Iowa State University Capstones, Theses and Creative Components Dissertations

Fall 2021

Barriers to Adoption of GM Crops

Madeline Esquivel

Follow this and additional works at: https://lib.dr.iastate.edu/creativecomponents

Part of the Agricultural Education Commons

Recommended Citation Esquivel, Madeline, "Barriers to Adoption of GM Crops" (2021). Creative Components. 731. https://lib.dr.iastate.edu/creativecomponents/731

This Creative Component is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Creative Components by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected].

Barriers to Adoption of GM Crops

By

Madeline M. Esquivel

A Creative Component submitted to the Graduate Faculty in partial fulfillment of the requirements for the degree of

MASTER OF SCIENCE

Major:

Program of Study Committee:

Walter Suza, Major Professor Thomas Lübberstedt

Iowa State University

Ames, Iowa

2021

1

Contents 1. Introduction ...... 3 2. What is a Genetically Modified Organism?...... 9 2.1 The Development of Modern Varieties and Genetically Modified Crops ...... 10 2.2 GM vs Traditional Breeding: How Are GM Crops Produced? ...... 12 2.3 Worldwide Perceptions and Acceptance of GM Crops ...... 13 3. Public Perception and Misinformation ...... 14 3.1 Framing ...... 15 3.2 Misinformation ...... 18 4.0 Potential Advantages of GM Crops ...... 21 4.1 Population Growth ...... 22 4.2 Decrease in Arable Land ...... 25 4.3 Climate Change ...... 27 5.0 Concerns around GM Crops ...... 29 5.1 Health Concerns ...... 30 5.2 Social and Economic Concerns ...... 35 5.3 Environmental Concerns ...... 38 6. Case Studies for GM Implementation, Regulation, and Perception ...... 40 6.1 Argentina ...... 40 6.2 ...... 45 7.0 Potential Routes towards Greater Acceptance ...... 50 7.1 Regulation and Political Support...... 51 7.2 Integrated Management Solutions ...... 53 7.3 Education ...... 54 References ...... 57

2

1. Introduction

Plant breeding is a multifaceted discipline; however, its definition varies in specificity based on educational and needs. According to Walter Fehr (1987), plant breeding is defined as “the art and science of the genetic improvement of plants”. Rex Bernardo (2010) builds on this by defining plant breeding as, “the genetic improvement of plants for human benefit”. More specifically, plant breeding, as a field, is a combination of both Fehr’s and

Bernardo’s definitions and is the collaborative, scientific, and strategic creation of genetically desirable plants for human benefit. Advancements in plant breeding and agriculture have been major catalysts for the rise and fall of civilizations, the development of culture, and the continual survival of the human population (Fuller and Stevens, 2019). For example, hunter gatherers’ effort to select edible kernels and larger ears made teosinte, a Mexican grass and wild relative of maize, into the modern corn we grow today (Benz, 2001). One of the largest, most well-known, and most controversial byproducts of these advancements are genetically modified organisms

(GMO).

GMO is a common term and acronym that most people have heard of, but not many people understand what a GMO is or how to define it. Ironically, some associate ‘GMO’ with human interference in the modification of organisms, yet humans have been genetically modifying organisms for thousands of years using selection breeding and crossbreeding to isolate and create plants with desirable traits (FDA, 2020a). This negative association is due to the term

GMO not being clearly defined, and freely used among consumers, the media, regulatory bodies, and even researchers. This is problematic because the term ‘GMO’ can encompass varying definitions, which may result in both negative and positive perception (Hallman et al., 2013).

Crops that have been modified using microorganisms, like , are often referred to

3 as ‘GMOs’, even though Agrobacterium mediated transformation is a much more complex and distinct process compared to a backcross scheme used in traditional breeding, or selection strategies used for domestication. In addition to ‘GMO’, crops that have been genetically engineered are commonly referred to as genetically modified (GM), genetically engineered (GE), or biotech crops (FDA, 2020a; ISAAA, 2020; USDA Economic Research Service, 2020a). For the purposes of this paper, genetically engineered crops will be referred to as GM crops.

GM crops are the product of , a process defined as the “manipulation of an organism’s by introducing, eliminating, or rearranging specific genes using the methods of modern ” (USDA, 2020). Genes are regions of DNA that encode , which in turn dictate functions, or characteristics of the organism (Nature Education, 2010a).

Traits are determined by one or more genes and genes are made of DNA (Nature Education,

2010a). Genetic engineering enables scientists to identify genes associated with desired traits, copy the DNA, transform it into another organism, and then grow that transformed organism

(FDA, 2020a). The primary advantage of genetic engineering is its ability to broaden genetic variation and diversity that may not have been possible with conventional breeding methods

(Acquaah, 2016). This is the case with the Project: rice does not contain β-carotene, the provitamin necessary for vitamin A conversion in the body and as a result, countries with high rice consumption suffer from varying levels of (Ye and Beyer, 2000).

This deficiency affects millions of children worldwide and causes blindness and exacerbates common afflictions (Ye and Beyer, 2000; Rice et al., 2004; Golden Rice Project, 2020a).

However, researchers were able to develop a genetically engineered rice to contain β-carotene, or

“Golden Rice” and engineer beneficial genetic variation that a traditional rice breeding program

4 would not have been able to create (Ye and Beyer, 2000). Figure 1 is a comparison between transgenically developed Golden Rice grains and white rice grains.

The way that GM crops are framed in the general public and represented by in the media can be problematic for researchers and confusing for consumers and regulators when assessing the safety and potential impacts of their use. Since their introduction, GM crops have generated health, environmental, social and economic concerns around safety and long-term effects.

Although some critiques raised against GM might be valid, many are based in inaccurate rhetoric and misinformation disseminated by the general public and the media. There are numerous articles claiming detrimental effects of GM crops, that have been later retracted or disproven

(McInerney, C. Bird and Nucci, 2004; Raman, 2018). For example, the Séralini experiment concluded that rats fed GM corn had an increased rate of tumor formation and was circulated

Figure 1. Golden Rice grain compared to white rice grain in screenhouse of Golden Rice plants. By International Rice Research Institute (IRRI) - https://www.flickr.com/photos/ricephotos/5516789000/in/set-72157626241604366, CC BY 2.0, https://commons.wikimedia.org/w/index.php?curid=14908001

5 widely prior to being retracted by scientific journals for being irreplicable and criticized for its experimental design and flawed findings (Séralini et al., 2012; Raman, 2017). Similarly, the

Monarch Butterfly experiment observed N4640-Bt reared larvae as eating less, growing slower, and having a higher mortality rate than normal, concluding that N4640-Bt maize could have significant off target effects (Losey et al., 1999). This study was meant to be preliminary and was later challenged for its validity and the “soundness of extrapolating laboratory assays to field testing” (Raman, 2017). The narrative underlying the framing of this study in the media was the potential loss of the Monarch butterfly due to GM crops (McInerney, C. Bird and Nucci, 2004).

Although both of these articles were published and peer reviewed, there are also numerous examples of media outlets or non-governmental organizations (NGOs) using press releases to report on studies before they are peer reviewed or published in a scientific journal (Schwartz et al., 2002). Anti-GM groups latch on to controversial studies, neglecting their shortcomings and circulating them widely to the public, pushing uncertainty and doubt into consumers about GM food technology (McInerney, C. Bird and Nucci, 2004). Groups like the Non-GMO Project have a mission of “encouraging a non-GMO seed supply” (The Non-GMO Project, 2016). This misrepresentation and negative framing of GM can be very detrimental to public perception of

GM, which can impact a customer’s willingness to purchase (WTP) GM material (Colson and

Rousu, 2013).

GM technology has the potential to be used beneficially, but there is also the potential for large companies, regulating bodies, and other entities to take advantage of the technology and use it in an unethical, irresponsible, or unsustainable manner. Critics point out that GM herbicide and pesticide resistant crops can lead to resistant weeds and pests; however, discussions over weed and pest resistance have shown that it is not the GM crop that is the problem; rather, it is

6 the mismanagement of herbicide and pesticide application that leads to resistance (Yang and

Chen, 2016). This is true particularly in the case of Roundup Ready soybeans, where growers’ reliance on as their primary weed control and poor management practices resulted in fifteen weed evolving resistance to glyphosate (Heap, 2014). Concerns have also been raised over large seed companies and their ability to use legal protection to control the GM market and dictate GM objectives (Russell and Sparrow, 2008). If large companies are the only institutions with the resources and capital to research and develop new GM crops, then the companies’ goals and objectives, normally driven by profit, might become the main breeding goals for GM material. This can overshadow GM trait research aimed at alleviating poverty and food insecurity, as they do not guarantee a return of profit (Chandrasekhara Rao and Mahendra

Dev, 2009). However, a lack of intellectual property protection can also be disadvantageous as it is in the case of India, where underregulating of its GM material has led to a flourishing illegal

GM seed market in which over half of Bt planted is with illegal seed, affecting not only the developers’ profits but the pest resistance and yield expectations of the GM material that growers are planting in their fields (Jayaraman, 2004).

The global population is expected to reach 8.5 billion in 2030 and 9.7 billion in 2050,

(United Nations Department of Economic and Social Affairs, 2019). In order to meet the demands of a growing population, global crop production needs to also increase, and this can only be achieved through agricultural extensification or intensification (Tilman et al., 2011).

Agricultural extensification will require bringing more land area into production, and intensification will require increasing yield on land already in production (Tilman et al., 2011).

Rising atmospheric CO2 levels along with changes in temperature and precipitation brought about by climate change has impacted how crops are managed, where they are grown, and its

7 effects on food security (Schmidhuber and Tubiello, 2007). These changes, brought on by climate change, will persist in creating challenges for agricultural production as temperatures continue to rise, causing negative ecological changes. The International Food Policy Research

Institute has listed agricultural adaptation to climate change as a key agenda point in meeting future challenges (Nelson et al., 2009). These challenges imply that future production will require crops that are able to exceed current yield demands on the same amount of land and adapted to warm environments and other stressors brought about by climate change. Genetic engineering and GM crops can potentially be important contributors for genetic improvement to meet many of the climate change challenges (Raman, 2018).

Based on information, perception, and framing, GM technology can take on various meaning and symbolization. For some people, GM symbolizes improvement and a hunger-free world, while for others they are unnatural and a risk to the environment and future generations

(GLP, 2020). In such a divided debate around GM material, their safety, and uses, it is important to reflect on one of the objectives of plant breeding as being beneficial to humans. GM material cannot benefit humans if growers do not adopt GM technology, and if consumers are not willing to buy them. There is a plethora of opportunities to utilize GM technology in creating more consistent yields, combatting pest and disease epidemics, and improving social and economic food inequalities in multiple ways that conventional breeding would be limited in. To obtain the desired and intended benefits, regulatory and societal obstacles need to be identified (Fedoroff,

2010). The objective of this literature review is to highlight barriers to GM adoption and provide potential routes towards greater GM acceptance.

8

2. What is a Genetically Modified Organism?

Understanding the processes of genetic modification is relevant in understanding the barriers to GM adoption. A genetically modified organism (GMO), from a literal perspective, is an organism whose genetic code has been modified in some way. However, definitions vary depending on country, regulating body, organization, and scientific convention (Edmisten,

2016). The inherent issue with this definition is its generalization. All crops have been genetically modified from their progenitor species through the process of domestication and human selection that has been done for thousands of years (ISAAA, 2020). The term GMO has not been fully defined or operationalized causing confusion between consumers, and even among researchers around what a ‘GMO’ could be referring to. Maize was cultivated nearly 9000 years ago in the Balsas region of Mexico and varies from its original form in vegetative, ear, and kernel morphology (Tenaillon and Charcosset, 2011). As seen in Figure 2, teosinte (top) contrasts significantly from a maize-teosinte hybrid (middle), and modern maize (bottom)

(Doebley, 2010). Transgenic maize genetically engineered to express (Bt)

Cry proteins, known as ‘Bt maize’ is toxic to specific and Coleoptera insects

(Hellmich, Richard; Hellmich, 2012). Without an operationalized definition, the term ‘GMO’

Figure 2. Photo By John Doebley - http://teosinte.wisc.edu/images.html, CC BY 3.0, https://commons.wikimedia.org/w/index.php?curid=3082050 (Doebley, 2010).

9 could be applied to the traditional cultivated maize as well as the transgenic maize because both have undergone genetic modification.

2.1 The Development of Modern Varieties and Genetically Modified Crops

Plant breeding has allowed for increased yields, enabling the creation and support of modern society and the global populations of today. Thousands of years ago, people in the

Middle East used genetic modification to improve crops such wheat, peas, and lentils (Pierce,

2014). The transition, from hunter-gatherers to more sedentary agricultural communities, is almost directly tied with peoples' turn towards crop domestication (Childe, 1950; Pierce, 2014).

Agricultural production guaranteed consistent food availability which allowed people time to focus on art, science, and the emergence of modern civilization (Diamond, 2002).

The concept and development of “modern crop varieties” (MVs) began in the 1950s and would serve as a kick starter for the Green Revolution of the 1960s to 1980s (Evenson et al.,

2003). The work of Dr. Norman E. Borlaug drove research at the time and he is often called the father of the Green Revolution (Kaur, 2015). Dr. Borlaug was able to cross local strains of wheat with a previously discovered dwarf wheat to create semi-dwarf varieties (Swaminathan, 2006).

He sped up the breeding process through his discovery of shuttle breeding, in which he could save breeding time by growing two generations a year versus one (Hesser, 2006). These new varieties of wheat and even rice stood out in this period due to their high-yielding capabilities and provided nations with an efficient solution to growing populations and outdated farming techniques (Kaur, 2015). The Green Revolution led to lower food prices and increased average caloric intake for consumers, along with providing productivity gains for many farmers through dramatic increases in food production and little to no increase in area planted (Evenson et al.,

2003).

10

Around the same time as the Green Revolution, a 1972 study by David Jackson, Robert

Symons and provided a scientific method for covalently joining duplex DNA molecules, launching research into genetic engineering (Jackson et al., 1972). This work helped pave the way for further genetic research, specifically a study creating the first transgenic organism, a recombinant containing DNA from both and

Staphylococcus, demonstrating that interspecies could be possible (Chang and Cohen, 1974). The first commercial product produced through recombinant DNA technology was human , which prior to its release had been derived from the pancreas glands of pork and beef as a by-product of the industry (Johnson, 1983). Human insulin from recombinant DNA technology provided a solution to supply shortage concerns and a reliable source of insulin for the increasing diabetic populations of the United States and world

(Johnson, 1983).

The discovery of Agrobacterium tumefaciens, the bacteria responsible for crown gall tumors, provided a framework for introducing recombinant DNA into plants (Chilton et al.,

1977). A dis-armed and noninfectious A. tumefaciens and part of its transfer-DNA (T-DNA) are used to introduce new genes into plant in a process called plant transformation (Chilton et al., 1980). A. tumefaciens proved to be a versatile tool for introducing foreign genes into plant genomes and has been studied extensively in varying crop species (Gelvin, 2003).

The first GM crop to be approved and sold in the United States was the ‘’ tomato in 1994 (Bruening, G.; Lyons, 2000). ‘Flavr Savr’ was developed to regulate the expression of polygalacturonase (PG), the thought to be responsible for fruit softening in ripe tomatoes (Kramer et al., 1992). Agrobacterium-mediated transformation was used to develop commercially viable breeding lines with a PG antisense mRNA, effectively lowering the

11

PG activity of the fruit (Kramer et al., 1992). Although developed to delay ripening and increase fruit quality, high production and distribution costs ultimately deemed the ‘Flavr Savr’ tomato as unsuccessful in the United States (Barben, 2010).

2.2 GM vs Traditional Breeding: How Are GM Crops Produced?

The plant breeding process can be generally summarized into four steps: set clear attainable objectives; create genetic variation; make selections based on evaluation of material; and seed certification and commercialization (Acquaah, 2016). The difference in the production of a GM crop vs a traditionally bred crop is normally in the second step of the plant breeding process: create genetic variation (Acquaah, 2016). Beneficial genetic variation can take years to research, identify, and isolate or can be impossible to create due to sexually incompatible organisms. Although GM crops are often portrayed as the product of a stand-alone process, their ability to introduce new genetic diversity makes them, “a technical advance that increases the efficiency of plant breeding” through time saved in assembling and isolating desirable genetic variation (Gepts, 2002).

Seed Certification Set Clear Attainable Create Genetic Make Selections and Objectives Variation Commercialization

Figure 3. The Plant Breeding Process. Conventional breeding, also known as traditional or classic, methods include development of cultivars through various crossing and selection strategies (Acquaah, 2016).

Goals of a breeding program will vary based on the type of crop and the program objectives, but most will involve creating variability and selecting desirable traits (Fehr, 1987). In a conventional breeding program, genetic variation is typically created by breeders who aim to

12 identify desirable crop traits and/or genes before providing this work to breeders to incorporate into their elite lines (Shimelis and Laing, 2012). Pre-breeding timelines can vary based on scale and financial support but involve years of research into characterization of landrace populations, creation of new parent populations, new trait introgression, and the creation of novel traits

(Shimelis and Laing, 2012). Reproductive barriers can hinder this process and might prevent interspecies crossing or mating (Tonosaki et al., 2016). Therefore, conventional breeding can only take advantage of genetic diversity present within these sexually crossable groups (Chahal,

G.S.; Gosal, 2002). Such reproductive barriers can limit our ability to generate the desired crop plant variability.

GM crops are often categorized based on their modification: a) increased input traits, b) enhanced output traits, and c) products beyond traditional food and fiber (Fernandez-Cornejo and

Caswell, 2006). Increased input traits consist of resistances, such as herbicide, insect, and environmental stress tolerances, while enhanced output traits add value to the crop like nutritionally enhanced crops, and products beyond traditional food and fiber apply to genetic modification of crops used for biofuels or pharmaceuticals (Fernandez-Cornejo and Caswell,

2006). Most genetic changes brought about by genetic engineering are gain of function traits, or increased input traits such as herbicide and pesticide resistance (Gepts, 2002). Once transfer has occurred, breeders use similar breeding methods to develop genetically engineered crops as they do for traditionally bred crops.

2.3 Worldwide Perceptions and Acceptance of GM Crops

Due to varying life experiences and worldviews, public consensus is hard to achieve with any subject, and the same is true of GM opposition or acceptance. A country like Argentina, with a nearly 100% adoption rate of GM technology, still has small factions of people that are

13 skeptical and vocally opposed to the country’s use of GM material (Cerier, 2018). The European

Union is often cited as one of the most strictly regulated bodies for GM material and accordingly has a large portion of its public that is anti-GM (Wunderlich and Gatto, 2015). A study assessing consumer acceptance of GMOs in Japan, Norway, Taiwan, and the United States found that

Americans were more willing to consume GM foods than Norwegians, Japanese, and Taiwanese people, who were willing to pay premiums on non-GM food of 33-40%, 55-69%, and 17-21%, respectively (Chern and Rickertsen, 2001). A meta-analysis of 51 studies analyzed average preferences and dispersion of preferences for GM food between regions and confirmed that the

EU is the most critical of GM crops with a steady increase in resistance since their introduction

(Dannenberg, 2009). Furthermore, the study found that opposition for GM food in the US is increasing at a slow rate and actually decreasing in the rest of the world (Dannenberg, 2009), which is interesting given how large countries with agricultural import power and anti-GM consumers could influence other countries willingness to grow GM material (Paarlberg, 2002).

3. Public Perception and Misinformation

Public perception of can be strongly influenced by the way in which press and media outlets choose to represent and disseminate information (Schwartz et al., 2002). A study conducted to determine consumer preference with access to GM information showed that

“positive (specific) information generated increased utility from a GM attribute while negative

(generic) information reduced utility from the presence of this attribute”, implying that different types of information provided to consumers will influence their choice behavior between GM and non-GM food (Hu et al., 2006b). Furthermore, it implies that negative information only needs to be generic in nature to cause a negative public perception, while positive information must be specific to influence consumer choices. A 2013 study conducted by Hallman et al.

14

(2013) found that despite the abundance of GM ingredients in the US, most consumers are oblivious of GM and that roughly 46% of Americans were unaware that GM material was for sale in supermarkets and only 26% believed that they had ever consumed food containing GM ingredients. This gap in knowledge and understanding of GM material is an opportunity for large players, like stakeholders, media, regulatory bodies, and even researchers, to provide information to fill this gap.

Misinformation and negative framing of GM material can have long term implications for its acceptance. The players, mentioned above, can also act as “amplification stations”, amplifying risks associated with new technologies by filtering, decoding, and processing the information, attaching social values to draw implications, interpreting and validating the information, assessing risk management, and “engaging in group or individual actions to accept, ignore, tolerate, or change the risk” (Kasperson et al., 1988).

3.1 Framing

Frames are ways in which people reflect and provide meaning to occurrences within their everyday life, but they can also be used by these amplification stations to shape interpretations of these events (Goffman, 1974). A study assessing how prior beliefs and new information from interested parties and third party sources affect consumers’ WTP for GM food found that consumers coming into the study with prior negative knowledge of GM food were less willing to accept new information and less likely to bid highly on GM food (Huffman et al., 2007).

Furthermore, their counterparts who came into the study with no prior knowledge of GM food were much more willing to accept new, positive information and apply that towards their willingness to pay for GM material (Huffman et al., 2007). Therefore, it is extremely important to educate consumers prior to their accrual of negative or biased information.

15

Effective framing can be extremely influential to the public’s opinion and perception of a topic, because it is normally done by tying peoples’ identities to the frame (Velardi and Selfa,

2021). This is the case in the framing of GM in the E.U., where, since the 1990s, GM critics have counterposed GM with ‘sustainable agriculture’, linking “GMO-free zones” with food sovereignty, ‘quality’ products, and agricultural practices that protect local farmers and environments, while framing GM as a threat to sustainable and organic farming (Levidow and

Boschert, 2008). An agricultural system does not need to be organic to be sustainable.

Sustainability is an extremely complex term to define in an operational and useful way, and assessing sustainability by focusing on the description and development of sustainable farming techniques without looking at the larger picture and socio-productive features of the specific farming system is unproductive and misguided (Rigby and Cáceres, 2001). Organic is not synonymous with sustainable, and there are cases where GM material could result in less pesticide or fertilizer usage, and less environmental degradation. For example, macozeb is a synthetic copper fungicide used to treat late blight and is considered environmentally healthier than its organic counterpart, copper sulphate, which is corrosive and toxic to humans and has harmful to very toxic ecotoxicity levels (Trewavas, 2004). GM material is capable of assisting a farming system in being more sustainable, just as it can be used in an unsustainable way

(Russell, 2008). But when GM is framed as unsustainable in the mass media, that is the context that consumers see them in.

Scientific literature on GM material had been released at a fairly stagnant rate through the

1990s, with little popular media coverage of GM crops until late 1998 and 1999 when the

Monarch Butterfly incident sparked increased media attention and articles around GM food

(McInerney, C. Bird and Nucci, 2004). The Monarch Butterfly incident began when the

16 scientific journal, Nature, published the letter “Transgenic pollen harms monarch larvae,” which presented findings indicating that monarch butterfly larvae exposed to Bt maize ate less and grew slower than those that weren’t exposed (Losey et al., 1999). The letter concluded that it was

“imperative that we gather the data necessary to evaluate the risks associated with this new agrotechnology and to compare these risks with those posed by pesticides and other pest-control tactics” (Losey et al., 1999). Given that this study was lab based and worked with small sample sizes, a call for more research was a justifiable conclusion, but it was not the conclusion that was widely circulated in the media.

Scientific publications are meant to be objective and reportorial, but popular press and media are more interpretive, sometimes oversimplifying and “putting into motion possible ripple effects of public concern” (McInerney, C. Bird and Nucci, 2004). These findings were presented as preliminary with the need for more research, due to the small sample size and scope of assessments, and was never intended to be a report of fully researched scientific findings

(McInerney, C. Bird and Nucci, 2004). However, on the same day the letter was published, the story was picked up by prominent newspapers like the New York Times, which released an article titled, “Altered corn may imperil butterfly, researchers say” (Yoon, 1999). The study’s findings, rather than being presented as preliminary, were presented as fact with an underlying narrative

Figure 4. Non-GMO Project logo By DeiFratelliTomatoes - Own work, CC BY-SA 4.0, https://commons.wikimedia.org/w/index.php?curid=89383433

17 implying the potential loss of the Monarch butterfly (McInerney, C. Bird and Nucci, 2004). This sparked immediate concern and controversy surrounding GM, and the Monarch butterfly was co- opted by anti-GM activists as a symbol of GM hazards and science going too far (McInerney, C.

Bird and Nucci, 2004). As can be seen in Figure 4, it is still currently in the label for the anti-GM group, The Non GMO Project (The Non-GMO Project, 2016).

Follow up research was conducted by independent scientists who found that on a field level, Losey et al.’s findings could not be confirmed and that the impact of Bt corn on monarch populations should remain low (Sears et al., 2001). Although this research alleviated some concern within the public about GM and its impacts on the Monarch butterfly, it was not picked up or reported on at the rate of Losey et al.’s initial study and their initial study is still a memorable event associated with the risks of GM (Marks et al., 2003). A study using content analysis to examine media coverage of GM crops found that within the UK and US, a majority of the focus in newspaper reporting has been on the risks of GM crops rather than the benefits

(Marks et al., 2003). If the news coverage of GM food is fairly minimal until a scandalous or risky event occurs (McInerney, C. Bird and Nucci, 2004), any significant news coverage or media representation of GM crops will be framed around the risks and negative, reckless, or unethical use of GM material.

3.2 Misinformation

Research and scientific findings are often presented by mass media in the form of press releases or abstracts and disseminated to the public prior to publication in a scientific journal

(Schwartz et al., 2002), meaning that scientific findings could be disseminated to the public before they are double checked or peer reviewed for authenticity. The Pusztai GM potato study is an example of how a scandalous, and later discredited, study on GM material can be picked up

18 and amplified by the media, creating long-term repercussions for GM. In 1998 Dr. Arpad Pusztai appeared on national television to report that he had conducted a study showing that GM potatoes fed to rats severely damaged their organs and overall development (Ewen and Pusztai,

1999; Marks et al., 2003). This study was not only generally discredited by many scientific societies, but was also discredited by Dr. Pusztai’s own research institute (Marks et al., 2003), yet Dr. Pusztai was given a huge platform to report on these findings and disseminate them to the public.

This broadcast severely affected GM crops, including the ‘Flavr Savr’ tomato, which, following its unsuccessful release in the US, was introduced into the UK as a processing tomato in 1996 (Bruening, G.; Lyons, 2000). Although initially well received, following the Pusztai broadcast, sales declined dramatically and distributors claimed they would no longer sell genetically engineered products (Bruening, G.; Lyons, 2000). These tomatoes were found “to be indistinguishable in almost every way from traditional tomatoes”, but they were effectively removed from shelves in the UK due to this broadcast (Bruening and Lyons, 2000). Pusztai’s study and subsequent broadcast is believed to have started “a crisis of confidence among British consumers about GM food products” (Marks et al., 2003). When people have no prior education or background in a new technology, like GM, media fills that gap, shaping their perception and giving context to what they are seeing (Best, 1991), regardless of it being true.

It is the media’s responsibility to choose which studies to amplify and how to frame them, but they are at times quick to pick up the scandalous story and slow to disseminate the attenuation. The Séralini study is cited by many anti-GM groups as an example of the detrimental effects of consuming GM food; however, it has been widely discredited by researchers and scientific groups as a flawed study with poor experimental design (Butler, 2012).

19

Originally published by Food and Chemical Toxicology in 2012, the controversial study by

Séralini et al. (2012), “RETRACTED: Long term toxicity of a Roundup herbicide and a

Roundup-tolerant genetically modified maize” concluded that there was increased tumor formation in rats fed genetically modified corn (Séralini et al., 2012). The response to this study was immediate: within hours, it had been blogged and tweeted about an estimated 1.5 million times, and within the month, anti-GM activists had destroyed a GM soybean shipment at the port of Lorient in France, a 10 year moratorium of GM crops was imposed in Peru, and all GM food imports were banned in Kenya (Huet, 2012; Bernhardt, 2012; Owino, 2012; Arjó et al., 2013).

This study had significant impacts on GM acceptance and regulation.

The Séralini study was previewed to journalists prior to its publication, on the condition that they sign a confidentiality agreement (Butler, 2012). Although confidentiality agreements are not abnormal, this agreement was strict, preventing journalists from discussing the paper with other scientists before the embargo on the agreement had expired (Butler, 2012). Journalists could not approach third party researchers for comment or ask for a second opinion (Ammann,

2016), something that could have tipped them off to the areas of the study deficient in scientific reasoning. The Organization for Economic Co-operation and Development (OECD) recommends that 10 rats of each sex per treatment are sufficient for regulatory tests when monitoring the rats for 90 days (OECD, 2018a). Séralini’s study followed this recommendation, but they monitored the rats for two years, even though rat lifespans in captivity are on average 2-3 years (Butler,

2012; Séralini et al., 2012). OECD recommends at least 20 rats per sex group for chemical- toxicity studies and 50 for carcinogenicity studies for tests of this long a duration (OECD, 2018a; b). In addition to this, Séralini used Sprague-Dawley rats, which are prone to developing spontaneous tumors, with one study exhibiting an 80% incidence of spontaneous tumors in the

20 rats, with most tumors developing after 90 days (Suzuki et al., 1979; Arjó et al., 2013).

Following many critical responses, the journal decided to review the study and assess it for statistical variability and authenticity before ultimately deciding to retract it for not reaching “the threshold of publication for Food and Chemical Toxicology” (Séralini et al., 2014). Although it was retracted, the article had already been published and communicated widely as conclusive evidence that GMOs were poison (Kuntz, 2019). Negative public perception and false narratives around GM can serve as barriers to the potential advantages they can have in addressing the world’s problems.

4.0 Potential Advantages of GM Crops

The ability to manipulate DNA and incorporate traits of pest resistance, stress resistance, increased yield, or nutritional value into crops is a powerful tool in a plant breeder’s toolbox

(Hall and Richards, 2013). More specifically, it is a tool that can be used to address many current and predicted challenges facing agronomic systems. Challenges surrounding food and food security are interwoven: predicted population growth will require more food to feed people, growing more food will require increased yield or increased cropland area, and sustainable intensive farming practices will be required to reduce agricultural practices currently contributing to global warming and climate change, both of which severely impact crop yield and food security.

Integrating GM material into current discussions around these issues can be advantageous to their future sustainable use and efficiently meeting these challenges. In Hawaii, researchers developed a transgenic resistant to the papaya ringspot (Gonsalves, 1998), a disease that infected plants quickly, was difficult to control using cultural methods, and that was ultimately responsible for a nearly 50% reduction in production (USDA & Department of

21

Agriculture State of Hawaii, 2012). As mentioned above, Golden Rice was transgenically developed to combat Vitamin A deficiency in populations with a dependence on rice as the predominant food in their diet (Golden Rice Project, 2020b). Countries with high rice consumption without meat or legumes or vegetables suffer from varying levels of Vitamin A deficiency, which can cause blindness, exacerbates afflictions such as diarrhea, respiratory diseases, and many childhood diseases like measles, and affects an estimated 124 million children worldwide (Ye and Beyer, 2000; Rice et al., 2004; Golden Rice Project, 2020a). Genetic engineering enabled researchers to quickly address a fast-spreading virus and a common nutrient deficiency in a way that would have been difficult to develop via conventional breeding. GM technology has the potential to effectively contribute to issues in agronomic systems. Below is a discussion on GM materials’ contribution to addressing the challenges of expanding human population, loss of arable land, and climate change.

4.1 Population Growth

Between 1994 and 2019, the world population grew by 2 billion to reach a total of 7.7 billion people (United Nations Department of Economic and Social Affairs, 2019). As can be seen in Figure 5, the world population is expected to continue rising (United Nations DESA

Population Division, 2019). According to the Food and Agriculture Organization of the United

Nations, of this current population, an estimated 2 billion people “do not have regular access to safe, nutritious and sufficient food” (United Nations Department of Economic and Social Affairs,

2019). To meet future population growth as well as maintain current population needs, studies have estimated that agricultural production will need to double by 2050, or at ~2.5% rate of crop production growth per year (Ray et al., 2013). Matching population demands means that many nations will need to find a way to close their yield gap (Godfray et al., 2010).

22

Figure 5. World Population Prospects 1950 - 2100. (United Nations DESA Population Division, 2019) A yield gap is the “difference between realized productivity and the best that can be achieved using current genetic material and available technologies and management,” in other words, actual yield versus yield potential (Godfray et al., 2010). Integral to closing the yield gap is improving all areas of crop management from varietal selection to best management practices based on resources, ecosystem, pest management, and knowledge (Godfray et al., 2010). Semi- dwarf wheat and rice varieties developed during the Green Revolution provided yield gains through an increased harvest index, the ratio between harvestable product and total crop biomass

(Cassman, 1999). China, India, and Indonesia are the top three rice and wheat producing nations in the world but also have low yield growth rates, with 0.7%, 1.0%, and 0.4% for rice, and 1.7%,

1.1%, and 0.8% for wheat, respectively (Ray et al., 2013). A study by Cassman (1999) estimated

23 that some parts of Asia had realized rice yields as little as 60% of the mean climate adjusted yield potential (Cassman, 1999).

In addition, Cassman (1999) estimated that most genetic yield gain in the past 30 years has been due to greater stress resistance in crops (Cassman, 1999), with approximately 70% of yield reduction being due to abiotic and biotic stress (Acquaah, 2016). In Mexico, wheat grain yield potential has increased significantly in Mexican spring wheat cultivars developed by

CIMMYT (The International Maize and Wheat Improvement Center), due to their slow rusting resistance (Sayre et al., 1998). Yield gains have also been seen in transgenic Bt corn introduced to control the Asian corn borer in the Philippines, where there was up to a 37% yield increase and 60% reduction by Bt corn farmers (Yorobe et al., 2006).

Yield increases due to genetically modified crops have been highest in low income countries in Asia, South America, and Central America (Brookes and Barfoot, 2020). This is significant because it is estimated that many of the world's most productive cropping systems in high income counties, such as the United States, are approaching their yield potential ceilings

(Cassman, 1999). A yield potential ceiling is the theoretical maximum yield that can occur in near perfect conditions (Glover, 2014); however, yield stagnation will normally occur at about

80% of the yield potential ceiling, forcing farmers to make yield gains through the elimination of small stresses (Cassman et al., 2003). Closing the yield gap through intensification of current land use is desirable (Neumann et al., 2010), but if yield does not increase at a rate sufficient for maintaining or increasing global food security, increased land usage would be needed to meet future population demands (Ray et al., 2013).

24

4.2 Decrease in Arable Land

Competition for land by human activities, urbanization, agriculture, and conservation of natural ecosystems will intensify as land availability decreases (Balmford et al., 2005).

Additionally, more and more land is being deemed unsuitable due to poor soil quality, erosion, or rapid urbanization (Nellermann et al., 2009). Soil degradation is a chief concern in the maintenance of soil quality for agricultural land and can occur through erosion by wind or water, damage by industrial pollutants and physical compaction, or even destruction due to rapid urbanization (The Royal Society, 2009). For decades, Africa, Asia, and South and Central

America contained nearly 80% of all degraded land (Cassman, 1999). Therefore, there will be a need for agricultural intensification to meet future food demands in parts of the world facing detrimental effects on land quality and the environment (Tilman et al., 2011).

Land extensification will require more land to be brought into agricultural production to meet yield demands, while land intensification requires that agriculture intensify on lands already in production. Land that is cleared for agricultural use as a function of extensification often comes from forests, grasslands, and other natural habitats (Tilman et al., 2001).

Extensification runs the major risk of exterminating vital ecosystems and incurring long-term costs greater than the short-term benefits of converting the land (Millennium Ecosystem

Assessment, 2005). In Ghana, it is estimated that over a 29-year period, 78% of the native forest was lost from deforestation due to crop expansion (Acheampong et al., 2019). Clearing new land for agriculture not only compromises native habitats, but releases CO2 previously stored in the soils and plant biomass, creating large “carbon debts” (Fargione et al., 2008). Deforestation for the expansion of crop land is occurring rapidly in the Amazon Basin and Southeast Asia and

25 contributes significantly to greenhouse gas emissions (Millennium Ecosystem Assessment,

2005).

Agricultural land intensification has been successful due to high yielding varieties, but also due to an increase in chemical fertilizers which can pollute groundwater and downstream water-ways (Matson et al., 1997). In heavily irrigated areas, solutes from irrigation water accumulate leading to heavily salinized soils (Koyro et al., 2006). High levels of soil salinity are not conducive to growth conditions of most conventional crops, and it is estimated that 45 million ha of the 230 million ha of irrigated land globally are salinized (Koyro et al., 2006). A turn towards sustainable intensification of agricultural lands may mitigate existing issues surrounding intensification and avoid the detrimental effects of expansion, such as greenhouse gas emissions and native ecosystem disruptions (Edgerton, 2009).

Land management strategies such as increasing agricultural production per unit land area, per unit fertilizer input, and per unit water consumed, along with increasing soil organic matter, have been proposed to combat some of the negative effects of extensive crop land use (Foley et al., 2005). Integrated and best management practices for nutrient use need to be established to reduce soil, air, and water pollution (Goulding et al., 2008). Transgenic herbicide tolerant crops have been able to directly affect land and soil quality through the facilitation of minimum and no-till soil management techniques which reduce soil erosion and increase organic soil matter

(Park et al., 2011). Efforts to bring previously unfit soil back into production are underway, with research being conducted to develop salt tolerant GM crops (Borsani et al., 2003). Fertilizers and soil amendments have been essential to agricultural intensification but can be an economic and environmental burden (Jewell et al., 2010). Specific biotechnological interventions in crop species, such as quick harvest lettuce with lower nitrogen requirements or rice with high affinity

26 phosphorus uptake and transporter genes, have been aimed at improving and enhancing crop efficiency with nutrient uptake and utilization to maintain and increase yield in nutrient-poor soils (Jewell et al., 2010). Integrated pest and soil management in conjunction with biotechnology advancements can have big impacts on yield and the sustainable intensification of current land (Tilman et al., 2002).

4.3 Climate Change

NASA defines climate change as the “long-term change in the average weather patterns that have come to define Earth’s local, regional, and global climates,” and global warming as the

“long-term heating of Earth’s climate system observed since the pre-industrial period (between

1850 and 1900) due to human activities” (NASA, 2020). Figure 6 shows the steep ascent of global temperatures between 1850 and 2020, with 2016 and 2020 tying for the hottest years on record (Lenssen et al., 2019; GISTEMP Team, 2021). These changes in temperature and precipitation have been primarily attributed to fossil fuel burning and heat trapping due to greenhouse gas emissions (NASA, 2020). For agriculture, climate change and global warming present several concerns and are projected to decrease water availability and quality, agricultural productivity, and exacerbate biodiversity losses (Millennium Ecosystem Assessment, 2005).

Climate change is predicted to adversely affect food security, particularly access to food, stability of food supplies, and food utilization, and to increase low income countries’ dependency on imports (Schmidhuber and Tubiello, 2007).

27

Figure 6. Global Mean Estimates based on Land and Ocean Data. Accessed 3/30/2021 GISTEMP v4 data (Lenssen et al., 2019; GISTEMP Team, 2021) The productivity used to maintain and improve agricultural yields can have an adverse effect of driving global warming through fossil fuel emissions from large scale agriculture (Park et al., 2011). The release of CO2 due to land conversion is significant and referred to as the

“carbon debt,” which cannot be repaid until biofuels from the converted land have net greenhouse gas emissions less than the life cycle emissions of the fossil fuels they displaced

(Fargione et al., 2008). Additionally, global warming and climate change can have a large impact on crop yields, with one study finding that between 1961 and 2002, the combined annual losses of three major crops, wheat, maize, and barley, due to global warming had accounted for approximately $5 billion per year (Lobell and Field, 2007).

Abiotic stress and yield can be significantly affected by warming temperatures (Lobell and Field, 2007), implying that more yield gains would have to overcome significant climate variability. Rising temperatures are expected to lead to higher rates of evapotranspiration and lower soil moisture levels (Intergovernmental Panel on Climate Change Working Group II,

28

2001). Considering that drought is one of the largest abiotic stresses affecting crop yield (Jewell et al., 2010), drought tolerance needs to be incorporated into plant genomes (Liang, 2016).

Furthermore, intensive cropping systems need to reduce agricultural contributions to global warming by decreasing net greenhouse gas emissions and/or increasing carbon storage (Park et al., 2011).

Maintaining and improving yields on a limited water supply, with higher salinity soils, and with lower nitrogen inputs will be essential in maintaining productivity in a warmer environment (Tester and Langridge, 2010). Drought resistance is a complex trait to breed for as there are a multitude of factors involved in a plant’s ability to yield well in a water limited environment (Nelson et al., 2007). Researchers found that the plant nuclear factor Y(NF-Y)B subunits confer drought tolerance in Arabidopsis and were able to develop a transgenic maize crop with increased tolerance to drought and improved yields (Nelson et al., 2007). In rice, research into drought tolerance has found that the overexpression of a stress responsive gene

(SNAC1) in transgenic rice can increase drought resistance while maintaining yields (Hu et al.,

2006a), but more intensive and integrated research needs to be conducted to further identify mechanisms of drought resistance in crops (Hu and Xiong, 2014). There have also been developments and research into GM crops adapted to high salinity soils, a complex trait to traditionally breed for, and lower nitrogen inputs, as is the case in quick harvest lettuce with low nitrogen requirements, which are both discussed above (Borsani et al., 2003; Jewell et al., 2010).

5.0 Concerns around GM Crops

Genetically modified crops’ release into the marketplace and their potential impact on human health, society and economic systems, and the environment has raised concerns with consumers, regulatory bodies, and environmental health groups (Nicolia et al., 2014). Many

29 environmental opponents to GM crops believe that they are unsustainable and a “treadmill phenomenon,” only fixing temporary problems and then needing a new solution in a few years

(Scott, 2005). Consumers note that GM technology has only been in the market since the mid-

1990s meaning that there are still long term effects yet to be seen (Yang and Chen, 2016). In addition to this, the cost and time needed to breed new varieties can serve as a deterrent to small scale competitors or public institutions, leaving most crop improvement and research goals in the hands of large companies (Yang and Chen, 2016). This is the case with projects like the Golden

Rice Project, which has consistently run into roadblocks and release delays due to high costs and strict regulations associated with bringing GM material to market (Potrykus, 2010). Many of these concerns are valid and have either been incorporated into the regulatory framework around

GM or need to be considered when assessing how GM could be beneficial to a particular challenge, like low-income population breeding objectives and pesticide and herbicide tolerance in pests and weeds.

5.1 Health Concerns

Desirable traits vary based on breeding objectives, but can be made up of one or more genes, and are referred to as upon into another organism (Nature Education,

2010b). Genes are expressed via , when the gene’s DNA sequence is copied to create RNA, and , when RNA is used to produce proteins, which determine the trait

(Nature Education, 2010b). Scientific and social debates around the safety of GM food and feed consumption focus on the safety of the inserted and the transcribed RNA, the safety of the encoded proteins by the transgene, and the safety of the intended, or unintended, effect of this transgene (Nicolia et al., 2014).

30

One of the fears surrounding the safety of the inserted transgene and the transcribed RNA is horizontal gene transfer (HGT), or lateral gene transfer, which is the non-sexual movement of genetic information between genomes of related or unrelated species (Keeling and Palmer,

2008). The potential risk of HGT between transgenic genes and bacteria present in the gastrointestinal tract (GIT) has led to concerns around these transgenes becoming incorporated into the cells of humans and livestock (Nicolia et al., 2014). Transgenic DNA transfer via the GI tract was found in rainbow trout fed genetically modified soybean meal (GM SBM), when a DNA fragment was detected in the leukocyte, head kidney and muscle of the fish

(Chainark et al., 2008). However, this DNA fragment was undetectable after a withdrawal period implying that “foreign DNA might not remain in fish following the uptake of DNA from GM

SBM, and GM SBM might be regarded as safe as non-GM SBM” (Chainark et al., 2008). In addition to this, this initial detection could have been due to a contamination in sampling.

HGT of transgenic DNA from GM plants to bacteria is generally considered a low frequency event (European Food Safety Authority, 2011), particularly the transfer of recombinant DNA from GM plants to bacteria or host cells in the GIT of mammals (Rizzi et al.,

2012). Most feed-ingested DNA is degraded upon digestion, particularly by the time it makes it to the colon or small intestine which are generally considered critical sites for HGT (Van Den

Eede et al., 2004). In addition to this, DNA of various origins such as plant, animal, microbial, and virus, have been present in human food and farm animal feed throughout history, implying that most sequences found in GM crop plants have entered the mammalian gut before now (Van

Den Eede et al., 2004). The somatic cells lining the gut and immune system, that are responsible for the uptake of this DNA have a rapid turnover, which lowers the risk of HGT (Van Den Eede et al., 2004).

31

Concerns with proteins encoded by transgenes center around the ingestion of transgenic proteins and their possible toxic or allergenic effects (Nicolia et al., 2014). Allergy risks to consumers were identified and distinguished as a) high risk with allergens or cross-reacting allergens being transferred into a food crop, b) intermediate risk with the potential for increasing the allergenicity of a crop, and c) low risk with the expression of novel proteins that may become allergens (Lehrer and Bannon, 2005). As a high risk example, a 1996 transgenic soybean study attempting to improve the nutritional quality of soybeans with the integration of a methionine- rich 2S albumin present in the Brazil nut (Bertholletia excelsa) proved that allergens can be transferred into a food crop via genetic engineering (Nordlee et al., 1996). The study noted the presence of the protein in the transgenic soybeans and the allergic reactions that it caused in subjects with nut allergies (Nordlee et al., 1996). A different soybean study, this one indicating intermediate risk, was conducted to compare the endogenous protein allergens previously identified in soybeans with Roundup Ready transgenic soybeans and to determine how genetic engineering affects these components (Burks and Fuchs, 1995). The study found that the transgenic soybeans were “qualitatively and quantitively indistinguishable from the results produced with the parental soybean variety” (Burks and Fuchs, 1995). The expression of novel proteins that may become allergens is considered low risk because it is based on the idea that proteins may be from sources which humans have not been exposed to before, meaning they could be potential new allergens (Lehrer and Bannon, 2005).

Many governmental agencies have worked to address the issue of allergenicity and GM crops. A joint consultation on the evaluation of allergenicity of GM food was organized by the

FAO and WHO and resulted in a robust allergy risk assessment process to determine if proteins developed by transgenes exhibit properties of known food allergens (Joint FAO/WHO Expert

32

Consultation on Allergenicity of Foods Derived from Biotechnology, 2001). All GM crops are assessed based on their allergenic or nonallergenic source, any proteins similar to known allergens, and their digestion stability in the GI tract (Lehrer and Bannon, 2005). In addition to this, the under or over-expression of specific allergenic proteins or the creation of new proteins are both evaluated when assessing GM food safety and unintended effects of recombinant DNA techniques (Joint FAO/WHO Expert Consultation on Allergenicity of Foods Derived from

Biotechnology, 2001). The Consultation found that after assessing the safety and nutritional aspects of GM food, that food “from modern biotechnology were inherently not less safe than those from traditional biotechnology” (Joint FAO/WHO Expert Consultation on Allergenicity of

Foods Derived from Biotechnology, 2001). Currently the FDA regulates GM products based on the nature of the product rather than the process by which it was produced and most studies have shown no long term human health impacts due to GM material (Yang and Chen, 2016). There are even studies being done to remove allergenic proteins endogenously present in some crops.

Transgene induced gene silencing was used to prevent the accumulation of an allergenic protein present in soybeans, effectively suppressing the allergenicity of the soybeans (Herman et al.,

2003).

Unintended effects of transgenes can be complicated to identify and evaluate particularly because these effects are unexpected. A bioinformatic analysis of overlap between a common viral promoter gene P35S, isolated from the cauliflower mosaic virus (CaMV), with the coding sequence of the CaMV viral gene VI identified the unintended production of a residual viral protein (Podevin and du Jardin, 2012). If the partial gene VI is transcribed, it could result in production of this residual viral protein, which was linked to leaf chlorosis, vein clearing, plant stunting, late flowering, and reduced fertility (Podevin and du Jardin, 2012). This product would

33 not advance to a commercial setting given these negative risks. There is also a concern about the combination of multiple GM traits (stacked events) into a single crop and whether this requires additional safety assessments (Nicolia et al., 2014). However, GM stacks have been found to not pose any additional risks in transgene stability and expression, and genomic changes are common in both non-GM and GM plants (Weber et al., 2012).

Safety assessments are required to determine compositional or nutritional equivalence to find in what ways a transgenic crop may be different from its traditionally bred counterpart and if any unintended effects have occurred (Nicolia et al., 2014). The 1996 FAO and WHO consultation assessing safety and nutritional aspects of genetically modified foods recommended substantial equivalence testing as “an important component in the safety assessment of foods and food ingredients derived from genetically modified plants intended for human consumption”

(Joint FAO/WHO Expert Consultation on Allergenicity of Foods Derived from Biotechnology,

2001). Equivalence testing assesses the similarities and differences between GM crops and their traditionally bred counterparts. Distinctions between GM crops modified for input traits versus modified with enhanced nutritional value will require different assessments in order to identify changes to the crop composition (European Food Safety Authority, 2011). Compositional analysis focuses on the agronomic performance and phenotypic characteristics between GM plants and their comparator, upon which differences are assessed from the perspective of potential impact on human health and animal health (European Food Safety Authority, 2011).

Bioinformatic tools have been sighted as effective tools to assess intra-species and inter-species similarities and risk potential for allergenicity and toxicity of GM food by the WHO/FAO and

EFSA (European Food Safety Authority, 2011; FAO, 2020). In the case of the CaMV promoter gene P35S overlap with gene VI, a bioinformatic study was able to determine that transcription

34 of the partial gene VI could have had negative consequences, therefore the use of the short variant of the promoter (P35S-343) was recommended (Podevin and du Jardin, 2012).

5.2 Social and Economic Concerns

Agriculture is one of the largest industries in the world and accounts for 4% of global gross domestic product (GDP) and as much as 25% of some developing countries GDP (The

World Bank, 2020). Within the United States, agricultural and food sectors account for 10.9% of all U.S. employment (USDA Economic Research Service, 2020b). However, there are growing concerns around food production becoming a commodity and source of profit and how this affects food security, research goals, and access to GM material. GM crops are regulated in a way that some believe reinforces private industry market dominance and that reduces the accessibility of GM crops to farmers in low income countries (Fischer et al., 2015).

There is a high cost associated with the development of agricultural GM technology.

Based on this fact, strong intellectual property protection is needed not only for companies to be able to recoup their investment but also to stimulate future investment into the research and development of GM technologies and products (Prasad et al., 2012). Intellectual property protection rights will vary by country and governing body but serve as a safeguard for developers, ensuring their product cannot be replicated and profited off by others for a given period. The International Convention for the Protection of New Varieties of Plants was held in

1961, 1972, and 1991, and served as a foundation for the establishment of plant breeders rights and the protection of new plant varieties (Prasad et al., 2012). Article 14 of this convention states that in respect to the new variety the following activities require the prior authorization of the breeder: “i) production or reproduction (multiplication), ii) conditioning for the purpose of propagation, iii) offering for sale, iv) selling or other marketing, v) exporting, vi) importing, vii)

35 stocking for any of the purposed mentioned in (i) to (vi), above” (UPOV, 1991). The private sector is undertaking research and development of GM material and genetic engineering because now they can profit off the technology. This gives them a large amount of power to drive where research is done, what traits are being researched, and how GM material will be distributed and marketed to the consumer.

Much of the debate around GM crops and biotechnology focuses on the environmental impact and biosafety issues rather than the issue of improving access to the technology and harnessing it to address poverty and food insecurity (Chandrasekhara Rao and Mahendra Dev,

2009). Current agricultural yields are and have been increasing, yet world hunger persists

(Magdoff et al., 2000), implying that although GM crops can serve as a tool to raise yields and meet future population demands, an increase in food production cannot be directly tied to an alleviation of hunger. GM material is not always accessible to farmers in certain countries due to the high seed costs and distribution constraints associated with GM crops (Fischer et al., 2015).

Research into beneficial traits for low income countries is significantly less than that of high income countries, with low income countries accounting for only 16% of total field trials for GM crop research in 2003 (Chandrasekhara Rao and Mahendra Dev, 2009). Low income countries have a higher potential of closing their yield gap with GM crops than high income countries that are approaching their yield ceilings (Cassman, 1999), meaning that it would be beneficial to invest in GM crop research in these countries.

There are a wide range of traits in all crops that can be developed transgenically for crop improvement, however because profit is the main motivator in GM crop research and development the scope of research is limited to that which can be “lucratively marketed”

(Chandrasekhara Rao and Mahendra Dev, 2009). Traits like insect resistance to combat local

36 pest populations, enhanced micronutrient content to reduce malnutrition, and abiotic and biotic stress resistances for yield stability are all valuable traits for poorer communities and farmers

(Food and Agriculture Organization of the United Nations, 2004). Agrochemical companies are willing to invest in creating transgenic herbicide tolerant soybean, maize and canola because not only are they effective forms of weed control but they also complement and improve the sales of the agrochemicals these companies produce (Rao, 2004). Humanitarian projects like the Golden

Rice Project do not guarantee a return of profit because the end product is meant to be distributed to low income populations freely, giving little incentive to investors to fund and develop these beneficial GM products (Potrykus, 2010). Public institutions have previously taken on the role of research rather than development and as a result are unprepared for the work of developing GM material that private companies are unwilling to invest in (Potrykus, 2010).

Critics of GM crops argue that root causes of hunger are not the non-availability of food but rather poverty, and that using GM to alleviate poverty and food insecurity is only possible with the investment in traits that are beneficial to more than just high income countries

(Chandrasekhara Rao and Mahendra Dev, 2009). Farmers do not all benefit equally when using

GM crops but studies have found that in looking at the global average, farmers have received a significant share of the profits of GM crops, as is the case with in India

(Chandrasekhara Rao and Mahendra Dev, 2009; Carpenter, 2013).

Intellectual property protection needs to be monitored by government agencies to encourage private-sector research and technology development and protect farmers from monopoly exploitation or excessive prices for their products (Food and Agriculture Organization of the United Nations, 2004). Many suggestions have been put forward to encourage breeding objectives that benefit food insecure communities and populations, including creating effective

37 regulatory framework to facilitate GM material and establishing public-private partnerships for local public sector organizations to adapt technologies developed by the private sector to combat local problems (Food and Agriculture Organization of the United Nations, 2004). Public-private partnerships would need to navigate public research goals of maximizing societal benefits and private firms goals of maximizing profits within acceptable levels of risks (Byerlee and Fischer,

2002).

5.3 Environmental Concerns

There are also concerns around sustainability and whether GM material results in less waste, less pesticide and herbicide usage, and less resource use. Long term consequences of GM crops are unpredictable but there are many concerns around the creation of weeds and pests resistant to herbicides and pesticides and how they could disturb native plant habitats or be detrimental to many beneficial insects (Maghari and Ardekani, 2011).

Weeds compete with agricultural crops for nutrients, moisture, and light, and can lead to billions of dollars in global crop losses annually (Mortensen et al., 2012). When herbicide resistant GM crops were released, growers were quick to adopt herbicides as their primary form of weed control, particularly because weed control had previously been fairly labor intensive to manage (Heap, 2014). However, the intense selection pressure exerted by these herbicides has resulted in the fast adaptive evolution of various weed populations and the development of herbicide resistant weeds (Mortensen et al., 2012). To address this, agrichemical companies have begun creating stacked herbicide resistant GM crops, combining glyphosate with other herbicide modes of action, so that weeds have to evolve multiple resistance traits in order to survive

(Wright et al., 2010). Not only can this lead to more herbicide use, as is the case with herbicide programs that recommend combining previous rates of glyphosate with 2,4-D or dicamba

38

(Mortensen et al., 2012), but there are, as of 2020, 104 weed species with resistance to multiple sites of action (Heap, 2021), meaning that weeds have been proven capable of evolving multiple resistance traits. Figure 7 shows the number of weed species with resistance to multiple sites of action (Heap, 2021).

Increase in the Number of Weeds Resistant to Two or More Herbicide Sites of Action 120

100

80

60 Two Three 40 Four 20 # of Species with Multiple Resistance Multiple with Species #of

0

Year

Figure 7. Adapted from: (Heap, 2021)

Inserting additional resistant traits and promoting weed control programs that predominantly rely on one or two herbicides is unsustainable and a reason that GM crops are criticized as being a treadmill solution to problems. Herbicide resistance is a valuable tool to have in combating weed pressure and can result in lower herbicide use when it is used in conjunction with an integrated weed management program which integrates various ecological management practices like crop rotation, cover crops, competitive crop cultivars, low till, and targeted herbicide application (Mortensen et al., 2012). A study conducted to compare corn and soybean yields in a low external input (LEI) cropping system and a conventional cropping system found that even with a 94% reduction in total herbicide use, yields matched or even

39 exceeded levels that were obtained from the conventionally managed crop (Liebman et al.,

2008). There are also examples of integrated weed management programs affectively controlling weeds that are already herbicide resistant, such as the case of glyphosate resistant horseweed that was effectively controlled through the integration of cover crops and soil-applied residual herbicides (Davis et al., 2009). GM crops that are herbicide or pesticide resistant should not be incorporated into management programs that predominantly rely on that resistance to manage their weed and pest populations, but rather should be incorporated into an integrated weed/pest management plan.

6. Case Studies for GM Implementation, Regulation, and Perception

GM material has been implemented, accepted, and used in varying ways throughout the world. This variation can lead to natural differences in crop behavior, sustainability, public acceptance, and regulation. Although there are examples of successful integration and adoption, these are not without faults or deficiencies in other areas. Below are case studies and examples of the various ways that GM has been integrated, regulated, accepted, and hindered in different countries and with different GM material.

6.1 Argentina

Argentina has been successful in its implementation of GM materials primarily due to its dynamic and transparent regulatory framework, and the usefulness of GM crops for their growers and the environment (Burachik et al., 2010). However, Argentine growers still struggle in managing weed resistance, and ensuring the efficacy of GM crops (Binimelis et al., 2009).

Argentina is currently ranked third in the world for biotech crop plantings, maintaining a 100% adoption rate, and in 2019, Argentina planted 24 million hectares of biotech crops including soybean, maize, cotton, and alfalfa (ISAAA, 2019a).

40

Argentina has been adaptable and science-based in its regulation of GM crops, and as a result has managed to stay ahead of GM developments and advancements, further facilitating their safe use within the country (Burachik, 2012). Argentina’s primary regulatory system for GE events, the National Advisory Commission of Agricultural Biotechnology (CONABIA) is nationally recognized and even considered “by the Organization for Food and Agriculture of the

United Nations (FAO) as a Center of Reference for the Biosafety of GE events” (Yankelevich,

2016). CONABIA was created to assess environmental impacts of biotech material from a technical and scientific perspective (Yankelevich, 2016). It is made up of representatives from public academia, government, and professional organizations along with representatives from private industries, making it both a multi-sector and multidisciplinary regulatory body (Burachik,

2012; Yankelevich, 2016). Argentina’s approach in creating CONABIA ensures that multiple perspectives and concerns are heard when facilitating and carrying out regulation of GM crops.

Approval for GM release is primarily based on a three-step procedure: “an assessment of the impact on agro-ecosystems, a food and feed safety evaluation, and an analysis of the production and marketing of a particular GM crop” (Adenle et al., 2017). CONABIA, with the

Advisory Technical Committee on Food Use of GMOs, CTAUOGM, and the Directorate of

Agricultural Markets, prepares the assessments of GM crops (Adenle et al., 2017), evaluating the safety of the crop to the agro-ecosystem, food supply, feed, and processing material in thorough confined release trials (Burachik et al., 2010). Confined release trials are carried out in greenhouses, field trials, and through production of regulated seed in winter nurseries to assess the risk of the gm crop in the agro-ecosystem in which it will be introduced (Burachik, 2012).

These regulatory procedures have been adjusted over time in a flexible, rational, and science- based manner (Burachik et al., 2010).

41

For instance, applications for GM approval have tripled since 1999 and regulators have remodeled their practices to match this increase (Yankelevich, 2016). Field trials previously focused on single applications for GE events, which required both time and space. An event in the context of transgenes is the insertion of a transgene at a specific location within the .

Argentina expanded these trials to include multiple applications for transformed plants with similar constructs but different insertions across the genome and different trait-related constructs

(Burachik et al., 2010). This expansion in trialing allowed researchers to evaluate multiple events within a trial versus focusing each trial on a single application, saving time and space. In addition to this development, in 2012, the Argentina’s government enacted further regulations aimed at cutting out bureaucracy and improving the efficiency of the GM crop approval timeline from four years to two (Cerier, 2018). According to a 2016 USDA study, this measure has met its expected goal in reducing the time required for approval and the bureaucracy involved in the process (Yankelevich, 2016). In 2019 the Argentinian government was able to approve nine events: six maize, two cotton and one soybean event (ISAAA, 2019a).

Argentina has also been quick to create new regulation in accordance with emerging technologies in crop biotechnology. In 2015, Argentina was one of the first countries to release official regulations as they apply to “New Breeding Techniques” (NBTs), emphasizing that all crops derived using modern biotechnology, or NBTs, would be reviewed on a case by case basis

(Schuttelaar & Partners, 2015). Debates whether crops derived from modern biotechnology warrant GM rules and regulations, even in the absence of a transgene, have led to the development of an Innovative Biotechnology, or NBT, category in GM crop regulation

(Yankelevich, 2016). In the absence of a transgene, crops produced using these new breeding techniques have the potential to be more similar to crops produced using conventional breeding

42 methods rather than transgenic crops (Orozco, 2018). For example, CRISPR-CAS9 has been effectively used for rice and wheat genome modification and is considered a new breeding technique because it results in organisms that have been modified using molecular biology but that are more alike their conventionally bred equivalents (Shan et al., 2013; Whelan and Lema,

2015). Argentina’s NBT policy ensures that new breeding techniques, like CRISPR-Cas9, will be evaluated based on their genetic modification with the potential to be regulated in the same way as conventionally bred crops.

GM crops proved to be extremely useful in addressing ongoing problems for Argentinian farmers including cutting down on management costs, effectively managing weed pressure, and increasing the profits going back to the farmer. According to a study done by Dr. Eduardo Trigo for the Argentina Council for Information and the Development of Biotechnology (ArgenBio), between 1996 and 2016, biotech crops were estimated to have generated roughly 127 billion dollars in Argentina, of which were “66% to the productive sector, 26% to the National State

(through export withholdings), and 8% to technology providers (seeds and herbicides)” (Trigo,

2016). Furthermore, Dr. Trigo estimates that over that 20 year period a total of 2,052,922 jobs were created (Trigo, 2016).

A high percentage (66% of soybeans, 45% of corn, and 95% of cotton) of the overall profit from individual GM crops goes back to the farmer, specifically (Trigo, 2016). Argentina has weak IP protection for plant varieties, with Argentina Patent Law No. 24.481/96 article 6, letter g, stating that “any kind of live material or substances already existing in nature” will not be considered an invention (The President of the Argentine Nation, 2011). Farm saved seeds in combination with weak IP protection has made seeds considerably less expensive for

Argentinian farmers, with only a 30% markup for Roundup Ready soybeans compared to a 43%

43 markup in the United States (Qaim and Traxler, 2005). These profits, along with improved yields and lower chemical pesticide inputs, were able to create enthusiasm and support from Argentine farmers towards GM crops (Yankelevich, 2016).

% OF OVERALL PROFIT FROM BIOTECH CROPS THAT GOES TO EACH SECTOR

Productive Sector National State Technology Providers

8%

26%

66%

Figure 8. Source: (Trigo, 2016) Glyphosate tolerant (“Roundup Ready”) soybeans, Argentina’s first GM crop, were approved in 1996 (Burachik et al., 2010). Crop management costs associated with hired labor and machinery went down as farmers reduced their weed control practices, like tilling (Qaim and

Traxler, 2005). Farmers adjusted their operations to low-till and no-till cultivation (Burachik et al., 2010), preventing soil erosion and the emission of greenhouse gases while also providing fuel savings (Creech, 2017). This was appealing to farmers as being more sustainable and better for soil conservation (Burachik et al., 2010).

Glyphosate has been able to almost completely replace previously used herbicides like atrazine, which was environmentally harmful and had high residual effects (Burachik et al.,

44

2010). This change is good for the safety and health of farmers and the environment, however, as discussed above, relying on a single herbicide for weed management will lead to herbicide resistant weeds. Consequently, the increased use of glyphosate has also lead to a growing number of glyphosate-tolerant or glyphosate-resistant weeds, like the glyphosate resistant biotype of johnsongrass, which now covers at least 10,000 ha of cropland in Argentina

(Binimelis et al., 2009).

6.2 India

Nearly 70% of the working population of India is in the agricultural sector (Shukla et al.,

2018). In 2002, Bt cotton was the first GM crop to be approved and introduced in India (Raman,

2018). Since then, farmer adoption rates of Bt cotton have increased from 36% to 94%, with 11.9 hectares of biotech cotton planted in 2019, the largest area of biotech crops of any other countries in Asia (ISAAA, 2019a; b). Despite high adoption rates in India, Bt cotton has become a major source of GM controversy for anti-GM groups and NGO narratives (Herring and

Chandrasekhara, 2012).

This controversy has resulted in inaccurate narratives being pushed to the public and ineffective policy being put in place to properly control GM material. Central to many of the anti-GM narratives is the accusation that Bt cotton does not improve yields or reduce pesticide use and is responsible for a dramatic increase in farmer suicides (Gruére and Sengupta, 2011).

These allegations are often due to a lack of information provided to farmers about growing conditions, the importance of planting proper seeds, and the expected earnings from using Bt cotton (Gruére and Sengupta, 2011).

Bt cotton seed is expensive, and because it is unregulated, an illegal market for unauthorized Bt seed emerged after the release of Bt cotton in 2002 (Ramani and Thutupalli,

45

2015). The illicit seeds are referred to as “stealth transgenics” and are “saved, cross-bred, repackaged, sold, exchanged, and planted,” flooding the market with non-transgenic seeds marketed as transgenic (Herring, 2007). Demand for these non-transgenic seeds is largely due to their perceived bollworm resistance, from the transgenic Bt trait, and their low price compared to legal Bt cotton (Jayaraman, 2004). In 2004, it was estimated that more than half of all GM cotton grown in the country was from illicit varieties (Jayaraman, 2004). Even worse, because illegal seed is outside of the regulation and standard of transgenic Bt cotton, many farmers believe their

Bt crop has failed when in reality they had mistakenly purchased illegal seed (Herring, 2007).

The failure of “stealth seeds” led to a misrepresentation of Bt cotton, its yield, and associated improved pesticide usage.

Prior to the introduction of Bt cotton, India's cotton yields were the lowest in the world and accounted for nearly 54% of the total pesticide consumption of the country (Ramani and

Thutupalli, 2015). Pesticides were the primary form of pest management for bollworm, the major insect pest of cotton (Raghuram, 2002). Areas with commercial crop production and good irrigation facilities accounted for most of the country’s pesticide usage, and although state agricultural departments or universities published best management practices for cultivation of crops and pesticide usage, many farmers followed their own spraying schedules, at times spraying up to double the optimal level to control for cotton pests (Shetty, 2004). For instance,

Table 1 shows the optimum and actual sprays of pesticides in hot spot districts with high pesticide usage, with the Guntur and Warangal district averaging 20-30 sprays, when the optimum recommended amount was only 15 (Shetty, 2004). Indiscriminate spraying has led to resistance in many of these hot spots, making Bt cotton less effective than it would be otherwise in controlling bollworm populations (Gruére and Sengupta, 2011).

46

Table 1. Optimum and Actual Sprays of Pesticides in Hot Spots.

District Crops Optimum Sprays Avg. No. of Sprays Followed Raichur and Bellary Paddy 8 15

Guntur and Warangal Cotton 15 20-30

Bathinda Cotton 10 15-20

Nashik Cole Crops 10 15-20

Source: (Shetty, 2004) Opponents of GM, particularly anti-GM organizations and NGOs, pushed the narrative that Bt cotton does not actually result in increased yield as promised (Herring and

Chandrasekhara, 2012). However, these narratives fail to consider how the emergence of an illegal seed market and Bt resistant pests can negatively impact Bt cotton yield and its effectiveness in combatting the bollworm pest. Additionally, they fail to acknowledge that India is a country of varying agro-ecology, and this environmental heterogeneity results in varying levels of success with Bt cotton (Herring and Chandrasekhara, 2012). There are many distinct cultivars of Bt cotton that are developed to perform best in certain conditions, so cultivars that work well in one region may not work well in another (Herring and Chandrasekhara, 2012). In such an unregulated seed market and with an innovation system primarily driven by private company goals, growers are not always given the guidance for what is best for them (Ramani and Thutupalli, 2015). However, a meta-analysis of 35 articles discussing the impact of Bt cotton in India found that on average Bt cotton has had a positive effect, and most farmers had increased yield and increased profits despite an increase in cultivation costs for Bt cotton

(Ramani and Thutupalli, 2015). These findings are not consistent with anti-GM narratives which imply that Bt cotton has failed in India, plunging farmers into insurmountable debt (Nadal,

2007).

47

In 2008, Prince Charles claimed that a rise in farmer suicides in India was a result of the failure of GM crops (Randerson, 2008). He was not the first to make this assertion, with other articles labeling Bt cotton as “seeds of destruction” and asserting that their ineffectiveness had led to the poor performance of farmers crops and their subsequent suicides (Nadal, 2007). Claims about an increase in farmer suicides due to Bt cotton fail to consider the potential of the technology versus its mismanagement. One study looked at the combined data on adoption of Bt cotton and the numbers of farmer suicides between 1997 and 2006 and found that the adoption of

Bt cotton was “not a sufficient condition for the occurrence of farmer suicides in India” (Gruére and Sengupta, 2011). Figure 9 demonstrates that the rate of farmer suicides from 1995 to 2015 has gone down since the introduction of Bt cotton in 2002. Gruére and Sengupta (2011) also pointed out that farmer suicides are “a more systematic problem in agriculture and society,” particularly the heavy indebtedness of farm households in states with high rates of suicide.

48

Figure 9. Farmer Suicides from 1995 to 2015. Bt cotton was introduced into India in 2002. By Datavizzer - Own work, CC BY-SA 4.0, https://commons.wikimedia.org/w/index.php?curid=98131313 Opposition between pro-GM and anti-GM groups have resulted in inconsistent policy around GM material. By the time Bt cotton was approved for commercial production in 2002, it was already controversial because of delays caused by anti-GM groups and NGOs, like the Gene

Campaign and Green Peace-India, protesting Bt cotton field testing and commercial release

(Ramani and Thutupalli, 2015). In 2009, the Genetic Engineering Approval Committee’s

(GEAC) approval of a new GM crop, Bt brinjal led to protests by anti-GM activists and civil society groups (Ramani and Thutupalli, 2015). Subsequently, the Ministry of Environment issued a decision imposing a moratorium on the cultivation and release of Bt brinjal (Ministry of

Environment and Forests, 2010). Initially following the moratorium, a “Technical Expert

Committee” was created to review existing regulations and open field tests of GM crops, subsequently concluding that field trials were not advisable until “gaps in the regulatory system are addressed” (Supreme Court of India, 2013). This meant that all trialing and development of

49

GM material for human consumption were halted. The moratorium on GM food crops was not lifted until 2014, yet it created a deadlocked regulatory system and the complete halting of GM crops field trials in the country for the five years it was in effect (Choudhary and Gaur, 2015).

The moratorium on GM testing created a gap in research, and even after the moratorium was lifted, there is widespread hesitation about the impacts of these new GM varieties on human and environmental health in India (Shukla et al., 2018).

When Bt cotton was introduced in the 1990s, India was unprepared to develop and monitor it, shifting the main drivers of agricultural innovation from the public to private sector

(Ramani and Thutupalli, 2015). The shift to the private sector, along with cuts in public agricultural research spending, hindered the ability of governmental agencies to advise on best practices, leaving many farmers with no means to verify claims made by private industry

(Chandrasekhara Rao and Mahendra Dev, 2009). Predictably, the lack in oversight led to an illegal seed market, indiscriminate spraying, and a general lack of farmer support and information around GM material, in some cases resulting in Bt cotton not yielding or reducing pesticide usage as much as promised. Although most Indian farmers do see increased yields and reduced pesticide usage (Chandrasekhara Rao and Mahendra Dev, 2009), cases to the contrary perpetuate anti-GM narratives and increase suspicions and animosity towards GM.

7.0 Potential Routes towards Greater Acceptance

For GM acceptance to happen on a larger scale, there must be a balance between the two predominant points of view pertaining to GM technology use. One predominant viewpoint is held by anti-GM consumers, organizations, and regulators with negative perceptions of this technology, and the other consists of dedicated researchers and scientists who view GM technology as potential contributors to solutions for many of the world’s problems. Regulation,

50 mismanagement, misinformation, and negative framing can all serve as barriers to GM adoption.

However, potential routes towards greater acceptance include effective regulation and political support, integrated management solutions, and proactive science-based education.

7.1 Regulation and Political Support

Effective regulation is difficult to achieve across countries but is an essential element in managing GM crops and their uses. Argentina is an example of how a country can proactively create regulation in step with technological developments. Not only were they one of the first countries to create a regulatory framework for NBTs, but they are also internationally recognized for their regulatory commission, CONABIA. However, there are still many organizations and anti-GM groups that believe Argentine policies to be insufficient. Taking these factors into consideration, a balance in creating GM technology regulation and political support that is sufficient for both anti-GM and pro-GM groups is difficult to achieve, but essential for overall

GM adoption.

There needs to be a standardized regulatory process on a global level that is transparent in its goals and intentions. Current international regulatory standards rely on comparative assessments as the core principle in assessing GM risks, but they are triggered at different stages and contain different approaches to field trial design, data requirements, and statistical tools

(Paoletti et al., 2008). This inconsistent framework across countries is understandable; however, a greater standardization of the values behind the regulatory policies and the instruments for assessing GM technology could help ensure greater trust and acceptance of GM products.

Furthermore, policies in one country can greatly impact the development of policies in another.

Developing and under-developed countries that are in the process of creating or revising their regulatory policies for GM material can be heavily influenced by developed and industrialized

51 nations, like the European Union, which have anti-GM sentiments (Paarlberg, 2002). From this perspective, perhaps a more consistent regulatory process would help these developing and less industrialized nations become more secure in developing GM solutions.

In addition, there is a need for greater investments into long term evaluations and monitoring for environmental and biosafety effects of GM crops, not only for scientific purposes, but to ensure that the public feels safe and informed on what they are eating (Ramani and

Thutupalli, 2015). In other words, long term evaluations and monitoring for safety measures creates a level of transparency, which allows for the consumer to trust and be more confident in

GM technology. Although this is done on a corporate level, this is not enough to gain public trust and confidence in GM technology, it must come from public or governmental, unbiased, sources.

These evaluations and studies need to be multi-disciplinary with representatives from multiple sectors to ensure that all points of view are being given a voice in the larger assessment.

Lastly, there needs to be more political support and funding for the research and development of GM foods. The cost required to gain regulatory clearance and bring a GM product to market is so expensive that it guarantees large companies will be the only drivers of

GM innovation (Kowalski et al., 2002; Potrykus, 2010). Golden Rice is a public sector product meant to be donated by its inventors for use in local rice breeding programs with no limitations on saving and replanting seed, selling seed, or selling grains (Dubock, 2019). Because the inventors are not expecting a return of profit, it is hard to raise funds for such a project. Public institutions are funded for research and “proof of concept” rather than product development, which is reserved for private institutions (Potrykus, 2010). This structure entails that GM product development is blocked for public good projects and can only be developed by large companies that can guarantee a return of their profit. New ways of recognizing merit and funding public

52 institutions for GM development need to be implemented (Potrykus, 2010). GM technology has the potential to benefit all people and yield creative and diverse solutions to global problems; therefore, it should not be limited to only the objectives of corporations and industrialized countries.

7.2 Integrated Management Solutions

GM material is an imperfect technology and should not be treated as an answer to every problem but rather an effective tool in contributing towards the solution. Furthermore, management of GM material needs to be regulated in a way that ensures their long-term efficacy which impacts environmental health and sustainability. Sustainability is achieved not by focusing on a particular farming practice, but by developing farming practices that are sustainable for the framework in which the material will be used (Rigby and Cáceres, 2001). Similarly, GM technology is only sustainable if incorporated into an agronomic system that prioritizes sustainability. In assessing a cotton producing community in Australia, Russell (2008) found that

GM has the inherent potential to become a pillar of sustainability by virtue of its “biological embeddedness,” but that its actual performance depends on the “mutual shaping” of both technology and the context in which it was implemented (Russell, 2008). The assessed growing community implemented GM Bt cotton into sustainable farming systems, where it was able to contribute further towards their sustainability, whereas farmers in this same community growing

GM HT cotton did not receive many sustainable benefits (Russell, 2008). The way that a GM crop is managed can make all the difference in its efficacy and usefulness to growers. This is the responsibility of both the policy makers and the manufacturers to make recommendations and best management practices to ensure that GM are managed in a way that ensures their long-term efficacy, rather than a quick and temporary solution. In most cases, incorporating GM materials

53 into a multi-method weed and pest management strategy is beneficial and maximizes their genetic potential.

7.3 Education

Education is the most important pathway towards greater acceptance of GM crops. GM crops have been around since the 1990s, but consumer knowledge of GM has not increased at the same rate as their adoption (Wunderlich and Gatto, 2015). A survey conducted to understand consumer attitudes and awareness towards GM foods found that 54% of American consumers knew very little about GM foods, and 25% had never heard of them (Hallman et al., 2013). Many consumers “hold at least some negative perceptions of GM food,” with only 45% agreeing that

GM foods were safe to eat (Hallman et al., 2013). The findings from this survey demonstrate the need for a more proactive effort by educators, researchers, people within the science community, and large companies to fill this gap in knowledge and be transparent about genetic modification processes and how they affect people.

Along with this, education needs to be science based and proactive. A failure to proactively educate others leaves most consumers impressionable to media and public framing of

GM. As stated by Hu et al. (2006b), consumer preferences to GM material can be influenced by the type of information they have access to, whether it be positive or negative, scientifically based or more general. Huffman et al.’s (2007) study comes to a similar conclusion, that consumers with a negative perception of GM are less likely to accept new information than consumers with no prior knowledge of GM (Huffman et al., 2007). Therefore, it is incredibly important to disseminate science-based information about GM prior to consumers’ accrual of negative, untrue, or biased information.

54

Effectively disseminating accurate information about GM technology through trusted consumer media sources is vital in gaining GM acceptance. Hu et al. (2006b) found that in assessing consumer preference with voluntary access to GM information, only one third of consumers surveyed accessed the information that was provided (Hu et al., 2006b). Further research into how consumers search for and obtain information that affects market behavior would be beneficial in determining how to disseminate more scientific based information on GM food (Hu et al., 2006b). Concurrently, policymakers, researchers, and GM industries need to assess what media or educational platforms are the most effective and trusted by consumers in order to disseminate accurate information about GM technology.

There are examples of outreach programs for GM education that are effective and can be used as models to create educational programs. The Australian Government Office of the Gene

Technology Regulator created GM kits for schools that are commercially available to help science students learn about (OGTR, 2018). Educational partnership can assist in demystifying some of the science behind GM technology and offer students a way to conceptualize what genetic modification is and ways in which GM crops are produced. “Feed

Your Mind” is a 2017 FDA education initiative focused on consumer outreach and education on biotechnology and GM crops through the distribution of science based information for consumers (FDA, 2020b). Similarly, the UC Davis Seed Biotechnology Center developed fact sheets to inform the public about the California seed industry, biotechnology, and information about GM crops (Seed Biotechnology Center, 2009).

Although these are all good examples of educational outreach, more must be done to reach more communities. Field days, public seminars, and learning modules in the form of educational videos could all help educate more people on GM crops. Proper use of social media

55 as an educational tool can also contribute to GM acceptance on a global level. An example of how this can be modeled is looking at how scientists are using the social media app “Tiktok” to creatively distribute accurate information about the COVID-19 , with one video receiving over 6 million views (Palca, 2021). Utilizing various social media platforms to disseminate information on GM can combat consumer fear and misinformation about the technology (Palca,

2021). Education is a responsibility and opportunity for scientists, researchers, and public and private institutions to reframe GM for what it is: an important technology that can be used for the betterment of society.

The purpose of this literature review is to contribute to the ongoing literature on GM technology and to build a repository of information for education purposes – specifically, to represent all the available, accurate information about GM material in a comprehensive document which can be useful to develop a better understanding and acceptance of GM technology for academic, industrial, and community purposes. This literature review is multi- faceted and can be utilized as a guide to educate and facilitate discussions about GM material outside of academia.

56

References

Acheampong, E.O., C.J. Macgregor, S. Sloan, and J. Sayer. 2019. Deforestation is driven by

agricultural expansion in Ghana’s forest reserves. Sci. African 5: e00146. doi:

10.1016/j.sciaf.2019.e00146.

Acquaah, G. 2016. Conventional plant breeding principles and techniques. Advances in Plant

Breeding Strategies: Breeding, Biotechnology and Molecular Tools. Springer International

Publishing. p. 115–158

Adenle, A.A., E.J. Morris, and D.J. Murphy. 2017. Genetically Modified Organisms in

Developing Countries: Risk Analysis and Governance.

Ammann, K. 2016. SCIENTIFIC POSITION ON SÉRALINIS THREE PAPERS FROM 2012

TO 2016. http://www.ask-force.org/web/AF-14-Seralini/Ammann-Scientific-Position-on-

Seralini-20160517.pdf (accessed 19 March 2021).

Arjó, G., M. Portero, C. Piñ Ol •, J. Viñ, X. Matias-Guiu, et al. 2013. Plurality of opinion,

scientific discourse and pseudoscience: an in depth analysis of the Séralini et al. study

claiming that Roundup TM Ready corn or the herbicide Roundup TM cause in rats.

Transgenic Res 22: 255–267. doi: 10.1007/s11248-013-9692-9.

Balmford, A., R.E. Green, and J.P.W. Scharlemann. 2005. Sparing land for nature: exploring the

potential impact of changes in agricultural yield on the area needed for crop production.

Glob. Chang. Biol. 11(10): 1594–1605. doi: 10.1111/j.1365-2486.2005.001035.x.

Barben, D. 2010. understanding of science and technology Analyzing acceptance politics:

Towards an epistemological shift in the public. doi: 10.1177/0963662509335459.

57

Benz, B.F. 2001. Archaeological evidence of teosinte domestication from Guilá Naquitz,

Oaxaca. Proc. Natl. Acad. Sci. U. S. A. 98(4): 2104–2106. doi: 10.1073/pnas.98.4.2104.

Bernhardt, A.G. 2012. 11/22/2012: Peru bans genetically modified plants. Junge Welt Dly.

Newsp.: 6. https://www.jungewelt.de/loginFailed.php?ref=/artikel/192568.peru-verbietet-

genmanipulierte-pflanzen.html (accessed 19 March 2021).

Best, J. 1991. ‘Road Warriors’ on ‘Hair-Trigger Highways’: Cultural Resources and the Media’s

Construction of the 1987. Sociol. Inq. 61(3): 327–345.

Binimelis, R., W. Pengue, and I. Monterroso. 2009. ''Transgenic treadmill": Responses to the

emergence and spread of glyphosate-resistant johnsongrass in Argentina. Geoforum. doi:

10.1016/j.geoforum.2009.03.009.

Borsani, O., V. Valpuesta, and M.A. Botella. 2003. Developing salt tolerant plants in a new

century : A molecular biology approach. Plant Cell Tissue Organ Cult. 73(May): 101–115.

doi: 10.1023/A.

Brookes, G., and P. Barfoot. 2020. GM crop technology use 1996-2018: farm income and

production impacts. GM Crops Food 11(4): 242–261. doi:

10.1080/21645698.2020.1779574.

Bruening, G.; Lyons, G.M. 2000. The Case of the FLAVR SAVR Tomato. Calif. Agric. 54(4):

6–7. http://calag.ucanr.edu/Archive/?article=ca.v054n04p6 (accessed 10 March 2021).

Burachik, M. 2012. Regulation of GM crops in Argentina. GM Crops Food 3(1): 48–51. doi:

10.4161/gmcr.18905.

Burachik, M., N. Biotechnology, B. Directorate, and B. Aires. 2010. Experience from use of

58

GMOs in Argentinian agriculture , economy and environment. N. Biotechnol. 27(5): 588–

592. doi: 10.1016/j.nbt.2010.05.011.

Burks, A.W., and R.L. Fuchs. 1995. Assessment of the endogenous allergens in glyphosate-

tolerant and commercial soybean varieties. J. Allergy Clin. Immunol. 96(6): 1008–1010.

doi: 10.1016/S0091-6749(95)70243-1.

Butler, D. 2012. Hyped GM maize study faces growing scrutiny. Nature 490(7419): 188. doi:

10.1038/490158a.

Byerlee, D., and K. Fischer. 2002. Accessing modern science: Policy and institutional options for

agricultural biotechnology in developing countries. World Dev. 30(6): 931–948. doi:

10.1016/S0305-750X(02)00013-X.

Carpenter, J.E. 2013. The socio-economic impacts of currently commercialised genetically

engineered crops. Int. J. Biotechnol. 12(4): 249–268. doi: 10.1504/IJBT.2013.059248.

Cassman, K.G. 1999. Ecological intensification of cereal production systems: Yield potential,

soil quality, and precision agriculture. Proc. Natl. Acad. Sci. U. S. A. 96(11): 5952–5959.

doi: 10.1073/pnas.96.11.5952.

Cassman, K.G., A. Dobermann, D.T. Walters, and H. Yang. 2003. MEETING CEREAL

DEMAND WHILE PROTECTING NATURAL RESOURCES AND IMPROVING

ENVIRONMENTAL QUALITY. Annu. Rev. Environ. Resour 28: 315–58. doi:

10.1146/annurev.energy.28.040202.122858.

Cerier, S. 2018. Argentina and GMOs: Exploring the nation’s long relationship with biotech

crops. Genet. Lit. Proj. https://geneticliteracyproject.org/2018/09/06/argentina-and-gmos-

59

exploring-the-nations-long-relationship-with-biotech-crops/ (accessed 10 March 2021).

Chahal, G.S.; Gosal, S.S. 2002. Principles and Procedures of Plant Breeding: Biotechnological

and Conventional Approaches. Nature.

https://books.google.com/books?hl=en&lr=&id=SmEdA8bYEB0C&oi=fnd&pg=PA1&dq=

Chahal+GS,+Gosal+SS+(2002)+Principles+and+procedures+of+plant+breeding.+Alpha+S

cience+International&ots=hCU6R_Di2A&sig=-2FpbG2-

CuET_lL2IjdPVAneoC8#v=onepage&q=conventional&f=false (accessed 10 March 2021).

Chainark, P., S. Satoh, I. Hirono, T. Aoki, and M. Endo. 2008. Availability of genetically

modified feed ingredient II: Investigations of ingested foreign DNA in rainbow trout

Oncorhynchus mykiss. Fish. Sci. 74(2): 380–390. doi: 10.1111/j.1444-2906.2008.01535.x.

Chandrasekhara Rao, N., and S. Mahendra Dev. 2009. Biotechnology and Pro-Poor Agricultural

Development. Econ. Polit. Wkly. 44(52): 56–64. doi: 10.2307/25663940.

Chang, A.C.Y., and S.N. Cohen. 1974. Genome Construction Between Bacterial Species In

Vitro: Replication and Expression of Staphylococcus Plasmid Genes in Escherichia coli

(transformation/R plasmid/antibiotic resistance/restriction endonuclease/recombination).

Proc. Natl. Acad. Sci. U. S. A. 71(4): 1030–1034. doi: 10.1073/pnas.71.4.1030.

Chern, W.S., and K. Rickertsen. 2001. Consumer Acceptance of GMO: Survey Results from

Japan, Norway, Taiwan, and the United States. Taiwan. Agric. Econ. Rev. 7(1): 1–28. doi:

https://doi.org/10.6196/TAER.2001.7.1.1.

Childe, V.G. 1950. THE URBAN REVOLUTION. Town Plan. Rev. 21(1): 3–17. doi:

108.202.115.112.

60

Chilton, M.D., M.H. Drummond, D.J. Merlo, D. Sciaky, A.L. Montoya, et al. 1977. Stable

incorporation of plasmid DNA into higher plant cells: the molecular basis of crown gall

tumorigenesis. Cell 11(2): 263–271. doi: 10.1016/0092-8674(77)90043-5.

Chilton, M.D., R.K. Saiki, and N. Yadav. 1980. T-DNA from Agrobacterium Ti plasmid is in the

nuclear DNA fraction of crown gall tumor cells. Proc. Natl. Acad. Sci. U. S. A. 77(7 II):

4060–4064. doi: 10.1073/pnas.77.7.4060.

Choudhary, B., and K. Gaur. 2015. Biotech Cotton in India, 2002 to 2014: Adoption, Impact,

Progress & Future. Ithica, NY.

Colson, G., and M.C. Rousu. 2013. What do consumer surveys and experiments reveal and

conceal about consumer preferences for genetically modified foods? GM Crops Food 4(3):

158–165. doi: 10.4161/gmcr.26322.

Creech, E. 2017. Saving Money, Time and Soil: The Economics of No-Till Farming | USDA.

Nat. Resour. Conserv. Serv. https://www.usda.gov/media/blog/2017/11/30/saving-money-

time-and-soil-economics-no-till-farming (accessed 10 March 2021).

Dannenberg, A. 2009. The dispersion and development of consumer preferences for genetically

modi fi ed food — A meta-analysis. Ecol. Econ. 68(8–9): 2182–2192. doi:

10.1016/j.ecolecon.2009.03.008.

Davis, V.M., K.D. Gibson, T.T. Bauman, S.C. Weller, and W.G. Johnson. 2009. Influence of

Weed Management Practices and Crop Rotation on Glyphosate-Resistant Horseweed (

Conyza canadensis ) Population Dynamics and Crop Yield-Years III and IV. Weed Sci.

57(4): 417–426. doi: 10.1614/WS-09-006.1.

61

Diamond, J. 2002. Evolution, consequences and future of plant and animal domestication. Nature

418(6898): 700–707. doi: 10.1038/nature01019.

Doebley, J. 2010. Doebley Lab. Univ. Wisconsin Madison. https://teosinte.wisc.edu/images.html

(accessed 31 March 2021).

Dubock, A. 2019. Golden Rice: To Combat Vitamin A Deficiency for Public Health. Vitamin A.

IntechOpen

Edgerton, M.D. 2009. Increasing crop productivity to meet global needs for feed, food, and fuel.

Plant Physiol. 149(1): 7–13. doi: 10.1104/pp.108.130195.

Edmisten, K. 2016. What Is the Difference Between Genetically Modified Organisms and

Genetically Engineered Organisms? | NC State Extension. NC State Extension, NC State

Univ. https://agbiotech.ces.ncsu.edu/q1-what-is-the-difference-between-genetically-

modified-organisms-and-genetically-engineered-organisms-we-seem-to-use-the-terms-

interchangeably/ (accessed 10 March 2021).

Van Den Eede, G., H. Aarts, H.J. Buhk, G. Corthier, H.J. Flint, et al. 2004. The relevance of

gene transfer to the safety of food and feed derived from genetically modified (GM) plants.

Food Chem. Toxicol. 42(7): 1127–1156. doi: 10.1016/j.fct.2004.02.001.

European Food Safety Authority. 2011. Guidance for risk assessment of food and feed from

genetically modified plants. EFSA J. 9(5): 2150. doi: 10.2903/j.efsa.2011.2150.

Evenson, R.E., D. Gollin, and N. Africa. 2003. Assessing the Impact of the Green Revolution ,

1960 to 2000. Science (80-. ). 300(May): 758–763. doi: 10.1126/science.1078710.

Ewen, S.W.B., and A. Pusztai. 1999. Effect of diets containing genetically modified potatoes

62

expressing Galanthus nivalis lectin on rat small intestine. Lancet 354(9187): 1353–1354.

doi: 10.1016/S0140-6736(98)05860-7.

FAO. 2020. FAO GM Foods Platform: Are we effectively evaluating food safety? Background

Paper for the Platform Community Meeting. Rome.

Fargione, J., J. Hill, D. Tilman, S. Polasky, and P. Hawthorne. 2008. Land clearing and the

biofuel carbon debt. Science (80-. ). 319(5867): 1235–1238. doi: 10.1126/science.1152747.

FDA. 2020a. Science and History of GMOs and Other Food Modification Processes | FDA.

https://www.fda.gov/food/agricultural-biotechnology/science-and-history-gmos-and-other-

food-modification-processes (accessed 10 March 2021).

FDA. 2020b. Agricultural Biotechnology | FDA. U.S. Food Drug Adm.

https://www.fda.gov/food/consumers/agricultural-biotechnology (accessed 29 March 2021).

Fedoroff, N. V. 2010. The past, present and future of crop genetic modification. N. Biotechnol.

27(5): 461–465. doi: 10.1016/j.nbt.2009.12.004.

Fehr, W.R. 1987. Principles of Cultivar Development: Theory and Technique (W.R. Fehr,

editor). McGraw-Hill, Inc.

Fernandez-Cornejo, J., and M. Caswell. 2006. Electronic Report Economic Information Bulletin

Number 11 The First Decade of Genetically Engineered Crops in the United States.

Fischer, K., E. Ekener-Petersen, L. Rydhmer, and K. Edvardsson Björnberg. 2015. Social

impacts of GM crops in agriculture: A systematic literature review. Sustain. 7(7): 8598–

8620. doi: 10.3390/su7078598.

Foley, J.A., R. DeFries, G.P. Asner, C. Barford, G. Bonan, et al. 2005. Global consequences of

63

land use. Science (80-. ). 309(5734): 570–574. doi: 10.1126/science.1111772.

Food and Agriculture Organization of the United Nations. 2004. THE STATE OF FOOD AND

AGRICULTURE 2003-2004: Agricultural Biotechnology Meeting the Needs of the Poor?

Fuller, D.Q., and C.J. Stevens. 2019. Between domestication and civilization: the role of

agriculture and arboriculture in the emergence of the first urban societies. Veg. Hist.

Archaeobot. 28(3): 263–282. doi: 10.1007/s00334-019-00727-4.

Gelvin, S.B. 2003. Agrobacterium-Mediated Plant Transformation: the Biology behind the

“Gene-Jockeying” Tool. Microbiol. Mol. Biol. Rev. 67(1): 16–37. doi:

10.1128/mmbr.67.1.16-37.2003.

Gepts, P. 2002. A Comparison between Crop Domestication, Classical Plant Breeding, and

Genetic Engineering. Crop Sci. 42(6): 1780–1790. doi: 10.2135/cropsci2002.1780.

GISTEMP Team. 2021. GISS Surface Temperature Analysis (GISTEMP), version 4. NASA

Goddard Inst. Sp. Stud. https://data.giss.nasa.gov/gistemp/ (accessed 30 March 2021).

Glover, D. 2014. Of yield gaps and yield ceilings: Making plants grow in particular places.

Geoforum 53: 184–194. doi: 10.1016/j.geoforum.2013.06.001.

GLP. 2020. Are GMOs safe? | Genetic Literacy Project. Genet. Lit. Proj.

https://geneticliteracyproject.org/gmo-faq/are-gmos-safe/ (accessed 10 March 2021).

Godfray, H.C.J., J.R. Beddington, I.R. Crute, L. Haddad, D. Lawrence, et al. 2010. Food

security: The challenge of feeding 9 billion people. Science (80-. ). 327(5967): 812–818.

doi: 10.1126/science.1185383.

Goffman, E. 1974. Frame Analysis: An essay on the Organization of Experience. Harvard

64

University Press, Cambridge, MA.

Golden Rice Project. 2020a. The Golden Rice Project. http://www.goldenrice.org/ (accessed 10

March 2021).

Golden Rice Project. 2020b. Why Golden Rice. http://www.goldenrice.org/Content3-

Why/why.php (accessed 10 March 2021).

Gonsalves, D. 1998. CONTROL OF PAPAYA RINGSPOT VIRUS IN PAPAYA: A Case

Study. Annu. Rev. Phytopathol. 36(1): 415–437. doi: 10.1146/annurev.phyto.36.1.415.

Goulding, K., S. Jarvis, and A. Whitmore. 2008. Optimizing nutrient management for farm

systems. Philos. Trans. R. Soc. B Biol. Sci. 363(1491): 667–680. doi:

10.1098/rstb.2007.2177.

Gruére, G., and D. Sengupta. 2011. Bt cotton and farmer suicides in India: An evidence-based

assessment. J. Dev. Stud. 47(2): 316–337. doi: 10.1080/00220388.2010.492863.

Hall, A.J., and R.A. Richards. 2013. Prognosis for genetic improvement of yield potential and

water-limited yield of major grain crops. F. Crop. Res. 143: 18–33. doi:

10.1016/j.fcr.2012.05.014.

Hallman, W.K., C.L. Cuite, X.K. Morin, and W.K. Hallman Professor. 2013. Public Perceptions

of Labeling Genetically Modified Foods Working Paper 2013-01.

Heap, I. 2014. Herbicide resistant weeds. Integrated Pest Management: Pesticide Problems,

Vol.3. Springer Netherlands. p. 281–301

Heap, I. 2021. Weed Populations Resistant to Multiple Sites of Action. Int. Herbic. Weed

Database. http://www.weedscience.org/Pages/Graphs/MultipleResistance.aspx (accessed 12

65

March 2021).

Hellmich, Richard; Hellmich, K. 2012. Use and Impact of Bt Maize. Nat. Educ. Knowl. 3(10):

1–9.

Herman, E.M., R.M. Helm, R. Jung, and A.J. Kinney. 2003. Genetic modification removes an

immunodominant allergen from soybean. Plant Physiol. 132(1): 36–43. doi:

10.1104/pp.103.021865.

Herring, R.J. 2007. Stealth seeds: Bioproperty, biosafety, biopolitics. J. Dev. Stud. 43(1): 130–

157. doi: 10.1080/00220380601055601.

Herring, R.J., and R.N. Chandrasekhara. 2012. On the “failure of Bt cotton”: Analysing a decade

of experience. Econ. Polit. Wkly. 47(18): 45–54.

https://www.researchgate.net/publication/289337160_On_the_’failure_of_Bt_cotton’_Anal

ysing_a_decade_of_experience (accessed 26 March 2021).

Hesser, L. 2006. Second Innovation: Shuttle Breeding. The Man Who Fed the World

Hu, H., M. Dai, J. Yao, B. Xiao, X. Li, et al. 2006a. Overexpressing a NAM, ATAF, and CUC

(NAC) enhances drought resistance and salt tolerance in rice. PNAS

29(35): 12987–12992. doi: 10.1073/pnas.0604882103.

Hu, W., M. Veeman, W. Adamowicz, and G. Gao. 2006b. Consumers’ Food Choices with

Voluntary Access to Genetic Modification Information. Can. J. Agric. Econ. Can.

d’agroeconomie 54(4): 585–604. doi: 10.1111/j.1744-7976.2006.00068.x.

Hu, H., and L. Xiong. 2014. Genetic Engineering and Breeding of Drought-Resistant Crops.

Annu. Rev. Plant Biol. 65(1): 715–741. doi: 10.1146/annurev-arplant-050213-040000.

66

Huet, F. 2012. Port of Lorient: Anti-GMO Operation. Fr. TV.

https://www.translatetheweb.com/?from=fr&to=en&dl=en&rr=UC&a=https%3A%2F%2Ff

rance3-regions.francetvinfo.fr%2Fbretagne%2F2012%2F11%2F23%2Florient-des-

militants-anti-ogm-detruisent-du-soja-importe-149823.html (accessed 19 March 2021).

Huffman, W.E., M. Rousu, J.F. Shogren, and A. Tegene. 2007. The effects of prior beliefs and

learning on consumers’ acceptance of genetically modified foods. doi:

10.1016/j.jebo.2005.04.019.

Intergovernmental Panel on Climate Change Working Group II. 2001. Climate Change 2001:

Impacts, Adaptation, and Vulnerability: Contribution of Working Group II to the Third

Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge Univ.

Press.

https://books.google.com/books?hl=en&lr=&id=RT7lQ24quc4C&oi=fnd&pg=PP9&dq=(2

001)+Climate+Change:+Impacts,+Adaptation+and+Vulnerability,+Contribution+of+Worki

ng+Group+II+to+the+Third+Assessment+Report+of+the+Intergovernmental+Panel+on+Cl

imate+Change+(Cambr (accessed 10 March 2021).

ISAAA. 2019a. Global Status of Commercialized Biotech/GM Crops: 2019 - ISAAA Brief 55-

2019 | ISAAA.org. ISAAA Br. 55. https://www.isaaa.org/resources/publications/briefs/55/

(accessed 10 March 2021).

ISAAA. 2019b. ISAAA Brief 55-2019: Executive Summary | ISAAA.org. ISAAA Br. 55.

https://www.isaaa.org/resources/publications/briefs/55/executivesummary/default.asp

(accessed 10 March 2021).

ISAAA. 2020. Q and A About Genetically Modified Crops | ISAAA.org. ISAAA.

67

https://www.isaaa.org/resources/publications/pocketk/1/ (accessed 10 March 2021).

Jackson, D.A., R.H. Symonst, P. Berg -, D.P. Lobban, and A.D. Kaiser. 1972. Biochemical

Method for Inserting New Genetic Information into DNA of Simian Virus 40: Circular

SV40 DNA Molecules Containing Genes and the Galactose Operon of

Escherichia coli (molecular hybrids/DNA joining//genetic transf. Proc.

Natl. Acad. Sci. U. S. A. 69(10): 2904–2909.

Jayaraman, K.S. 2004. Illegal seeds overtake India’s cotton fields. Nat. Biotechnol. 22(11):

1333–1334. doi: 10.1038/nbt1104-1333.

Jewell, M.C., B.C. Campbell, and I.D. Godwin. 2010. Transgenic Plants for Abiotic Stress

Resistance. Transgenic Crop Plants. Springer-Verlag, Berlin Heidelberg. p. 67–132

Johnson, I.S. 1983. Human Insulin from Recombinant DNA Technology on JSTOR.

https://www.jstor.org/stable/1690508?refreqid=excelsior%3Aef306c7907eb895736adc6ba6

1171b4d&seq=1 (accessed 10 March 2021).

Joint FAO/WHO Expert Consultation on Allergenicity of Foods Derived from Biotechnology.

2001. Evaluation of Allergenicity of Genetically Modified Foods Report of a Joint

FAO/WHO Expert Consultation on Allergenicity of Foods Derived from Biotechnology.

Kasperson, R.E., O. Renn, P. Slovic, H.S. Brown, J. Emel, et al. 1988. The Social Amplification

of Risk: A Conceptual Framework. Risk Anal. 8(2): 177–187. doi: 10.1111/j.1539-

6924.1988.tb01168.x.

Kaur, K. 2015. Plant Breeding in the 20th Century.

http://plantbreeding.coe.uga.edu/index.php?title=2._History_of_Plant_Breeding#Plant_Do

68

mestication (accessed 10 March 2021).

Keeling, P.J., and J.D. Palmer. 2008. Horizontal gene transfer in eukaryotic evolution. Nat. Rev.

Genet. 9(8): 605–618. doi: 10.1038/nrg2386.

Kowalski, S.P., R. V. Ebora, R.D. Kryder, and R.H. Potter. 2002. Transgenic crops,

biotechnology and ownership rights: what scientists need to know. Plant J. 31(4): 407–421.

doi: 10.1046/j.1365-313X.2002.01367.x.

Koyro, H.-W., N. Geissler, S. Hussin, and B. Huchzermeyer. 2006. Mechanisms Of Cash Crop

Halophytes To Maintain Yields And Reclaim Saline Soils In Arid Areas. Ecophysiology of

High Salinity Tolerant Plants. Springer Netherlands. p. 345–366

Kramer, M., R. Sanders, and C. Waters. 1992. Postharvest evaluation of transgenic tomatoes

with reduced levels of polygalacturonase : processing , firmness and disease resistance. 1:

241–255.

Kuntz, M. 2019. The Seralini affair. The dead-end of an activist science - Fondapol. Fond. pour

L’innovation Polit. https://www.fondapol.org/en/study/the-seralini-affair-the-dead-end-of-

an-activist-science/ (accessed 21 March 2021).

Lehrer, S.B., and G.A. Bannon. 2005. Risks of allergic reactions to biotech proteins in foods:

perception and reality. Allergy 60(5): 559–564. doi: 10.1111/j.1398-9995.2005.00704.x.

Lenssen, N.J.L., G.A. Schmidt, J.E. Hansen, M.J. Menne, A. Persin, et al. 2019. Improvements

in the GISTEMP Uncertainty Model. J. Geophys. Res. Atmos. 124(12): 6307–6326. doi:

10.1029/2018JD029522.

Levidow, L., and K. Boschert. 2008. Coexistence or contradiction? GM crops versus alternative

69

agricultures in Europe. Geoforum 39(1): 174–190. doi: 10.1016/j.geoforum.2007.01.001.

Liang, C. 2016. Genetically modified crops with drought tolerance: Achievements, challenges,

and perspectives. Drought Stress Tolerance in Plants, Vol 2: Molecular and Genetic

Perspectives. Springer International Publishing. p. 531–547

Liebman, M.Z., L.R. Gibson, D.N. Sundberg, A.H. Heggenstaller, and P.R. Westerman. 2008.

Statistics Publications Agronomic and Economic Performance Characteristics of

Conventional and Low-External-Input Cropping Systems in the Central Corn Belt. doi:

10.2134/agronj2007.0222.

Lobell, D.B., and C.B. Field. 2007. Global scale climate-crop yield relationships and the impacts

of recent warming Related content Global scale climate-crop yield relationships and the

impacts of recent warming. Environ. Res. Lett 2: 14002–14009. doi: 10.1088/1748-

9326/2/1/014002.

Losey, J.E., L.S. Rayor, and M.E. Carter. 1999. Transgenic pollen harms monarch larvae [5].

Nature 399(6733): 214. doi: 10.1038/20338.

Magdoff, F., J.B. Foster, and F.H. Buttel. 2000. Hungry for Profit: The Agribusiness Threat to

Farmers, Food, and the Environment. NYU Press.

Maghari, B.M., and A.M. Ardekani. 2011. Genetically modified foods and social concerns.

Avicenna J. Med. Biotechnol. 3(3): 109–117. /pmc/articles/PMC3558185/ (accessed 13

March 2021).

Marks, L.A., N. Kalaitzandonakes, K. Allison, and L. Zakharova. 2003. Media Coverage of

Agrobiotechnology: Did the Butterfly Have an Effect? J. Agribus. 21(1): 1–20. doi:

70

10.22004/AG.ECON.14674.

Matson, P.A., W.J. Parton, A.G. Power, and M.J. Swift. 1997. Agricultural intensification and

ecosystem properties. Science (80-. ). 277(5325): 504–509. doi:

10.1126/science.277.5325.504.

McInerney, C. Bird, N., and M. Nucci. 2004. The Flow of Scientific Knowledge from Lab to the

Lay Public: The Case of Genetically Modified Food . Sci. Commun. 26: 44–74.

http://scx.sagepub.com/ (accessed 11 March 2021).

Millennium Ecosystem Assessment. 2005. Ecosystems and Human Well-being: Biodiversity

Synthesis.

Ministry of Environment and Forests. 2010. Bt Brinjal: Note by Ministry of Environment and

Forests - The Hindu. The Hindu. https://www.thehindu.com/news/national/Bt-Brinjal-Note-

by-Ministry-of-Environment-and-Forests/article16578296.ece (accessed 11 March 2021).

Mortensen, D.A., J.F. Egan, B.D. Maxwell, M.R. Ryan, and R.G. Smith. 2012. Navigating a

Critical Juncture for Sustainable Weed Management. 62(1): 75–84. doi:

10.1525/bio.2012.62.1.12.

Nadal, A. 2007. , Cereal Killer GM and Agrarian Suicides in India. Znet India.

https://library.wur.nl/WebQuery/file/cogem/cogem_t4730e331_001.pdf (accessed 26 March

2021).

NASA. 2020. Global Warming vs. Climate Change | Resources – Climate Change: Vital Signs of

the Planet. NASA. https://climate.nasa.gov/resources/global-warming-vs-climate-change/

(accessed 11 March 2021).

71

Nature Education. 2010a. Essentials of Cell Biology: Differential Control of Transcription and

Translation Underlies Changes in Cell Function. Nat. Educ.

https://www.nature.com/scitable/ebooks/essentials-of-cell-biology-14749010/122996928/

(accessed 11 March 2021).

Nature Education. 2010b. Essentials of Cell Biology: . Nat. Educ.

https://www.nature.com/scitable/topicpage/gene-expression-14121669/ (accessed 11 March

2021).

Nellermann, C., M. MacDevette, T. Manders, B. Eickhout, B. Svihus, et al. 2009. THE

ENVIRONMENTAL FOOD CRISIS THE ENVIRONMENT’S ROLE IN AVERTING

FUTURE FOOD CRISES A UNEP RAPID RESPONSE ASSESSMENT.

Nelson, D.E., P.P. Repetti, T.R. Adams, R.A. Creelman, J. Wu, et al. 2007. Plant nuclear factor

Y (NF-Y) B subunits confer drought tolerance and lead to improved corn yields on water-

limited acres. Proc. Natl. Acad. Sci. U. S. A. 104(42): 16450–16455. doi:

10.1073/pnas.0707193104.

Nelson, G.C., M.W. Rosegrant, J. Koo, R. Robertson, T. Sulser, et al. 2009. REPORT F O O D P

O L I C Y CLIMATE CHANGE Impact on Agriculture and Costs of Adaptation. Am.

Assoc. Adv. Sci. 219(4585): 632–637. doi: 10.2499/0896295354.

Neumann, K., P.H. Verburg, E. Stehfest, and C. Müller. 2010. The yield gap of global grain

production: A spatial analysis. Agric. Syst. 103(5): 316–326. doi:

10.1016/j.agsy.2010.02.004.

Nicolia, A., A. Manzo, F. Veronesi, and D. Rosellini. 2014. An overview of the last 10 years of

genetically engineered crop safety research. Crit. Rev. Biotechnol. 34(1): 77–88. doi:

72

10.3109/07388551.2013.823595.

Nordlee, J.A., S.L. Taylor, J.A. Townsend, L.A. Thomas, and R.K. Bush. 1996. Identification of

a Brazil-Nut Allergen in Transgenic Soybeans. N. Engl. J. Med. 334(11): 688–692. doi:

10.1056/NEJM199603143341103.

OECD. 2018a. OECD/OCDE 453 OECD GUIDELINE FOR THE TESTING OF CHEMICALS

COMBINED CHRONIC TOXICITY\CARCINOGENICITY STUDIES.

OECD. 2018b. OECD GUIDELINE FOR THE TESTING OF CHEMICALS Repeated Dose 90-

day Oral Toxicity Study in Rodents.

OGTR. 2018. GM Kits in Schools. Aust. Gov. Dep. Heal. Off. Gene Technol. Regul.

http://www.ogtr.gov.au/internet/ogtr/publishing.nsf/Content/9AA09BB4515EBAA2CA257

D6B00155C53/$File/05 - GMOs in Schools.pdf (accessed 29 March 2021).

Orozco, P. 2018. Argentina and Brazil merge law and science to regulate new breeding

techniques. Cornell Alliance Sci.

https://allianceforscience.cornell.edu/blog/2018/01/argentina-and-brazil-merge-law-and-

science-to-regulate-new-breeding-techniques/ (accessed 11 March 2021).

Owino, O. 2012. Scientists torn over Kenya’s recent GM food ban. Nature. doi:

10.1038/nature.2012.11929.

Paarlberg, R.L. 2002. The real threat to GM crops in poor countries: Consumer and policy

resistance to GM foods in rich countries. Food Policy 27(3): 247–250. doi: 10.1016/S0306-

9192(02)00014-3.

Palca, J. 2021. The Viral TikTok That Explains Vaccine Science - And Makes You Laugh. Natl.

73

Public Radio. https://www.npr.org/sections/goatsandsoda/2021/04/01/983397422/the-viral-

tiktok-video-that-explains-vaccine-science-and-makes-you-laugh (accessed 6 April 2021).

Paoletti, C., E. Flamm, W. Yan, S. Meek, S. Renckens, et al. 2008. GMO risk assessment around

the world: Some examples. Trends Food Sci. Technol. 19(SUPPL. 1): S70–S78. doi:

10.1016/j.tifs.2008.07.007.

Park, J.R., I. McFarlane, R. Hartley Phipps, and G. Ceddia. 2011. The role of transgenic crops in

sustainable development. Plant Biotechnol. J. 9(1): 2–21. doi: 10.1111/j.1467-

7652.2010.00565.x.

Pierce, B.A. 2014. Genetics: A Conceptual Approach (L. Schultz, A. Bristow, A. Michael, G.L.

Hadler, J. Furino, et al., editors). 5th ed. W.H. Freeman and Company, New York.

Podevin, N., and P. du Jardin. 2012. Possible consequences of the overlap between the CaMV

35S promoter regions in plant transformation vectors used and the viral gene VI in

transgenic plants. GM Crops Food 3(4): 296–300. doi: 10.4161/gmcr.21406.

Potrykus, I. 2010. Lessons from the “Humanitarian Golden Rice” project: Regulation prevents

development of public good genetically engineered crop products. N. Biotechnol. 27(5):

466–472. doi: 10.1016/j.nbt.2010.07.012.

Prasad, R., U.S. Bagde, and A. Varma. 2012. An overview of intellectual property rights in

relation to agricultural biotechnology. AFRICAN J. Biotechnol. 11(73). doi:

10.5897/ajb12.262.

Qaim, M., and G. Traxler. 2005. Roundup Ready soybeans in Argentina: farm level and

aggregate welfare effects. Agric. Econ. 32(1): 73–86. doi: 10.1111/j.0169-

74

5150.2005.00006.x.

Raghuram, N. 2002. India joins the GM club. Trends Plant Sci. 7(7): 322–323. doi:

10.1016/S1360-1385(02)02286-0.

Raman, R. 2017. The impact of Genetically Modified (GM) crops in modern agriculture: A

review. GM Crops Food 8(4): 195–208. doi: 10.1080/21645698.2017.1413522.

Raman, R. 2018. Biotechnology in Agriculture and the Food Chain The impact of Genetically

Modified ( GM ) crops in modern agriculture : A review The impact of Genetically

Modified ( GM ) crops in modern agriculture : A review. 5698. doi:

10.1080/21645698.2017.1413522.

Ramani, S. V., and A. Thutupalli. 2015. Emergence of controversy in technology transitions:

Green Revolution and Bt cotton in India. Technol. Forecast. Soc. Change 100: 198–212.

doi: 10.1016/j.techfore.2015.06.018.

Randerson, J. 2008. Indian farmer suicides not GM related, says study | GM | The Guardian.

Guard. https://www.theguardian.com/environment/2008/nov/05/gmcrops-india (accessed 26

March 2021).

Rao, N.C. 2004. Plant Biotechnology and the Emerging Scenario. Rev. Dev. Chang. 9(1): 69–92.

doi: 10.1177/0972266120040103.

Ray, D.K., N.D. Mueller, P.C. West, and J.A. Foley. 2013. Yield Trends Are Insufficient to

Double Global Crop Production by 2050 (J.P. Hart, editor). PLoS One 8(6): e66428. doi:

10.1371/journal.pone.0066428.

Rice, A.L., K.P. West Jr., and R.E. Black. 2004. Vitamin A Deficiency. Comparative

75

Quantification of Health Risks. p. 211–256

Rigby, D., and D. Cáceres. 2001. Organic farming and the sustainability of agricultural systems.

Agric. Syst. 68(1): 21–40. doi: 10.1016/S0308-521X(00)00060-3.

Rizzi, A., N. Raddadi, C. Sorlini, L. Nordgrd, K.M. Nielsen, et al. 2012. The Stability and

Degradation of Dietary DNA in the Gastrointestinal Tract of Mammals: Implications for

Horizontal Gene Transfer and the Biosafety of GMOs. Crit. Rev. Food Sci. Nutr. 52(2):

142–161. doi: 10.1080/10408398.2010.499480.

Russell, A.W. 2008. GMOs and their contexts: A comparison of potential and actual

performance of GM crops in a local agricultural setting. Geoforum 39(1): 213–222. doi:

10.1016/j.geoforum.2007.04.001.

Russell, W.A., and R. Sparrow. 2008. The Case for Regulating Intragenic GMOs. J. Agric.

Environ. Ethics 21: 153–181. https://robsparrow.com/wp-content/uploads/The-case-for-

regulating-intragenic-GMOs.pdf (accessed 12 March 2021).

Sayre, K.D., R.P. Singh, J. Huerta‐Espino, and S. Rajaram. 1998. Genetic Progress in Reducing

Losses to Leaf Rust in CIMMYT‐Derived Mexican Spring Wheat Cultivars. Crop Sci.

38(3): 654–659. doi: 10.2135/cropsci1998.0011183X003800030006x.

Schmidhuber, J., and F.N. Tubiello. 2007. Global food security under climate change. Proc. Natl.

Acad. Sci. U. S. A. 104(50): 19703–19708. doi: 10.1073/pnas.0701976104.

Schuttelaar & Partners. 2015. The regulatory status of New Breeding Techniques in countries

outside the European Union Developed.

Schwartz, L.M., S. Woloshin, and L. Baczek. 2002. Media coverage of scientific meetings: Too

76

much, too soon? J. Am. Med. Assoc. 287(21): 2859–2863. doi: 10.1001/jama.287.21.2859.

Scott, D. 2005. The magic bullet criticism of agricultural biotechnology. J. Agric. Environ.

Ethics 18(3): 259–267. doi: 10.1007/s10806-005-0632-4.

Sears, M.K., R.L. Hellmich, D.E. Stanley-Horn, K.S. Oberhauser, J.M. Pleasants, et al. 2001.

Impact of Bt corn pollen on monarch butterfly populations: A risk assessment. Proc. Natl.

Acad. Sci. U. S. A. 98(21): 11937–11942. doi: 10.1073/pnas.211329998.

Seed Biotechnology Center. 2009. California Seed Association fact sheets - Seed Biotechnology

Center. Univ. California, Davis. http://sbc.ucdavis.edu/Outreach_-

_Public_Service_pages/California_Seed_Association_fact_sheets/ (accessed 1 April 2021).

Séralini, G.E., E. Clair, R. Mesnage, S. Gress, N. Defarge, et al. 2012. Long term toxicity of a

Roundup herbicide and a Roundup-tolerant genetically modified maize. Food Chem.

Toxicol. 50(11): 4221–4231. doi: 10.1016/j.fct.2012.08.005.

Séralini, G.E., E. Clair, R. Mesnage, S. Gress, N. Defarge, et al. 2014. Retraction notice to Long

term toxicity of a Roundup herbicide and a Roundup-tolerant genetically modified maize

[Food Chem. Toxicol. 50 (2012) 4221-4231]. Food Chem. Toxicol. 63: 244. doi:

10.1016/j.fct.2013.11.047.

Shan, Q., Y. Wang, J. Li, Y. Zhang, K. Chen, et al. 2013. Targeted genome modification of crop

plants using a CRISPR-Cas system. Nat. Biotechnol. 31(8): 686–688. doi:

10.1038/nbt.2650.

Shetty, P.K. 2004. Socio-Ecological Implications of Pesticide Use in India. Econ. Polit. Wkly.

39(49): 5261–5267. doi: 108.202.115.112.

77

Shimelis, H., and M.D. Laing. 2012. Timelines in conventional crop improvement: Pre-breeding

and breeding procedures. Aust. J. Crop Sci. 6(11): 1542–1549.

https://www.researchgate.net/publication/256086199_Timelines_in_conventional_crop_imp

rovement_Pre-breeding_and_breeding_procedures (accessed 11 March 2021).

Shukla, M., K.T. Al-Busaidi, M. Trivedi, and R.K. Tiwari. 2018. Status of research, regulations

and challenges for genetically modified crops in India. 9(4): 173–188. doi:

10.1080/21645698.2018.1529518.

Supreme Court of India. 2013. Final report of the Technical Expert Committee (TEC) set up by

the Supreme Court in a Public Interest Litigation on Genetically Modified Organisms

(GMOs). India Environ. Portal.

http://www.indiaenvironmentportal.org.in/content/380047/final-report-of-the-technical-

expert-committee-tec-set-up-by-the-supreme-court-in-a-public-interest-litigation-on-

genetically-modified-organisms-gmos/ (accessed 11 March 2021).

Suzuki, H., U. Mohr, and G. Kimmerle. 1979. Spontaneous Endocrine Tumors in Sprague-

Dawley Rats*. Springer-Verlag.

Swaminathan, M.S. 2006. An Evergreen Revolution. Crop Sci. 46(5): 2293–2303. doi:

10.2135/cropsci2006.9999.

Tenaillon, M.I., and A. Charcosset. 2011. A European perspective on maize history. Comptes

Rendus - Biol. 334(3): 221–228. doi: 10.1016/j.crvi.2010.12.015.

Tester, M., and P. Langridge. 2010. Breeding technologies to increase crop production in a

changing world. Science (80-. ). 327(5967): 818–822. doi: 10.1126/science.1183700.

78

The Non-GMO Project. 2016. Mission – The Non-GMO Project. Non-GMO Proj.

https://www.nongmoproject.org/about/mission/ (accessed 11 March 2021).

The President of the Argentine Nation. 2011. Law on Patents and Utility Models. (20): 1–50.

The Royal Society. 2009. Reaping the benefits: Science and the sustainable intensification of

global agriculture.

The World Bank. 2020. Agriculture Overview. World Bank.

https://www.worldbank.org/en/topic/agriculture/overview (accessed 11 March 2021).

Tilman, D., C. Balzer, J. Hill, and B.L. Befort. 2011. Global food demand and the sustainable

intensification of agriculture. PNAS 108(50): 20260–20264. doi:

10.1073/pnas.1116437108.

Tilman, D., K.G. Cassman, P.A. Matson, R. Naylor, and S. Polasky. 2002. Agricultural

sustainability and intensive production practices. Nature 418(6898): 671–677. doi:

10.1038/nature01014.

Tilman, D., J. Fargione, B. Wolff, C. D’Antonio, A. Dobson, et al. 2001. Forecasting

agriculturally driven global environmental change. Science (80-. ). 292(5515): 281–284.

doi: 10.1126/science.1057544.

Tonosaki, K., K. Osabe, T. Kawanabe, and R. Fujimoto. 2016. The importance of reproductive

barriers and the effect of allopolyploidization on crop breeding. Breed. Sci. 66(3): 333–349.

doi: 10.1270/jsbbs.15114.

Trewavas, A. 2004. A critical assessment of organic farming-and-food assertions with particular

respect to the UK and the potential environmental benefits of no-till agriculture. Crop Prot.

79

23(9): 757–781. doi: 10.1016/j.cropro.2004.01.009.

Trigo, E.J. 2016. VEINTE AÑOS DE CULTIVOS GENÉTICAMENTE MODIFICADOS EN

LA AGRICULTURA ARGENTINA.

United Nations Department of Economic and Social Affairs. 2019. World Population Prospects

2019 Highlights. New York.

United Nations DESA Population Division. 2019. World Population Prospects - 2019. United

Nations Dep. Econ. Soc. Aff. Popul. Dyn.

https://population.un.org/wpp/Graphs/DemographicProfiles/Line/900 (accessed 6 April

2021).

UPOV. 1991. International Convention for the Protection of New Varieties of Plants

International Union for the Protection of New Varieties of Plants. Geneva.

USDA. 2020. Agricultural Biotechnology Glossary | USDA. USDA.

https://www.usda.gov/topics/biotechnology/biotechnology-glossary (accessed 11 March

2021).

USDA & Department of Agriculture State of Hawaii. 2012. NATIONAL AGRICULTURAL

STATISTICS SERVICE In cooperation with Hawaii Farm Facts. USDA.

www.nass.usda.gov/hi (accessed 11 March 2021).

USDA Economic Research Service. 2020a. USDA ERS - Recent Trends in GE Adoption.

USDA. https://www.ers.usda.gov/data-products/adoption-of-genetically-engineered-crops-

in-the-us/recent-trends-in-ge-adoption.aspx (accessed 11 March 2021).

USDA Economic Research Service. 2020b. USDA ERS - Ag and Food Sectors and the

80

Economy. USDA. https://www.ers.usda.gov/data-products/ag-and-food-statistics-charting-

the-essentials/ag-and-food-sectors-and-the-economy/ (accessed 11 March 2021).

Velardi, S., and T. Selfa. 2021. Framing local: an analysis of framing strategies for genetically

modified organism (GMO) labeling initiatives in the northeastern U.S. Agroecol. Sustain.

Food Syst. 45(3): 366–389. doi: 10.1080/21683565.2020.1818159.

Weber, N., C. Halpin, L. Curtis Hannah, J.M. Jez, J. Kough, et al. 2012. Editor’s Choice: Crop

genome plasticity and its Relevance to food and feed safety of genetically engineered

breeding stacks. Plant Physiol. 160(4): 1842–1853. doi: 10.1104/pp.112.204271.

Whelan, A.I., and M.A. Lema. 2015. Regulatory framework for gene editing and other new

breeding techniques (NBTs) in Argentina. GM Crops Food 6(4): 253–265. doi:

10.1080/21645698.2015.1114698.

Wright, T.R., G. Shan, T.A. Walsha, J.M. Lira, C. Cui, et al. 2010. Robust crop resistance to

broadleaf and grass herbicides provided by aryloxyalkanoate dioxygenase transgenes. Proc.

Natl. Acad. Sci. U. S. A. 107(47): 20240–20245. doi: 10.1073/pnas.1013154107.

Wunderlich, S., and K.A. Gatto. 2015. Consumer perception of genetically modified organisms

and sources of information. Adv. Nutr. 6(6): 842–851. doi: 10.3945/an.115.008870.

Yang, Y.T., and B. Chen. 2016. Governing GMOs in the USA: science, law and public health. J.

Sci. Food Agric. 96(6): 1851–1855. doi: 10.1002/jsfa.7523.

Yankelevich, A. 2016. USDA Foreign Agricultural Service - Argentina Annual Biotechnology

Report.

Ye, X., and P. Beyer. 2000. Engineering the provitamin A (β-carotene) biosynthetic pathway into

81

(carotenoid-free) rice endosperm. Science (80-. ). 287(5451): 303–305. doi:

10.1126/science.287.5451.303.

Yoon, C.K. 1999. Altered Corn May Imperil Butterfly, Researchers Say - The New York Times.

New York Times. https://www.nytimes.com/1999/05/20/us/altered-corn-may-imperil-

butterfly-researchers-say.html (accessed 18 March 2021).

Yorobe, J., C. Quicoy, J.M. Yorobe, C.B. Quicoy, J.M. Yorobe, et al. 2006. Economic Impact of

Bt Corn in the Philippines. Philipp. Agric. Sci. 89(3).

https://www.researchgate.net/publication/279560297 (accessed 11 March 2021).

82