, C. (1998).
Growth rates, salt tolerance and water use characteristics of native and invasive riparian
plants from the delta of the Colorado River, Mexico. Journal of Arid
Environments, 40(3), 281-294.
Glenn, E. P., & Nagler, P. L. (2005). Comparative ecophysiology of Tamarix ramosissima and
native trees in western US riparian zones. Journal of Arid Environments, 61(3), 419-446.
Ghimire, S. K., Akashi, Y., Masuda, A., Washio, T., Nishida, H., Zhou, Y. H., ... & Kato, K.
(2006). Genetic diversity and phylogenetic relationships among East Asian common
wheat (Triticum aestivum L.) populations, revealed by the analysis of five
isozymes. Breeding Science, 56(4), 379-387.
24
Gressel, J. (2005). Introduction-the challenges of ferality. Crop ferality and volunteerism, 1-7.
Guo, Y., Qiu, Q. S., Quintero, F. J., Pardo, J. M., Ohta, M., Zhang, C., ... & Zhu, J. K. (2004).
Transgenic evaluation of activated mutant alleles of SOS2 reveals a critical requirement
for its kinase activity and C-terminal regulatory domain for salt tolerance in Arabidopsis
thaliana. The Plant Cell, 16(2), 435-449.
Hails, R. S., & Morley, K. (2005). Genes invading new populations: a risk assessment
perspective. Trends in Ecology & Evolution, 20(5), 245-252.
Halfter, U., Ishitani, M., & Zhu, J. K. (2000). The Arabidopsis SOS2 protein kinase physically
interacts with and is activated by the calcium-binding protein SOS3. Proceedings of the
National Academy of Sciences, 97(7), 3735-3740.
Hancock, J. F., R. Grumet, and S. C. Hokanson. (1996). The opportunity for escape of
engineered genes from transgenic crops. HortScience 31:1080–1085.
Hancock, J. (2011). Factors influencing the genetic diversity of plant species and the potential
impact of transgene movement. In Environmental safety of genetically engineered crops.
Michigan State University Press.
Hancock, J. F. (2012). Plant Evolution and the Origin of Crop Species. Cabi.
Harker, K. N., Clayton, G. W., Blackshaw, R. E., O'Donovan, J. T., Johnson, E. N., Gan, Y., ...
& Van Acker, R. C. (2006). Persistence of glyphosate-resistant canola in western
Canadian cropping systems. Agronomy Journal, 98(1), 107-119.
Harlan, J. R. (1965). The possible role of weed races in the evolution of cultivated
plants. Euphytica, 14(2), 173-176.
Haygood, R., Ives, A. R., & Andow, D. A. (2004). Population genetics of transgene
containment. Ecology Letters, 7(3), 213-220.
Heiser Jr, C. B. (1947). Hybridization between the sunflower species Helianthus annuus and H.
petiolaris. Evolution, 1(4), 249-262.
Horn, G., Hofweber, R., Kremer, W., & Kalbitzer, H. R. (2007). Structure and function of
bacterial cold shock proteins. Cellular and molecular life sciences,64(12), 1457-1470.
Hornsey, K. G., & Arnold, M. H. (1979). The origins of weed beet. Annals of Applied
Biology, 92(2), 279-285.
Jaglo-Ottosen, K. R., Gilmour, S. J., Zarka, D. G., Schabenberger, O., & Thomashow, M. F.
(1998). Arabidopsis CBF1 overexpression induces COR genes and enhances freezing
tolerance. Science, 280(5360), 104-106.
25
James, C. (2012). Global Status of Commercialized Biotech/GM Crops: 2012. ISAAA Brief
No.44. ISAAA: Ithaca, NY.
Kamei, A., Seki, M., Umezawa, T., Ishida, J., Satou, M., Akiyama, K., ... & Shinozaki, K.
(2005). Analysis of gene expression profiles in Arabidopsis salt overly sensitive mutants
sos2− 1 and sos3− 1. Plant, Cell & Environment, 28(10), 1267-1275.
Kasuga, M., Liu, Q., Miura, S., Yamaguchi-Shinozaki, K., & Shinozaki, K. (1999). Improving
plant drought, salt, and freezing tolerance by gene transfer of a single stress-inducible
transcription factor. Nature biotechnology, 17(3), 287-291.
Katiyar-Agarwal, S., Zhu, J., Kim, K., Agarwal, M., Fu, X., Huang, A., & Zhu, J. K. (2006). The
plasma membrane Na+/H+ antiporter SOS1 interacts with RCD1 and functions in
oxidative stress tolerance in Arabidopsis. Proceedings of the National Academy of
Sciences, 103(49), 18816-18821.
Klinger, T., Elam, D. R., & Ellstrand, N. C. (1991). Radish as a model system for the study of
engineered gene escape rates via crop‐weed mating. Conservation Biology, 5(4), 531535.
Knispel, A. L., McLachlan, S. M., Van Acker, R. C., & Friesen, L. F. (2008). Gene flow and
multiple herbicide resistance in escaped canola populations. Weed Science, 56(1), 72-80.
Ladizinsky, G., & Zohary, D. (1971). Notes on species delimination, species relationships and
polyploidy in Avena L. Euphytica, 20(3), 380-395.
Liu, Q., Kasuga, M., Sakuma, Y., Abe, H., Miura, S., Yamaguchi-Shinozaki, K., & Shinozaki, K.
(1998). Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA
binding domain separate two cellular signal transduction pathways in drought-and lowtemperature-responsive gene expression, respectively, in Arabidopsis. The Plant
Cell, 10(8), 1391-1406.
Longden, P. C. (1989). Effects of increasing weed‐beet density on sugar‐beet yield and
quality. Annals of Applied Biology, 114(3), 527-532.
Longden, P. C. (1993). Weed beet: a review. Aspects of Applied Biology. 35:185-194.
Maruyama, K., Sakuma, Y., Kasuga, M., Ito, Y., Seki, M., Goda, H., ... & Yamaguchi‐Shinozaki,
K. (2004). Identification of cold‐inducible downstream genes of the Arabidopsis
DREB1A/CBF3 transcriptional factor using two microarray systems. The Plant
Journal, 38(6), 982-993.
Mittler, R. (2006). Abiotic stress, the field environment and stress combination. Trends in Plant
Science, 11(1), 15-19.
26
Mücher, T., P. Hesse, M. Pohl-Orf, N.C. Ellstrand, and D. Bartsch. 2000. Characterization of
weed beets in Germany and Italy. Journal of Sugar Beet Research, 37: 19–38.
National Research Council (NRC) Staff & Board on Agriculture and Natural Resources Staff.
(2002). Environmental effects of transgenic plants: the scope and adequacy of regulation.
National Academies Press.
Navarro, M., Martinez, Y., Laur, J., El Kayal, W., Marque, and C., Teulieres, C. (2011). Two
EguCBF1 genes overexpressed in Eucalyptus display a different impact on stress
tolerance and plant development. Plant Biotechnology Journal, 9(1), 50-63.
Nehra, N.S. & Pearson, L. (2011). Petition for Determination of Non-regulated Status for Freeze
Tolerant Hybrid Eucalyptus Lines. ArborGen Inc. Summerville, SC.
Qiu, Q. S., Guo, Y., Dietrich, M. A., Schumaker, K. S., & Zhu, J. K. (2002). Regulation of SOS1,
a plasma membrane Na+/H+ exchanger in Arabidopsis thaliana, by SOS2 and
SOS3. Proceedings of the National Academy of Sciences, 99(12), 8436-8441.
Reeves, W. R. (2010). Petition for the Determination of Non‐Regulated Status for MON 87460.
The Monsanto Company, St. Louis MO.
Rieseberg, L. H., & Seiler, G. J. (1990). Molecular evidence and the origin and development of
the domesticated sunflower (Helianthus annum, Asteraceae). Economic Botany, 44(3),
79-91.
Rieseberg LH, Beckstrom-Sternberg S, Doan K (1990a) Helianthus annuus ssp. texanus has
chloroplast DNA and nuclear ribosomal RNA genes of Helianthus
debilis ssp. cucumerifolius. Proceedings of the National Academy of Science USA,
87:593–597
Rieseberg LH, Carter R, Zona S (1990b) Molecular tests of the hypothesized hybrid origin of
two diploid Helianthus species (Asteraceae). Evolution 44:1498–1511
Rieseberg LH, Beckstrom-Sternberg S, Liston A, Arias D (1991a) Phylogenetic and systematic
inferences from chloroplast DNA and isozyme variation in Helianthus sect. Helianthus.
Systematic Botany 16:50–76.
Rieseberg LH, Choi H, Ham D (1991b) Differential cytoplasmic versus nuclear gene flow
in Helianthus. Journal of Heredity, 82:489–493.
Rieseberg, L. H., Kim, S. C., Randell, R. A., Whitney, K. D., Gross, B. L., Lexer, C., & Clay, K.
(2007). Hybridization and the colonization of novel habitats by annual
sunflowers. Genetica, 129(2), 149-165.
27
Roux, F., Camilleri, C., Bérard, A., & Reboud, X. (2005). Multigenerational versus single
generation studies to estimate herbicide resistance fitness cost in Arabidopsis
thaliana. Evolution, 59(10), 2264-2269.
Sanvido, O., Romeis, J., Gathmann, A., Gielkens, M., Raybould, A., & Bigler, F. (2012).
Evaluating environmental risks of genetically modified crops: ecological harm criteria for
regulatory decision-making. Environmental Science & Policy, 15(1), 82-91.
Seki, M., Narusaka, M., Abe, H., Kasuga, M., Yamaguchi-Shinozaki, K., Carninci, P., ... &
Shinozaki, K. (2001). Monitoring the expression pattern of 1300 Arabidopsis genes under
drought and cold stresses by using a full-length cDNA microarray. The Plant Cell, 13(1),
61-72.
Shao, Q., LI, C. and Basang, C. (1983) Semi-wild wheat from Xizang (Tibet). Proc. 6th Int.
Wheat Genet. Symp. Kyoto, 111-114.
Shen, B., Jensen, R. G., & Bohnert, H. J. (1997). Mannitol protects against oxidation by
hydroxyl radicals. Plant Physiology, 115(2), 527-532.
Shi, H., Ishitani, M., Kim, C., & Zhu, J. K. (2000). The Arabidopsis thaliana salt tolerance gene
SOS1 encodes a putative Na+/H+ antiporter. Proceedings of the National Academy of
Sciences, 97(12), 6896-6901.
Shi, H., Quintero, F. J., Pardo, J. M., & Zhu, J. K. (2002). The putative plasma membrane
Na+/H+ antiporter SOS1 controls long-distance Na+ transport in plants. The Plant
Cell, 14(2), 465-477.
Shi, H., Lee, B., Wu, S.J. and Zhu, J.K. (2003) Overexpression of a plasma membrane Na+ ⁄ H+
antiporter gene improves salt tolerance in Arabidopsis thaliana. Nature Biotechnology.
21, 81–85.
Sickler, C. M., Edwards, G. E., Kiirats, O., Gao, Z., & Loescher, W. (2007). Response of
mannitol-producing Arabidopsis thaliana to abiotic stress. Functional Plant
Biology, 34(4), 382-391.
Simard, M. J., Légère, A., Séguin-Swartz, G., Nair, H., & Warwick, S. (2005). Fitness of double
vs. single herbicide-resistant canola. Weed Science, 53(4), 489-498.
Slatkin, M. (1985). Gene flow in natural populations. Annual Review of Ecology and
Systematics, 16, 393-430.
Smedema, L. K., & Shiati, K. (2002). Irrigation and salinity: a perspective review of the salinity
hazards of irrigation development in the arid zone. Irrigation and Drainage
Systems, 16(2), 161-174.
28
Sreenivasulu, N., Sopory, S. K., & Kavi Kishor, P. B. (2007). Deciphering the regulatory
mechanisms of abiotic stress tolerance in plants by genomic approaches. Gene, 388(1), 1.
Stockinger, E. J., Gilmour, S. J., & Thomashow, M. F. (1997). Arabidopsis thaliana CBF1
encodes an AP2 domain-containing transcriptional activator that binds to the Crepeat/DRE, a cis-acting DNA regulatory element that stimulates transcription in
response to low temperature and water deficit. Proceedings of the National Academy of
Sciences, 94(3), 1035-1040.
Sun, Q., Ni, Z., Liu, Z., Gao, J., & Huang, T. (1998). Genetic relationships and diversity among
Tibetan wheat, common wheat and European spelt wheat revealed by RAPD
markers. Euphytica, 99(3), 205-211.
Taji, T., Seki, M., Satou, M., Sakurai, T., Kobayashi, M., Ishiyama, K., ... & Shinozaki, K.
(2004). Comparative genomics in salt tolerance between Arabidopsis and Arabidopsisrelated halophyte salt cress using Arabidopsis microarray. Plant Physiology, 135(3),
1697-1709.
Tiedje, J. M., Colwell, R. K., Grossman, Y. L., Hodson, R. E., Lenski, R. E., Mack, R. N., &
Regal, P. J. (1989). The planned introduction of genetically engineered organisms:
ecological considerations and recommendations. Ecology, 70(2), 298-315.
Thompson, J. D. (1991). The biology of an invasive plant. BioScience, 41(6), 393-401.
Thuiller, W., Lavorel, S., Araújo, M. B., Sykes, M. T., & Prentice, I. C. (2005). Climate change
threats to plant diversity in Europe. Proceedings of the National Academy of Sciences of
the United States of America, 102(23), 8245-8250.
Verslues, P. E., Batelli, G., Grillo, S., Agius, F., Kim, Y. S., Zhu, J., ... & Zhu, J. K. (2007).
Interaction of SOS2 with nucleoside diphosphate kinase 2 and catalases reveals a point of
connection between salt stress and H2O2 signaling in Arabidopsis thaliana. Molecular
and Cellular Biology, 27(22), 7771-7780.
Viard, F., Bernard, J., & Desplanque, B. (2002). Crop-weed interactions in the Beta vulgaris
complex at a local scale: allelic diversity and gene flow within sugar beet
fields. Theoretical and Applied Genetics, 104(4), 688-697.
Viard, F., Arnaud, J. F., Delescluse, M., & Cuguen, J. (2004). Tracing back seed and pollen flow
within the crop–wild Beta vulgaris complex: genetic distinctiveness vs. hot spots of
hybridization over a regional scale. Molecular Ecology, 13(6), 1357-1364.
Vogel, J. T., Zarka, D. G., Van Buskirk, H. A., Fowler, S. G., & Thomashow, M. F. (2005).
Roles of the CBF2 and ZAT12 transcription factors in configuring the low temperature
transcriptome of Arabidopsis. The Plant Journal, 41(2), 195-211.
29
Wang, W., Vinocur, B., & Altman, A. (2003). Plant responses to drought, salinity and extreme
temperatures: towards genetic engineering for stress tolerance. Planta, 218(1), 1-14.
Warwick, S. I., Legere, A., Simard, M. J., & James, T. (2008). Do escaped transgenes persist in
nature? The case of an herbicide resistance transgene in a weedy Brassica rapa
population. Molecular Ecology, 17(5), 1387-1395.
Warwick, S. I., Beckie, H. J., & Hall, L. M. (2009). Gene flow, invasiveness, and ecological
impact of genetically modified crops. Annals of the New York Academy of
Sciences, 1168(1), 72-99.
Weaver, S. A., & Morris, M. C. (2005). Risks associated with genetic modification:–An
annotated bibliography of peer reviewed natural science publications. Journal of
Agricultural and Environmental Ethics, 18(2), 157-189.
Whitlock, M. C. (2000). Fixation of new alleles and the extinction of small populations: drift
load, beneficial alleles, and sexual selection. Evolution, 54(6), 1855-1861.
Whitton, J., Wolf, D. E., Arias, D. M., Snow, A. A., & Rieseberg, L. H. (1997). The persistence
of cultivar alleles in wild populations of sunflowers five generations after
hybridization. Theoretical and Applied Genetics, 95(1-2), 33-40.
Wright, S. (1969). Evolution and the genetics of populations: Vol. 2. The theory of gene
frequencies. University of Chicago Press.
Wu, S. J., Ding, L., & Zhu, J. K. (1996). SOS1, a genetic locus essential for salt tolerance and
potassium acquisition. The Plant Cell, 8(4), 617-627.
Yamaguchi-Shinozaki, K., & Shinozaki, K. (1994). A novel cis-acting element in an Arabidopsis
gene is involved in responsiveness to drought, low-temperature, or high-salt stress. The
Plant Cell, 6(2), 251-264.
Yang, T., Zhang, L., Zhang, T., Zhang, H., Xu, S., & An, L. (2005). Transcriptional regulation
network of cold-responsive genes in higher plants. Plant Science, 169(6), 987-995.
Zhang, J. Z., Creelman, R. A., & Zhu, J. K. (2004). From laboratory to field. Using information
from Arabidopsis to engineer salt, cold, and drought tolerance in crops. Plant
Physiology, 135(2), 615-621.
Zhifang, G., & Loescher, W. H. (2003). Expression of a celery mannose 6‐phosphate reductase
in Arabidopsis thaliana enhances salt tolerance and induces biosynthesis of both
mannitol and a glucosyl‐mannitol dimer. Plant, Cell & Environment, 26(2), 275-283.
Zhu, J. K., Liu, J., & Xiong, L. (1998). Genetic analysis of salt tolerance in Arabidopsis:
evidence for a critical role of potassium nutrition. The Plant Cell, 10(7), 1181-1191.
30
Zohary, D. (1971). Orgin of south-west Asiatic cereals: wheats, barley, oats and rye. In Plant life
in south-west Asia. Botanical Society of Edinburgh.
31
Chapter 2
Multigenerational study of the establishment of abiotic stress tolerance transgenes in
Arabidopsis thaliana populations under competitive field conditions.
Introduction
Reducing crop losses due to abiotic stresses is a major target of agricultural
biotechnology. However, concerns have been raised about the potential impact of such traits on
native ecosystems (Davis and Shaw 2001, Ellstrand 2003, Thuiller et al. 2005). A prerequisite
for ecological impacts is establishment of self-sustaining populations containing the transgene,
which in turn depends on transgene movement rates, population sizes, and fitness effects.
Although gene flow has been the subject of numerous modeling and empirical studies, transgene
establishment has received less attention.
Establishment could occur by two processes. Transgenic plants could become feral and
form a self-sustaining population outside of agricultural plantings, perhaps first along field
margins and later expanding into natural areas. Alternatively, the transgene could enter a
compatible wild relative through hybridization and introgression and then be maintained due to a
selective advantage conferred by the transgene. Establishment alone, however, does not
constitute harm; the transgenic populations would need to behave differently from the
untransformed crop or wild relative (EFSA 2012). If the transgenic crop behaves the same as
wild-type or its relative, with no increase in invasiveness or toxicity, the risk of a feral/hybrid
population is not unique to the presence of the transgene, but is instead, a risk posed by
cultivation of the crop in general.
In the case of transgenes increasing abiotic stress tolerance, environmental concerns are
generally linked to two possible effects (Tiedje et al. 1989, Hancock et al. 1996, Dale et al.
2002, Ellstrand 2003, Hancock 2003, Hails and Morley 2005, Snow et al. 2005, Weaver and
Morris 2005, Weebadde & Maredia 2011). First, expression of the transgene will confer a
competitive advantage under stressful conditions allowing plants to outcompete non-transgenic
32
plants within that environment. Second, the transgenic plants, feral or hybrid, could invade
regions where that plant was previously constrained due to environmental stresses. In either case,
the fitness effects of the transgene would have an important influence on the success of transgene
establishment outside of cultivation and would be a key value in any environmental risk
assessment seeking to prevent that outcome.
In this study, I sought to determine the long-term fitness effects of abiotic stress tolerance
genes and their potential for transgene establishment within competitive populations. Extensive
research over the past two decades on plant responses to their environments has identified a wide
variety of genes with possible applications for enhancing abiotic stress tolerance (BhatnagarMathur et al. 2008). Examples include chaperone proteins, membrane stabilization proteins,
metabolic and detoxification enzymes, stress signaling pathway genes, and transcriptional
activators. For this study, three genes representing diverse mechanisms shown to confer salinity
tolerance were selected for analysis: (i) an abiotic stress associated transcription factor [C-repeat
binding factor 3/drought responsive element binding factor 1a (CBF3/DREB1a)]; (ii) a plasma
membrane Na+/H+ antiporter [Salt Overly-Sensitive 1 (SOS1)]; and (iii) a mannitol biosynthesis
enzyme [mannose-6-phosphate reductase (M6PR)].
The CBF (C repeat binding factor)/DREB (drought responsive element binding factor)
genes are a family of transcription factors, that induce a regulon of cold responsive (COR) or
drought responsive (DR) genes upon exposure to abiotic stresses such as cold or dehydration
(Baker et al. 1994, Yamaguchi-Shinozaki et al. 1994, Stockinger et al. 1997, Liu et al. 1998,
Kasuga et al. 1999, Gilmour et al. 2000, Yang et al. 2005). Expression of CBF3/DREB1a results
in the induction of COR gene expression, phenotypic changes in growth and development, and
increased plant tolerance to drought, cold, and salinity (Gilmour et al. 2004). Over-expression of
CBF3/DREB1a results in a cascade of gene activations and repressions that influence over 1300
genes (Jaglo-Ottosen et al. 1998, Kasuga et al. 1999, Gilmour et al. 2000, Seki et al. 2001,
Fowler and Thomashow 2002, Maruyama et al. 2004, Vogel et al. 2005, Chan et al. 2012). This
cascade of gene expression changes results in increased freezing, drought and salinity tolerance
33
(Jaglo-Ottosen et al. 1998, Kasuga et al. 1999). These changes, however, also result in
significant phenotypic changes in transgenic Arabidopsis, including varying degrees of
dwarfism, delayed flowering, reduced seed production, and altered plant architecture (Kasuga et
al. 1999, Gilmour et al. 2000, Chan et al. 2012). These phenotypic and transcriptomic effects
could have impacts on plant fitness and the ability to establish under field conditions.
The SOS1 gene encodes a plasma membrane Na+/H+ antiporter (Shi et al. 2000, Shi et
al. 2002). High Na+ concentrations induce the release of a Ca
2+
signal which binds to the
calcium binding protein SOS3, allowing it to complex with the protein kinase SOS2 (Halfter et
al. 2000). The SOS2-3 complex then phosphorylates SOS1, which extrudes cytosolic Na+ ions
into the apoplast (Shi et al. 2000, Qiu et al. 2002, Guo et al. 2004). SOS1 can also be activated
by RCD1, an oxidative-stress regulator, bringing together salinity and oxidative stress response
pathways (Katiyar-Agarwal et al. 2006). SOS1 transcript levels are post-transcriptionally
regulated resulting in increased mRNA stability in the presence of reactive oxygen species
produced under salinity stress (Shi et al. 2003, Chung et al. 2008). Transcriptomic analysis of
SOS1 overexpression lines detected roughly 600 genes with altered expression profiles in the
absence of salinity stress, with the down regulation of oxidative and redox related genes the most
characteristic impact (Chan et al. 2012). SOS1 over-expression plants had reduced levels of
sodium ions in cellular tissues and sap after salt exposure, as well as greater ability to germinate,
grow and set seed under salinity stress than wild-type plants (Shi et al. 2003, Chan et al. 2012).
In the absence of salt stress, no phenotypic differences were observed between the transgenic and
WT plants. By maintaining cellular homeostasis under abiotic stress and with no apparent cost of
resistance, SOS1 overexpression could alter plant fitness and establishment in the field.
Mannose-6-phosphate reductase (M6PR) catalyzes the first committed step in mannitol
biosynthesis. The M6PR gene was cloned from celery, Apium graveolens L., a salt tolerant plant
whose tolerance has been linked to its production of mannitol as a primary product of
photosynthesis (Everard et al. 1997, Zhifang and Loescher 2003). Mannitol can act as a
compatible solute, counterbalancing the osmotic effects of sodium and chloride ions, as well as
34
an osmoprotectant, reducing the damage caused by free radicals produced under salinity stress
(Shen et al. 1997, Sickler et al. 2007, Zhifang and Loescher 2003). Additionally, transcriptomic
analysis of M6PR overexpression lines in the absence of salt found over 1700 genes with altered
expression levels, with strong overlap with genes affected by CBF3 overexpression and an
additional up regulation of biotic stress responsive genes (Chan et al. 2011, Chan et al. 2012).
Constitutive expression of M6PR in A. thaliana resulted in the accumulation of mannitol and
increased ability to germinate, grow and set seed under salinity stress without visible phenotypic
effects in the absence of salinity stress (Zhifang and Loescher 2003, Chan et al. 2011).
Transgenic lines were also able to maintain their photosystem II efficiency at wild-type stressfree levels in the presence of salinity stress and showed reduced salt stress impacts on CO2
assimilation (Sickler et al. 2007). These experiments demonstrated that M6PR confers primary
and secondary effects which could significantly alter plant fitness and confer advantages that
might increase the likelihood of transgene establishment in the natural environment.
In this study we compared the fitness effects of these three different types of abiotic
stress resistance genes and their influence on, and ability to predict, transgene establishment
using the model species Arabidopsis thaliana. We measured fitness relative to wild-type in noncompetitive populations and tracked the transgene frequency within competitive, transgenic
verses wild-type, populations over six generations in the field. Both non-competitive and
competitive fitness was modeled to determine the predictive ability of those fitness values. When
in competition with wild-type plants, CBF3 plants reached near-extinction within four
generations, SOS1 plants showed reduced fitness relative to wild type, and M6PR lines showed
either selective neutrality or a competitive advantage over wild-type plants. Modeling of
expected gene frequency based on early generation testing in competition allowed for greater
accuracy in predicting observed gene frequency after six generations than did relative fitness
assessments from pure line populations. These results illustrate the role of competition in
influencing gene frequency and indicate that competitive field trials may facilitate evaluation of
ecological risk in cases where there are concerns regarding transgene establishment.
35
Materials and Methods
Arabidopsis lines for field experiment
To determine the long-term fitness effect of transgene expression under field conditions,
pure line and mixed competitive populations were grown for six generations. Arabidopsis
thaliana over-expression lines A28, A30 and A40 for the CBF3/DREB1a transcription factor in
the Wassilewskija ecotype (WS) (Gilmour et al. 2000) were provided by Dr. Michael
Thomashow of Michigan State University. Lines 1-1 and 7-6, overexpressing SOS1 in the
Columbia glabrous ecotype (Col(gl)) (Shi et al. 2003), were provided by Dr. Huazhong Shi from
Texas Technological University. The M6PR overexpression lines M2-1, M5-1, and M7-6 in the
Columbia ecotype (Col) (Zhifang and Loescher 2003) were provided by Dr. Wayne Loescher of
Michigan State University. All lines represented unique transformation events. Each of the
abiotic stress resistance transgenes was expressed constitutively using the Cauliflower Mosaic
Virus (CaMV) 35S promoter. Each line also contains the neomycin phosphotransferase II
(NPTII) gene for kanamycin resistance as a selectable marker expressed under the nopaline
synthase (NOS) promoter. The wild-type backgrounds are all rapid-cycling Arabidopsis thaliana
ecotypes and each was used in the experiment for both competitive and non-competitive
phenotypic comparisons between background and transgenic populations.
The initial parent plants used for seed production were first verified to contain and
express their transgene (and only their intended transgene) by PCR, Southern and northern blot
analysis (Figure 2.1, Chan et al. 2012). All seeds used for the initial multigenerational fitness
experiments were from greenhouse grown seed stocks produced during the winter of 2008.
Experimental design for field experiment
The experimental design for the multigenerational field study was adapted from a
greenhouse experiment by Roux et al. 2005 who examined the long-term fitness effects of
Arabidopsis mutations under competitive conditions. For our study, gene frequency of the CBF3,
M6PR and SOS1 transgenes was monitored over six generations from a starting allelic frequency
36
Figure 2.1. Verification of transgenic Arabidopsis thaliana lines via Southern (A) and northern
analysis (B). SOS1 and CBF3 lines show the respective transgene and the endogenous gene,
while M6PR lines show only the transgene since the gene is not endogenous to Arabidopsis
thaliana. Verification was performed by Zhulong Chan.
37
of 50% transgenic and 50% wild-type within populations of ~180 individuals. The planting
2
density, 2600 seeds/m , was equivalent to that used in Roux et al. 2005 and comparable to the
natural density of an Arabidopsis thaliana population studied in 2008 (appendix). In each
subsequent generation, each population was started with enough viable seeds for 180 seedlings.
The number of seeds planted for each population was based upon viability tests of seeds derived
from the prior generation. Fourteen replicate competitive populations were established for each
transgenic line and background wild-type mix. In addition to the mixed competitive populations,
five pure populations of each genotype were included for comparative purposes. A total of 167
populations were grown for each of the six generations: 112 mixed populations and 55 pure
populations (Table 2.1).
Seeds were counted using custom-made counting plates, 20 gauge galvanized steel plates
with multiple 5/64 inch (1.98 mm) holes drilled into them. Calibration counts were performed to
determine the number of seeds that fit into one hole of that size for each genotype at each
generation, and this calculation was then used to determine the correct number of holes of seed
needed for proper planting density of the replicates. This calibration step compensated for
changes in seed size due to genotypic and/or environmental effects. The counted seeds were then
mixed with sterile dry white laboratory sand which aided planting by improving seed scatter and
seeding visibility on the dark soil.
All seeds were first planted in the greenhouse utilizing 26.0x26.0 cm trays, (L-HFT NCR,
Landmark Plastics Corporation; Akron, OH) with perforated bottoms filled with Baccto potting
soil (Michigan Peat Company), pre-moistened and lightly compacted to create a firm seed bed.
Soil choice was based on preliminary growth studies performed in the greenhouse (appendix).
Slow-release (3-4 month) 14-14-14 Osmocote Classic® fertilizer (The Scotts Company LLC,
Marysville, Ohio) was mixed into the Baccto potting soil at the rate of 2 g/L soil prior to
planting. Seeds were stratified at 4°C for 48-60 hours and then directly seeded onto the soil
surface. Once planted, all watering was performed via subsoil irrigation. The trays were placed,
two per flat, into 27.4x53.9 cm greenhouse flats (L-1020NCRN, Landmark Plastics Corporation;
38
Table 2.1. Transgenic Arabidopsis thaliana lines engineered
with abiotic stress tolerance genes and the number of replicate
populations tested over six generations in the field.
Wild1
type
Transgene
Transgenic
line
Number
of Reps
Initial
Mix
Ratio
WS
5
Unmixed
Col
5
Unmixed
Col(gl)
5
Unmixed
CBF3
A28
5
Unmixed
CBF3
A30
5
Unmixed
CBF3
A40
5
Unmixed
SOS1
S1-1
5
Unmixed
SOS1
S7-6
5
Unmixed
M6PR
M2-1
5
Unmixed
M6PR
M5-1
5
Unmixed
WS
CBF3
A28
14
1:1
WS
CBF3
A30
14
1:1
WS
CBF3
A40
14
1:1
Col(gl)
SOS1
S1-1
14
1:1
Col(gl)
SOS1
S7-6
14
1:1
Col
M6PR
M2-1
14
1:1
Col
M6PR
M5-1
14
1:1
1
Wild-types are Wassilewskija (WS) ecotype, Columbia glabrous (Col(gl))
ecotype and Columbia (Col) ecotype
39
Akron, OH) filled with water to approximately 2.5 cm. This depth was maintained daily to
ensure sufficient soil moisture for germination. After germination, watering was reduced to 2030 minutes of subsoil irrigation on an as needed basis, with any excess water dumped to prevent
water-logging of the soil.
Once the study populations reached the rosette stage (five to six true leaves) they were
transported from the greenhouse to the field in an enclosed vehicle. A tray-in-flat potting method
was developed for the field, in which a single planted 26x26 cm tray was placed in the center of
a 27.4x53.9 cm flat, thus allowing for subsoil irrigation as the flat filled with water via trickle
hose (Fig. 2.2a, appendix). The larger flat also allowed for secure anchoring to the ground, using
15.2 cm landscaping stakes, to prevent movement or tipping of the planted trays. To protect the
exposed plants during extreme weather conditions and prevent movement of plants or seeds from
the trial site, the trays were covered with clear plastic lids secured to the anchor stakes by bungee
cords. The lids were deployed no more than six hours ahead of damaging weather and were
removed once the weather passed. Four drainage holes were drilled into the large flat
approximately 2.5 cm from the bottom to allow excess rainwater to drain off. This system
successfully protected the plants from driving rains, sleet, hail, early frosts and snows, and
sustained high winds. Plants were grown in the field until approximately 75% of siliques had
begun to lose chlorophyll and yellow, a change immediately prior to maturation (the beginning
of stage 18 as determined by Roeder and Yanofsky 2006). The trays were then returned to the
greenhouse for final maturation and dry down prior to harvest. This system was designed to
reduce seed loss prior to harvest which aided in both the accuracy of seed yield measurements
and seed containment.
Developmental data collected on all populations included germination rate, days to
germination, days to two true leaves, days to rosette formation (5-6 true leaves), days to initial
bolting and flowering, days to ~75% bolting and flowering, days to 75% of siliques maturing,
and percentage of plants that survived to population maturity. At harvest, all aboveground plant
40
Figure 2.2. The tray-in-flat field planting method showing the trickle hose for irrigation and the
high density of planting for interplant competition (A) and the screening of progeny seed on ½
MS 1% agar 100mg/L kanamycin containing media (B). Germination rates and transgene
frequencies can be calculated from the number of: bleaching of wild-type seedlings (blue
arrows), healthy transgenic seedlings (green arrows) and non-germinating seeds (red arrows).
For interpretation of the references to color in this and all other figures, the reader is referred to
the electronic version of this dissertation.
41
material was removed and temporarily stored in 10lb-sized brown paper bags for transport back
to campus. In lab, the dried plant materials were cleaned of seed and total plant dry weight and
seed mass were recorded. Prior to any further analysis, a subset of the seed produced by each
replicate population was set aside for planting the next generation. This protocol allowed each
replicate population to be separately maintained and carried forward through six generations.
Climatic data was recorded at the Horticultural Teaching and Research Center weather
station, which was located less than 400m from the field site. The weather station recorded air
temperature, precipitation, solar radiation, potential evapotranspiration, and wind speed across all
six field growing seasons.
Genotyping field grown progeny via kanamycin screening
Kanamycin screening of progeny seed was performed on each replicate population in
every generation to determine relative gene frequency of the NPTII selectable marker gene. A
~75 uL aliquot of seed was taken from each mixed competitive population to test for kanamycin
resistance. Each seed aliquot was sterilized with 95% EtOH for 10 minutes followed by three
cycles of 15% commercial bleach plus 0.000025% SDS for 10 minutes. The seed samples were
vortexed during each 10 minute period to ensure uniform sterilization. Sterilized seed were then
washed 4 times in sterile dH2O. These seeds were then screened by being plated, via 20uL
pipette, onto ½ MS media + 1% agar containing 100 mg/L kanamycin, with the exception of
lines A40 and S1-1 which were plated onto media containing 75mg/L kanamycin due to a lower
level of resistance than the other transgenic lines. The kanamycin screenings were performed in
triplicate with the objective of scoring 100 seedlings in each replicate. For each mixed
population screening, seed from the respective WT were included on the same plate as negative
controls to verify kanamycin effectiveness. Wild-type seedlings showed complete bleaching
while transgenic seedlings grew normally. Resistant and susceptible individuals were easily
distinguished from each other and from any seeds that failed to germinate (Fig. 2.2b). In cases of
low germination rates, the plating was repeated with increased numbers of seeds to achieve a
total of three replicates of 100 scored seedlings.
42
The competitive fitness of each transgenic line in each generation was calculated from
the mean of the transgene frequencies of the 14 replicate populations as determined by
kanamycin screening. The relative fitness of transgenic lines was derived from comparison of the
mean seed yields of the pure transgenic populations to the mean yields of the corresponding WT
background (each with 5 replicate pure populations).
qPCR verification of transgene frequency calculated from kanamycin screening
Due to the possibility of transgene silencing, the transgene frequencies of the competitive
populations determined via kanamycin screening were verified by qPCR analysis. Progeny seed
were sterilized as described above and plated via pipette on to ½ MS media + 1% agar. A set of
standards was created using confirmed transgenic or WT seed. Seedlings were grown to the two
true leaves stage in the growth room under fluorescent lights for 16 hours light and 8 hours dark,
~8-12 days depending on genotype. Seedlings were then removed by hand in batches of 100,
weighed and flash frozen in liquid nitrogen, yielding ~150-200 mg fresh weight. For the
standards, seedlings were removed and mixed by hand to result in 100 seedling batches with 0%
transgenic, 10%, 20%, 50% and 100% transgenic seedlings. These standardized batches were
then frozen and processed following the same protocol as the seedlings from competitive
populations.
The frozen material was processed using a modified version of the protocol described by
Vorwerk in 2001, the protein purification step was performed twice, using the Wizard®
Genomic DNA Purification kit (Promega Corporation, Madison, WI, USA). DNA quality was
confirmed via gel electrophoresis and DNA quantity was calculated using the Qubit®
fluorometric quantification system (Life Technologies Corporation, Invitrogen™, Grand Island,
NY, USA). The qPCR was performed using primers for the NPTII gene on a Stratagene Mx4000
(now Agilent Technologies Inc., Santa Clara, California) (forward 5’CGGCTGCATACGCTTGATC-3’ and reverse 5’-GATGCGATGTTTCGCTTGGT-3’) and
SYBR® Green master mix (Life Technologies Corporation, Applied Biosystems®, Grand
Island, NY, USA). Transgene frequency within the mixed populations was calculated by
43
comparison of each replicates Ct values to the Ct values of the known transgene frequency
standards.
Determination of selective pressure under competitive conditions
The variance in mean allele, or transgene, frequencies predicted to be associated with
random genetic drift at the t generation was calculated based on theoretical drift distributions
using the following formula; Vqt = q0*p0*(1-(1-1/(2Ne))^t) (Falconer & Mackay 1996). The
initial frequencies of transgenic and wild-type plants are q0 and p0 respectively and Ne is the
effective population size. Confidence intervals can then be derived as the expected mean ±
[1.96*(Vqt/14)^0.5] with 14 being the number of replicate populations and the expected mean
equal to the starting frequency, 0.5, due to the assumption of no selective pressure.
Modeling transgene frequency
2
The PopGene.S modeling program (Hamilton 2011) was used to model transgene
frequency within 100 theoretical populations across six generations. Each population began at a
starting transgene frequency of 0.5. The number of plants within each population was 100, the
mean number of plants surviving to maturity per population across the six field seasons. Due,
however, to the highly selfing nature of Arabidopsis thaliana, 98-99% (Snape & Lawrence
1971), the effective population size was calculated based on the following formula, Ne =
N/[1+(β/(2- β))] (Caballero 1994) where β is the proportion of selfing (0.98) and N is the
observed population size (100), for an effective population size of 51.
The program incorporated both genetic drift, through a Monte Carlo simulation to
introduce random chance into the probability of producing progeny, and natural selection based
on user-entered fitness values. For this study, measured fitness values for two transgenic lines
per transgene under three growing conditions were used: non-competitive relative fitness values
from yield data from growth chamber pure populations (Chan et al. 2012), non-competitive
relative fitness values from field generation 1 pure populations, and competitive fitness measured
from selectable marker screening of progeny seed from field generation 1 competitive
populations.
44
Statistical Analysis
All statistical calculations and comparisons were performed with the SAS version 9.2 and
R version 2.14.0 statistical programs (SAS Institute Inc., Cary, NC and R Foundation for
Statistical Computing, Vienna, Austria). Differences were considered statistically significant at
P<0.05 and Duncan’s Multiple Range Test was used for mean separation.
Results
These experiments examined the fitness effects of three abiotic stress tolerance
transgenes over six generations in the field in the presence and absence of competition from their
respective wild-type background. All seed used to initiate the experiments were from plants that
were verified by Southern and northern analysis for presence and expression of each transgene
(Figure 2.1, Chan et al. 2012). Continued presence and expression of the transgene in pure-line
populations, as determined by kanamycin resistance, was verified after each of the six
generations in the field. The absence of the transgenes in wild type lines also was verified after
each of the six generations.
2
The planting density (2600 seeds/m ) was selected to ensure a high level of competition
between plants. This density had been used previously by Roux et al. 2005 to assess the fitness
impacts of herbicide tolerance mutations on Arabidopsis thaliana and was also determined to be
within the range of natural population densities we observed growing in ruderal habitat south of
2
East Lansing, Michigan (769-3254 plants/m ). Although Arabidopsis thaliana is generally
classified as a winter annual (Pigliucci et al. 2002), i.e. germinating and forming a rosette in the
fall and flowering and setting seed the following spring, the observed Michigan populations
contained individuals with both spring and fall annual growth habits. This ‘bet-hedging strategy’
of plant development has been previously observed in Arabidopsis by Montesinos-Navarro et al.
(2012), who found such life-history polymorphisms within 17 natural populations in Spain. Our
field experiments used both spring and fall plantings.
Effects of seasonal differences on plant development and productivity
45
Environmental influences on plant development and productivity were observed across
the six growing seasons (Tables 2.2 and 2.3). Plants grown in the late fall (field generation 2)
were the slowest to reach every developmental stage, vegetative or reproductive, regardless of
genotype (Table 2.2). This season was also marked by the lowest average maximum and
minimum air temperatures of the six field seasons (Figure 2.3a).
The different seasons also showed a 4-fold range in yield capacity measured by average
seed yield across all genotypes and 2.4-fold differences in biomass capacity, as measured by total
aboveground dry weight, and seed partitioning, the proportion of total aboveground dry weight
that is seed (Table 2.2, Figure 2.4). For example, field generation 3 spring-grown plants out
yielded those of all other seasons (all P<0.001), while another spring-growing season (generation
5) tied with a fall season (generation 2) for the worst yield regardless of genotype (Table 2.2).
Seasonal influences on productivity were not correlated with temperature or solar radiation levels
2
(all R values <0.1); thus, other unmeasured environmental factors or combinations of factors
likely contributed to the seasonal effects on productivity. Additionally, not all seasonal
influences had uniform effects on the performance of populations containing the three different
transgenes, indicating significant genotype by environment (GxE) interactions (P<0.01, Table
2.3). Seed yield of the SOS1 and M6PR transgenic lines largely paralleled yield capacity of the
environment (Figure 2.4). There did not appear to be a yield cost to the expression of these
transgenes in high-yielding environments under the conditions encountered in these experiments.
The CBF3 transgenic lines underperformed except under the least favorable conditions (Figure
2.4).
Effects of transgene expression on plant development
Overexpression of the transcription factor CBF3 delayed plant reproductive development
across all six field seasons from initial bolting through to maturation (stages 4-8, P<0.001, Figure
2.5a). No delays in vegetative development from germination to rosette formation (stages 1-3)
occurred but CBF3 plants did show the characteristic dwarf phenotype observed in previous
46
Table 2.2. Seasonal effect on development and productivity averaged across all lines and populations.
Measures of
Days to reach Lifecycle Stage
Productivity
MajorField
Germi
5-6
ity
First Majority Mature
PartitioGenera2 True True First Bolting Flowe- Flowering (75% dry Dry Seed ning to
tion
Time in field Season nation Leaves Leaves Bolting (75%) ring (75%%) siliques) Weight Yield Seed
1
2
3
4
5
6
June - July 08
Sept - Nov 08
May - June 09
Sept - Oct 09
May - June 10
Sept - Oct 10
Summer
Fall
Spring
Fall
Spring
Fall
1
4.31c
7.82d
4.02c
3.06a
4.0c
3.48b
12.13b
17.51d
12.9c
11.58a
12.38c
12.91c
16.25a
28.02c
19.06b
18.88b
18.64b
18.5b
20.18a
37.31b
20.74a
20.96a
22.95a
19.74a
NA
23.58a
55.64c 46.02c
27ab 24.66a
25.22a 26.22a
28.85b 32.74c
27.48ab 27.53a
1
NA
67.52d
32.14ab
30.98a
37.47c
34.66b
44.71a
118.8e
53.18b
55.0b
57.83c
68.25d
12.33b
10.6bc
17.36a
9.95c
7.45d
18.72a
1.33d
0.91e
3.56a
2.34c
0.83e
2.72b
0.11c
0.09c
0.21a
0.21a
0.1c
0.15b
Each value is the mean of all non-competitive populations grown during a season (n=50). Means with the same letter are not
significantly different from each other at P<0.05 (Duncan’s).
47
Table 2.3. Analysis of variance showing the effects of transgene, transformation
event, environment and their interactions on seed yield from non-competitive pure
populations grown for six seasons in the field.
Source
Degrees
of
freedom
Sum of
Squares
Mean
Square
F
Value
Pr>F
58
487.24
8.4
10.43
<0.0001
1
5
285.36
57.07
70.83
<0.0001
2
5
71.95
14.39
17.86
<0.0001
Transgenic line
4
29.47
7.37
9.14
<0.0001
Genotype*Environment
25
72.44
2.9
3.6
<0.0001
Transgenic
line*Environment
19
28.02
1.47
1.83
0.02
Error
230
185.33
0.81
Corrected Total
288
672.58
Model
Environment
Genotype
1
Environment includes seasonal effects.
2
Includes transgenes and wild-type ecotypes
48
Figure 2.3. Seasonal differences in maximum and minimum air temperature (A) and daily total
solar flux density (B) across the six field generations. Each value is the mean ± SE across all
days populations were in the field for that growing season. Bars with the same letter were not
significantly different from each other at P<0.05 (Duncan’s).
49
Figure 2.3 (cont’d)
50
Figure 2.4. Mean genotype yield in each season in relation to the environmental yield capacity of
that season. The environmental yield capacity was the mean yield for all genotypes grown in
pure line plots (n=50) (dark line). The CBF3, SOS1 and M6PR values are the mean of 15, 10 and
10 populations, respectively (three, two and two transgenic lines per transgene, with five
populations per line).
51
Figure 2.5. Mean days to reach various lifecycle stages for pure populations of wild-type (solid)
and transgenic plants (dashed) CBF3 overexpression lines (A40, A30) and their background
ecotype WS (A), M6PR lines (M2-1, M5-1) and their wild-type Col (B), and SOS1 lines S1-1
and S7-6 with their wild-type Col(gl) (C). Values are the mean of five field generations with
five replicate populations per genotype. The stages are days to: first germination (1), 75% of the
population with two true leaves (2), 75% of the population with five to six true leaves forming a
rosette (3), first bolting (4), first flowering (5), 75% reaching bolting (6) , 75% flowering and
75% mature (75% of siliques drying down). The transition from vegetative to reproductive
growth is indicated (black arrow), highlighting the effect of CBF3 overexpression.
52
Figure 2.5 (cont’d)
53
studies (Chan et al. 2012). The other two transgenes studied, the Na+/H+ antiporter SOS1 and
M6PR, for mannitol production, had no effect on the rate of vegetative or reproductive
development across the six field growing seasons (P>0.8, Figure 2.5b,c).
Effects of transgene expression on non-competitive plant productivity
To ensure that all productivity parameters for pure populations were accurate, seed
samples from all pure populations, transgenic and wild-type, were screened on kanamycincontaining media. This process would detect via phenotype, contamination by seed or
outcrossing between wild-type and transgenic populations (and vice versa) or silencing of the
transgene in the transgenic lines.
The three transgenic CBF3 lines exhibited reduced aboveground biomass and seed yield,
resulting in reduced fitness relative to the wild-type WS background (Table 2.4). The semi-dwarf
CBF3 line A40 was less severely impacted in productivity compared to fully dwarfed lines A28
and A30, indicating transgene position effects. Similar to the lack of observable effect on
development, overexpression of SOS1, did not significantly impact any measure of plant
productivity when averaged across lines (all P>0.8, Table 2.4). While seed yield was not
significantly increased for line S1-1, the derived value of relative fitness was increased compared
to wild-type and the other SOS1 line, S7-6 (Table 2.4). Expression of M6PR did not
significantly change dry weight or seed yield (P>0.8), however it did significantly increase
partitioning to seed and fitness relative to wild-type in both lines (P<0.03, Table 2.4).
Productivity, fitness and transgene frequency in competition with wild-type plants
Each of the transgenic lines was also grown in competitive populations with their wildtype background starting with an initial transgene frequency of 50%. Fourteen replicate
populations of each line were maintained separately over six generations. Transgene frequencies,
tracked by selectable marker screening of progeny seed, were determined for each population at
every generation. Observed changes in transgene frequency from the initial 50% could indicate
either selection for or against the presence of the transgene, or the effects of genetic drift. A
mean transgene frequency either above or below 95% confidence intervals based on theoretical
54
Table 2.4. Transgene effect on productivity as observed in noncompetitive populations averaged across all field seasons.
Line
Dry
Seed
Partitioning Relative
2
fitness
Genotype weight (g) yield (g) to seed
WS
A28
A30
A40
WT
CBF3
CBF3
CBF3
16.32a
7.29d
10.18c
13.78ab
3.26a
0.73c
1.17c
2.22b
0.20a
0.10f
0.11ef
0.17b
0.22*
0.36*
0.68c*
Col(gl) WT
S1-1
SOS1
S7-6
SOS1
13.91ab
12.54b
13.51b
1.81b
2.10b
1.70b
0.13cde
0.15bcd
0.13cde
1.16*
0.94
Col
M2-1
M5-1
13.71ab
13.18b
11.17bc
1.82b
2.06b
2.10b
0.12def
0.15bc
0.16b
1.13*
1.15*
1
WT
M6PR
M6PR
1
Each value is the mean of 5 replicate populations per genotype per
season. Mean values with the same letter are not significantly
different from each other at P<0.05 (Duncan’s).
2
The relative fitness value for each transgenic line is its seed yield
relative to the seed yield of the corresponding wild-type
background. Mean transgene relative fitness values significantly
higher or lower than wild-type at P<0.05 are indicated (*).
55
drift distributions would be significantly different from predicted means influenced by drift
alone, indicating selective pressure on the transgene within those replicate populations (Falconer
& Mackay 1996).
In competition with WT, the proportion of plants overexpressing CBF3 decreased quickly
and went extinct in most replicate populations within four generations in all three transgenic
lines (Figure 2.6abc). To rule out the possibility that apparent reduction in transgene frequency
as measured by kanamycin screening was an artifact due to gene silencing, progeny seed from
the sixth field generation was tested by qPCR analysis. All samples from CBF3 lines under
competition showed genomic levels of the nptII selectable marker gene below the detection limit
of 1%, confirming the near extinction of this transgene in all three lines by the sixth generation
(Table 2.5). Thus the dwarf phenotype, delayed transition to reproductive growth, and reduced
relative seed production seen in the pure population studies of the CBF3 lines is manifested in a
very low competitive fitness and strong selective disadvantage.
Changes in transgene frequency were also observed in the competitive populations with
the SOS1 transgene and its Col(gl) background. The mean values for both lines fell below the
95% confidence interval indicating negative selection against the transgenic plants over the six
generations (Figure 2.7ab).
Unlike CBF3 and SOS1, M6PR showed positive or neutral selection depending on the
transgenic line involved. M6PR line M2-1 maintained a high mean frequency, above the 95%
confidence interval, for six generations, indicating a positive selection for this line across a range
of field environmental conditions (Figure 2.8a). In contrast, the transgene frequencies of the
individual replicate populations of M5-1 diverged after the first generation and continued to
separate in subsequent generations (Figure 2.8b). When averaged across all 14 replicate
populations the mean fell within the 95% confidence interval indicating that genetic drift, and
not selection, was affecting transgene frequency as might be expected for small population sizes
in the absence of selection.
56
Figure 2.6. Transgene frequency within mixed populations of wild-type WS and three CBF3
overexpression lines, A40 (A), A30 (B) and A28 (C), as determined by selectable marker
screening. All populations began at 50% starting frequency (FG0) and were maintained
separately in subsequent generations. Changes in transgene frequency from one generation to the
next in each of the fourteen replicate mixed populations are shown by the black lines, with the
mean of all fourteen replicates indicated by the gray line. Dashed lines indicate 95% confidence
intervals predicted for mean transgene frequency undergoing solely genetic drift.
57
Figure 2.6 (cont’d)
A
B
C
58
Table 2.5. Comparison of transgene frequencies estimated
by phenotypic selectable marker screening (nptII) and
qPCR analysis.
Transgene Sampled
1
populations
M6PR
Positive control
None (WT) Negative control
CBF3
CBF3
CBF3
SOS1
SOS1
SOS1
SOS1
SOS1
SOS1
M6PR
M6PR
M6PR
M6PR
1
A30 #7
A28 #5
A40 #2
S1-1 #6
S1-1 #11
S1-1 #14
S7-6 #1
S7-6 #5
S7-6 #6
M2-1 #14
M5-1 #3
M5-1 #8
M5-1 #14
nptII
qPCR
2
3
screening analysis
100.0
>90.0
0.0
0.007
0.0
0.027
0.003
0.0
0.003
0.007
0.05
0.133
0.117
0.08
0.127
0.107
0.05). For all subsequent analyses, unless specifically mentioned, all data
were pooled between the repeated experiments.
The timing of early reproductive stages, bolting through flowering, were not significantly
affected by salinity stress in the greenhouse (all P>0.05); however, salinity-stressed plants
exhibited a more rapid maturation of siliques, observed as a decrease in the number of days until
89
75% of siliques had begun to dry down (Table 3.2). In the growth chamber, salinity stress
delayed reproductive development for both days to bolting as well as flowering (Chan et al.
2012). Regardless of genotype, wild-type or transgenic, salinity stress decreased above ground
dry weight and seed yield in the greenhouse (Figure 3.2ab, Table 3.2, all P<0.05). Mean dry
weight for all genotypes was reduced by 52%, while overall mean seed yield decreased 73%.
Mean biomass partitioning to seed was also decreased by salinity stress by 66% (Figure 3.2c,
Table 3.2, all P<0.001). Progeny seed viability was also markedly reduced by salinity stress,
regardless of genotype (Figure 3.2d, Table 3.2, P<0.01).
Salinity stress in the growth chamber (100mM NaCl) resulted in decreases in dry weight
(by 64%) and seed yield (by 72%), comparable to those observed under 75mM NaCl stress in the
greenhouse (Figure 3.3ab, all P<0.05). Mean partitioning of biomass to seed decreased 45%
(Figure 3.3c). Progeny seed viability in the growth chamber was reduced under 100mM NaCl
salinity stress for all genotypes, except CBF3 line A40 and SOS1 line S7-6 (Figure 3.3d, P<0.01
and P>0.05 respectively). Averaged across all genotypes, the reduction in progeny seed viability
due to salinity stress in the growth chamber was considerably less severe (an 11% reduction)
than that observed in the greenhouse (a 47% reduction).
Effects of transgene expression on plant development and productivity in the absence of
salinity stress
The three CBF3 overexpression lines exhibited delayed reproductive development at all
stages from initial bolting to maturation of the siliques under greenhouse and growth chamber
conditions in the absence of salinity stress (Figure 3.4a, Table 3.3, Chan et al. 2012, all P<0.05).
The delayed development, dwarf phenotype and change in leaf architecture described in previous
CBF studies was strongest in the fully dwarfed A28 and A30 lines, while the semi-dwarf A40
90
Figure 3.1. Viable seed yield of pure populations of transgenic and wild type plants in relation to
yield capacity of the environment for six growing conditions, three seasons ±salt (closed and
open symbols respectively). The environmental yield capacity was calculated as the mean viable
seed yield of all genotypes wild-type and transgenic (n=50, gray solid line). The yield capacity
for specific genotypes was the mean viable seed yield for each transgenic line (n=5, black dashed
lines) and wild-type background (n=5, black solid line). The three CBF3 lines and their WS
wild-type background (A), the two SOS1 lines and their wild-type Col(gl), and the two M6PR
lines and their wild-type Col (C).
91
Figure 3.1 (cont’d)
92
Table 3.2. Salinity treatment effect on development and productivity averaged across all lines, populations and experiments.
Treatment
First
bolting
Control
Salt
1
38.6a
40.1a
Days after planting to reach lifecycle stage
Majority
First
Majority
bolting
flowering
flowering
(75%)
(75%)
49.1a
45.3a
59.7a
48.8a
45.1a
61.2a
1
Mature
(75% dry
siliques)
94.3b
85.5a
Dry
weight
(g)
11.73a
5.61b
1
Measures of productivity
Seed yield Partitioning
Seed
(g)
to seed
viability
1.58a
0.43b
0.14a
0.05b
0.81a
0.40b
Each value is the mean of five replicate populations per genotype per treatment per experiment. Data are pooled from three replicate
experiments; equivalent trends were observed in each experiment. Mean values with the same letter were not significantly different
from each other at P<0.05.
93
Figure 3.2. Mean productivity measures for greenhouse-grown pure line populations under
control and salinity stress (75mM NaCl); aboveground dry weight (A), viable seed yield based
on germination rates of progeny seed (B), partitioning to seed (C) and progeny seed germination
rate (D). Salt treatment reduced dry weight, seed yield, partitioning to seed and progeny seed
viability across all genotypes (P<0.01). Each value is the mean from five replicate populations
per genotype per treatment per experiment. Data are pooled from three replicate experiments;
equivalent trends were observed in each experiment. Mean values with the same letter were not
significantly different from each other at P<0.05. (Duncan’s).
94
Figure 3.2 (cont’d)
95
Figure 3.3. Mean productivity measures for growth chamber grown pure line populations under
control and salinity stress (100mM NaCl); aboveground dry weight (A), viable seed yield based
on germination rates of progeny seed (B), partitioning to seed (C) and progeny seed germination
rate (D). Each value is the mean relative fitness from three replicate populations per genotype per
treatment . Mean values with the same letter were not significantly different from each other at
P<0.05. (Duncan’s). The dry weight data for Col and M6PR lines were previously included in
Chan et al. 2011 and were gathered by Zhulong Chan.
96
Figure 3.3 (cont’d)
97
Figure 3.4. Mean days to reach various lifecycle stages under control (black line) and salt treated
(75mM NaCl, grey lines) conditions for pure populations of wild-type (solid) and transgenic
plants (dashed). The development of CBF3 overexpression lines (A40, A30, A28) and their
background ecotype WS (A), M6PR lines (M2-1, M5-1, M7-6) and their wild-type Col (B), and
SOS1 lines (S1-1, S7-6) with their wild-type Col(gl) (C). Values are the mean of three
experiments with five replicate populations per genotype. The stages are days to: first
germination (1), 75% of the population with two true leaves (2), five to six true leaves forming a
rosette (3), first bolting (4), first flowering (5), 75% reaching bolting , flowering and
maturation(6, 7 & 8 respectively).
98
Figure 3.4 (cont’d)
99
line was more intermediate to the WS wild-type (Figure 3.4a). The mean seed yield from pure
CBF3 populations in each of the six greenhouse environments (three repeated experiments ±salt)
only marginally increased as environmental conditions improved, falling well below both the
environmental yield capacity and their wild-type background (Figure 3.1, P<0.05). In almost all
cases mean dry weight, viable seed yield and partitioning to seed of the CBF3 overexpression
lines were reduced compared to WS wild-type in both the greenhouse and growth chamber
(Table 3.3, Figures 3.2abc & 3.3abc, all P<0.05). Progeny seed viability was equal to wild-type
under control conditions in the greenhouse, while line A30 was reduced relative to wild-type in
the growth chamber (Figures 3.2d & 3.3d, Table 3.3, P>0.05 and P<0.05 respectively).
The sodium antiporter SOS1overexpression lines had wild-type development except for a
delay in initial bolting and days to maturity (Figure 3.4b, Table 3.3, P<0.05). In the growth
chamber SOS1 lines were unaffected in development compared to their respective wild-type
Col(gl) (Chan et al. 2012). The seed yield of SOS1 lines was not significantly different from the
environmental yield capacity measured across the six greenhouse growth conditions (Figure 3.1,
P>0.05). Averaged over all environments, lines overexpressing the salt antiporter
SOS1maintained wild-type levels of dry weight, seed partitioning and progeny seed viability in
the greenhouse (Figures 3.2acd, Table 3.3, P>0.05). Viable seed yield trended to be lower than
Col(gl) wild-type (Figures 3.2b, Table 3.3, P<0.1). In the growth chamber, in the absence of
salinity stress, SOS1lines were comparable to wild-type across all measures of productivity
(Figure 3.3abcd, all P>0.05). Under the highest yielding environment in the absence of salinity
stress, the wild-type line Col(gl) exceeded the yield of the SOS1 lines (Figure 3.1b, all P<0.05).
The mannitol biosynthetic enzyme M6PR had no effect on development under control
conditions in either the greenhouse (Figures 3.4c, Table 3.3, all P>0.05) or growth chamber
(Chan et al. 2012). M6PR seed yield was not significantly different from the environmental yield
capacity measured across the six greenhouse growth conditions (Figure 3.1, P>0.05). Likewise
M6PR expression lines were overall not significantly different from their Columbia wild-type in
the absence of salinity stress in dry weight, viable seed yield, or partitioning to seed in either the
100
Table 3.3. Transgene effect on development and productivity averaged across lines and populations.
Treatment
Trans-gene
First
bolting
Control
WS-WT
CBF3
31.7a
44.3d
1
Days to reach lifecycle stage
Majority
First
Majority
bolting
flowerflower(75%)
ing
ing
(75%)
39.0a
38.7a
46.1a
60.8d
52.5c
75.5f
Col(gl)-WT
SOS1
35.5ab
40.2c
43.7ab
46.1b
41.6ab
45.8b
53.5bc
54.8bcd
Col-WT
M6PR
36.7bc
34.6abc
44.5b
44.5b
42.7ab
41.6ab
WS-WT
CBF3
31.8a
49.3e
39.0a
56.1c
Col(gl)-WT
SOS1
35.0ab
38.4bc
Col-WT
M6PR
38.3bc
35.2bc
Salt
Majority
mature
(75% dry
siliques)
80.4a
120.2f
1
Measures of productivity
Dry
Seed
Partitio Seed viability
weight yield
ning to
(g)
(g)
seed
15.1a
5.4c
2.6a
0.54c
0.184a
0.101bc
0.70a
0.74a
85.9abcd
92.9e
14.5a
13.0a
2.3ab
1.46b
0.158ab
0.146ab
0.88a
0.87a
54.8bcd
55.4bcd
89.4cde
89.4cde
14.9a
15.3a
2.76ab
1.89ab
0.174a
0.155a
0.78a
0.88a
38.3a
53.4c
50.1ab
75.4f
82.2ab
91.2de
7.9b
1.7d
0.61c
0.05d
0.076cd
0.014e
0.49b
0.26c
48.4b
47.4b
41.5ab
44.8b
60.2de
58.1cde
83.5ab
86.1bcd
8.1b
8.4b
0.75c
0.64c
0.059cd
0.056d
0.5b
0.43b
47.8b
46.1b
44.8b
41.8ab
63.5e
60.9de
82.7ab
85.1abc
6.0bc
6.4bc
0.38c
0.62c
0.056d
0.064cd
0.42b
0.39b
1
Each value is the mean of five replicate populations per genotype per treatment per experiment. Data are pooled from three replicate
experiments, equivalent trends were observed in each experiment. Mean values with the same letter were not significantly different
from each other at P<0.05 (Duncan’s).
101
greenhouse or growth chamber (Figures 3.2abcd, 3.3abc, Table 3.3, all P>0.05). As was
observed for SOS1, in the highest yielding environment, seed yield of wild-type Col exceeded
the M6PR lines (Figure 3.1c, all P<0.05). In the growth chamber, progeny seed viability was
increased relative to wild-type (Figure 3.3d, P<0.05).
Viable progeny seed yield is the measure for plant fitness in a self-pollinated, seedpropagated species like Arabidopsis thaliana. The relative fitness of the transgenic lines was
determined by comparing the viable seed yield between pure transgenic and pure wild-type
populations. In the absence of salinity stress in the greenhouse and growth chamber, CBF3 lines
overall had reduced fitness relative to their WS wild-type (Figures 3.5ab, all P<0.05). SOS1 lines
trended to be lower in relative fitness in the greenhouse (P<0.1) while maintaining wild-type
fitness in the growth chamber (P>0.05). In the greenhouse, M6PR lines had a relative fitness
comparable to wild-type under the same conditions (all P>0.05); however, in the growth
chamber line M2-1 exceeded wild-type fitness (P<0.05).
Effects of transgene expression on plant development and productivity in the presence of
salinity stress
In the presence of salinity stress, whether imposed in the greenhouse (at 75mM NaCl) or
growth chamber (at 100mM NaCl), CBF3 lines were delayed in all reproductive stages (Figures
3.4a, Table 3.3, Chan et al. 2012, P<0.05). Under salinity stress in the greenhouse, the three
CBF3 overexpression lines also had reduced mean dry weight, viable seed yield and partitioning
to seed relative to WS wild-type (Figure 3.2abc, Table 3.3, all P<0.05). CBF3 lines were the only
transgenic lines to have greater reductions in progeny seed viability under salinity stress in the
greenhouse than their wild-type counterpart (Figure 3.2d, Table 3.3, P<0.05). Performance under
salinity stress in the growth chamber showed some differences from the greenhouse, where the
semi-dwarf line A40 had increased dry weight, viable seed yield and
102
Figure 3.5. Mean relative fitness for pure populations of CBF3 (lines A40, A30 and A28 ), SOS1
(lines 1-1 and 7-6) , and M6PR (lines M2-1 and M5-1) transgenic plants under control and salt
treated (75mM NaCl) conditions in the greenhouse (A) and control and salt treated condions in
the growth chamber (100 mM NaCl) (B). Relative fitness was calculated as the proportion of
viable transgenic seed yield to viable wild-type seed yield under the same treatment. The relative
fitness of SOS1 and M6PR lines under salinity stress in the growth chamber exceeded 20 and
were cut off for figure clarity. Each value is the mean relative fitness from replicate populations,
with n=15 in the greenhouse and n=3 in the growth chamber. Mean fitness values which differ
from wild-type at P<0.1 or P<0.05 are marked by ( .) or (*) respectively.
103
Figure 3.5 (cont’d)
104
partitioning to seed relative to wild-type and had no significant reduction in progeny seed
viability (Figure 3.3abcd).
In the greenhouse in the presence of salinity stress, pure populations of lines
overexpressing the Na+/H+ antiporter SOS1 had wild-type development as well as wild-type
levels of dry weight, viable seed yield, partitioning to seed and progeny seed viability (Figure
3.4b, Figure 3.2abcd, Table 3.3, all P>0.05). In the growth chamber under salinity stress, line S76 had increased bolting and flowering relative to wild-type Col(gl) while line S1-1 had wild-type
development (Chan et al. 2012). In contrast to greenhouse results, in the growth chamber
SOS1confered an advantage in the presence of salinity stress with markedly increased dry
weight, viable seed yield, partitioning to seed and progeny seed viability relative to wild-type
Col(gl) (Figure 3.3abcd, all P<0.05).
Similar to control conditions, under salinity stress in the greenhouse M6PR lines were not
significantly different from Col wild-type in development, dry weight, viable seed yield,
partitioning to seed or progeny seed viability (Figures 3.4c and 3.2abcd, Table 3.3, all P>0.05).
Under salinity stress in the growth chamber, the M6PR lines had increased bolting and flowering
relative to Col wild-type (Chan et al. 2012) and produced over 4-fold more biomass than wildtype and yielded viable seed while Col was unable to do so (Figure 3.3abd, all P<0.05).
The relative fitness, as measured by mean viable seed yield of pure transgenic
populations compared to viable yield from pure wild-type populations, of all the transgenic lines
was increased in the presence of salinity stress. While the CBF3 lines under salinity stress in the
greenhouse continued to have reduced fitness relative to the WS wild-type (Figure 3.5a, all
P<0.05), in the growth chamber CBF3 line A40 had increased fitness relative to WS (Figure
3.5b, P<0.05). SOS1lines under salinity stress in the greenhouse had fitness levels equal to wildtype (Figure 3.5a, all P>0.05), but had significantly increased relative fitness in the growth
chamber under salt stress (Figure 3.5b, all P<0.05). In the greenhouse with salt, M6PR line M2-1
had increased relative fitness while line M5-1 had wild-type fitness (Figure 3.5a, P<0.05 &
105
P>0.05 respectively). In the growth chamber with salt both M6PR lines had increased relative
fitness (Figure 3.5b, P<0.05).
Effects of salinity and transgene expression on competitive ability
Competitive advantages conferred by the transgene in the presence of salt stress were
predicted from pure population studies in the growth chamber (Chan et al. 2012). To quantify
competitive advantage each transgenic line was grown in competition in the greenhouse with
their wild-type counterpart in populations (14 replicates/line/treatment in each of 3 repeated
experiments) that were established as 1:1 mixes of each transgenic line and their wild-type
background. Transgene frequencies were determined by selectable marker screening of progeny
seed from all salinity and control populations. To assess whether changes in observed transgene
frequency from the initial 50% were likely the result of selection or genetic drift, 95%
confidence intervals were calculated based on theoretical drift distributions (Falconer & Mackay
1996). Mean transgene frequencies above or below those thresholds indicated a competitive
advantage or disadvantage due to the transgene.
In competition under no salt stress control conditions, CBF3 overexpression lines showed
a significant competitive disadvantage against their WS wild-type background (Figure 3.6,
P<0.05). In contrast, neither SOS1or M6PR lines experienced a competitive advantage or
disadvantage under control conditions in competition (Figure 3.6, P>0.05). Under salinity stress
no transgenic line showed a competitive advantage against wild-type plants. In the presence of
salinity stress, CBF3 lines continued to be disadvantaged (Figure 3.6, P<0.05). SOS1 transgenic
plants under salinity stress experienced a disadvantage in competition with Col(gl), while M6PR
transgenic plants were competitively the equal of Col (Figure 3.6, P<0.05 and P>0.05
respectively).
To determine whether any transgene conferred a competitive advantage under specific
growth environments in the greenhouse, transgenic seed yield was compared to 50% of the mean
seed yield of all mixed populations in that environment (i.e. if all plants from the starting 1:1 mix
106
Figure 3.6. Mean competitive fitness of transgenic lines, CBF3 lines (A40, A30 and A28), SOS1
(lines 1-1 and 7-6) and M6PR (lines M2-1, M5-1), in mixed populations under control and salt
treated (75mM NaCl) conditions. All populations were planted as 1:1 wild-type:transgenic
mixes, with each transgenic line planted with its respective wild-type background: WS with
CBF3 lines, Col(gl) with SOS1 lines and Col with M6PR lines. Competitive fitness was
calculated from selectable marker screening of progeny seed. Each value is the mean fitness
from fourteen replicate populations per treatment in three repeated greenhouse experiments.
Mean competitive fitness values outside the 95% confidence intervals, calculated based on
theoretical drift distributions, indicate negative selective pressures (marked by *).
107
Figure 3.7. Seed yield of competitive mixed populations of transgenic plants and their wild type
backgrounds in relation to yield capacity of the environment for six growing conditions, 3
seasons ±salt (closed and open symbols respectively). The three CBF3 lines in competition with
their WS wild-type background (A), the two SOS1 lines in competition with wild-type Col(gl),
and the two M6PR in competition with wild-type Col (C). All values are the mean of 14 replicate
populations, each planted with an initial transgene frequency of 50%. Overall yield capacity was
calculated as mean seed yield of all genotypes for each of the three experiments multiplied by
50% (the starting transgene frequency). Yield capacity of each transgenic line was calculated as
the mean seed yield multiplied by the mean transgene frequency measured by selectable marker
phenotyping of progeny seed.
108
Figure 3.7 (cont’d)
109
yielded equivalently; half the seed would be transgenic) (Figure 3.7). CBF3 lines were shown to
be at greater disadvantages vs. their wild-type as the environment improved, while SOS1 and
M6PR lines generally averaged around that capacity measurement. Thus the fitness increases
under salinity stress seen in some SOS1 and M6PR lines in pure populations in the greenhouse
and growth chamber were not observed while in competition with wild-type plants in the
greenhouse.
Discussion
This study compared the fitness effects of three abiotic stress tolerance transgenes in salt
stress experiments in the greenhouse in the presence and absence of competition with wild-type
parents. Prior field experiments with these transgenic lines indicated the importance of assessing
transgene fitness effects in the presence of competition rather than pure line performance
(Chapter 2). However, the field competitive experiments were not performed with salinity stress.
As expected from previously published growth chamber studies (Chan et al. 2011, 2012),
salinity stress impacted reproductive development rates, and reduced vegetative and reproductive
productivity. Another impact of salinity stress was a marked reduction (>50%) in progeny seed
viability, to the extent that it was not possible to obtain sufficient viable seed to perform multigenerational fitness assessments as had been done in the field experiments (Chapter 2). As is
most common in the literature, our previous studies in the growth chamber did not measure seed
viability; however, the low viability observed in progeny seed from salt stressed plants in the
greenhouse indicates the importance of this criterion in making accurate fitness assessments.
Transgene impacts on plant development in the greenhouse in the presence and absence
of salt stress were consistent with previous non-competitive growth chamber experiments (Chan
et al. 2011, 2012); however, differences were observed in the effect of the transgenes on
productivity. While CBF3 line A40 showed enhanced dry weight and seed yield under salinity
stress in the growth chamber, it did not in the greenhouse. The significant gains in productivity,
both vegetative and reproductive, from SOS1 and M6PR overexpression shown under salinity
110
stress in the growth chamber, were reduced in the greenhouse. SOS1 lines were comparable to
wild-type, and the M6PR plants showed elevated relative fitness in only one line. Interestingly
this was the same line observed to have a competitive advantage over wild-type across six
generations in the field (Chapter 2).
Observed differences between growth chamber- and greenhouse-grown pure populations
may be due to the combinations of stresses that can occur in the less tightly controlled
environment of the greenhouse. Light levels in the greenhouse were 4-fold higher than those in
the growth chamber and peak light levels also corresponded to increased air temperatures in the
greenhouse. Research has shown that the effects of combinations of abiotic stresses can be more
severe than those suffered under the stresses individually (reviewed by Mittler 2006). A recent
transcriptomic analysis of Arabidopsis thaliana ecotypes under paired combinations of abiotic
stresses (such as high light and salt stress or high temperature and salt) revealed that over 60% of
expression patterns under stress combinations were not predicted by their individual stress
expression profiles (Rasmussen et al. 2013). These findings may indicate that the combined
conditions of salinity, high light and temperature contributed to the lower relative fitness of the
greenhouse-grown transgenic plants than was observed in the tightly controlled and relatively
mild environment of the growth chamber.
In competition with their wild-type background in the greenhouse, the performance of the
transgenic lines differed from both growth chamber and greenhouse pure line assessments. No
transgene conferred a competitive advantage over wild-type plants in the presence or absence of
salinity stress. CBF3 and SOS1 lines were less competitive in the presence of salinity stress than
they were under control conditions, indicating that competition remains an important factor on
transgene fitness even under significant salinity stress. Competitive assessments of transgenic
plants have been performed previously for traits such as herbicide, virus, and insect resistance.
Transgenic insect resistant hybrids of cultivated, Oryza sativa L., and weedy red rice, O. sativa f.
spontanea, showed an increased seed yield under natural insect pressures in pure populations;
however, in competition with non-transgenic hybrids the transgenic hybrids were equal to wild111
type (Yang et al. 2011, 2012). Similarly, the increased productivity in disease-stressed pure
populations of virus-resistant sugar beet was reduced when planted in competition with
Chenopodium album (Bartsch et al. 1996). Herbicide tolerant oilseed rape, Brassica napus, has
been shown under interspecific competition in field plots (Fredshavn et al. 1995), field margins
(Warwick et al. 2008) and roadsides (Knispel et al. 2008), to have productivity equal to or less
than wild-type. Thus, in general, first generation transgenic traits conferred significant fitness
gains relative to wild-type under the target stress (reviewed by Warwick et al. 2009), but relative
fitness values were reduced in the presence of competition whether in the presence or absence of
the target stress.
The reduced relative fitness observed from the salt-stressed transgenic lines when in
competition with wild-type plants may be linked to the natural responses of Arabidopsis thaliana
to intraspecific competition, which has been found to be significantly different from abiotic or
biotic stress responses (Geisler et al. 2012). Transcriptomic analysis of Columbia ecotype
Arabidopsis plants at varying levels of competition (i.e. planting densities), found that in
response to increasing competition, the plants down-regulated genes involved in abiotic stress
response, pathogen defense and secondary metabolism, and up-regulated photosynthesis related
genes. Thus, Arabidopsis plants appear to attempt to outgrow their competition at the expense of
defense. These changes in transcriptional responses relative to the absence of competition were
2
2
seen in densities ranging from 1400 seeds/m to over 15,000 seeds/m ; our plant density of 2600
2
2
seeds/m was within this range. Given that densities higher than 2600 seeds/m were observed in
natural populations in MI, these transcriptional changes in response to competition likely
occurred in our field (Chapter 2) and greenhouse studies as well in natural populations. It is
possible that our transgenic lines, due to constitutive transgene expression, were transcriptionally
locked into an abiotic stress response (Chan et al. 2012), and thus may have been unable to
compete effectively with wild-type plants.
While the transgenic pure line performance vs. competitive performance trends observed
in this study were similar to those observed in transgenic plants expressing first generation traits
112
(e.g herbicide or insect resistance), there are some important differences between these traits and
traits increasing abiotic stress tolerance. Herbicide, virus and insect resistance in current
transgenic crops have been conferred by genes encoding alternate forms of existing plant
enzymes, viral coat proteins, or small RNAs, or Bt proteins respectively (Warwick et al. 2009).
These gene products function directly to confer the trait of interest; the introduced genes encode
enzymes that degrade or are unaffected by herbicide, or RNAs and proteins that only interact
with their target pest. Expression of these traits has resulted in few secondary effects on the
transformed plant as measured by the timing of life-cycle parameters, plant architecture,
biomass, and seed yield (Thies and Devare 2007, Cheng et al. 2008, Zolla et al. 2008). However,
microarray analysis of the genes in this study showed that expression of the three transgenes had
complex transcriptional effects; CBF3 and M6PR both affected genes involved in ABA
signaling, cell wall biosynthesis, ion and ABC transport, redox, and osmoprotectant biosynthesis,
while the effect of SOS1caused a general down-regulation of oxidative stress response genes
(Chan et al. 2012). Thus transgenes enhancing abiotic stress tolerance can result in complex
changes to regulatory, signaling, and metabolic pathways that could impact plant fitness.
Furthermore, although both disease and insect pressures can, like abiotic stresses, limit
plant populations (reviewed by Warwick et al. 2009), the fitness gains from resistance to biotic
stresses are density dependent.
Each resistant plant, by not serving as a host, reduces the
pressure on its neighbors, as has been observed for insect resistance (Yang et al. 2012). This
effect does not occur in abiotic stress conditions, as all plants in an affected area experience the
stress regardless of how many of their neighbors may be more tolerant to it. This lack of density
dependence in abiotic stress tolerance has important implications on transgene establishment and
further supports the need for competitive assessment of genes enhancing this trait.
Overall, these studies indicate the importance of assessment of abiotic stress tolerance
enhancing transgenes under a range of environmental conditions and corroborate the effects of
competition observed over six generations in the field in the absence of salinity stress (Chapter
2). The secondary fitness effects of transgene expression under salinity stress were influenced by
113
the presence of other abiotic stresses, and by the presence of competition. Although significant
fitness gains in the presence of salt stress were observed in pure population relative to wild-type
in the growth chamber, these gains were reduced in greenhouse pure populations. These
differences in fitness relative to wild-type may be the result of possibly differing transcriptomic
and physiological responses under combined abiotic stresses (as occurs in a greenhouse) than
when a stress is experienced individually (as occurs in a growth chamber). In direct competition
with wild-type plants, these small gains were lost, resulting in no transgenic line showing a
competitive advantage under salt stress. These results indicate that risk assessments and
modeling based on non-competitive assessments, especially those under tightly regulated
environmental conditions, may poorly estimate the risks of abiotic stress tolerance enhancing
transgene spread and establishment. Future studies on the secondary fitness impacts of abiotic
stress tolerance enhancing transgenes should strive to more closely mimic natural conditions by
including competition and a range of environmental conditions to gain more accurate and useful
results for risk assessments.
114
LITERATURE CITED
115
LITERATURE CITED
Andow, D. A., & Hilbeck, A. (2004). Science-based risk assessment for nontarget effects of
transgenic crops. Bioscience, 54(7), 637-649.
Bartsch, D., Schmidt, B., Pohl‐Orf, M., Haag, C., & Schuphan, I. (1996). Competitiveness of
transgenic sugar beet resistant to beet necrotic yellow vein virus and potential impact on
wild beet populations. Molecular Ecology, 5(2), 199-205.
Bhatnagar-Mathur, P., Vadez, V., & Sharma, K. K. (2008). Transgenic approaches for abiotic
stress tolerance in plants: retrospect and prospects. Plant Cell Reports, 27(3), 411-424.
Chan, Z., Grumet, R., & Loescher, W. (2011). Global gene expression analysis of transgenic,
mannitol-producing, and salt-tolerant Arabidopsis thaliana indicates widespread changes
in abiotic and biotic stress-related genes. Journal of Experimental Botany, 62(14), 47874803.
Chan, Z., Bigelow, P. J., Loescher, W., & Grumet, R. (2012). Comparison of salt stress
resistance genes in transgenic Arabidopsis thaliana indicates that extent of transcriptomic
change may not predict secondary phenotypic or fitness effects. Plant Biotechnology
Journal, 10(3), 284-300.
Cheng, K. C., Beaulieu, J., Iquira, E., Belzile, F. J., Fortin, M. G., & Strömvik, M. V. (2008).
Effect of transgenes on global gene expression in soybean is within the natural range of
variation of conventional cultivars. Journal of Agricultural and Food Chemistry, 56(9),
3057-3067.
Davis, M. B., & Shaw, R. G. (2001). Range shifts and adaptive responses to Quaternary climate
change. Science, 292(5517), 673-679.
Ellstrand, N. C. (2003). Dangerous Liaisons?. Johns Hopkins University Press.
FAO (Food and Agriculture Organization of the United Nations), (2011). The state of the
world’s land and water resources for food and agriculture: managing systems at risk.
http://www.fao.org/docrep/017/i1688e/i1688e.pdf (accessed March 2013).
th
Falconer, D. S., & Mackay, T. F. C. 1996. Introduction to quantitative genetics,4
Press.
ed. Longman
Fredshavn, J. R., & Poulsen, G. S. (1996). Growth behavior and competitive ability of transgenic
crops. Field Crops Research, 45(1), 11-18.
Geisler, M., Gibson, D. J., Lindsey, K. J., Millar, K., & Wood, A. J. (2012). Upregulation of
photosynthesis genes, and downregulation of stress defense genes, is the response of
Arabidopsis thaliana shoots to intraspecific competition. Botanical Studies, 53(1), 85-96.
116
Gilmour, S. J., Sebolt, A. M., Salazar, M. P., Everard, J. D., & Thomashow, M. F. (2000).
Overexpression of the Arabidopsis CBF3 transcriptional activator mimics multiple
biochemical changes associated with cold acclimation. Plant Physiology, 124(4), 18541865.
Hoffman, C. A. (1990). Ecological risks of genetic engineering of crop plants. BioScience, 40(6),
434-437.
Kasuga, M., Liu, Q., Miura, S., Yamaguchi-Shinozaki, K., & Shinozaki, K. (1999). Improving
plant drought, salt, and freezing tolerance by gene transfer of a single stress-inducible
transcription factor. Nature Biotechnology, 17(3), 287-291.
Knispel, A. L., McLachlan, S. M., Van Acker, R. C., & Friesen, L. F. (2008). Gene flow and
multiple herbicide resistance in escaped canola populations. Weed Science, 56(1), 72-80.
Mantri, N., Patade, V., Penna, S., Ford, R., & Pang, E. (2012). Abiotic stress responses in plants:
present and future. In Abiotic Stress Responses in Plants (pp. 1-19). Springer New York.
Mittler, R. (2006). Abiotic stress, the field environment and stress combination. Trends in Plant
Science, 11(1), 15-19.
Rasmussen, S., Barah, P., Suarez-Rodriguez, M. C., Bressendorff, S., Friis, P., Costantino, P., ...
& Mundy, J. (2013). Transcriptome responses to combinations of stresses in Arabidopsis
thaliana. Plant physiology. Plant Physiology, 161(4), 1783-1794.
Ricroch, A.E., Berge, J.B. and Kuntz, M. (2011) Evaluation of genetically engineered crops
using transcriptomic, proteomic, and metabolomic profiling techniques. Plant Physiology,
155, 1752–1761.
Shi, H., Ishitani, M., Kim, C., & Zhu, J. K. (2000). The Arabidopsis thaliana salt tolerance gene
SOS1 encodes a putative Na+/H+ antiporter. Proceedings of the National Academy of
Sciences, 97(12), 6896-6901.
Shi, H., Quintero, F. J., Pardo, J. M., & Zhu, J. K. (2002). The putative plasma membrane
Na+/H+ antiporter SOS1 controls long-distance Na+ transport in plants. The Plant
Cell, 14(2), 465-477.
Shi, H., Lee, B., Wu, S.J. and Zhu, J.K. (2003) Overexpression of a plasma membrane Na+ ⁄ H+
antiporter gene improves salt tolerance in Arabidopsis thaliana. Nature Biotechnology.
21, 81–85.
Sickler, C. M., Edwards, G. E., Kiirats, O., Gao, Z., & Loescher, W. (2007). Response of
mannitol-producing Arabidopsis thaliana to abiotic stress. Functional Plant
Biology, 34(4), 382-391.
117
Thies, J. E., & Devare, M. H. (2007). An ecological assessment of transgenic crops. Journal of
Development Studies, 43(1), 97-129.
Thuiller, W., Lavorel, S., Araújo, M. B., Sykes, M. T., & Prentice, I. C. (2005). Climate change
threats to plant diversity in Europe. Proceedings of the National Academy of Sciences of
the United States of America, 102(23), 8245-8250.
Wang, W., Vinocur, B., & Altman, A. (2003). Plant responses to drought, salinity and extreme
temperatures: towards genetic engineering for stress tolerance. Planta, 218(1), 1-14.
Warwick, S. I., Legere, A., Simard, M. J., & James, T. (2008). Do escaped transgenes persist in
nature? The case of an herbicide resistance transgene in a weedy Brassica rapa
population. Molecular Ecology, 17(5), 1387-1395.
Warwick, S. I., Beckie, H. J., & Hall, L. M. (2009). Gene flow, invasiveness, and ecological
impact of genetically modified crops. Annals of the New York Academy of
Sciences, 1168(1), 72-99.
Yang, X., Xia, H., Wang, W., Wang, F., Su, J., Snow, A. A., & Lu, B. R. (2011). Transgenes for
insect resistance reduce herbivory and enhance fecundity in advanced generations of
crop–weed hybrids of rice. Evolutionary Applications, 4(5), 672-684.
Yang, X., Wang, F., Su, J., & Lu, B. R. (2012). Limited fitness advantages of crop-weed hybrid
progeny containing insect-resistant transgenes (Bt/CpTI) in transgenic rice field. PloS
One, 7(7), e41220.
Zhifang, G., & Loescher, W. H. (2003). Expression of a celery mannose 6‐phosphate reductase
in Arabidopsis thaliana enhances salt tolerance and induces biosynthesis of both
mannitol and a glucosyl‐mannitol dimer. Plant, Cell & Environment, 26(2), 275-283.
Zolla, L., Rinalducci, S., Antonioli, P., & Righetti, P. G. (2008). Proteomics as a complementary
tool for identifying unintended side effects occurring in transgenic maize seeds as a result
of genetic modifications. Journal of Proteome Research, 7(5), 1850-1861.
118
Chapter 4
Crop improvement utilizing abiotic stress tolerance enhancing transgenes and the
implications for ecological risk assessments.
Introduction
The next generation of transgenic crops will express a range of traits designed to reduce
crop yield loss from the abiotic stresses that can devastate fields today, and to meet the future
challenges of a changing global climate over the next century. The Food and Agriculture
Organization of the United Nations has estimated a world population of 9.1 billion people by the
year 2050 which will require an estimated increase of 70% in total crop yield over 2007 levels
(Bruinsma 2009). Climate modeling predicts that during this same time period farmers will
experience significant increases in the severity and frequency of extreme weather events (e.g.
heatwaves, droughts, or heavy precipitation) (Goodness 2013).
The 2012 drought across much of the United States affected 80% of US agricultural land,
making it the most severe American drought in a half century, with significant reductions in
yield observed in corn, soybean, and sorghum as well as negative impacts on meat and dairy
livestock (ERS 2013). In the 2012 drought, the world’s first commercialized abiotic stress
tolerance enhanced crop, Monsanto’s DroughtGard corn, faced a natural trial by fire alongside
millions of acres of conventional and genetically engineered corn expressing first generation
traits like herbicide and insect resistance. The farms of the next century are predicted to face
even greater challenges from abiotic stress, as climate change is expected to increase the
frequency and severity of extreme weather events such as heat waves, droughts, floods, and
frosts (Goodness 2013). Combining with these climate-imposed threats will be increases in
agricultural soil salinity that will occur in arid and semi-arid regions due to farm irrigation
applied to protect against drought, and from rising sea levels that will push saline waters farther
into coastal regions (Smedema and Shiati 2002, Chinnusamy et al. 2005). Facing these threats
119
and meeting the additional challenge of feeding a growing global population, will require the
planting of crops able to sustain yield under these extreme conditions.
Approaches to developing such crops incorporate both conventional and molecular
breeding techniques, such as marker assisted selection, as well as genetic engineering approaches
(Mittler and Blumwald 2010). While progress has been made in identifying quantitative trait loci
and selectable markers for abiotic stress tolerance traits (Collins et al. 2008, Mittler and
Blumwald 2010, Varshney et al. 2011), breeding for improved varieties has been hindered by the
complex nature of abiotic stress tolerance, the yield drag that commonly results from crosses
with more stress tolerant wild relatives, and the germplasm diversity constraints caused by
reproductive incompatibility (Varshney et al. 2011). Since crops developed using marker
assisted selection do not require ecological risk assessments under the biosafety regulatory
systems of most countries, this chapter will focus on the implications on ecological risk
assessments of abiotic stress tolerant crops developed with genetic engineering.
Analysis of implications for ecological risk assessments
The current status of abiotic stress tolerant crops
Although abiotic stress results from a range of climatic and soil conditions, there is
significant overlap in impacts on plant tissues, resulting in osmotic and oxidative stress at the
cellular level (Wang et al. 2003, Warwick et al. 2009, Mittler and Blumwald 2010). Low water
availability in cells causes a loss of turgor pressure, negative conformational changes to proteins
and RNA, and a reduction in chemical reactions including photosynthesis. Abiotic stress
conditions, such as high temperatures, drought and high salinity, can affect cellular metabolism
resulting in elevated production of damaging reactive oxygen species. These impacts at the
cellular level exert significant selective pressure on plants and have led to complex, targeted and
overlapping plant abiotic stress response mechanisms. This overlap in response indicates the
potential for enhanced resistance to multiple stresses through regulatory, signaling or metabolic
changes.
120
Significant interest, both private and public, in the development of enhanced abiotic
stress tolerance is evident in the increasing number of submissions to the USDA for field trials of
abiotic stress related traits over the previous decade (Chapter1). These field trials have examined
a range of candidate genes for abiotic stress tolerance, with an even greater array being assessed
in growth chambers and greenhouse. The functions of these genes range from membrane, protein
and RNA stabilization to stress signaling and transcriptional regulation (Wang et al. 2003, Zhang
et al. 2004a, Sreenivasulu et al. 2007, Bhatnagar-Mathur et al.2008, Munn & Tester 2008,
Warwick et al. 2009, Hirayama & Shinozaki 2010, Mittler and Blumwald 2010). A number of
these abiotic stress tolerance genes have been transformed into crop species and expressed either
constitutively or under stress inducible promoters resulting in elevated abiotic stress tolerance
relative to untransformed controls or conventional cultivars; a proof of concept critical for the
movement of lab discoveries to the farmer’s field (reviewed by Mittler & Blumwald 2010,
Deikman et al. 2012, Mantri et al. 2012).
Experimentally assessed abiotic stress tolerance enhanced crops include monocots and
dicots, transformed with a broad range of genes with differing resistance mechanisms. The use of
stress responsive transcription factors in crop species to enhanced abiotic stress has been
demonstrated in rice (Hu et al. 2006 &2008, Oh et al. 2009, Xiao et al. 2009), canola (Jalgo et
al. 2001, Savitch et al. 2005), tomato (Hsieh et al. 2002, Zhang et al. 2004b), tobacco (Kasuga et
al. 2004), wheat (Pellegrineschi et al. 2004), potato (Behnam et al. 2006 & 2007) and forage
grass (James et al. 2008). Transgenes encoding ion transporters have been utilized to increase
salinity tolerance in tobacco (Gao et al. 2006, Yue et al. 2012), tomato (Olías et al. 2009),
sweetpotato (Gao et al. 2012), barley (Mian et al. 2011), and rice (Xiao et al. 2009, Batelli et al.
2007). Introgression of a sodium transporter from a wild relative into durum wheat increased
yields 25% in fields with saline soil compared to the non-introgressed background line (Munn et
al. 2012). Transgenes increasing production of compatible solutes, such as proline, inositol,
sorbitol and glycine betaine, have increased abiotic stress tolerance in tobacco (Sheveleva et al.
1997 & 1998, Konstantinova et al. 2002), wheat (Abebe et al. 2003), and maize (Wei et al.
121
2011). While some of these stress tolerant transgenic crops have been tested in the growth
chamber, greenhouse and field, only two crops have been submitted to the USDA for
deregulation.
The first and currently only commercialized abiotic stress tolerant crop cultivar is
Monsanto’s MON 87460 (APHIS 2011ab). This line expresses the gene CspB which encodes a
cold shock protein from Bacillus subtilis (Reeves 2010). This cold shock protein is naturally
upregulated in response to environmental stress and acts as an RNA chaperone preventing the
misfolding of RNA that can disrupt translation (reviewed by Horn et al. 2007). CspB has been
demonstrated to increase cold, heat and drought tolerance when constitutively expressed in
Arabidopsis thaliana and rice (Castiglioni et al. 2008). Transgenic maize lines expressing CspB
yielded significantly higher than non-transformed lines by maintaining higher chlorophyll
content and photosynthetic rates under drought stress in the field, while exhibiting no cost of
resistance in high yielding environments (Castiglioni et al. 2008). In spring 2013, DroughtGard
maize was made available to farmers for wide-scale planting across the Western Great Plains
region of the United States.
At the time of writing, a petition for deregulation had been submitted to the USDA for
only one other abiotic stress tolerant GE crop: freeze-tolerant hybrid Eucalyptus (Eucalyptus
grandis x E. urophylla) (Nehra & Pearson 2011). Eucalyptus is grown in tropical regions around
the world, largely for its high quality fiber used in paper production. Eucalyptus is also being
planted as a source of biomass for celluloisic biofuel production. In temperate climates, its low
freezing tolerance has prevented its planting in managed forest plantations. Constitutive overexpression of Eucalyptus CBF1a and CBF1b homologs has been shown to increase freezing
tolerance (Navarro et al. 2011), but these lines also showed significant negative phenotypic
effects. To overcome these negative effects, the abiotic stress response transcription factor Crepeat binding factor 2 (CBF2) gene was expressed under the stress inducible promoter rd29A.
The transgenic Eucalyptus lines submitted for deregulation also express a barnase gene under the
anther-specific promoter, PrMC2, to cause the production of sterile pollen (Nehra & Pearson
122
2011). After 5 years, transgenic lines grown at multiple sites were 40-50 ft tall, while the
background line continued to experience severe winter die-back. The transgenic lines did show a
small but significant cost of resistance, visible as a slower rate of growth under freeze-free
conditions compared to background, as well as the intended male sterile phenotype. The USDA
has begun to assemble an Environmental Impact Statement, indicating that freezing tolerant
Eucalyptus may be the second deregulated abiotic stress tolerant crop in the US (APHIS 2013).
Challenges in the development and implementation of abiotic stress tolerant crops
A number of issues have delayed the successful development and implementation of
transgenic crops with enhanced abiotic stress tolerance. Barriers to market for genetically
engineered crops have been described as: demonstration of trait efficacy in the field,
development of product concepts, intellectual property concerns, establishing support of
stakeholders, recordkeeping requirements, regulatory approval, and consumer communication
and acceptance (Rommens 2010). Importantly only two of these barriers can be directly
addressed through scientific data, demonstrating trait efficacy and regulatory approval.
Regulatory approval, however, does not hinge on scientific evidence alone, as it, like the other
hurdles on the way to market, incorporates legal and socioeconomic concerns. Althought these
barriers are factors affecting the commercialization of all transgenic crops, they have also slowed
the commercialization of abiotic stress tolerant crops.
Abiotic stress tolerance traits face significant hurdles from the first of these barriers to
market, demonstrating trait efficacy in the field. Abiotic stress field trials are expensive to
conduct and suffer from inconsistent stress levels across field sites (Mittler & Blumwald 2010,
Richards et al. 2010, Deikman et al. 2012, Weber et al. 2012). Although greenhouse stress trials
can be used to pre-screen transformed lines, promising transformed lines may not show the same
level of tolerance in the field that was observed in the greenhouse (Deikman et al. 2012, Saint
Pierre et al. 2012,Tavakkoli et al. 2012, Witt et al. 2012). The difficulty in phenotyping for
abiotic stress tolerance traits has hindered plant breeders ability to select resistant germplasm,
although recent technological advances in field-based phenomics (e.g. spectral imaging for leaf
123
temperature, water, nitrogen and chlorophyll content) may aid breeders in finding new
selectable phenotypes strongly linked to stress tolerance (White et al. 2012). The next scientific
barrier to the commercialization of abiotic stress tolerant crops is the focus of this chapter, the
ecological risk assessments that occur during the regulatory approval process.
Ecological risks of abiotic stress tolerant crops
The ecological implications of a genetically engineered crop considered during the risk
assessment process are the result of interactions between four factors: the gene inserted, the trait
conferred by that gene, the crop transformed, and the location where the crop will be grown
(Grumet et al. 2011). Environmental risks could result if the transgene affects: non-target
species, the ferality of the transformed crop, the genetic diversity or invasiveness of compatible
wild or weedy relatives, or the ecological range of recipient plants (Tiedje et al. 1989, Hoffman
1990, Hancock et al. 1996, Conner et al. 2003, Ellstrand 2003, Weaver and Morris 2005, Hails
and Morley 2005). Non-target impacts, a concern for insect resistant traits, are not an ecological
risk for abiotic stress tolerance enhancing transgenes. The remaining environmental risks are all
related to the fitness and ecological range of the recipient plants.
Abiotic stress tolerance enhancing transgenes could enable recipient plants to better
tolerate one or multiple abiotic stresses, and thus could confer a fitness advantage over wild-type
plants (Hancock et al. 1996, Ellstrand 2003). While increased survival and fitness, in terms of
seed yield under agricultural production, is a desired transgene impact in the transformed crop
variety, this attribute could also have implications outside of crop fields. Increased fitness under
abiotic stress could alter the ferality of the crop variety and the invasiveness of recipient wild
relatives that acquire the transgene through pollen-mediated gene flow. Either result could lead
to the displacement of wild, non-recipient competitors reducing ecosystem biodiversity and
possibly resulting in localized extinctions of non-recipient plants (Hancock, 2011). Additionally,
a transgene conferring a strong selective advantage could cause a selective sweep of compatible
wild germplasm. This sweep could reduce the diversity of the recipient plants’ gene pool as the
transgene and other linked crop genes or alleles became introgressed. A transgene conferring a
124
significant negative fitness effect (presumably only under non-agricultural conditions or else it
would be unlikely to be planted) could reduce the fitness and genetic diversity of nearby ruderal
recipient populations of compatible wild or weedy relatives resulting in a ‘demographic
swamping’ of the wild germplasm.
Abiotic stress tolerance transgenes could, depending the promoter used, alter the plant’s
response to the environmental conditions experienced from germination to reproduction, and so
could alter the ecological range the recipient plant is able to successfully inhabit (Ellstrand
2003). This change in ecological range could be an intended impact on the transformed crop, as
is the case for freezing-tolerant Eucalyptus, or an unintentional effect on recipient plants if the
transgene enters wild germplasm. The ability of introduced genetic material, through
hybridization and introgression, to cause ecological range expansion has been documented and is
attributed to the acquisition of locally adapted germplasm (Rieseberg et al. 2007). Thus abiotic
stress tolerance transgenes by increasing plant survival and fitness under previously rangelimiting conditions, could confer increased ecological range in recipient plants. Given the
pressures of a changing global climate and rising market demand for food, feed, fiber and fuel
due to a growing world population, the ability to increase crop yield and planting range under
abiotic stressed conditions will be critical to future farmers. On the other hand, these abiotic
stress tolerance transgenes will pose potentially greater environmental risks than the first
generation of genetically engineered crops.
Implications of abiotic stress tolerant GE crops for current ecological risk assessment
methodologies
Around the globe governmental policies on genetically engineered crops are implemented
by regulators who must weigh the potential ecological risks (and in some systems, benefits) of
deregulating a new transgenic crop for the commercial market. Ecological risk assessment alone
does not encompass the full range of factors considered as socio-economic considerations are
also included in the decision-making process for the deregulation of transgenic crops (Devos et
al. 2008). The considered factors differ from country to country, particularly socio-economic
125
ones, resulting in a variety of regulatory frameworks for genetically engineered crops.
Nonetheless, all these systems are all built around the same general scientific principles for the
assessment and management of risk (Regal 1994, Hill & Sendashonga 2003, Jaffe 2004).
Risk is defined as the combination of hazard, harms which could result from a given
action, and exposure, which influences the likelihood of the hazard occurring (Suter 1993, EPA
1998, Thies & Devare 2007). The first step in a risk assessment is problem formulation, during
which regulators must define environmental harms and set assessment endpoints which define
what is to be protected from harm (EPA 1998, Suter 2000, Nickson 2008, Wolt et al. 2010).
Although ecological harm has not always been clearly defined, one of the most frequently
addressed harms is the reduction of biodiversity, at both the species and ecosystem level
(Sanvido et al. 2012). As described earlier, the environmental risks from abiotic stress tolerant
crops could negatively impact biodiversity with a scope potentially larger than that of the first
generation transgenes whose target stresses are generally specific to agricultural systems.
Regardless of the regulatory system, the hazard characterization stage of an ecological
risk assessment of transgenic crops attempts to predict the secondary effects of the transgene on
recipient plants, both the transformed crop itself and any future hybrids from intercrosses with
wild or weedy relatives (EPA 1998, EFSA 2010). These predictions incorporate measurable
phenotypic effects of the transgene as well as baseline data pertaining to the crop species. This
baseline data includes the biology of the crop plant, the cropping system currently used with
conventional varieties of that crop and gene flow and hybridization data between the crop and
any compatible wild or weedy relatives. The risk characterization for a transgenic crop
introduction incorporates both the characterized potential hazards and the characterized exposure
level for each of those hazards. In addition, regulators consider risk management; whether the
characterized risks could be reduced or mitigated by certain management practices (e.g. refugia
in insect resistant crop plantings to slow the evolution of resistance in target insect species).
These methods have been highly successful in guiding the deregulation and commercialization of
the first generation of transgenic crops expressing herbicide, insect and virus resistance with little
126
negative environmental impact (Thies & Devare 2007). The traits which will be present in the
next generation of transgenic crops, such as abiotic stress tolerance, may however, confer more
complex secondary and fitness effects on the transformed plants than has been examined
previously (Grumet et al. 2011). Questions about the ability of current risk assessment
methodologies to assess the ecological risks from these potentially more complex traits, have led
to suggestions for regulatory changes. These proposed changes include suggestions for altered
field trial requirements to better assess fitness effects, experimental testing of lab created cropwild hybrids, and utilizing ‘omic’ approaches to characterize transgene impacts at the transcript,
protein and metabolite level (EFSA 2008). Regulatory changes may be needed address the new
more complex traits; but, as regulatory systems adapt to these new crops they must also continue
protecting their countries’ ecosystems while not placing an undue burden on plant breeders and
farmers. This can be done by ensuring that the regulatory system’s data requirements are
proportionate with the level of potential risk posed by a new submission (Jaffe 2004).
European Food Safety Association guidance report on the ERA of transgenic plants
In June of 2007, a colloquium was held by the European Food Safety Association
(EFSA) to examine the future challenges and approaches for the ecological risk assessment of
transgenic crops (EFSA 2008). The colloquium gathered one hundred stakeholder participants,
including regulators, public and private researchers, industry experts and members of nongovernmental organizations, from 19 EU countries to discuss topics related to risk assessment
and the current EU regulatory system for genetically engineered crops which is based on 2001
legislation (EC 2001). These topics examined how current scientific research addressed aspects
of risk assessment and included: measuring non-target effects, modeling the impacts of GE
cultivation, assessing long-term and large-scale environmental effects, and the consideration of
environmental benefits during risk assessments (EFSA 2008). The recommendations made
during this colloquium could have implications for the ecological risks assessments of future
abiotic stress tolerant crops in the EU.
127
Based in part on the recommendations from the 2007 colloquium, the European Food
Safety Association has issued new guidance on the ecological risk assessment of genetically
engineered plants (EFSA 2010). The risk assessment methodology changes addressed in this
report could significantly impact the regulatory burden on the developers of abiotic stress
tolerant crops. The guidance document outlines multiple broad areas to be considered in an
ecological risk assessment: persistence and invasiveness including the effects of plant to plant
gene flow, interactions with target organisms, interactions with non-target organisms, impacts of
cultivation and harvest methods, and effects on biogeochemical processes. Each of these areas of
risk will be examined as they relate to the ecological risk assessment of abiotic stress tolerance
enhanced crops.
Persistence, invasiveness and plant-to-plant gene flow
The 2010 guidance report outlines problem formulation and hazard identification steps to
be used by EU regulators in considering the environmental harm from transgene persistence and
gene flow to compatible wild species. The report establishes a tiered assessment method for this
area of risk which requires more information to be submitted for transgenic crops which
potentially place the ecosystems of the European Union at greater risks (Figure 4.1). According
to the guidance report, in addition to the standard molecular characterization of the transgenic
plant, all applications should provide basic species-specific information including reproductive
biology, known weediness traits or invasiveness, range limiting factors and the presence of
known compatible relatives within the EU (EFSA 2010). This information is then to be used to
determine whether the transformed plant could survive, reproduce, and possibly hybridize, under
the environmental conditions present within the EU (described as stage 1).
For plants that are potentially able to do so, additional transformation event-specific
information is then required (Figure 4.1). This information compares the transformed plant to its
conventional counterpart in terms of: seed viability, germination, and persistence; plant
phenotype under agronomic conditions; response to abiotic and biotic stress; and reproductive
128
Figure 4.1. The tiered hazard characterization system used by EFSA to address ecological
concerns related to persistence, ferality, and invasiveness of transgenic plants, whether the
genetically engineered crop itself or recipient wild plants due to gene flow. Reproduced from the
EFSA 2010 report, Guidance on the environmental risk assessment of genetically modified
plants.
129
Figure 4.1 (cont’d)
130
Figure 4.1 (cont’d)
131
biology. These data are used to answer questions related to the transgenic crop’s fitness and
persistence in agricultural settings and thus its ability to become feral and/or hybridize (stage 2).
If the transformed plant has altered fitness or is more able to persist than its counterpart,
the protocol then examines trait-specific information to address the next two stages of questions.
Stage 3 questions examine whether the transgene would confer altered fitness in semi-natural
settings to either feral crops or recipient plants (hybrids or introgressed wild relatives) (Figure
4.1). This portion also acknowledges that fitness effects may differ outside of agricultural
settings and in the presence of intra- and inter-specific competitors. Multi-year (2+) field
experiments mimicking the disturbed sites common to ruderal areas are recommended. In
addition, field and greenhouse treatments that include biotic and abiotic stresses should be
performed to test for selective advantages in certain situations with data collection focused on
plant survival and fecundity. Modeling is then incorporated to examine worst case scenarios.
When fitness effects are detected in semi-natural conditions or where the transgene could
increase the range of recipient plants, additional modeling and experiments are suggested to
determine whether the transgene could alter the population size of feral or recipient wild
populations (stage 4, Figure 4.1). The report acknowledges effects on population dynamics from
both enhanced and decreased fitness. Population sizes could increase as recipient plants become
better competitors or gain access to a larger range due to increased fitness under adverse
conditions, while populations could diminish due to outbreeding depression in semi-natural areas
undergoing genetic swamping from nearby agricultural fields. The combination of growth
chamber, field, modeling and background ecological data is then used to assess potential impacts
on ecosystems from changing population sizes in recipient wild species.
Interactions with target and non-target species
As described in the guidance document, target species are generally pests or pathogens of
the transformed crop, so these sections are specifically written for insect and pathogen resistant
transgenic crops. Since generally abiotic stress tolerance transgenes do not impact pest or
pathogen response, this section will not be reviewed in detail in this chapter. However, some
132
abiotic stress transgenes involved in regulatory or signaling pathways could activate pest or
pathogen responsive genes. The production of the sugar alcohol mannitol in transgenic
Arabidopsis thaliana lines expressing the biosynthetic enzyme M6PR, was shown to activate
pathogen responsive genes, an unintended impact proposed to be due to the mannitol being
perceived as a signal of fungal infection (Chan et al. 2011). Transcriptomic analysis could be
utilized to determine whether biotic stress response genes have been affected by abiotic stress
tolerance transgenes, and guide whether biotic stress hypothesis testing should be conducted.
This aspect will be discussed below in greater detail.
Cultivation and management changes
Environmental concerns related to changes in crop cultivation and management could be
significant for abiotic stress tolerant crops that have been engineered to allow successful
cultivation of a crop in a region where it has not been grown previously. Using the freezing
tolerant Eucalyptus example mentioned earlier, regulators would have to consider the
environmental effects of not just the crop itself, but of managed short rotation forestry. Thus
while the transgenic crop and its management system may be new to a country or region,
cultivation, harvest and management practices are likely well established and documented
elsewhere with known and predictable environmental impacts. The EFSA Guidance report
specifies that regulators are to consider scientific and technical literature, field trials, data from
cultivation in other countries and modeling studies in predicting the environmental impacts due
to changes in cultivation methods due to the planting of the genetically engineered crop (EFSA
2010). These possible cultivation impacts are to then be considered in the context of existing
known cultivation impacts on the environment from crop management systems across the EU.
The 2010 EFSA Guidance report on the ERA of genetically engineered plants provides
specific guidelines for each step in an ecological risk assessment from problem formulation to
risk management across multiple broad categories of sources of ecological harm with significant
alterations from the prior system under the 2001 EC directive (EFSA 2010, EC 2001). In 2012
the EFSA GMO Panel issued its first recommendation on an abiotic stress tolerant crop,
133
Monsanto’s DroughtGard maize (MON 87460) (EFSA 2012). While the panel did recommend
this crop to the EU member states, the application was for food and feed only and so a full
ecological risk assessment was not performed. Based on the examples discussed in the 2010
Guidance report, although the ecological risk assessment of abiotic stress tolerance enhanced
crops would address multiple sources of environmental risk as detailed earlier, the ERA would
likely focus on the possibility and possible harms from transgene persistence or gene flow to
wild species (EFSA 2010).
The tiered hazard characterization system detailed in the guidance report section
pertaining to persistence, invasiveness and gene flow from transgenic crops is a significant
change from the past method (EC 2001), which has been termed the “bucket” method for the
sheer quantity of detailed information required in a submission for the environmental release of a
transgenic crop (Raybould 2010). This critical mass approach developed as regulators dealt with
the first generation of transgenic crops which were generally one of four crops, maize, soybean,
cotton and oilseed rape, transformed with a limited number of transgenes conferring either
herbicide, insect, or virus resistance (Wilkinson & Tepfer 2009). But, with newer more complex
traits, such an approach of gathering specific detailed data pertaining to every possible risk of
release would significantly slow or halt commercialization efforts and increase the cost of
completing the regulatory process. A tiered regulatory approach has been advocated for at least
the past decade (Hancock 2003, Wilkinson et al. 2003) to focus data collection and analysis on
crops that for which the combination of trait, gene, crop and location present elevated risks to the
environment. For abiotic stress tolerance enhancing transgenes, one of the most critical aspects
for the risk assessment will be the determination of the transgene’s fitness impacts, whether
detrimental, neutral or advantageous (Wilkinson & Tepfer 2009). The importance of this
characterization is stressed in the guidance report as well as prior proposed tiered risk assessment
systems (Hancock 2003, Wilkinson et al. 2003).
Assessing risk of abiotic stress tolerance enhancing transgenes by the measurement of
secondary fitness effects
134
Given the importance of fitness to the environmental risks described earlier, such as the
biodiversity, ferality, and invasiveness of recipient plants, the accurate measurement of transgene
fitness impacts is, and will remain, critical for ecological risk assessments regardless of the
method used, whether under current regulatory systems or revised systems like the one proposed
by the EFSA Guidance report. First generation transgenes directly encoded the conferred traits
(e.g. Cry genes encoded Bt toxins for insect resistance, altered enzymes for herbicide tolerance,
viral coat proteins for virus resistance) (reviewed by Warwick et al. 2009). These traits were also
of limited advantage outside of agricultural monocultures that generate the high insect, weed and
disease pressures that necessitated the introduction of the transgenes. Unlike those transgenes,
abiotic stress tolerance enhancing transgenes can operate indirectly by manipulating regulatory,
signaling and metabolic pathways and target the environmental stresses that decrease plant
survival and fitness in both agricultural and natural ecosystems (Grumet et al. 2011). Fitness
assessments of first generation transgenic crops used a number of developmental and
reproductive parameters to compare the performance of transgenic lines relative to
untransformed lines and conventional cultivars (NRC 2002).
A second approach to measuring fitness involves measuring directly the contribution to
the next generation from known genotypes growing in mixed competitive populations (Bourguet
et al. 2004). This method requires knowing the starting genotypic frequency (i.e. the percent of
each homozygous positive and negative) within the population and a method for determining the
genotype of progeny seed. Used previously to assess the fitness effects of specific T-DNA
insertions and EMS mutations (Gilliland et al. 1998, Roux et al. 2005), this fitness measurement
method was also utilized to assess three transgenes which confer abiotic stress tolerance through
differing methods (Chapter 2, 3).
The ability of these two fitness measurement approaches to predict the long-term risk of
transgene establishment by three abiotic stress tolerance genes; CBF3, SOS1 and M6PR, was
compared using the model species Arabidopsis thaliana (Chan et al. 2012, Chapter 2, Chapter 3).
These three transgenes were selected for their differing modes of action in conferring increased
135
salinity stress tolerance. CBF3 is an AP2 family transcription factor which activates an abiotic
stress response regulon (Gilmour et al. 2000). SOS1 is a plasma membrane bound Na+/H+
antiporter which is activated in response to elevated levels of sodium ions (Shi et al.2000), and
M6PR is a biosynthetic enzyme allowing the production of the compatible solute and
osmoprotectant mannitol (Zhifang & Loescher 2003). Multiple independent lines for each
transgene were assessed for fitness impacts in pure non-competitive and mixed populations,
where transgenic competed against wild-type plants, across six field environments and in the
three repeated experiments in the presence and absence of salinity stress. Pure line performance
indicated significantly reduced fitness relative to wild-type in CBF3 overexpression lines, while
SOS1 had relative fitness equal to WT. The M6PR lines had relative fitness levels which
averaged better than wild-type across the six field seasons (Chapter 2). Either of these transgene
effects on relative fitness, increased or decreased, would necessitate experiments to address stage
3 questions in environmental risk assessment in the EU.
Competition with wild-type plants was shown to alter the observed the fitness effects of
the transgenes, however, with measured competitive fitness lower than predicted by pure line
performance in all CBF3 and SOS1 transgenic lines assessed in the field (Chapter 2). In
competition in the field one M6PR line was neutral while the other was observed to confer a
competitive advantage across a range of field conditions. A similar competition effect was
observed in the presence of salinity stress, where fitness gains observed in pure populations were
not observed in competition with wild-type under the same conditions, such that none of the
transgenes showed a competitive advantage against their wild-type competition in repeated
greenhouse studies (Chapter 3). This difference in observed transgene fitness effects highlights a
key factor in ERA field trial designs, and how significantly a difference in design can impact the
data collected.
Given the importance of assessing transgene fitness impacts to the ecological risk
assessment process, regulators will need to carefully consider and provide input on the design of
field trials. The EFSA guidance report details suggestions for field trial location selection, study
136
design, choice of conventional comparator, and data analysis (EFSA 2010). It also acknowledges
that the relative fitness effect of a transgene influenced by environment, the presence of interand intra-specific competition and the presence of abiotic or biotic stresses. To examine stage 3
questions pertaining to ferality, the report mandates the inclusion of disturbance field trials, in
which perennial plant species have been removed prior to planting. Further examination of
fitness effects on hybrid or introgressed wild relatives and transgene establishment is also based
on modeling incorporating data from pure line assessments and the disturbance trials.
The study of CBF3, SOS1 and M6PR overexpression lines also compared observed
changes in transgene frequency over six generations to the predicted frequencies from stochastic
models incorporating a range of measured fitness values (Chapter 2). Non-competitive pure line
performance fitness measurements were found to poorly predict transgene frequency, while
models incorporating fitness values observed under competition, produced frequencies similar to
observed trends. Given that transgenic plants that have been characterized to stages 3 and 4 are
considered by the EFSA to be the most ecologically hazardous, these results have strong
implications about the source of the data to be used in the suggested modeling. Based on the data
collected about the fitness effects of the three abiotic stress tolerance transgenes (Chan et al.
2012, Chapters 2 & 3), field trials and modeling that substantially rely on pure line performance
measures could miss critical information needed to predict the long-term performance of
transgenic plants in competitive wild populations. Competitive fitness assessments could be
crucial to predicting whether an abiotic stress tolerance transgene will increase the ferality of the
transgenic crop or the invasiveness of recipient wild relatives, both unintended effects with
significant environmental risks.
Assessing risk of abiotic stress tolerance enhancing transgenes using ‘omic’ approaches
Given the possibility of complex and unintended secondary effects of transgene
expression, a variety of ‘omic’ analyses have been proposed for non-targeted molecular profiling
and assessment of genetically engineered crops (Kuiper et al. 2003, Cellini et al. 2004,
Hoekenga 2008, Ricroch et al. 2011). Concerns that these secondary effects could manifest at the
137
transcript, protein or metabolic levels has led some to suggest that ecological risk assessments
incorporate transcriptomic, proteomic and metabolomic analyses as unbiased approaches to
examine for unintended effects. These techniques have also been considered as means to confirm
at the molecular level, the ‘substantial equivalence’ of transgenic lines undergoing the risk
assessment process prior to deregulation (Millstone et al. 1999, Baudo et al. 2009, Beale et al.
2009).While not required in petitions for deregulation by current regulations, if utilized these
‘omic’ analyses would need to be carefully considered within the context of the varietal range
and phenotypic impacts upon which selection could act.
First generation transgenic crops which utilized simple directly encoded traits were
observed to have less impact at the molecular level than the differences between existing
commercial varies. Transcriptomic analysis, via 9K cDNA microarray, of glutenin
overexpressing wheat lines at multiple stages of development found the differences between the
transformed and untransformed wheat lines was significantly less than the difference between the
untransformed line and a commercial variety (0.06% and 5.59% respectively) (Baudo et al.
2006). Similar results were observed in transcriptional analysis of two transgenic glyphosate
resistant soybean varieties and three conventional varieties that showed greater differences
between the conventional varieties than between the transgenic varieties and the closest related
conventional variety (Cheng et al. 2008). Greater transcriptomic effects were detected between
radiation mutagenised rice lines and the non-mutagenised control lines, than between
transformed rice lines and their respective controls (Batista et al. 2008). Differences in gene
expression were greater between non-transgenic maize varieties than between insect resistant
cryIA(b) expressing maize and its untransformed background (Coll et al. 2009, 2010).
Expression differences due to environment (high and low nitrogen) were also found to exceed
the effect of the transgene. Proteomic comparison of 12 independent transgenic Arabidopsis
thaliana lines to their untransformed background in the presence and absence of cold stress
found that the number of protein differences due to transgene insertion and expression was lower
than the effect of the cold treatment on the wild-type background (Ren et al. 2009). Thus,
138
overall, the first generation transgenic crops caused lower molecular effects than the differences
between conventionally bred varieties, lower than differences due to environment and less than
that caused by mutagenesis. While these results do not directly apply to abiotic stress tolerant
crops, they do indicate the importance of placing all ‘omic’ changes in the correct context.
The next generation of transgenic crops will express traits like abiotic stress tolerance
using genes involved in regulatory, signaling and metabolic pathways, increasing the potential
for unintended secondary effects which could alter a plant’s molecular profile. Overexpression of
the abiotic stress response transcription factor CBF3 in Arabidopsis thaliana results in changes
in expression level for 1350 genes in the absence of salinity stress and 1037 genes in the
presence of salt stress (Chan et al. 2012). Overexpression of the sodium antiporter SOS1 under
the same conditions altered, 619 and 845 transcript levels respectively, corresponding to previous
evidence that the antiporter is stabilized and activated in response to salt stress (Qiu et al. 2003,
Chung et al. 2008). The mannitol biosynthetic enzyme M6PR influenced the expression of 1719
genes in the absence of salt and 1001 in its presence (Chan et al. 2011). In comparison, salt stress
on three wild-type ecotypes WS, Col and Col(gl) altered the expression levels of 1093, 2552 and
138 genes respectively, indicating that in terms of the number of transcripts affected all three
transgenes were within the range caused by environmental effects (Chan et al. 2012). For CBF3
and M6PR substantial overlap was observed in the genes affected by the transgene and those
affected by salt stress in wild-type plants, again stressing the importance of context, and not
assuming risk purely based on the number of transcriptional changes.
For CBF3 and SOS1, the pathways of genes influenced by transgene expression matched
known target genes and previously researched interactions (Chan et al. 2012). The biosynthetic
enzyme M6PR, however, was shown to activate even more pathways than the transcription
factor CBF3, including pathogen responsive genes. As mentioned earlier, this connection may be
due to mannitol acting as a molecule signaling a fungal attack on the transformed plant (Chan et
al. 2011). Whether the up-regulation of these pathogen responsive genes would confer a
selective advantage under disease pressure has not yet been determined. In addition, the total
139
number of transcript expression differences was compared to changes in plant performance to
examine whether transcriptomic analysis could predict the level of secondary phenotypic impact
from transgene expression (Chan et al. 2011, 2012). But, the magnitude of transcriptomic
changes relative to wild-type in the presence or absence of salinity stress was not indicative of
the performance of a transgenic line relative to wild-type under the same conditions.
The use of transcriptomic analysis to test for unintended effects has also been utilized in
the assessment of drought tolerant Arabidopsis thaliana overexpressing the abiotic stress
responsive transcription factor ABF3 (Abdeen et al. 2010). This analysis found that only under
drought stress did the transcriptome of the transgenic lines differ from wild-type. Significant
overlap was observed between the genes affected by drought stress in wild-type plants and those
affected by drought in the ABF3 transgenic lines, with most showing an enhanced response in
the transgenic plants compared to wild-type. The transgene-affected genes were shown to be in
expected pathways responsive to abiotic stress, indicating no unintended pathway activations or
repressions. The studies of ABF3 and M6PR transgenic plants demonstrate the capacity of
transcriptomic analysis to detect unintended expression changes at the pathway and gene level.
These results indicate that molecular profiling through ‘omic’ approaches may be able to guide
further assessment inquiry but do not by themselves indicate levels of risk, and so must be
considered in the context of the natural range of differences that exist between cultivars and
within the same cultivar across differing environmental conditions.
Discussion
The agricultural challenges of the next century, both in terms of climate change and a
growing world population, will require a diverse tool set. While significant advances in plant
breeding have allowed selection to be performed at the molecular and genetic level and advanced
imaging allows in field selection on a host of new trait linked phenotypes, these approaches are
still limited by the complexity of stress responses and the availability and speed at which stress
tolerance traits can be introgressed from sexually compatible germplasm (Mittler and Blumwald
2010, Varshney et al. 2011). The last decade of research has unveiled a broad suite of abiotic
140
stress tolerance genes which operate over many differing cellular and tissue level mechanisms to
confer tolerance (Wang et al. 2003, Zhang et al. 2004a, Sreenivasulu et al. 2007, BhatnagarMathur et al. 2008, Munn & Tester 2008, Warwick et al. 2009, Hirayama & Shinozaki 2010,
Mittler and Blumwald 2010).
Concerns have been raised that the various national regulatory procedures, which capably
addressed the ecological risks of the first generation of transgenic crops (Thies & Devare 2007),
might not be able to handle to the more complex secondary effects of abiotic stress tolerance.
This analysis examined the implications of abiotic stress tolerance crops on the ecological risk
assessments critical to regulatory decision on transgenic crops. The EU system, considered one
of the most stringent in the world, was examined for how the regulations would address the
ecological risks from abiotic stress tolerant crops, namely the risks of ferality in the transformed
crop, negative impacts on the genetic diversity of compatible wild relatives, enhanced
invasiveness in wild or weedy relatives due to gene flow, or range expansion due to ecological
release of recipient transgenic plants (Tiedje et al. 1989, Hoffman 1990, Hancock et al. 1996,
Conner et al. 2003, Ellstrand 2003, Weaver and Morris 2005, Hails and Morley 2005).
The EU system addressed perceived weaknesses in its regulatory practices related to next
generation traits and implemented, at least in part, a tiered ecological risk assessment system,
long called for by scientific experts (Hancock 2003, Wilkinson 2003, Raybould 2010). The
implementation of the 2010 system in the ecological risk assessment of abiotic stress tolerant
crops is as of yet untested, as the recent EFSA GMO Panel approval of drought tolerant maize
was a submission for import for feed and food only and so did not require a full ERA. In
addition, the EU system allows for member states to object to EFSA approvals under the
safeguard clause, resulting in protracted legal disputes over the legal status of genetically
engineered crops in specific countries. For example, Austria has invoked the safeguard clause
multiple times over the past five years to prevent the importation of genetically engineered
oilseed rape lines, each time requiring substantial time and resources to examine their concerns,
and each time their concerns have been found to be without scientific merit (EFSA 2013). Thus,
141
although the risk assessment system may have improved under the 2010 Guidance report on the
environmental risk assessment of genetically modified plants, the EU system as a whole still
leaves much to be desired from a biotechnological crop developer’s perspective.
The use of ‘omic’ approaches to perform non-targeted review of transgene secondary
effects has merit in guiding further lines of inquiry, but not in directly assessing ecological risk
from phenotypic changes. For CBF3, SOS1 and M6PR the total number of genes influenced by a
transgene’s expression in the presence or absence of salinity stress was not indicative of the
performance of the transgenic plant under the same condition (Chan et al. 2012). This
phenotypic buffering was also observed in an assessment of 162 recombinant inbred Arabidopsis
thaliana lines bred from a cross between two ecotypes which differ genetically by at least
500,000 SNPs (Fu et al. 2009). These lines were assessed for over 40,000 molecular traits
including transcript, protein and metabolite levels and examined for phenotypic changes across
139 morphological, developmental and stress response traits; however, only six loci for systemic
change were found, indicating widespread buffering. The effects of this phenotypic buffering
could decrease the phenotypic impact of almost any genetic change whether due to mutation or
DNA insertion. On the other hand, by detecting the unintended up-regulation of biotic stress
response genes in M6PR lines, transcriptomic analysis did demonstrate the potential for ‘omic’
approaches to guide inquiries into secondary transgene effects (Chan et al. 2011).
Field trials have been the most critical aspect of ecological risk assessments since the first
transgenic crop was deregulated. The new European system has addressed changing
environmental concerns due to new traits by altering the field trial requirements to better address
transgene fitness effects on both the transformed crop and on wild interfertile recipients (EFSA
2010). Fitness effects are to be assessed across multiple sites and a minimum of two years, and
transgenes which show fitness impacts are required to be examined under a range of field
treatments including disturbance to mimic the effects of growing in disturbed habitats such as
field margins or road sides. Although this treatment addresses interspecific competition, which
affects the persistence and ferality of transgenic crops and has been shown to influence observed
142
transgene fitness effects (Fredshavn et al. 1995, Bartsch et al. 1996), it does not address the
fitness effects under intra-specific competition from wild-type plants of the same species.
Performance under intra-specific competition determines the long-term population dynamics that
will occur after a transgene has introgressed into a wild species (Ellstrand 2003). In regulatory
terms, for any environmental harm to occur the exposure rate has to increase (i.e. the transgene
frequency within the wild population will need to increase). Under the EU system, transgene
fitness effects on wild relatives, due to hybridization or introgression, are to be assessed in
controlled environments (growth chamber or greenhouse) and predicted using models which
incorporate the data collected in the field, greenhouse or growth chamber (EFSA 2010). Given
the fitness differences observed in the study of three abiotic stress tolerance transgenes between
field and controlled environments and between non-competitive and competitive assessments,
and the effects those differences had on the accuracy of model predictions (Chapter 2 & 3),
regulators will need to carefully consider and interpret the results of fitness studies in the context
of realistic agricultural scenarios.
The abiotic stress tolerance traits that have the potential to aid farmers in feeding a
growing world population while encountering increasing climatic uncertainty will pose greater
challenges to the regulatory systems conducting ecological risk assessment than were faced from
the limited and relatively low impact traits of first generation transgenic crops. With the use of
ERAs that incorporate phenotypic and molecular approaches to assess for secondary transgene
effects, the potential ecological risks in terms of ferality, genetic diversity, invasiveness and
range expansion can be manageable. Tiered risk assessments should allow the focusing of
resources on the gene/trait/crop/location combinations that are most likely to cause unintended
environmental effects, without greatly hindering commercialization of transgenic crops which
pose insignificant risks to the environment and potentially substantial benefits to farmers in need
of crops capable of enduring a changing climate.
143
LITERATURE CITED
144
LITERATURE CITED
Abdeen, A., J. Schnell, and B. Miki, 2010: Transcriptome analysis reveals absence of unintended
effects in drought-tolerant transgenic plants overexpressing the transcription factor ABF3.
BMC Genomics 11, 69.
Abebe, T., Guenzi, A. C., Martin, B., & Cushman, J. C. (2003). Tolerance of mannitolaccumulating transgenic wheat to water stress and salinity. Plant Physiology, 131(4),
1748-1755.
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2011a). Monsanto Company Petition (07-CR-191U) for determination of
non-regulated status of event MON 87460. Draft Environmental Assessment
http://www.aphis.usda.gov/brs/aphisdocs/09_05501p_dea.pdf (accessed March 2013).
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2011b). Determination of Nonregulated Status for MON 87460 Corn (Zea
mays L). http://www.aphis.usda.gov/brs/aphisdocs/09_05501p_det.pdf (accessed March
2013).
APHIS (Animal and Plant Health Inspection Service of the United States Department of
Agriculture), (2013). Notice of intent to prepare an environmental impact statement for
determination of nonregulated status of freeze tolerant Eucalyptus lines and notice of
virtual public meetings. Federal Register, 78 (39) 13309- 13312.
http://www.aphis.usda.gov/brs/fedregister/11_01901p_fr_noi.pdf (accessed April 2013).
Bartsch, D., Schmidt, B., Pohl‐Orf, M., Haag, C., & Schuphan, I. (1996). Competitiveness of
transgenic sugar beet resistant to beet necrotic yellow vein virus and potential impact on
wild beet populations. Molecular Ecology, 5(2), 199-205.
Batelli G., Verslues P.E., Agius F., Qiu Q., Fujii H., Pan S., Schumaker K.S., Grillo S., Zhu J.K.
(2007). SOS2 promotes salt tolerance in part by interacting with the vacuolar H+-ATPase
and upregulating its transport activity. Molecular and Cellular Biology 27: 7781–7790.
Batista, R., Saibo, N., Lourenço, T., & Oliveira, M. M. (2008). Microarray analyses reveal that
plant mutagenesis may induce more transcriptomic changes than transgene
insertion. Proceedings of the National Academy of Sciences, 105(9), 3640-3645.
Baudo, M. M., Lyons, R., Powers, S., Pastori, G. M., Edwards, K. J., Holdsworth, M. J., &
Shewry, P. R. (2006). Transgenesis has less impact on the transcriptome of wheat grain
than conventional breeding. Plant Biotechnology Journal, 4(4), 369-380.
Baudo, M. M., Powers, S. J., Mitchell, R. A., & Shewry, P. R. (2009). Establishing substantial
equivalence: transcriptomics. In Transgenic Wheat, Barley and Oats (pp. 247-272).
Humana Press.
145
Beale, M. H., Ward, J. L., & Baker, J. M. (2009). Establishing substantial equivalence:
metabolomics. In Transgenic Wheat, Barley and Oats (pp. 289-303). Humana Press.
Behnam, B., Kikuchi, A., Celebi-Toprak, F., Yamanaka, S., Kasuga, M., Yamaguchi-Shinozaki,
K., & Watanabe, K. N. (2006). The Arabidopsis DREB1A gene driven by the stressinducible rd29A promoter increases salt-stress tolerance in proportion to its copy number
in tetrasomic tetraploid potato (Solanum tuberosum). Plant Biotechnology, 23(2), 169177.
Behnam, B., Kikuchi, A., Celebi-Toprak, F., Kasuga, M., Yamaguchi-Shinozaki, K., &
Watanabe, K. N. (2007). Arabidopsis rd29A:: DREB1A enhances freezing tolerance in
transgenic potato. Plant Cell Reports, 26(8), 1275-1282.
Bhatnagar-Mathur, P., Vadez, V., & Sharma, K. K. (2008). Transgenic approaches for abiotic
stress tolerance in plants: retrospect and prospects. Plant Cell Reports, 27(3), 411-424.
Bourguet, D., Guillemaud, T., Chevillon, C., & Raymond, M. (2004). Fitness costs of insecticide
resistance in natural breeding sites of the mosquito Culex pipiens. Evolution, 58(1), 128135.
Bruinsma, J. (2009). The resource outlook to 2050: by how much do land, water and crop yields
need to increase by 2050?. In Proceedings of the Expert Meeting on How to Feed the
World in 2050 24-26 June 2009, FAO Headquarters, Rome.
ftp://ftp.fao.org/docrep/fao/012/ak542e/ak542e00.pdf (accessed August 2012).
Chan, Z., Grumet, R., & Loescher, W. (2011). Global gene expression analysis of transgenic,
mannitol-producing, and salt-tolerant Arabidopsis thaliana indicates widespread changes
in abiotic and biotic stress-related genes. Journal of Experimental Botany, 62(14), 47874803.
Chan, Z., Bigelow, P. J., Loescher, W., & Grumet, R. (2012). Comparison of salt stress
resistance genes in transgenic Arabidopsis thaliana indicates that extent of transcriptomic
change may not predict secondary phenotypic or fitness effects. Plant Biotechnology
Journal, 10(3), 284-300.
Castiglioni, P., Warner, D., Bensen, R. J., Anstrom, D. C., Harrison, J., Stoecker, M., ... &
Heard, J. E. (2008). Bacterial RNA chaperones confer abiotic stress tolerance in plants
and improved grain yield in maize under water-limited conditions. Plant
Physiology, 147(2), 446-455.
Cellini, F., Chesson, A., Colquhoun, I., Constable, A., Davies, H. V., Engel, K. H., ... & Smith,
M. (2004). Unintended effects and their detection in genetically modified crops. Food
and Chemical Toxicology, 42(7), 1089-1125.
Cheng, K. C., Beaulieu, J., Iquira, E., Belzile, F. J., Fortin, M. G., & Strömvik, M. V. (2008).
Effect of transgenes on global gene expression in soybean is within the natural range of
146
variation of conventional cultivars. Journal of Agricultural and Food Chemistry, 56(9),
3057-3067.
Chinnusamy, V., Jagendorf, A., & Zhu, J. K. (2005). Understanding and improving salt tolerance
in plants. Crop Science, 45(2), 437-448.
Chung, J. S., Zhu, J. K., Bressan, R. A., Hasegawa, P. M., & Shi, H. (2008). Reactive oxygen
species mediate Na+‐induced SOS1 mRNA stability in Arabidopsis. The Plant Journal,
53(3), 554-565.
Coll, A., Nadal, A., Collado, R., Capellades, G., Messeguer, J., Melé, E., ... & Pla, M. (2009).
Gene expression profiles of MON810 and comparable non-GM maize varieties cultured
in the field are more similar than are those of conventional lines. Transgenic Research,
18(5), 801-808.
Coll, A., Nadal, A., Collado, R., Capellades, G., Kubista, M., Messeguer, J., & Pla, M. (2010).
Natural variation explains most transcriptomic changes among maize plants of MON810
and comparable non-GM varieties subjected to two N-fertilization farming practices.
Plant Molecular Biology, 73(3), 349-362.
Collins, N. C., Tardieu, F., & Tuberosa, R. (2008). Quantitative trait loci and crop performance
under abiotic stress: where do we stand?. Plant Physiology,147(2), 469-486.
Conner, A.J., Glare, T.R. and Nap, J.P. (2003). The release of genetically modified crops into the
environment. II. Overview of ecological risk assessment. Plant Journal 33, 19–46.
Deikman, J., Petracek, M., & Heard, J. E. (2012). Drought tolerance through biotechnology:
improving translation from the laboratory to farmers’ fields. Current Opinion in
Biotechnology, 23(2), 243-250.
Devos, Y., Demont, M., & Sanvido, O. (2008). Coexistence in the EU—return of the moratorium
on GM crops?. Nature Biotechnology, 26(11), 1223-1225.
EC (European Community). (2001). Directive 2001/18/EC of the European Parliament and of the
Council of 12 March 2001 on the deliberate release into the environment of genetically
modified organisms and repealing Council Directive 90/220/EEC. Official Journal of the
European Communities L106, 1-39.
Ellstrand, N. C. (2003). Dangerous Liaisons?. Johns Hopkins University Press.
EFSA (European Food Safety Authority). (2008). EFSA scientific colloquium summary report:
environmental risk assessment of genetically modified plants - challenges and
approaches. http://www.efsa.europa.eu/en/search/doc/colloquiagmoera.pdf (acessed April
2013).
147
EFSA (European Food Safety Authority). (2010). Guidance on the environmental risk
assessment of genetically modified plants.
http://www.efsa.europa.eu/en/efsajournal/pub/1879.htm (accessed April 2013).
EFSA (European Food Safety Authority). (2012). Scientific Opinion on an application (EFSAGMO-NL-2009-70) for the placing on the market of genetically modified drought
tolerant maize MON 87460 for food and feed uses, import and processing under
Regulation (EC) No 1829/2003 from Monsanto.
http://www.efsa.europa.eu/en/efsajournal/doc/2936.pdf (accessed April 2013).
EFSA (European Food Safety Authority). (2013). Scientific Opinion on a request from the
European Commission related to the prolongation of prohibition of the placing on the
processing and feed uses in Austria.
http://www.efsa.europa.eu/en/efsajournal/doc/2878.pdf
EPA (Environmental Protection Agency). (1998). Guidelines for Ecological Risk Assessment.
EPA 630/R-95-002F, Washington, D.C.
ERS (Economic Research Service of the United States Department of Agriculture), (2013). U.S.
Drought 2012: Farm and Food Impacts. http://www.ers.usda.gov/topics/in-the-news/usdrought-2012-farm-and-food-impacts.aspx (accessed April 2013).
Fredshavn, J. R., & Poulsen, G. S. (1996). Growth behavior and competitive ability of transgenic
crops. Field Crops Research, 45(1), 11-18.
Fu, J., Keurentjes, J. J., Bouwmeester, H., America, T., Verstappen, F. W., Ward, J. L., ... &
Jansen, R. C. (2009). System-wide molecular evidence for phenotypic buffering in
Arabidopsis. Nature genetics, 41(2), 166-167.
Gao, F., Gao, Q., Duan, X., Yue, G., Yang, A., & Zhang, J. (2006). Cloning of an H+-PPase
gene from Thellungiella halophila and its heterologous expression to improve tobacco
salt tolerance. Journal of Experimental Botany, 57(12), 3259-3270.
GAO, S., YUAN, L., ZHAI, H., LIU, C. L., HE, S. Z., & LIU, Q. C. (2012). Overexpression of
SOS genes enhanced salt tolerance in Sweetpotato. Journal of Integrative
Agriculture, 11(3), 378-386.
Gilliland, L. U., McKinney, E. C., Asmussen, M. A., & Meagher, R. B. (1998). Detection of
deleterious genotypes in multigenerational studies. I. Disruptions in individual
Arabidopsis actin genes. Genetics, 149(2), 717-725.
Gilmour, S. J., Sebolt, A. M., Salazar, M. P., Everard, J. D., & Thomashow, M. F. (2000).
Overexpression of the Arabidopsis CBF3 transcriptional activator mimics multiple
biochemical changes associated with cold acclimation. Plant Physiology, 124(4), 18541865.
148
Goodess, C. M. (2013). How is the frequency, location and severity of extreme events likely to
change up to 2060?. Environmental Science & Policy. 27(s1), s4–s14.
Grumet, R., Wolfenbarger, L., & Ferenczi, A. (2011). Future possible genetically engineered
crops and traits and their potential environmental impacts. In Environmental safety of
genetically engineered crops. Michigan State University Press.
Hails, R. S., & Morley, K. (2005). Genes invading new populations: a risk assessment
perspective. Trends in Ecology & Evolution, 20(5), 245-252.
Hancock, J. F., R. Grumet, and S. C. Hokanson. (1996). The opportunity for escape of
engineered genes from transgenic crops. HortScience 31:1080–1085.
Hancock, J. F. (2003). A framework for assessing the risk of transgenic crops. BioScience, 53(5),
512-519.
Hancock, J. (2011). Factors influencing the genetic diversity of plant species and the potential
impact of transgene movement. In Environmental safety of genetically engineered crops.
Michigan State University Press.
Hill, R. A., & Sendashonga, C. (2003). General principles for risk assessment of living modified
organisms: lessons from chemical risk assessment. Environmental Biosafety Research,
2(2), 81.
Hirayama, T., & Shinozaki, K. (2010). Research on plant abiotic stress responses in the
post‐genome era: past, present and future. The Plant Journal,61(6), 1041-1052.
Hoffman, C. A. (1990). Ecological risks of genetic engineering of crop plants. BioScience, 40(6),
434-437.
Hoekenga, O. A. (2008). Using metabolomics to estimate unintended effects in transgenic crop
plants: problems, promises, and opportunities. Journal of Biomolecular Techniques,
19(3), 159.
Horn, G., Hofweber, R., Kremer, W., & Kalbitzer, H. R. (2007). Structure and function of
bacterial cold shock proteins. Cellular and Molecular Life Sciences, 64(12), 1457-1470.
Hsieh, T. H., Lee, J. T., Charng, Y. Y., & Chan, M. T. (2002). Tomato plants ectopically
expressing Arabidopsis CBF1 show enhanced resistance to water deficit stress. Plant
Physiology, 130(2), 618-626.
Hu H, DaiM, Yao J, Xiao B, Li X, Zhang Q, Xiong L (2006) Overexpressing a NAM, ATAF,
and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in
rice. Proceedings of the National Academy of Sciences USA, 35: 12987– 12992.
149
Hu H, You J, Fang Y, Zhu X, Qi Z, Xiong L (2008) Characterization of transcription factor gene
SNAC2 conferring cold and salt tolerance in rice. Plant Molecular Biology, 67: 169–181.
Jaffe, G. (2004). Regulating transgenic crops: a comparative analysis of different regulatory
processes. Transgenic Research, 13(1), 5-19.
Jaglo K.R., Kleff S., Amundsen K.L., Zhang X., Haake V., Zhang J.Z., Deits T., Thomashow
M.F. (2001). Components of the Arabidopsis C-repeat/dehydration-responsive element
binding factor cold-response pathway are conserved in Brassica napus and other plant
species. Plant Physiology, 127:910–917.
James V.A., Neibaur I., Altpeter F. (2008). Stress inducible expression of the DREB1A
transcription factor from xeric, Hordeum spontaneum L in turf and forage grass
(Paspalum notatum Flugge) enhances abiotic stress tolerance. Transgenic Research, 17,
93–104.
Kasuga, M., Miura, S., Shinozaki, K., & Yamaguchi-Shinozaki, K. (2004). A combination of the
Arabidopsis DREB1A gene and stress-inducible rd29A promoter improved drought-and
low-temperature stress tolerance in tobacco by gene transfer. Plant and Cell
Physiology, 45(3), 346-350.
Konstantinova T, Parvanova D, Atanassov A, Djilianov D. (2002). Freezing tolerant tobacco,
transformed to accumulate osmoprotectants. Plant Science, 163:157–164
Kuiper, H. A., Kok, E. J., & Engel, K. H. (2003). Exploitation of molecular profiling techniques
for GM food safety assessment. Current Opinion in Biotechnology, 14(2), 238-243.
Mantri, N., Patade, V., Penna, S., Ford, R., & Pang, E. (2012). Abiotic stress responses in plants:
present and future. In Abiotic Stress Responses in Plants (pp. 1-19). Springer New York.
Mian, A., Oomen, R. J., Isayenkov, S., Sentenac, H., Maathuis, F. J., & Véry, A. A. (2011).
Over‐expression of an Na+‐and K+‐permeable HKT transporter in barley improves salt
tolerance. The Plant Journal, 68(3), 468-479.
Millstone, E., Brunner, E., & Mayer, S. (1999). Beyond ‘substantial equivalence’. Nature,
401(6753), 525-526.
Mittler, R., & Blumwald, E. (2010). Genetic engineering for modern agriculture: challenges and
perspectives. Annual Review of Plant Biology, 61, 443-462.
Munns, R. and Tester, M. 2008. Mechanism of salinity tolerance. Annual Review of Plant
Biology. Vol.59, 651-681.
Munns, R., James, R. A., Xu, B., Athman, A., Conn, S. J., Jordans, C., ... & Gilliham, M. (2012).
Wheat grain yield on saline soils is improved by an ancestral Na+ transporter
gene. Nature Biotechnology, 30(4), 360-364.
150
Navarro, M., Martinez, Y., Laur, J., El Kayal, W., Marque, and C., Teulieres, C. (2011). Two
EguCBF1 genes overexpressed in Eucalyptus display a different impact on stress
tolerance and plant development. Plant Biotechnology Journal, 9(1), 50-63.
Nehra, N.S. & Pearson, L. (2011). Petition for Determination of Non-regulated Status for Freeze
Tolerant Hybrid Eucalyptus Lines. ArborGen Inc. Summerville, SC.
Nickson, T. E. (2008). Planning environmental risk assessment for genetically modified crops:
problem formulation for stress-tolerant crops. Plant Physiology, 147(2), 494-502.
National Research Council (NRC) Staff & Board on Agriculture and Natural Resources Staff.
(2002). Environmental effects of transgenic plants: the scope and adequacy of regulation.
National Academies Press.
Oh, S. J., Kim, Y. S., Kwon, C. W., Park, H. K., Jeong, J. S., & Kim, J. K. (2009).
Overexpression of the transcription factor AP37 in rice improves grain yield under
drought conditions. Plant Physiology, 150(3), 1368-1379.
Olías, R., Eljakaoui, Z., Pardo, J. M., & Belver, A. (2009). The Na+/H+ exchanger SOS1
controls extrusion and distribution of Na+ in tomato plants under salinity
conditions. Plant signaling & behavior, 4(10), 973-976.
Pellegrineschi, A., Reynolds, M., Pacheco, M., Brito, R. M., Almeraya, R., YamaguchiShinozaki, K., & Hoisington, D. (2004). Stress-induced expression in wheat of the
Arabidopsis thaliana DREB1A gene delays water stress symptoms under greenhouse
conditions. Genome, 47(3), 493-500.
Qiu, Q. S., Barkla, B. J., Vera-Estrella, R., Zhu, J. K., & Schumaker, K. S. (2003). Na+/H+
exchange activity in the plasma membrane of Arabidopsis. Plant Physiology, 132(2),
1041-1052.
Raybould, A. (2010). The bucket and the searchlight: formulating and testing risk hypotheses
about the weediness and invasiveness potential of transgenic crops. Environmental
Biosafety Research, 9(03), 123-133.
Reeves, W. R. (2010). Petition for the Determination of Non‐Regulated Status for MON 87460.
The Monsanto Company, St. Louis MO.
Regal, P. J. (1994). Scientific principles for ecologically based risk assessment of transgenic
organisms. Molecular Ecology, 3(1), 5-13.
Ren, Y., Lv, J., Wang, H., Li, L., Peng, Y., & Qu, L. J. (2009). A comparative proteomics
approach to detect unintended effects in transgenic< i> Arabidopsis. Journal of
Genetics and Genomics, 36(10), 629-639.
151
Ricroch, A. E., Bergé, J. B., & Kuntz, M. (2011). Evaluation of genetically engineered crops
using transcriptomic, proteomic, and metabolomic profiling techniques. Plant Physiology,
155(4), 1752-1761.
Rieseberg, L. H., Kim, S. C., Randell, R. A., Whitney, K. D., Gross, B. L., Lexer, C., & Clay, K.
(2007). Hybridization and the colonization of novel habitats by annual
sunflowers. Genetica, 129(2), 149-165.
Richards, R. A., Rebetzke, G. J., Watt, M., Condon, A. T., Spielmeyer, W., & Dolferus, R.
(2010). Breeding for improved water productivity in temperate cereals: phenotyping,
quantitative trait loci, markers and the selection environment. Functional Plant
Biology, 37(2), 85-97.
Rommens, C. M. (2010). Barriers and paths to market for genetically engineered crops. Plant
Biotechnology Journal, 8(2), 101-111.
Roux, F., Camilleri, C., Bérard, A., & Reboud, X. (2005). Multigenerational versus single
generation studies to estimate herbicide resistance fitness cost in Arabidopsis
thaliana. Evolution, 59(10), 2264-2269.
Saint Pierre, C., Crossa, J. L., Bonnett, D., Yamaguchi-Shinozaki, K., & Reynolds, M. P. (2012).
Phenotyping transgenic wheat for drought resistance. Journal of experimental botany,
63(5), 1799-1808.
Sanvido, O., Romeis, J., Gathmann, A., Gielkens, M., Raybould, A., & Bigler, F. (2012).
Evaluating environmental risks of genetically modified crops: ecological harm criteria for
regulatory decision-making. Environmental Science & Policy, 15(1), 82-91.
Savitch, L. V., Allard, G., Seki, M., Robert, L. S., Tinker, N. A., Huner, N. P., ... & Singh, J.
(2005). The effect of overexpression of two Brassica CBF/DREB1-like transcription
factors on photosynthetic capacity and freezing tolerance in Brassica napus. Plant and
Cell Physiology, 46(9), 1525-1539.
Sheveleva E, Chmara W, Bohnert HJ, Jensen RG (1997) Increased salt and drought tolerance by
D-ononitol production in transgenic Nicotiana tabacum L. Plant Physiol 115:1211–1219
Sheveleva EV, Marquez S, Chmara W, Zegeer A, Jensen RG, Bohnert HJ (1998) Sorbitol-6phosphate dehydrogenase expression in transgenic tobacco. Plant Physiol 117:831–839
Shi, H., Ishitani, M., Kim, C., & Zhu, J. K. (2000). The Arabidopsis thaliana salt tolerance gene
SOS1 encodes a putative Na+/H+ antiporter. Proceedings of the National Academy of
Sciences, 97(12), 6896-6901.
Smedema, L. K., & Shiati, K. (2002). Irrigation and salinity: a perspective review of the salinity
hazards of irrigation development in the arid zone. Irrigation and drainage systems, 16(2),
161-174.
152
Sreenivasulu, N., Sopory, S. K., & Kavi Kishor, P. B. (2007). Deciphering the regulatory
mechanisms of abiotic stress tolerance in plants by genomic approaches. Gene, 388(1), 1.
Suter, G.W. II. (1993). Ecological Risk Assessment. Lewis Publishers.
Suter, G.W. II. (2000). Generic assessment endpoints are needed for ecological risk assessment.
Risk analysis, 20(2), 173-178.
Tavakkoli, E., Fatehi, F., Rengasamy, P., & McDonald, G. K. (2012). A comparison of
hydroponic and soil-based screening methods to identify salt tolerance in the field in
barley. Journal of Experimental Botany, 63(10), 3853-3867.
Tiedje, J. M., Colwell, R. K., Grossman, Y. L., Hodson, R. E., Lenski, R. E., Mack, R. N., &
Regal, P. J. (1989). The planned introduction of genetically engineered organisms:
ecological considerations and recommendations. Ecology, 70(2), 298-315.
Thies, J. E., & Devare, M. H. (2007). An ecological assessment of transgenic crops. Journal of
Development Studies, 43(1), 97-129.
Varshney, R. K., Bansal, K. C., Aggarwal, P. K., Datta, S. K., & Craufurd, P. Q. (2011).
Agricultural biotechnology for crop improvement in a variable climate: hope or
hype?. Trends in Plant Science, 16(7), 363-371.
Wang, W., Vinocur, B., & Altman, A. (2003). Plant responses to drought, salinity and extreme
temperatures: towards genetic engineering for stress tolerance. Planta, 218(1), 1-14.
Warwick, S. I., Beckie, H. J., & Hall, L. M. (2009). Gene flow, invasiveness, and ecological
impact of genetically modified crops. Annals of the New York Academy of
Sciences, 1168(1), 72-99.
Weber, V. S., Melchinger, A. E., Magorokosho, C., Makumbi, D., Bänziger, M., & Atlin, G. N.
(2012). Efficiency of managed-stress screening of elite maize hybrids under drought and
low nitrogen for yield under rainfed conditions in Southern Africa. Crop Science, 52(3),
1011-1020.
Weaver, S. A., & Morris, M. C. (2005). Risks associated with genetic modification:–An
annotated bibliography of peer reviewed natural science publications. Journal of
Agricultural and Environmental Ethics, 18(2), 157-189.
Wei, A., He, C., Li, B., Li, N., & Zhang, J. (2011). The pyramid of transgenes TsVP and BetA
effectively enhances the drought tolerance of maize plants. Plant Biotechnology
Journal, 9(2), 216-229.
153
White, J. W., Andrade-Sanchez, P., Gore, M. A., Bronson, K. F., Coffelt, T. A., Conley, M. M.,
... & Wang, G. (2012). Field-based phenomics for plant genetics research. Field Crops
Research, 133, 101-112.
Wilkinson, M. J., Sweet, J., & Poppy, G. M. (2003). Risk assessment of GM plants: avoiding
gridlock?. Trends in Plant Science, 8(5), 208-212.
Wilkinson, M., & Tepfer, M. (2009). Fitness and beyond: Preparing for the arrival of GM crops
with ecologically important novel characters. Environmental Biosafety Research, 8(1), 1.
Witt, S., Galicia, L., Lisec, J., Cairns, J., Tiessen, A., Araus, J. L., ... & Fernie, A. R. (2012).
Metabolic and phenotypic responses of greenhouse-grown maize hybrids to
experimentally controlled drought stress. Molecular Plant, 5(2), 401-417.
Wolt, J. D., Keese, P., Raybould, A., Fitzpatrick, J. W., Burachik, M., Gray, A., ... & Wu, F.
(2010). Problem formulation in the environmental risk assessment for genetically
modified plants. Transgenic Research, 19(3), 425-436.
Xiao, B. Z., Chen, X., Xiang, C. B., Tang, N., Zhang, Q. F., & Xiong, L. Z. (2009). Evaluation
of seven function-known candidate genes for their effects on improving drought
resistance of transgenic rice under field conditions. Molecular Plant, 2(1), 73-83.
Yue, Y., Zhang, M., Zhang, J., Duan, L., & Li, Z. (2012). SOS1 gene overexpression increased
salt tolerance in transgenic tobacco by maintaining a higher K+/Na+ ratio. Journal of
Plant Physiology,169(3), 255-261.
Zhang, J. Z., Creelman, R. A., & Zhu, J. K. (2004a). From laboratory to field. Using information
from Arabidopsis to engineer salt, cold, and drought tolerance in crops. Plant
Physiology, 135(2), 615-621.
Zhang, X., Fowler, S. G., Cheng, H., Lou, Y., Rhee, S. Y., Stockinger, E. J., & Thomashow, M.
F. (2004b). Freezing‐sensitive tomato has a functional CBF cold response pathway, but a
CBF regulon that differs from that of freezing‐tolerant Arabidopsis. The Plant
Journal, 39(6), 905-919.
Zhifang, G., & Loescher, W. H. (2003). Expression of a celery mannose 6‐phosphate reductase
in Arabidopsis thaliana enhances salt tolerance and induces biosynthesis of both
mannitol and a glucosyl‐mannitol dimer. Plant, Cell & Environment, 26(2), 275-283.
154
Conclusions and future work
The changing global climate will pose significant environmental challenges to the
farmers of the next century. Whether bred through conventional methods including molecular
breeding techniques, or produced using transgenesis, crops better able to tolerant abiotic stress
will be critical to feeding a growing human population under these difficult conditions. While
there have been environmental concerns about all transgenic crops, the genes conferring abiotic
stress tolerance function in more indirect manners (by altering regulatory, signaling or metabolic
pathways) than the transgenes which have been previously commercialized (e.g. herbicide, insect
and virus resistance). These indirect modes of action could result in more complex secondary
effects than those observed with prior transgenic traits. Both the primary and possible secondary
effects of abiotic stress tolerance transgenes raise a number of environmental concerns about the
potential for the transgene to alter the fitness or ecological range of the crop or, through gene
flow, recipient compatible wild species.
This dissertation examined the fitness effects of three abiotic stress tolerance transgenes
which function by differing modes of action to confer increase salinity tolerance to Arabidopsis
thaliana and assessed whether the observed transgene fitness impacts differed between a variety
of environmental conditions and the presence or absence of competition from wild-type plants.
The field studies of Chapter 2 determined that the fitness effects observed in lines from all three
transgenes were significantly reduced in competition from what would be predicted from their
performance relative to wild-type in pure populations. These differences led to significantly
different outcomes in models that predicted the frequency of transgenic plants in simulated
competitive populations. Models incorporating fitness values from non-competitive fitness
values overestimated transgene frequency compared to the transgene frequency trends observed
155
across six generations in the field. Ecological risk assessments models designed to predict the
likelihood of transgene establishment could have limited accuracy if they utilize transgene fitness
measurements assessed under non-competitive conditions.
Chapter 3 examined whether the presence of salinity stress would increase the
competitive ability of the transgenic lines. Although significantly increased fitness relative to
wild-type was observed in pure populations subjected to salt stress in the growth chamber (by
Zhulong Chan) and in the greenhouse, the presence of wild-type competitors again limited
transgene fitness, resulting in no transgenes conferring a competitive advantage under
competitive conditions. The lower transgene fitness observed in pure greenhouse populations
compared to pure growth chamber populations may be due to the combinations of stresses (i.e.
salt, high light and temperature) that co-occur in the less tightly controlled greenhouse
environment. Recent transcriptomic analyses indicate that the reduction in fitness observed in the
transgenic lines under competition in the field and greenhouse may be due transcriptional and
physiological changes which occur in Arabidopsis thaliana plants in response to competition,
these include an up-regulation of growth and photosynthesis related genes and the down
regulation of stress response genes. Due to constitutive overexpression of abiotic stress response
genes, our transgenic lines likely would have been unable perform the same changes as wild-type
plants, possibly resulting in differing growth patterns in competition than in pure populations
where all plants have the same genotype.
The implication of this research and abiotic stress tolerance traits in general for
ecological risk assessments were examined in chapter 4 as part of the requirements for my
USDA National Needs Fellowship to examine genetic engineering for abiotic stress tolerance in
a broader societal context. Given the calls for regulatory changes to better assess next generation
156
biotechnological traits, the recently updated EFSA system for the environmental risk assessment
of genetically modified plants was examined in the context of abiotic stress tolerant crops. The
recommended adoption of tiered ecological risk assessment systems are intended to allow
greater focus on gene/trait/crop/location combinations that are most risky to the receiving
environment while allowing transgenic crops which pose negligible risks to be commercialized
in a reasonable time span. For example, M6PR overexpression was shown to have the potential
to increase plant fitness compared to wild-type. However, transgenic M6PR maize for planting in
Europe would likely pose insignificant environmental risks due the inability of maize to persist
and become feral and the lack of compatible wild relatives, while planting the same transgenic
cultivar in Mexico, the crop’s center of origin, should necessitate further risk assessment.
Although assessing transgene fitness impacts on recipient plants is critical for an ecological risk
assessment, these results must be considered within the context of all four environmental risk
factors; gene, trait, crop and location.
The results of this dissertation research indicate significant competitive impacts on the
observed fitness of transgenic plants with enhanced abiotic stress tolerance. Although this affect
was observed in the field across six seasons and in the greenhouse in the presence and absence of
salt stress, these results will need to be confirmed in a crop species. The methodology utilized in
this dissertation, designed to mimic a worst case scenario of transgene introgression into a wild
population, can be easily adapted to most crops. The most critical aspect is the ability to
genotype progeny seed from competitive populations. Given that most transgenic plants coexpress some form of selectable marker (e.g antibiotic or herbicide resistance) it should be fairly
simple to adapt this method to a new crop/transgene combination. Corroboration of these
findings in a crop species would strengthen the need to more closely consider the effects of
157
competition in ecological risk assessments. Given that, in this study, competition reduced the
observed fitness effects of all three transgenes compared to pure line performance, abiotic stress
tolerance may be a less hazardous trait under competitive non-agronomic conditions than has
previously been assumed from non-competitive assessments.
This methodology has the additional potential for plant breeders utilizing molecular
breeding strategies to enhance abiotic stress tolerance in crop species. Traditional breeding
strategies for this trait have been limited by the uneven abiotic stress levels present in field plots.
This stress heterogeneity and variability reduces the gains from selection and increases the
chance that promising material may be culled due to where and when its plot was located. By
planting with a mix of genotypes at known proportions, all plants will experience the full range
of stress present in that field, regardless of underlying soil variations. The field could then be
harvested normally and a subset of the bulked seed genotyped by either ‘seed chip’ technologies
or screening of progeny seedlings. Changes in the genotypic ratio of progeny seed relative to the
known planting ratios would indicate a yield advantage under the environmental conditions
present in that field site.
Lastly, given that it was the only one of three transgenes which showed beneficial effects
across six field seasons and in the greenhouse, M6PR warrants further inquiry. Pure populations
overexpressing this transgene consistently showed enhanced partitioning to seed resulting in
increased fitness relative to wild-type. M6PR lines were also the only lines which performed
equal to or better than wild-type in direct competition. The prior transcriptomic analysis of
M6PR lines indicated significant changes in gene expression, resulting in the pre-activation of
stress response genes and an unexpected up-regulation of biotic stress response genes. Whether
this up-regulation confers a disease resistant phenotype remains to be determined, but M6PR, by
158
demonstrating increased abiotic stress tolerance with no observed cost of resistance and
enhancement of harvest index, demonstrated its potential for crop improvement and the need for
further research.
159
APPENDIX
160
Supplemental material related to field study design
The measurement of plant density in natural Arabidopsis thaliana populations
To determine whether the planting density utilized in the field and greenhouse
experiments (see Chapter 2, 3) was within the range of densities found in natural populations, ten
natural Arabidopsis thaliana populations were located with the assistance of Dr. Brainard in late
September 2009. These populations were distributed in disturbed field margins surrounding a
field of broccoli. All populations were at most ~3 months old, as the entire site was sprayed with
glyphosate prior to planting the vegetable field in mid-June. Since the purpose of the density
measurements was to determine the highest natural population densities, the tem most dense
patches of Arabidopsis thaliana were selected. A 26x26 cm quadrant, matching the pot size
utilized in the field and greenhouse studies, was placed down and all Arabidopsis thaliana plants
from cotyledon to senescence were counted. Population densities ranged from 770 to 3250
2
plants/m , with a mean density of 1845±219 SEM (Figure S.1). All counted populations
contained plants at multiple stages of vegetative and reproductive development (Figure S.2).
Some populations were visibly younger containing no mature plants (3/10 populations), however
most contained mature plants which were already setting seed. Seedlings were visible beneath
some senesced plants indicating that two generations of plants were present within the
population. Given the time constraint from the vegetable field preparation in June, these
populations contained at least some individuals with no vernalization requirement and low seed
dormancy.
161
Figure S.1. The proportion of plants within ten natural Arabidopsis thaliana populations at three
broad developmental and reproductive stages. These categories were: pre-bolting, which
included all plants from the cotyledon to rosette stages of vegetative development; bolted, which
included plants which had bolted but not yet flowered and flowering plants without developed
siliques; and mature, which included flowering plants with well developed siliques and plants
undergoing senescence.
162
Figure S.2. The density in plants per square meter of ten natural Arabidopsis thaliana
populations near East Lansing, MI. All ten populations were in disturbed ground along the
margins of a broccoli field at the ‘Sand Hill site’ near the Michigan State University Tree
Research Facility. Populations were distributed over several acres. Population density was
determined by counting all Arabidopsis thaliana plants within a 26x26 cm quadrant.
163
Variability study to improve the growth of Arabidopsis thaliana in the growth chamber and
greenhouse
To optimize lab protocols for the growth Arabidopsis thaliana in the growth chamber and
greenhouse, an experiment was performed using two wild-type ecotypes Wassilewskija (WS)
and Columbia (Col), two soil mixtures Baccto potting soil (Michigan Peat Company) and RediEarth Plug & Seedling mix (Sun Gro Horticulture Canada Ltd.), and three pot sizes ranging from
a standard greenhouse flat (L-1020, Landmark Plastics) to small pots which fit 18 (L-1801,
Landmark Plastics) or 36 pots per flat (L-3601, Landmark Plastics) . The seeds were sterilized
using 1 mL 75% ethanol for 1 minute followed by 1 mL 15% bleach with 0.0025% SDS for 15
minutes. After washing five times with sterile dH2O, all seed was stratified at 4°C for 4days
prior to planting. Seed from each ecotype was planted into each soil/pot combination. Five seeds
were direct seeded to damp soil in each small pot and then thinned to one after the development
of four true leaves. The experiment was also performed in the growth chamber, excluding the
full sized flats. For the growth chamber seeds were germinated on sterile ½ MS media and
transplanted at the four true leaf stage into damp soil. All watering was performed via subsoil
irrigation using 1/2 strength Hoagland’s solution. Pots were allowed to saturate prior to removal
of excess solution, with watering performed as needed. All pots were randomized after each
watering to reduce location effects.
The days till germination, four-leaf stage, bolting, flowering, and senesce were recorded
for plant. Rosette diameter was measured twice weekly and seed yield determined at harvest.
WS wild-type seed production was found to exceed that of Col by 1.5-2 fold under the same
condition (location, pot size and soil type). While WS performed equally well on either soil mix,
Col yielded ~50% less when grown in Red-Earth than in Baccto. After controlling for the affects
of soil type, pot-size was found to affect fecundity and plant growth with the larger pot-sizes,
164
full-tray and 18 pot flats out growing and out yielding the 36 pot flats. To prevent ecotype bias
Baccto potting soil was utilized in all greenhouse and field experiments.
Preliminary field study to assess viability of experimental design and methods
To develop the necessary methodology to perform a large scale field study with
Arabidopsis thaliana, a three-quarters scale field study was performed under APHIS permit in
the summer and fall of 2007 at the Michigan State University Horticultural Teaching and
Research Farm. The design for the greenhouse experiments was adapted for field conditions.
Three CBF3/DREB1a lines (A28, A30, and A40) and three M6PR lines (M2-1, M5-1, M7-6)
were used in the pilot experiment. Lines from a third transgene, the vacuolar sodium transporter
NHX1, were also assessed however these lines were later determined to have been silenced by
Zhulong Chang. Seed from each transgenic line was mixed at a 50/50 ratio with seed from its
corresponding wild-type background, WS for CBF3/DREB1a and Col for M6PR, for a total of
100 seeds per population and a total of 14 replicates were planted for each mixed population. All
seed was stratified at 4°C for 3 days prior planting. Seed mixes were scattered randomly on to
26x26 cm trays in pre-moistened standard Baccto potting soil mix on 7/28/07. Seeded trays were
germinated in a hoop house and then transported to the field after reaching the four-leaf stage of
development.
A tray-in-tray potting method was developed for the field to allow for subsoil irrigation
via trickle hose (Figure S.3a). The larger lower tray also allowed for secure anchoring to the
ground, using 6 inch landscaping stakes, to prevent movement or tipping of the trays. To protect
the exposed plants from extreme weather conditions and prevent movement of plants or seeds
from the trial site, the trays were covered with clear plastic lids secured to the anchor stakes by
bungee cords. The lids were deployed no more than six hours ahead of damaging weather and
165
removed once the weather broke. Four drainage holes were drilled into the irrigation tray
approximately 1 inch from the bottom to allow excess rainwater to drain off. This system
successfully protected the plants from driving rains, sleet, early frost, and sustained high winds.
Plants were allowed to mature in the field; mature stems were harvested by hand prior to
shattering of the siliques to prevent in-field seed dispersal. Each tray was harvested 3 to 4 times
from 9/12/07 to the final harvest on 10/24/07. Harvested plant matter was placed into paper bags
and allowed to dry for at least two weeks before the seed was removed. Each bag represented the
mature seed for a single mixed population from a specific date. The seed from each bag of dried
plant material was harvested individually and the seed stored in separate Eppendorf tubes. The
separate seed collections were scored individually to develop a series of time points for each
mixed population to determine the relative seed production of the wild-type and transgenic lines
over time in competitive environments. These data give a record of the average production of
seed over time and enable us to determine the most appropriate time(s) to harvest future studies
for each line.
Soil overheating was considered to be a possible design problem, so a set of trays planted
with wild-type plants, Columbia ecotype, were buried into the soil up to the rims of the trays
while a control set was set up in the manner described earlier. Temperatures of the soil and of the
base trays were recorded on sunny days. The largest difference between buried and unburied
trays was found to be soil temperature difference of 0.5°C. This difference was not considered
significant to necessitate the labor intensive burial process which would significantly slow field
setup if used on the full experiment.
The summer/fall of 2007 field trial demonstrated that the tray-in-tray system could be
utilized to grow populations of Arabidopsis thaliana under field conditions and with some
166
modification to the experimental methods was used in all other field trials (see Chapter 2). Due
to the time intensive and potentially damaging process of in-field harvest of mature siliques, the
methods were revised to have populations removed from the field and transferred to the
greenhouse when 75% of the siliques begin turning yellow, the first sign of senescence, to
prevent seed loss in the field. Harvesting of the mixes containing either of the late-maturing
dwarf lines, A28 or A30, would include the removal of mature WS wild-type biomass, prior to
allowing the dwarf material further time to mature. This method was selected to prevent
sampling bias against the dwarf lines which would lead to an underestimation of their
competitive fitness. The removal of mature populations and the in greenhouse harvest method
also reduced in field seed loss, and thus volunteers.
Due to the preliminary nature of this study, only three transgenic lines were selected for
kanamycin screening to determine the transgene frequency of progeny seed, CBF3 lines A28 and
A40 and M6PR line M2-1. All CBF3 populations showed significantly reduced fitness in
competition with WS wild-type (Figure S.4ab), while M6PR line M2-1 showed significant gains
in fitness (Figure S.4c).
167
A.
B.
Figure S.3. Photographs of the summer/fall 2007 preliminary field trial. The field layout with the
tray-in-tray design allowing subsoil watering by trickle hose (A) and the protective lids
temporarily deployed prior to inclement weather which could endanger the plot (B).
168
Figure S.4. Fitness of transgenic plants in competition with their wild-type background ecotype
in a preliminary field study fall 2007. Fitness was calculated based on selectable marker
screening of progeny seed from mixed populations of CBF3 lines A28 and wild-type WS and
line A40 and WS (A and B respectively) and M6PR line M2-1 and wild-type Col (C). A fitness
of 1 would indicate seed production equal to wild-type (red line). Both CBF3 lines showed
significantly reduced fitness in competition (P<0.05), while M6PR line M2-1 showed increased
fitness (P<0.05).
169
Figure S.4 (cont’d)
170