Anda di halaman 1dari 123

Thesis for the Degree of Master

Transitioning to Sustainable Agriculture in


Argentina’s Pampa-region: a Mixed Method
Analysis

by

JONAS ADRIAENSENS

Department of Food and Resource Economics

Graduate School

Korea University

June 2019
Transitioning to Sustainable Agriculture in
Argentina’s Pampa-region: a Mixed Method
Analysis

A Thesis
by
JONAS ADRIAENSENS

Submitted to the Graduate School of


Korea University in Partial Fulfillment of the
Requirements for the Degree of

Master of Economics

Department of Food and Resource Economics


Korea University

June 2019
KOREA UNIVERSITY
GRADUATE COMMITTEE APPROVAL

of a thesis submitted by

JONAS ADRIAENSENS

Approved as to style and content by:

Lim, Song Soo


(The Chair of Committee)

Suh, Dong Hee

(Committee Member)

D’Haese, Marijke

(Committee Member)

June 2019
Son cosas chiquitas. They are small things.
No acaban con la pobreza They do not end poverty,
no nos sacan del subdesarrollo, do not remove us from underdevelopment,
no socializan los medios de producción do not socialize the means of production
y de cambio, no expropian las cuevas de Alí Babá. and in turn, do not expropriate the caves of Ali Baba.

Pero quizá desencadenen la alegría de hacer, But they may spark the joy of doing,
y la traduzcan en actos. and translate it into acts.

Y al fin y al cabo, actuar sobre la realidad And after all, acting on reality
y cambiarla aunque sea un poquito, and changing it, even a little,
es la única manera de probar is the only way to prove
que la realidad es transformable. that reality is transformable.

EDUARDO GALEANO
Son cosas chiquitas
Transitioning to Sustainable Agriculture in
Argentina’s Pampa-region: a Mixed Method
Analysis
JONAS ADRIAENSENS

Department of Food and Resource Economics


Chair of Advisory: Lim, Song Soo

Acknowledgements

First and foremost, I want to thank all the people who have guided me on this journey. I would like to
thank Lim Song-Soo, who gave me the freedom to venture into my academic interests but also guided
me whenever I needed it. Also, I would like to thank to both committee members, Suh Dong-Hee and
D’Haese Marijke for taking the time to analyse my thesis and engage with it critically.

I want to also thank Bie, Jos and Staf, who gave me invaluable advice of areas that I still needed to work
on. Although I am certain that this thesis still contains errors and shortcomings, your feedback made it
a lot better. Thank you so much.

I want to thank Enrique Goites from INTA, who spent time and effort to accompany me and present
me to the farmers, union members and academics who are involved in agroecology. Also, a big thanks
to all of the people who welcomed me at INTA for the two weeks I spent in La Plata. I also want to
thank Claudia, Carlos, Nazarena, Franchu, Simon, Gisela, Pety, Ian and Andy who offered me not only
a house, but a true home in Las Rosas. I love all of you.

I would like to thank all the lovely individuals that I interviewed. Santiago Sarandón, Luis Perez, Juan
Amador, Camila, Andrea, the amazing people from the Activando project who helped me out with the
interview (Emince, Fausto, Agustino, Franco) but also those who could not be there for the interview
(Javi, Dara, Ceci), Antonio Lattuca, Facundo, Diego Fernandez, and Enrique Goites, again. Thank you

I
all for giving me hope and showing me that there are people out there fighting for a better world. I admire
you all and am very thankful to have met each and every one of you.

I would also like to thank Mr. Delbridge and Mr. King, who provided me with their dynamic
programming model, which allowed me to carry out the quantitative part of this thesis and without
whom none of this would have been possible.

I would like to thank my family, Bie, Isha, Laura, Dirk, Ulla and Nolle, who have been very patient with
my academic career (I know at times it has been a struggle), but who have always supported me
whenever I needed it. I also want to thank Patrick, my brother and best friend, for challenging me
intellectually whenever I needed to be. I want to thank Itayosara, a person who inspired me to always
defend the rights of the most vulnerable in society. And I want to thank Lorenzo, who gave me the
experience of a lifetime by going to Uruguay and who reminded me of the importance of reason and
philosophy.

Also, a big thank you to everyone at the Ghent University, the Cordoba University, the Pisa University
and Korea University for accompanying me on this trip and assisting me whenever I needed assisting. I
am so very grateful for the opportunities the IMRD programme has offered me, a programme that has
changed my life. Finally, I want to thank all my classmates and everyone who I met on this incredible
journey but forgot to mention.

To all of you, thank you. This would not have been possible without you.

II
Transitioning to Sustainable Agriculture in
Argentina’s Pampa-region: a Mixed Method
Analysis

JONAS ADRIAENSENS

Department of Food and Resource Economics


Chair of Advisory: Lim, Song Soo

Abstract

Sustainable agriculture is a theme that has been gaining a lot of attention in last years as being a paradigm
on the crossroads of several disciplines like economics, ecology and political science. However, it is
quite unclear what sustainability really means. In this work, both organic and agroecological agriculture
were chosen as embodiments of sustainable agriculture and were examined through a mixed methods
research approach in the Argentinean Pampa-region. First, a dynamic programming model was
performed to determine: 1) whether small-scale farms are more likely to make the transition than
medium-scale farms; 2) to what extent the transition depends on the organic price premium; and 3)
whether there are differences between provinces. Small-scale farms were proven more likely to make
the decision, based on the differences in acreage, and the organic price premium played an important
role, leading to very volatile transition probabilities. There were no noteworthy differences between
provinces. Afterwards, a qualitative study was performed based on interviews conducted in the Pampa-
region through a Marxian Ecology – Ecological Economics framework. These interviews were
performed to 1) better understand agroecology as a paradigm and its relation to the dominant capitalist
framework; and 2) to see what this means for the development of the paradigm. The interviews showed
that transitions to sustainable agriculture depend on social, political and ecological aspects and that the
farmers who decide to make the transition take these aspects into account. Conflicts with conventional
agriculture arise out of soil deterioration, damage to biodiversity, health hazards, limited access to land,

III
poverty issues, among others. The biggest challenge to agroecology as a paradigm is to avoid its
introduction into the global market, its transformation and ultimately co-optation by global capitalism.
Agroecological actors are aware of this, but debate exists whether coexistence with the dominant
framework is possible and whether this co-optation can be avoided. Education and political struggle
seem to be two crucial elements of agroecology in this regard if it wishes to maintain its identity as a
science, movement and practice.

Keywords: agroecology, mixed methods, Marxian ecology, ecological economics, Pampa-region

IV
아르헨티나 팜파(Pampa) 지역에서 지속 가능한 농업으로의 전환:
혼합분석론을 기반으로

요나스 아드리앤센스

초록
지속가능한 농업은 경제학과 생태학 및 정치과학 등 여러 학문이 관련한 패러다임으로서
최근에 주목을 받고 있다. 그러나 지속가능성이 의미하는 바가 아주 명확하지 않은
상태이다. 이 논문은 유기농업과 농생태학적 농업을 지속가능한 농업의 전형으로
선택하여 혼합분석 접근방식을 통해 아르헨티나 팜파(Pampa) 지역 사례를 분석하였다.
첫째, 동적 계획모형(dynamic programming model)을 활용하여, 1) 소농이 중규모 농가에
비해 전환이 용이한지, 2) 유기농 가격 프리미엄의 정도가 전환에 영향을 미치는지, 3)
지역 간 차이가 있는지 등을 분석하였다. 소농은 면적 차이에 기초하여 전환 결정을 하는
경향을 나타냈고 유기농 가격 프리미엄이 중요한 역할을 하여 매우 불안정한 전환
가능성을 초래했다. 지역 간에는 신뢰할 만한 차이가 나타나지 않았다. 이후 정성적인
분석은 막스주의(Marxian) 생태학, 곧 생태 경제학의 틀에서 팜파 지역 농가들을
대상으로 한 인터뷰에 기반을 두었다. 이 인터뷰는, 1) 패러다임으로서 농생태학과,
주류인 자본주의와 관계를 더 잘 이해하고, 2) 이것이 패러다임 개발에 시사하는 점을
살펴보기 위함이다. 인터뷰 결과, 지속가능한 농업으로 전환은 사회적과 정치적 및 생태적
측면에 달려 있고, 실제로 전환을 결정한 농가들이 이러한 측면을 고려하는 것으로
나타났다. 토양 붕괴, 생물다양성 훼손, 건강 위험, 토지에 대한 제한된 접근, 빈곤문제
등에서 기존 농업과 갈등이 발생한다. 패러다임으로서 농생태학에 가장 큰 도전은
세계시장으로 진출, 그와 같은 전환 및 궁극적으로 세계 자본주의에 의한 조작에서
벗어나는 것이다. 농생태학에 참여하는 사람들은 이를 인지하고 있으나, 자본주의의
지배적 틀과 공존이 가능한지, 이러한 조작을 피할 수 있을지 등에 관해 논쟁이 존재한다.

V
만약 과학과 운동 및 실천방식으로서 그 정체성을 유지하길 원한다면, 이와 관련한
농생태학의 결정적인 두가지 요소로 교육과 정치적 갈등을 꼽을 수 있을 것이다.

키워드: 농생태학, 혼합분석 방법론, 막스주의 생태학, 생태 경제학, 팜마 지역

VI
Table of Contents
Acknowledgements............................................................................................................................................I

Abstract.............................................................................................................................................................III

초록 ...................................................................................................................................................................V

Table of Contents........................................................................................................................................... VII

List of Tables.....................................................................................................................................................X

List of Figures ................................................................................................................................................. XI

Abbreviations ................................................................................................................................................. XII

1 Introduction...............................................................................................................................................1

1.1 Introduction to Organic and Agroecological Agriculture..........................................................1

1.1.1 Problems with the Conventional Paradigm............................................................................1

1.1.2 Defining Organic and Agroecological Agriculture ...............................................................2

1.1.3 The Question of Food Sovereignty .........................................................................................5

1.2 Introduction to the Pampa-region .................................................................................................6

1.3 Objectives, Methods and Aims ....................................................................................................7

2 Literature review ....................................................................................................................................10

3 Model ......................................................................................................................................................14

3.1 Dynamic programming approach ..............................................................................................14

3.1.1 Introduction to the Model .......................................................................................................14

3.1.2 Transition Status ......................................................................................................................15

3.1.3 Return Processes......................................................................................................................16

3.1.4 Current Net Returns ................................................................................................................18

3.1.5 Formulation of the Model.......................................................................................................20

VII
4 Results .....................................................................................................................................................22

4.1 Interpretation of the Results ........................................................................................................22

4.2 Limitations of the Study and the Need for a New Perspective ...............................................26

5 Introduction to Qualitative Analysis ....................................................................................................28

5.1 Data Collection .............................................................................................................................28

5.2 Marxian Ecology..........................................................................................................................30

5.2.1 Rationale for Use of a Marxian Ecology Framework .........................................................30

5.2.2 Materialism and the Conception of Nature and Man ..........................................................33

5.2.3 Metabolism as an Ecological and Societal Concept............................................................37

5.2.4 Commodities and the Creation of Value ..............................................................................41

5.3 Ecological Economics .................................................................................................................48

5.3.1 On the Character of Economics.............................................................................................48

5.3.2 The Incommensurability of Values .......................................................................................52

5.3.3 Social Institutions and the Economy .....................................................................................54

5.4 Towards a ME-EE Nexus ...........................................................................................................55

6 Results and Discussion..........................................................................................................................61

6.1 Ecological Analysis......................................................................................................................61

6.1.1 Soil ............................................................................................................................................61

6.1.2 Biodiversity ..............................................................................................................................63

6.2 Economic Analysis ......................................................................................................................64

6.2.1 Value Issue ...............................................................................................................................64

6.2.2 Local Production and Consumption......................................................................................68

6.3 Social Analysis .............................................................................................................................69

VIII
6.4 Political and Conflict Analysis....................................................................................................72

6.4.1 Producer-State Interactions and Perspectives.......................................................................72

6.4.2 Agroecology as a Movement .................................................................................................78

6.5 Capitalism and the Development of Agroecology ...................................................................79

7 Conclusion ..............................................................................................................................................85

8 Reference List.........................................................................................................................................88

9 Appendices .......................................................................................................................................... 102

Appendix A Interview 1 .................................................................................................................. 102

Appendix B Interview 2 .................................................................................................................. 103

IX
List of Tables
Table 3.1 Reversion Parameters .......................................................................................... 17
Table 3.2 Yield Ratio Used per Crop ................................................................................... 18
Table 3.3 Production Costs for Each Cropping System and Farm Size ($/acre) ................. 19
Table 3.4 Interaction between State and Control Variables ................................................ 20
Table 3.5 Used Sources and Their Function in the Model ................................................... 21
Table 5.1 Profiles of Respondents ........................................................................................ 28

X
List of Figures
Figure 1.1 Division of Total Emissions per Sector ....................................................................2
Figure 1.2 The Pampa-region and its provinces .......................................................................7
Figure 2.1 Sustainability Performance of Organic and Conventional Production................ 12
Figure 4.1 Transition probability to organic cultivation for small-scale farmers in the Pampa-
region.................................................................................................................... 23
Figure 4.2 Transition probability to organic cultivation for medium-scale farmers in the
Pampa-region ....................................................................................................... 23
Figure 4.3 Amount of certified organic producers per province (1997 – 2017) ..................... 24
Figure 5.1 A representation of the traditional view on sustainability (a) and the view proposed
by ecological economics (b) .................................................................................. 48
Figure 5.2 The Economic System in Ecological Economics................................................... 51
Figure 6.1 Soil Use in the Pampa-region: annual cultivation (a), natural forests (b) and natural
pastures (c)........................................................................................................... 62
Figure 6.2 Standardized Scale of Comparative Analysis of Sustainability Between
Conventional and Agroecological Agriculture .................................................... 66
Figure 6.3 Comparison La Aurora and Farms in the Area, Average Production (a),
Environmental Indicators (b) .............................................................................. 67

XI
Abbreviations
AFOLU Agriculture, Forestry and Other Land Uses
EE Ecological Economics
FAO Food and Agricultural Organization
INTA National Institute for Agrarian Technology
IPAF Institute for Research and Development of Family Farming
ME Marxian Ecology
MTE Movement of Excluded Workers (Movimiento de Trabajadores Excluidos)
SENASA National Service of Agrifood Health and Quality
SOCLA Latin American Society of Agroecology
UTT Union of Workers of the Land (Unión de Trabajadores de Tierra)

XII
1 Introduction

1.1 Introduction to Organic and Agroecological Agriculture


1.1.1 Problems with the Conventional Paradigm

Agriculture today is facing unprecedented challenges. The emergence of climate change as a threat to
food production systems significantly limits the ability to feed an ever-growing global population,
especially in low-to middle-income countries. The effects of climate change on agriculture are plethora:
rising temperatures, changes in levels and frequency of precipitation, extended droughts, rising sea levels,
the salinization of water and arable land and the increased intensity of extreme weather events are all
consequences of climate change and provide a challenge to agriculture. In their 2016 report, the FAO
stated that in order to feed the global population by 2050, food production would need to increase by
60% compared to 2006 levels.

The issue of climate is not only a challenge to agriculture, but also a partial consequence of conventional
agriculture. The cultivation of extensive monocultures and the excessive application of chemical
fertilizers and pesticides have severely impacted biodiversity, soil quality, public health and other socio-
ecological factors. Agriculture is also responsible for the emission of three of the most dangerous
greenhouse gases: carbon dioxide, methane and nitrous dioxide. According to the FAO (2016),
agriculture, forestry and other land uses (AFOLU) are responsible for 21% of total global emissions.
This is lower than the 27% during the 1990s, but this is mainly due to the sharp increase of emissions in
other sectors. A division of emissions per sector can be seen in Figure 1.1. Note that emissions from
processing industries are not included under the AFOLU-sector. This division also varies strongly
between countries and regions, as the pre- and post-production processes account for more emissions in
high-income economies than in low-to middle-income economies. Energy is by far the largest sector in
emissions (47%), followed by AFOLU (21%); transport (11%); residential, commercial and
institutional (8%); and industrial processes and solvent use (7%). 6% of emissions come from diverse
other sectors.

1
Figure 1.1 Division of Total Emissions per Sector

Industrial
Processes and Others, 6%
Solvent Use, 7%
Residential,
Commercial and
Institutional, 8%

Energy, 47%
Transport, 11%

AFOLU, 21%
Source: own image based on FAO (2016)

Not only the ecological aspect forms cause for concern, there has been strong criticism uttered towards
the social consequences of conventional agriculture and its focus on economic efficiency. Such
consequences include land grabbing, the acquisition of massive agricultural parcels of land inhabited
by indigenous communities by foreign investors; migration of large rural populations; a lack of legal
protection for small-scale farmers; and the health costs that result from conventional agriculture. The
latter two are especially relevant in the case of the Pampa-region and will be elaborated upon further in
the analysis section below.

1.1.2 Defining Organic and Agroecological Agriculture

Organic and agroecological agriculture are a response to these problems and both propose a more
sustainable approach to agriculture. They are often used interchangeably and mistaken for synonyms;
defining these two is thus difficult but it is a crucial exercise to understand the importance of agroecology
as a paradigm, rather than just a cultivation method. Both organic and agroecological agriculture will be
analysed, the former in the quantitative analysis and the latter in the qualitative analysis.

2
There consists a variety of definitions for organic agriculture, as contemporary organic farming is based
on a number of different approaches which have blended over time to produce the current school of
thought (Rigby & Cáceres, 2001), but for our purposes the definition by Lampkin (1994) is utilized,
stating that the aim of organic agriculture is

to create integrated, humane, environmentally and economically sustainable production


systems, which maximise reliance on farm-derived renewable resources and the management
of ecological and biological processes and interactions, so as to provide acceptable levels of
crop, livestock and human nutrition, protection from pests and disease, and an appropriate
return to the human and other resources. (Lampkin & Padel, 1994, p. 4)

Lampkin, an authority on organic agriculture, goes on to state that “sustainability lies at the heart of
organic farming and is one of the major factors determining the acceptability or otherwise of specific
production practices” (Lampkin & Padel, 1994, p. 5). There is no real discussion about the fact that
organic agriculture and sustainable agriculture are closely related, but there is discussion as to which
degree organic agriculture is truly sustainable. This warrants a definition of what sustainable agriculture
is. Here the definition of Ikerd (1993) is chosen, defining sustainable agriculture as “capable of
maintaining its productivity and usefulness to society over the long run. […] it must be environmentally-
sound, resource-conserving, economically viable and socially supportive, commercially competitive,
and environmentally sound” (Ikerd, 1993, p. 30). This definition is broad and does not make a hierarchy
of different substainabilities, nor does it state how to identify sustainability. This is because sustainability
is extremely difficult to identify, partly due to the amount of parties involved in the struggle to being
identified as sustainable. The debate over how to achieve sustainability is thus plagued by fundamental
disputes and disagreements over which elements of production are acceptable and which are not, and
this makes it a very problematic issue (Rigby & Cáceres, 2001). This debate also problematizes the
definition of organic agriculture as sustainable.

There are some aspects to organic agriculture which have been subject to criticism. The fact that organic
agriculture is governed by very specific laws, as “a complete set of certification procedures governs
organic farming, from the soil to the dining table” (MacCormack, 1995, p. 60). This makes it easier to
define organic agriculture, but also severely limits its ability to adapt to regional differences and consider

3
socio-political aspects. As organic agriculture is included in the globalized market mechanisms, they are
(possibly) subject to long-distance consumption and large food value chains. This is where agroecology
differs. Its foundations are not only rooted in ecology, but also in socio-political and structural realities.

Agroecology is often defined as either a science, a movement or a practice. It would however be an error
to separate the meaning from each of these elements, as they influence each other in continuous cycles
of development as the paradigm evolves. The science of agroecology cannot be separated from its
politics and practices. Any attempt at defining agroecology should thus take these three aspects into
account. For the purposes of our research, the definition by Sevilla Guzmán & Woodgate (1997) best
fulfils this condition:

Agroecology promotes the ecological management of biological systems through collective


forms of social action, which redirect the course of coevolution between nature and society in
order to address ‘the crisis of modernity’. This is to be achieved by systemic strategies that
control the development of the forces and relations of production in order selectively to change
modes of human production and consumption that have produced this crisis. Central to such
strategies is the local dimension where we encounter endogenous potential encoded within
knowledge systems (local, peasant or indigenous) that demonstrate and promote both
ecological and cultural diversity. Such diversity should form the starting point of alternative
agricultures and the establishment of dynamic yet sustainable rural societies. (Sevilla Guzmán
& Woodgate, 1997, pp. 93–94)

Agroecology thus encapsulates more than just an ecological production system, it takes into account
social, cultural and political aspects and arose out of the problems of modernization of agriculture and
the resistance against these issues. Agroecology conceives that the current model is not equipped to
address the needs of the population and that a change in paradigm is necessary. Some core elements,
which separate it from organic agriculture, are: 1) political structures based on radical democracy, as
expressed in the horizontality of decision-making processes which are based on consensus; 2) local roots,
as indicated in their focus on local consumption and the embedding of this local consumption in an
agricultural system that is in harmony with the surrounding ecosystems.; and 3) focus on traditional
knowledge and indigenous crops, as conveyed through research in communities around the world and

4
through emphasis on variability in cultivation, depending on the region and/or the knowledge systems
that this cultivation is enclosed in.

After inspecting these definitions, we can thus deduce that organic agriculture can be, but is not
necessarily agroecological. In the view of agroecology, considering the ecological aspects of
sustainability is not sufficient, the social and political aspects of sustainability should also be addressed.
And, of course, the economic aspect of sustainability, which is a necessary condition for these cultivation
methods to sustain themselves. Nonetheless, even in agroecological studies organic agriculture is often
placed on the same line as agroecological agriculture and they do have a lot in common. This makes it
difficult for policymakers, but also academics, to tackle the subject of sustainable agriculture.

1.1.3 The Question of Food Sovereignty

Agroecology is inherently linked to food sovereignty and has been regarded as an instrument to reach
the latter. To understand the emergence of food sovereignty, it is first important to consider food security
as a concept. Food security was defined by the FAO in 2001, and repeated in their 2003 paper, as:

Food security [is] a situation that exists when all people, at all times, have physical, social and
economic access to sufficient, safe and nutritious food that meets their dietary needs and food
preferences for an active and healthy life. (FAO, 2003, p. 28)

Many academics related to agroecology criticized the concept of food security, claiming that power
relations behind global food trade were made invisible through these definitions and that the main
instrument that was considered to resolve this issue was the market. In their view, the definition of food
security entirely avoided discussing the social control over food systems. Or, as Patel states, “the terms
on which food is, or is not, made available by the international community has been taken away from
institutions that might be oriented by concerns of ‘food security’, and given to the market, which is
guided by an altogether different calculus” (Patel, 2009, p. 664). A broadening of the food security
definition to consider social control, public health and nutrition was mainly the result of the introduction
of food sovereignty by La Via Campesina (LVC) in 1996 (Patel, 2009). They defined food sovereignty
as

5
Long-term food security depends on those who produce food and care for the natural
environment. As the stewards of food producing resources we hold the following principles as
the necessary foundation for achieving food security. […] Food is a basic human right. This
right can only be realized in a system where food sovereignty is guaranteed. Food sovereignty
is the right of each nation to maintain and develop its own capacity to produce its basic foods
respecting cultural and productive diversity. We have the right to produce our own food in our
own territory. Food sovereignty is a precondition to genuine food security. (La Via Campesina,
1996, p. 1)

Food sovereignty thus asks questions about food security and its neglect of power relations and social
structures that are in place and govern food systems. The concept of food sovereignty emphasized that
this aspect of food security needed to feature in the discussion and that there needed to be a discussion
about how to achieve food security, for which food sovereignty is a necessary requirement. It opened
up a discussion about rights and democracy and their relation to food security. This concept is crucial to
agroecology and helps explain the social and political perspectives that lie at the heart of it.

1.2 Introduction to the Pampa-region


The Pampa-region was chosen as a case for agroecology for several reasons. It has a temperate, humid
climate and consists of five provinces: Buenos Aires (1), La Pampa (2), Córdoba (3), Santa Fe (4), and
Entre Ríos (5). They are numbered in Figure 1.2 below, with the Pampa-region highlighted in light green.

The Pampa-region is the most important economic region of Argentina. The most important sector is
agriculture and the main agricultural activities consist of cattle breeding and soy, wheat and maize
cultivation. Because Argentina’s economy is not very diversified, its economy largely depends on
returns from agriculture, for which the Pampa-region is the most important region as it accounts for 74.1%
of exports. Buenos Aires (32.1%), Santa Fe (23.9%) and Córdoba (14.6%) are the most important
provinces in this region (National Treasury of Argentina, 2017).

Because of the conflict between large-scale landowners and small-scale family farmers, this region
becomes important to the topic of this study. Due to soy monocultures in the region, family farmers run
the risk of being marginalized and access to land is restricted. The agroecological movement is quite

6
active here, as the first university programme in agroecology in Argentina was founded in the National
University of La Plata.

Figure 1.2 The Pampa-region and its provinces

5
4
3

2
1

Source: Cabrera (1976)

1.3 Objectives, Methods and Aims


The methods used in this paper are both quantitative and qualitative. First, quantitative data based on
reports from the National Ministry of Agriculture, the National Service of Agrifood Health and Quality
(SENASA) and the FAO were used to run a dynamic programming model and analysed to better
understand the economic conditions for transition from conventional agriculture to organic agriculture.
Afterwards, qualitative data was gathered from interviews with farmers, academics and actors involved
in the agroecological movement in the Pampa-region in Argentina.

The method this study makes use of is commonly referred to as an Explanatory Sequential Mixed
Methods Design, which is used when a researcher aims to follow up quantitative results with qualitative

7
data. In this framework, the quantitative part of the study might be more, less or equally important as the
qualitative part (Edmonds & Kennedy, 2017). This study assigns equal importance to both the
quantitative and qualitative data used, as the analytical value of both data is considered to be crucial to
better understanding the research subject.

After running the dynamic programming model to answer the different hypotheses (1), this study takes
a closer look to the development of sustainable agriculture and attempts to answer the qualitative
research question (2). Both the hypotheses and he research question form the basis of the paper and are
formulated as follows:

1) H01 = small-scale farms are as inclined towards an organic transition as medium-sized farms
H02 = the transition decision depends strongly on the applied organic price premium
H03 = there are no substantial differences between provinces

After this, the research question is specified:

2) How can sustainable agriculture develop given current socio-economic and political
constraints?

The issue which this research question suggests is twofold: first, it demands a specification of the socio-
economic and political constraints that exist. For this, an ecological economics – Marxian ecology
framework will be used to analyse the context in which sustainable agriculture is situated. Afterwards,
some projections for the agroecological paradigm are given, based on qualitative data gathered in
Argentina. The objective is thus to get more insight into the issue of sustainable agriculture by using
both quantitative and qualitative methods.

The aim of this research is twofold as well; to at one hand better understand the issue of transitioning to
sustainable agriculture, and on the other to contribute to the academic field. The goal is here not only to
come up with the right answers, but also to raise the right questions. In order to better understand the
topic and taking into account the interdisciplinary nature of sustainable agriculture, an interdisciplinary
approach using mixed methods is warranted.

8
The decision to use an interdisciplinary approach comes from the conviction that the ontological and
epistemological reality of the economy is embedded within a social structure, which influences decision-
making processes and thus the economy as a whole. This approach follows the view of ecological
economics, which states that to analyse the economic system by itself without taking into account the
social and ecological spheres that it is embedded in, is misguided and leads to wrongful conclusions
about the economy. Or, as Splash (2017) states it:

They [neoliberal scholars] make recommendations for society on the basis of a discipline that
has no theory of society, nor indeed any conception of social structure, but rather merely
regards society as an aggregation of individual agents, each pursuing their own self-interest.
(Splash, 2017, p. 5)

First, a quantitative analysis gives more insight in the specific decision-making processes of individual
farmers to make the transition to organic agriculture, after which a qualitative analysis is conducted to
analyse the specific social and ecological structures that they make this decision in.

9
2 Literature review
In the following literature review, a short overview will be given of academic advances in agroecology,
organic agriculture, Marxian ecology and ecological economics.

Research concerning the economic and political viability of agroecology is limited. A recent study by
Bellamy & Ioris (2017) examined the potential of agroecology to transform the production food system
and the knowledge gaps that currently exist in the field.

The first to introduce agroecology into the academic sphere was Stephen Gliessman. He established the
first university agroecology programme at the University of California, Santa Cruz in 1981. He saw
agroecology as a food production method that was more sustainable than the industrial paradigm. The
focus here was ecological, which is where agroecology as a science started. The political and economic
notions of agroecology, among others, were absent in academia at that time.

Agroecology found its academic roots in ecology but has transformed as a discipline in the last decennia.
The conflict that many initiatives experienced with agro-industry became the basis for a more political
focus on agroecology in the scientific literature and for the development of a new discipline called
political agroecology, which Bellamy & Ioris (2017) define as follows:

[…] [political agroecology] draws attention to broader food production and consumption
systems, and which foregrounds power and politics. Latin American agroecology, for instance,
has its roots in social movements explicitly aimed at agrarian empowerment, which emerged
as a response to economic exclusion produced by agricultural modernization. (Bellamy &
Ioris, 2017, p. 3)

This political agroecology can be traced back to rural sociology. This academic field looked at the
peasantry as a sociological group in society and examined its struggles against capitalist modernity and
agricultural industrialization. This trend, partially contributed to the ‘crises’ of IMF Structural
Adjustment Programmes and pressure to introduce a free market system in developing countries,
created a political movement which political agroecology is now a part of. Recent works like
Agroecology: Foundations in Agrarian Social Thought and Sociological Theory (Sevilla Guzmán &

10
Woodgate, 2013) and Agroecology and Politics. How to Get Sustainability? About the Necessity for a
Political Agroecology (Gonzalez de Molina, 2013) document this evolution and have been a guideline
for research into this field, also due to their experience in the field of rural sociology. Also, a recent study
concerning the science and politics behind agroecology was written together with Altieri (2017). Of
course, rural sociology has long included these political aspects into its analyses, which the works of the
authors mentioned above are clearly proof of, but experts on agroecology are increasingly introducing
these sociological and political dimensions in their works, indicating the growing awareness of the
importance to construe and conceptualize agroecology in its social context (Bartolomé, Casado, de
Molina, & Guzmán, 2001).

As the political notion of agroecology is growing, research into the economics of agroecology is rather
limited. Many authors have published about the need to include a broader economic context when
analyzing agroecology and the questions that still exist surrounding its viability (Bellamy & Ioris, 2017;
Enríquez, 2013; Herrera, Domené-Painenao, & Cruces, 2017; Paz, 2011; Rosset & Altieri, 2017).
Indeed, some critiques on agroecology like The debate on food sovereignty theory: agrarian capitalism,
dispossession and agroecology (Jansen, 2015) and Potatoes Made of Oil: Eugene and Howard Odum
and the Origins and Limits of American Agroecology (Madison, 1997) have created doubts concerning
the productivity possibilities and consequential economic viability of agroecological systems, naming
the absence of modern science as a big problem for agroecology today. Notwithstanding these critiques,
other authors have documented promising results from implementing agroecological innovations as
studies like Agroecology and the design of climate change-resilient farming systems (Altieri, Nicholls,
Henao, & Lana, 2015) and ‘Sustainable Intensification’ (Williams, Toulmin, & Pretty, 2011) are an
example of. Discussion still exists in the academic community and this is an important point of conflict
concerning agroecology.

Authors like Altieri, Rosset, Sarandón, Sevilla Guzmán and Tittonell have defined agroecology as it is
today, each from a different perspective. Their works have proven instrumental in the development of
agroecology not just as a production method, but as a paradigm. Still, it has a long way to go and many
questions remain unanswered; the discussions within the agroecological paradigm in the coming years
will determine its future.

11
Organic agriculture has had a similar trajectory as agroecology, which is not surprising as they share
some of the same core values. Although being strongly rejected in the 1970s and 1980s by many
scientists and farmers, organic production has sharply increased in the last decades and has become
mainstream. Nonetheless, the discipline still faces a lot of criticism from observers who doubt in its
potential to compete with or replace the current conventional model (Kirchmann & Torvaldsson, 2000;
Pickett, 2013; Trewavas, 2001). While authors like De Ponti (2012), Badgley (2007), Pas (2014) and
Seufert (2012) have shown that organic yields underperform when compared to conventional yields,
cases have been made for organic agriculture, proving that it has been profitable, as shown by an
increasing number of farmers making the transition (Kumar, Saini, & Chopra, 2018; Lotter, 2003;
Reganold & Wachter, 2016). Indeed, in their meta-analysis concerning sustainability of organic and
conventional production systems published in Nature, Reganold & Wachter (2016) make a case for
organic agriculture, as displayed in Figure 2.1 – with the length of the flower petals determining the level
of performance of specific sustainability metrics:

Figure 2.1 Sustainability Performance of Organic and Conventional Production

Source: Reganold & Wachter (2016)

In this study, the school of Marxian ecology was combined with ecological economics. Marxian theory
was analysed for theory on ecology through the different works of Marx and Engels. Although Engels
is more often regarded as the ecological thinker, both were deeply involved in the study of natural
sciences and were aware of the latest developments. Marx’s Grundrisse and The German Ideology are
both works where this ecological element first appears. Engels’ unfinished study Dialectics of Nature

12
was instrumental in establishing materialism firmly within the theory of change in society. Two authors
first came to this insight at the late 90s and are responsible for the rediscovery of ecology in Marxism:
John Bellamy Foster with his work Marx’s Ecology: Materialism and Nature (2000), and Paul Burkett
with his work: Marx and Nature: a Red and Green Perspective (2014). Many of the insights of this
study were based on their findings.

Ecological economics as a field has many scholars, of whom Martínez-Alier belongs to the most
prominent. His influential book Handbook of Ecological Economics (2015) and articles such as Social
Metabolism and Environmental Conflicts (2007) have placed him firmly at the heart of scholarship on
ecological economics. Other notable scholars include Splash, editor of the Routledge Handbook on
Ecological Economics, Honkasalo, and Howarth & Baumgärtner, editors of the journal Ecological
Economics. Many of these authors have experience in research into agroecology as well.

13
3 Model

3.1 Dynamic programming approach


3.1.1 Introduction to the Model

This model was created by Delbridge & King (2016), for more information concerning the model,
please refer to their paper1. This section is based on their explanation of the model. There are a lot of
similarities between this section and their 2016 paper, all credit thus goes to these authors.

This study made use of a dynamic programming approach, considering a model in which a risk-neutral
farm manager faces two options: either adopting a conventional approach with a maize-soybean crop
rotation, or an organic approach with a maize-soybean-wheat-sorghum crop rotation. In case the organic
cultivation option is chosen, two distinct stages are possible: a transitional stage or a certified organic
stage. The choice for organic agriculture implies a transitional period of two years, where it is assumed
that yields and productions costs are organic, but prices are conventional (there are no organic price
premiums). After this period of two years, the farmer certifies his organic production and receives the
organic price premium.

The model contains a single control variable, xt, which refers to the production decision of either
switching to organic production methods (assuming the value 1) or to maintain conventional production
methods (assuming the value 0). In our model we consider that all farmers start from conventional
production methods and take into account only economic factors. There are two state variables in this
model: st and crt. The transition status variable st indicates whether a farm is certified organic, in
transition, or neither. The current per acre revenue, crt, is modeled as a single stochastic process, rather
than as a joint distribution of separate crop yields and prices. Due to a lack of consistent and long-term
data concerning organic production, the estimation of similar stochastic processes for organic and
conventional cropping systems are impossible and the organic and transitional revenues are modeled as
functions of conventional revenues. This leads to the inherent impossibility to derive critical conversion
or yield points from the model. Although cases have been made in the academic literature that organic

1
The model in question is used with the permission of the authors.

14
and conventional cropping systems may be independent of one another, they are inextricably linked
through external impulses such as weather events and short-or long-term climate change.

The decision problem is modelled using data from the FAO and the Argentinean Ministry of Agriculture.
The organic cropping system consists of a four-year rotation (wheat-corn-soybean-sorghum), the
conventional system consists of a two-year rotation (corn-soybean). These rotations were chosen based
on the practices and most prevalent crops in the Pampa-region.

Below, different elements of the model are explained and the reasoning behind them is clarified.

3.1.2 Transition Status

In this model, two farm-size scenarios are considered: a small-size farm, reflecting the average size of a
family agriculture farm in Argentina, and a medium-sized farm, reflecting the median size of certified
organic farms (for organic cultivation) and the average size of agribusiness farms (for conventional
cultivation) in Santa Fe, a province in the Pampa-region. This province was chosen for consistency, as
the production costs used in this study stem from a comparative study conducted in Santa Fe. The small-
sized farms are determined at 264 acres and the medium-sized farms at 744 acres (for organic cultivation)
and 996 acres (for conventional cultivation). These surface sizes are based on reports and datasets from
both the FAO (Salcedo & Guzmán, 2014) and the National Service of Agrifood Health and Quality
(SENASA, 2018). These farm sizes are used for the other Pampa-regions as well, as it is assumed that
the different provinces share an overall socio-economic character in agriculture and that applying the
Santa Fe farm sizes on the other four provinces does thus have no distorting effect on the results.

These farm sizes also matter because of the transition possibilities and their dependence on acreage. It
is considered that organic agriculture tends to smaller surfaces than conventional agriculture. For a
transition of conventional to organic, the saturated land market in the region makes it easy for a farmer
to reduce the amount of land managed by renting it out; a transition of organic to conventional however,
requires the acquisition of more land, which in the tight land market causes farmers to gradually increase
their acreage. This is also included in the model.

Three distinct transition situations are possible for organic farmers: first year transition (𝑠𝑡 ≤
4 𝑎𝑛𝑑 𝑥 = 1) , second-year transition (𝑠𝑡 = 5 𝑎𝑛𝑑 𝑥 = 1) , and certified organic (𝑠𝑡 =

15
6 𝑎𝑛𝑑 𝑥 = 1) . Five distinct transition situations are possible for conventional farmers: full
conventional acreage (𝑠𝑡 = 4 𝑎𝑛𝑑 𝑥 = 0) and the four levels of transition between organic and
conventional acreage (𝑠𝑡 > 4 𝑎𝑛𝑑 𝑥 = 0, 𝑎𝑛𝑑 𝑠𝑡 = [1,2,3] 𝑤𝑖𝑡ℎ 𝑥 = 0). The transition status
state variable and its dynamics are described as follows, following the notation by Delbridge & King
(2016):

1 𝑖𝑓 𝑠𝑡 > 4 𝑎𝑛𝑑 𝑥𝑡 =0
𝑚𝑖𝑛(𝑠𝑡 + 1,4) 𝑖𝑓 𝑠𝑡 < 5 𝑎𝑛𝑑 𝑥𝑡 =0
𝑠𝑡+1 ={ (3. 1)
5 𝑖𝑓 𝑠𝑡 < 5 𝑎𝑛𝑑 𝑥𝑡 =1
𝑚𝑖𝑛(𝑠𝑡 + 1,6) 𝑖𝑓 𝑠𝑡 > 4 𝑎𝑛𝑑 𝑥𝑡 =1

3.1.3 Return Processes

Revenues produced by conventional production are assumed to follow a mean-reverting Ornstein-


Uhlenbeck process

𝑑𝑐𝑟 = 𝜂(𝑐𝑟
̅ − 𝑐𝑟)𝑑𝑡 + 𝜎𝑑𝑧 (3. 2)

where 𝜂 is the reversion parameter, 𝑐𝑟


̅ and 𝑐𝑟 are the long-run mean revenues per acre for a
conventional cropping system with corn-soybean rotation; 𝜎 is a variance parameter and 𝑑𝑧 is an
increment of the Wiener process. The discrete approximation can be noted as follows:

𝑐𝑟𝑡 − 𝑐𝑟𝑡−1 = 𝜂𝑐𝑟


̅ − 𝜂𝑐𝑟𝑡−1 + 𝜎𝜀
= 𝛼0 + 𝛼1 𝑐𝑟𝑡−1 + 𝜎𝜀 (3. 3)

̂
𝛼
̂ = − 0 and 𝜂̂ = −𝛼̂1; the standard error for the regression 𝜀~𝑁(0,1) is the estimate of the
where 𝑐𝑟
̅ ̂
𝛼 1

variance parameter 𝜎.

Due to lack of numbers on historical crop yield and crop price data from the Pampa region, the estimates
based on Delbridge & King (2016) are adopted into this model. Although the estimates in their case
were calculated based on a seventy-one year series of obsessed gross revenues from the US Department
of Agriculture, it is assumed that the reversion parameters are suitable for the Argentinean case, as an
application of Argentinean yields between 1961 – 2014 rendered similar results. For reasons of
consistency they are not used however, as the absence of prices on Argentinean crops might distort the

16
results. The estimates of the parameters are observed in Table 1 and are an exact copy of the parameters
used in Delbridge & King (2016).

Table 3.1 Reversion Parameters

Parameter Estimate Standard Error P-value


𝛼̂0 85.02 54.23 0.122
𝛼̂1 -0.095 0.049 0.056
Long-run mean ̂
𝑐𝑟
̅ 899.33
Reversion rate 𝜂̂ 0.095
Variance 𝜎̂ 201.99
Source: Delbridge & King (2016)

Calculating return processes for the Argentinean Pampa-region based on reversion parameters from the
Midwest in the US might seem distorting at first, but becomes logical after analysing the similarities and
differences between both areas. The Pampa-region and the Midwest are both characterized by plain
landscapes, rich soils, similar climates and a high degree of intensification. In both areas the cultivation
of cereals (wheat, soy and corn) and the raising of cattle dominates. Differences between both regions
include the growing seasons and the destination for production, as Midwestern cereals and cattle are
mainly aimed at the domestic market, while most of production in the Pampa-region is exported. Still,
it is here argued that both regions share enough similarities to assume identical return processes.

Revenues from organic and transitional states are modeled as linear functions of the conventional
revenues:

𝐺𝑅𝑖 = 𝛽0𝑖 + 𝛽1𝑖 𝑐𝑟 + 𝜀𝑖 𝑓𝑜𝑟 𝑖 = 𝑜, 𝑟 (3. 4)

where 𝐺𝑅𝑖 is the per acre revenue for organic and transitional cropping methods; 𝑐𝑟 is the per acre
revenue for conventional cropping methods; 𝜀𝑖 is the error term for revenue distribution 𝑖; and the
organic and transitional cropping methods are denoted by 𝑖 = 𝑜 and 𝑖 = 𝑟 , respectively.

Estimations of the parameters of equation (3.4) are based on detrended and inflation-corrected data from
the Argentinean Ministry of Labour and Production and the FAO. Due to a lack of trustworthy data on

17
organic yields and prices, yield ratios were extracted from a meta-analysis conducted by Ponisio et al.
(2014), who conducted an analysis into the so-called yield gap that exists between organic and
conventional agriculture based on academic literature. According to several authors, including
Delbridge & King (2016), organic yields have the potential of exceeding conventional yields – although
this strongly depends on the production conditions and most academic literature suggests otherwise.
Sorghum has the highest comparative yields, which complies with recent studies on organic yields
(Amujoyegbe, Opabode, & Olayinka, 2007; Serme et al., 2018; Shuaibu, Bala, Kawure, & Shuaibu,
2018). The used ratios for each crop in our study is given in Table 3.2.

Table 3.2 Yield Ratio Used per Crop

Crop Organic:conventional yield ratio


Corn 0.8822
Sorghum 0.9920
Soybean 0.8318
Wheat 0.8295
Source: Ponisio et al. (2014)

The yield sample was limited to 20 years (1993 – 2012) and the price data to seven years (2006 – 2012),
creating a total of 140 revenue states for each province and system (20 × 7 = 140). The model
consists of five different organic premiums: 0% (conventional price), 35%, 40%, 45% and 50%. This is
based on the assumption made by several reports concerning organic price premiums in Argentina that
price premiums are often somewhere between 30% - 50% (FAO, 2001; Rodríguez, Gentile, Lupín, &
Garrido, 2002) and approximates the premiums used in Delbridge & King (2016).

3.1.4 Current Net Returns

The farm manager’s return in a single period is formulated as

𝑓(𝑥𝑡 , 𝑠𝑡 , 𝑐𝑟𝑡 ) = (𝐺𝑅(𝑥𝑡 , 𝑠𝑡 , 𝑐𝑟𝑡 ) − 𝑐(𝑥𝑡 ))𝑎(𝑥𝑡 , 𝑠𝑡 ) (3. 5)

where 𝐺𝑅(𝑥𝑡 , 𝑠𝑡 , 𝑐𝑟𝑡 ) is the revenue in a single period, depending on the management decision (𝑥𝑡 ),
the organic transition status (𝑠𝑡 ) and the current revenue generated from conventional production

18
methods (𝑐𝑟𝑡 ); 𝑐(𝑥𝑡 ) refer to the production costs and 𝑎(𝑥𝑡 , 𝑠𝑡 ) is the acreage available to the
manager.

Production costs are made up of variable operating costs and fixed overhead costs. This data was taken
from a comparative study in 1997 between organic and conventional production costs in the province
of Santa Fe (National Ministry of Agriculture and Rural Development, 2004). These costs are assumed
constant over the 20-year period that is analysed and over the different provinces. Table 3.3 gives an
overview of the different costs as used in this study. A representation of the different interactions
between state (𝑠𝑡 ) and control (𝑥𝑡 ) variables is given in Table 3.4. Acreage, revenue and production
costs all depend on the management decision and the organic transition status (Delbridge & King, 2016).

Table 3.3 Production Costs for Each Cropping System and Farm Size ($/acre)

Conventional Organic
Small Medium Small Medium
Crop acreage 264 996 264 744

Variable costs 111.02 106.17 119.67 113.49


Fixed costs 275.49 275.49 280.97 280.97
Total costs 386.51 381.66 400.64 394.46

19
Table 3.4 Interaction between State and Control Variables

s x Acreage Revenue Production costs


1 0 𝑎0 + 0.4(𝑎𝑐 − 𝑎𝑜 ) 𝐺𝑅𝑐 conventional
2 0 𝑎0 + 0.6(𝑎𝑐 − 𝑎𝑜 ) 𝐺𝑅𝑐 conventional
3 0 𝑎0 + 0.8(𝑎𝑐 − 𝑎𝑜 ) 𝐺𝑅𝑐 conventional
4 0 𝑎𝑐 𝐺𝑅𝑐 conventional
5 0 𝑎0 + 0.2(𝑎𝑐 − 𝑎𝑜 ) 𝐺𝑅𝑐 conventional
6 0 𝑎0 + 0.2(𝑎𝑐 − 𝑎𝑜 ) 𝐺𝑅𝑐 conventional

1 1 𝑎𝑜 𝐺𝑅𝑟 organic
2 1 𝑎𝑜 𝐺𝑅𝑟 organic
3 1 𝑎𝑜 𝐺𝑅𝑟 organic
4 1 𝑎𝑜 𝐺𝑅𝑟 organic
5 1 𝑎𝑜 𝐺𝑅𝑟 organic
6 1 𝑎𝑜 𝐺𝑅0 organic
Source: Delbridge & King (2016)

3.1.5 Formulation of the Model

The model can be written in the form of the Bellman equation as

𝑉(𝑠, 𝑐𝑟) = max 𝑓(𝑥𝑡 , 𝑠𝑡 , 𝑐𝑟𝑡 ) + 𝛿𝐸𝑉(𝑔1 (𝑥𝑡 , 𝑠𝑡 ), 𝑔2 (𝑐𝑟𝑡 ))


𝑥𝑡 ∈{0,1}

subject to

1 𝑖𝑓 𝑠𝑡 > 4 𝑎𝑛𝑑 𝑥𝑡 =0
𝑚𝑖𝑛(𝑠𝑡 + 1,4) 𝑖𝑓 𝑠𝑡 < 5 𝑎𝑛𝑑 𝑥𝑡 =0
𝑠𝑡+1 = 𝑔1 (𝑥𝑡 , 𝑠𝑡 ) = {
5 𝑖𝑓 𝑠𝑡 < 5 𝑎𝑛𝑑 𝑥𝑡 =1
𝑚𝑖𝑛(𝑠𝑡 + 1,6) 𝑖𝑓 𝑠𝑡 > 4 𝑎𝑛𝑑 𝑥𝑡 =1

𝑐𝑟𝑡+1 = 𝑔2 (𝑐𝑟𝑡 ) = 𝛼0 + (1 + 𝛼1 )𝑐𝑟𝑡 + 𝜀𝑡

𝐺𝑅𝑖𝑡 = ℎ𝑖 (𝑐𝑟𝑡 ) = 𝛽0𝑖 + 𝛽1𝑖 𝑐𝑟𝑡 + 𝜀𝑖𝑡 𝑓𝑜𝑟 𝑖 = 𝑟, 𝑜 (3. 6)

20
where 𝑉(𝑠, 𝑐𝑟) is the present value of a farm given the values of state variables 𝑠 and 𝑐𝑟; 𝛿 is a discount
factor that is set at 0.04 throughout the analysis and the dynamics of the transition status state variable
𝑠𝑡+1 are the ones given in equation (3.1). This equation was solved using the dpsolve function from the
COMPECON Toolbox developed for MATLAB by Miranda & Fackler (2004). The dpsolve routine
solves discrete time, continuous state decision models by approximating the value function 𝑉(𝑠, 𝑐𝑟)
from equation (3.6), using collocation methods. Afterwards, the probability of a farm manager making
the transition to organic cultivation methods is determined using Monte Carlo simulations. The
probability of a farm ending with organic certification is calculated over a simulated period of 30 years
(Delbridge & King, 2016). Below, in Table 3.5, the different sources are summarized for clarification
purposes.

Table 3.5 Used Sources and Their Function in the Model

Data Function Source Reference


Yields and Prices FAO and Ministry of Agriculture FAO (n.d.); Ministero de Producción y
Trabajo (n.d.)
Farm Acreage FAO and SENASA Salcedo & Guzmán (2014); SENASA
(2018)
Return Processes Delbridge & King Delbridge & King (2016)
Yield Ratio Ponisio et al. Ponisio et al. (2014)
Production Costs Ministry of Agriculture National Ministry of Agriculture and
Rural Development (2004)

21
4 Results

4.1 Interpretation of the Results


The results of the first simulation are displayed in Table 4.1 and Table 4.2. The results are consistent
with the result of Delbridge & King. There are several phenomena to be noticed: small-scale farmers
are more willing to switch over to organic farming, which partially has to do with the possibility of
farming on a greater number of acres in conventional agriculture in the medium-sized farm scenario.
For hypothetical purposes, the experiment was repeated with a 70% price premium for medium-sized
farms, because of the absence of meaningful results with the commonly applied price premiums. The
70% price premium is very high, however, and most studies report cereal price premiums to be lower
(European Commission, 2019; Islam, 2013); it is thus reasonable to assume that in this scenario hardly
any medium-sized farmholder would make the transition under these parameters.

There seem to be little differences between the different provinces, with Santa Fe and La Pampa
showing the highest transition decision probabilities. Another remarkable result is the apparent volatility
that is caused by shifts in the price premiums. An increase or decrease of 5% in price premiums can
have drastic effects on the decision of a farmer to make the transition. This volatility is reflected in data
provided by SENASA, showing the amount of certified organic producers over a period of 20 years
(1997-2017), as shown in Figure 4.1.

22
Figure 4.1 Transition probability to organic cultivation for small-scale farmers in the
Pampa-region

94.15%

100%
100%
100%

100%
100%

100%

100%
100%

100%
53.92%

44.20%
0%
0%
0%

0%
0%
0%

0%
0%
0%

0%
0%

0%
0%
Buenos Aires Cordoba Entre Rios La Pampa Santa Fe
50% 100% 100% 100% 100% 100%
45% 100% 100% 94.15% 100% 100%
40% 0% 0% 0% 53.92% 44.20%
35% 0% 0% 0% 0% 0%
0% 0% 0% 0% 0% 0%
50% 45% 40% 35% 0%

Figure 4.2 Transition probability to organic cultivation for medium-scale farmers in the
Pampa-region
100%

100%
93.71%

94.22%

85.08%
0%

0%

0%

0%
0%
0%
0%
0%

0%
0%
0%

0%

0%
0%
0%
0%

0%
0%
0%

0%

0%
0%
0%
0%
0%

Buenos Aires Cordoba Entre Rios La Pampa Santa Fe


50% 0% 0% 0% 0% 0%
45% 0% 0% 0% 0% 0%
40% 0% 0% 0% 0% 0%
30% 0% 0% 0% 0% 0%
0% 0% 0% 0% 0% 0%
70% 93.71% 94.22% 85.08% 100% 100%

50% 45% 40% 30% 0% 70%

23
La Pampa and Santa Fe seem to be the provinces where farmers are most inclined to make the transition
to organic agriculture, both in the small-scale scenario as in the medium-scale scenario. However, the
overall differences between the provinces seem neglectable, only showing minimal divergences over
the different price premiums. As mentioned above, Figure 4.1 shows the amount of certified organic
producers per province in the Pampa region. Aside from Buenos Aires, which is one of the Argentinean
provinces with the highest amount of certified organic producers over time – along with Rio Negro and
Misiones – the provinces do seem to show a similar trend in the transition to organic agriculture, with
La Pampa consistently ending up at the bottom and Cordoba and Entre Rios manifesting the highest
amount of organic farmers of the four provinces. The volatility in this time series coincides with the
results from the dynamic programming model. The difference in the amount of organic producers over
time is substantial, with some years demonstrating big drops in organic certification and other years
showing significant increases. There are plenty of causes for this volatility, of which the unpredictability
in organic price premiums is one. Other factors like changes in applied policies and environmental
impacts undoubtedly play a role too.

Figure 4.3 Amount of certified organic producers per province (1997 – 2017)
250

200

150

100

50

0
2012

2015
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011

2013
2014

2016
2017

Buenos Aires Cordoba Entre Rios La Pampa Santa Fe

Source: own image based on data from SENASA (2018)

Taking into account the results outlined above, two out of three null hypotheses are adopted and one is
rejected. These results coincide with the results in Delbridge & King (2016) and thus build further on
the main consensuses in academic literature.

24
H01 = small-scale farms are as inclined towards organic transition as medium-sized farms

This null hypothesis is to be rejected based on the clear difference in transition probability in each of the
provinces between small-sized and medium-sized farms. A big reason for this is the difference in
available acreage and the higher (fixed and variable) costs for organic cultivation. Intuitively, it seems
straightforward that a transition on a small-scale farm requires less effort than a transition on a medium-
sized farm due to differences in investments, possible loss of a part of the consumer base due to lower
yields and the higher financial losses during conversion. Based on this information, the null hypothesis
cannot be adopted and thus has to be rejected.

H02 = the transition decision depends strongly on the applied organic price premium

This might well be the most obvious result of the dynamic programming model. The transition
probabilities drop drastically with slight decreases in price premiums, reflecting the risk that is associated
with adopting organic cultivation methods. This model is based on risk-neutral decision-making units,
so the result likely depends on the nature of farmers making the decision, but in this scenario the null
hypothesis is adopted: the transition decision depends strongly on the applied price premium. In 0% –
5% of all cases, the farmers found themselves in the transition stage at the end of the 30 – year period,
which does give some room for interpretation of these transition probabilities and thus contain a certain
error margin.

H03 = there are no substantial differences between provinces

Any attempt at adopting or rejecting this hypothesis is based on interpretation of the results. There are
differences between provinces, with La Pampa and Santa Fe showing the highest transition probabilities
of the Pampean provinces. However, to determine whether these differences are ‘substantial’, an
intuitive decision has to be made. Aside from the fact that La Pampa and Santa Fe are the only provinces
where small-scale farmers would consider a transition at a 40% price premium, the other results are
quite similar, if not identical. That is why this null hypothesis too, is adopted, as the differences between
provinces are not big enough to be considered ‘substantial’.

25
4.2 Limitations of the Study and the Need for a New Perspective
When interpreting these results, it is important to take into account the assumptions that this model is
based on and the shortcomings it has to explain the complexity that characterizes individual decision-
making processes in making the transition to sustainable agriculture. These shortcomings are lined out
in this section.

The first shortcoming of this study is the assumption of perfect comparability between organic and
conventional agriculture. The assumption of a four-year crop rotation model for organic agriculture,
using cereals as the cultivated crops, does not align completely with reality: not only cereals, but also
vegetables and fruits are often cultivated in organic agriculture and more than four crops might be used.
The difference between organic and conventional producers not only entails applied methods, but also
produced crops. While conventional producers are mainly focused on cultivating cereals and raising
cattle, organic producers are focused on highly diversified production systems.

Not only the crop difference might complicate a transition, the inherent difference in culture that
characterizes both methods influences the decision to make this transition. Conventional agriculture in
the Pampa-region is mostly aimed towards intensified, large-scale operations, making use of highly
developed machinery and thus requiring high investments. Organic agriculture is mostly characterized
by small-scale farming and often includes an element of collective action, as many farmers sell their
produce together. Often, the profitability of the organic production method is thus tightly linked to the
social fabric that organic farmers form part of. Although many of the certified farmers do export their
produce, questions should be asked about the certification process and the amount of organic farmers
who engage in this production method without certification (exceeding the amount of certified farmers).
This should thus be taken into account when analysing the transition from conventional to organic
agriculture.

Another shortcoming concerns the assumption that the decision-making process is guided only by
economic drivers. This model analyses how a farmer’s decision is influenced by economic stimuli, but
fails to take into account the social, ecological and political influences that determine this decision. This
is the reason why a qualitative assessment is necessary and this assessment will thus focus not only on
the economic, but also on the social, ecological and political drivers to better understand the motivations

26
of farmers engaged in sustainable agriculture on one hand; and to get insight into the development of
sustainable agriculture on the other.

The last shortcoming is one that results from the different sources that were used to calculate the
profitability. The lack of data on organic agriculture caused different datasets to be used to calculate the
model. This means that the results of this model should thus be taken with a grain of salt. Until more
data concerning sustainable agriculture is available, every assessment analysing it is inherently biased.

These shortcomings thus induce the need for a deeper look into the transition decision. In what follows,
a qualitative method will be used to analyse the social and ecological context that Pampean agriculture
is situated in and to give new insights into the decision to make the transition on one hand, and the
development of sustainable agriculture as a whole on the other. Here, agroecology is used, mainly
because of its focus on social and ecological aspects and its political foundation as a social movement.
Although sustainable agriculture is difficult to define, one could argue that it consists of more than only
ecological sustainability. Agroecology argues that any production system is inherently embedded in a
economic, social and ecological context and that for a cultivation method to be sustainable, it should
take these different contexts into account (see above). A nexus of Marxian ecology and ecological
economics was chosen to analyse the agroecological paradigm in the Pampean region and to identify
the challenges it faces, the struggles it goes through and the developments that ensue from these struggles.

27
5 Introduction to Qualitative Analysis

5.1 Data Collection


Data was collected through semi-structured in-depth interviews and one focus group. In what follows,
the profile of the respondents is provided, together with a rationale why these people were selected. Most
contacts with farmers was made through the National Institute for Agrarian Technology (INTA), who
helped with the preparation and execution of the interviews. It should be clear that although these
respondents have differing views, they all identify as agroecological actors, and that despite their
differences they do agree on the main aspects of agroecology. They all thus represent the agroecological
paradigm, be it from different perspectives.

Table 5.1 Profiles of Respondents

N° Name Function Type of interview


1 Santiago Sarandón President of SOCLA In-depth interview
2 Luis Perez Farmer of Association 1610 In-depth interview
3 Juan Amador Representative UTT In-depth interview
4 Camila Coordinator INTA In-depth interview
5 Andrea Representative MTE In-depth interview
6 Activando Project Farmers in Zavalla Focus group
7 Antonio Lattuca Farmer in Rosario In-depth interview
8 Facundo Activist in Malvinas Argentinas In-depth interview
9 Diego Fernandez Farmer El Roble In-depth interview
10 Enrique Goites Coordinator INTA In-depth interview

28
[1] is the president of the Latin American
Society of Agroecology (SOCLA). He is an
agronomist by training and is one of the
founders of the programme of agroecology in 8

the National University of La Plata.


9 67

[2] is a farmer active in the association 1610 in


Florencio Varela, which consists of 17 families
2 1 10
45 3
who sell their agroecological produce together.

[3] is a representative in the Unión de


Trabajadores de Tierra (UTT) in La Plata.

[4] Camila is a coordinator at the National


Institute for Agrarian Technology (INTA) in La
Plata.

[5] is a representative for the Movimiento de


Source: SOS Geografía
Trabajadores Excluidos (MTE) in La Plata.

[6] are a group of young agronomy students who started their own agroecological project in Zavalla,
just outside Rosario. Four people participated in the focus group.

[7] is one of the pioneers of urban agriculture in Argentina and has set up numerous activities related to
agroecology in and around Rosario.

[8] is an activist who used to be active in the student protest against Monsanto in Malvinas Argentinas,
now he works at the new agroecology project that was developed in the same village.

[9] is the owner of El Roble, an agroecological farm which also processes its own produce and is a
certified organic producer in Bouquet.

29
[10] is coordinator at INTA and has a lot of experience in agroecological projects in La Plata and Berisso.
He has experience in both social sciences as agronomy and coordinates research at the Institute for
Research and Development of Family Farming (IPAF). He assisted in the conducting of interviews.

These persons will hereinafter be appointed with their number, in order to avoid confusion and to
ascertain that repetition of names is avoided. When in doubt, it is recommended to consult Table 5.1.
The interview templates were added as appendices (Appendix A, B) at the end of this study. Both were
used interchangeably, depending on the profile of the respondent.

5.2 Marxian Ecology


5.2.1 Rationale for Use of a Marxian Ecology Framework

It is […] true that the radical critique of commodity production is indispensable for
understanding the environmental crisis as a crisis of the relationship between humanity and
nature, and thus as social crisis. (Tanuro, 2010, p. 92)

The value in Marxian ecology lies in its ability to go deeper into the structural and political imbalances
of capitalism as a system and link it to the environmental context. It gives a political and social dimension
to the discipline of ecological economics and is thus not necessarily contradictive to ecological
economics but can instead be complementary.

The main critique on Marxian theory is that it has no conception of either Nature or the finite character
of resources and that it follows a mainly productivist approach – this is the Green view, as it is often
called – arguing that Marx and Engels did not consider the relationship between the conceptual
framework that they conceived and biophysical realities (Douai, 2017). However, in Capital can be read:

When man engages in production, he can only proceed as nature does herself, i.e. he can only
change the form of the materials. Furthermore, even in this work of modification he is
constantly helped by natural forces. Labour is therefore not the only source of material wealth,
i.e. of the use-values he produces. As William Petty says, labour is the father of material wealth,
the earth is its mother. (Marx, 1976a, pp. 133–134)

The Marxian school is grounded in ecology on three fronts:

30
1) Marx wrote about historical materialism and his vision on society as a ‘social metabolism’
could be viewed as a step towards a natural or ecological materialism. The man-nature
metabolism is characterized by determinant rules set by Nature on one hand; and by the social
dynamics that are created by the capitalist production process on the other. In part four of
Volume I of Capital, Marx states that:

Capitalist production collects the population together in great centres, and causes the urban
population to achieve an ever-growing preponderance. This has two results. On the one hand
it concentrates the historical motive power of society; on the other hand, it disturbs the
metabolic interaction between man and the earth, i.e. it prevents the return to the soil of its
constituent elements consumed by man in the form of food and clothing; hence it hinders the
operation of the eternal natural condition for the lasting fertility of the soil. (Marx, 1976a, p.
637)

This is a recognition of the metabolism of human transformation of nature and furthermore,


Marx understood that capitalist social relations in agriculture and urbanisation interrupt nutrient
cycling and result in fertility loss (Douai, 2017).

2) Another characteristic of Marxian ecology is the double division of labour. At the start of
Capital, he distinguishes between two aspects of human labour: on one hand the
anthropological dimension of social life; on the other hand the process of capitalist production,
which creates economic value for capital. This is what he called use-value and exchange-value
of labour. Use-value of labour is incommensurable, exchange-value is characterized by its
creation of economic value. In the latter case, Marx speaks of abstract labour, where all labour
is reduced to the same type and they cannot be distinguished anymore. The danger here is that
production of use-values ceases to be an end in itself and is merely a means for creating
economic value. Therefore, the possibility exists that real social needs will not be satisfied and
that social counter-utilities will be generated by a mode of production focused on profit
(Harribey, 2008). This division of labour is important to analysing the relationship of man and
nature. The creation of use-value produced by labour is a result of a transformation of matter
and energy and forms the core of the human-nature metabolism. The creation of economic

31
value by abstract labour relates to an immaterial social relation in capitalism, be it that the
production and exchange of commodities have an energetic and material quality (Douai, 2017).
Or, as Altvater (2007) stated:

In Marx's “Critique of Political Economy” the category of the double character (use
value/exchange value) of labor, production, and the commodity as a result of production is of
utmost importance (Marx himself called it the “Springpunkt”, the core question), because it
bridges the rift between society and nature, between the openness of social practice and hard
natural laws. (Altvater, 2007, p. 58)

3) A third characteristic of Marxian ecology relates to the inherent ecological contradiction in


capitalism. In this issue, different writers have different opinions as to where Marxian theory
refers to this contradiction. Harribey (2008) refers to this contradiction by stating that “without
the exploitation of nature, exploitation of labour would have no material support; and without
the exploitation of labour, exploitation of nature could not have been extended and generalised”
(Harribey, 2008, p. 195), thus referring to the exploitation of nature as being inherent to
capitalism; Altvater (2004) draws on the laws of thermodynamics to explain the ecological
contradiction, as he quite subtly combines ecological economics and Marxian ecology.
Commodity production is simultaneously the quantitative transformation of economic value
and the qualitative transformation of materials and energy, which leads to the conclusion that
there is an inherent tension between the expansionary requirements of capitalist growth and the
necessary increase in entropy (Douai, 2017). Burkett, finally, identifies two main intertwined
crises that capitalism struggles with: the crisis of capital accumulation on one hand; and the
crisis of natural wealth as a condition of human development. The latter refers to the fact that
in Marx’s dialectical view, ecological and social costs of economic value accumulation are
endogenous to the metabolic process of human–nature reproduction in its specifically capitalist
form (Burkett, 2006, pp. 171–172).

32
5.2.2 Materialism and the Conception of Nature and Man

In order to make a consistent and truthful Marxian analysis about the ecological question and the role
that agriculture plays in it, a thorough understanding of materialism is essential. A common
misconception about Marxian theory is the idea that scientific materialism as it was developed by Marx
necessarily entails a rigid, mechanical determinism which has no eye for ecological processes. In fact,
Marx states in his Grundrisse that:

“It is not the unity of living and active humanity with the natural, inorganic conditions of their
metabolic exchange with nature, and hence their appropriation of nature, which requires
explanation or is the result of a historic process, but rather the separation between these
inorganic conditions of human existence and this active existence, a separation which is
completely posited only in the relation of wage labour and capital.” (Marx, 1973, p. 489)

Marx’s conception of the relationship between nature and man is central to materialism as a scientific
theory and to is crucial in understanding Marxian contributions to the ecological debate.

Materialism as a scientific theory has its roots in Greek philosophy, more notably in the works of
Lucretius and Epicurus, the latter of which was the subject of Marx’s doctoral thesis and whose
philosophy played an influential role in the development of the materialism of the French and English
Enlightenment. The former, Lucretius, is regarded by authors like evolutionary biologist Michael Rose
as “the greatest classical forerunner to modern science” (Foster, 2000). Central to their view was the
fact that life was born from the earth and not from the skies (or the result of creation by the gods).
Lucretius wrote in his De Rerum Natura (1969) that “certainly living creatures cannot have dropped
from heaven, nor can terrestrial animals have emerged from the briny gulfs of the sea. So it follows that
the earth has deservedly gained the name of mother, since from the earth all things have been created”
(Lucretius, 1969, p. 158).

The central premise of materialist theory is the idea that reality is dependent on nature and, necessarily,
matter. It is centrally opposed to idealism; the main difference between the two being the assumption of
either matter or idea as the origin of existence. Materialism has been equivalent with most scientific
advances in academia in the 20th century, with its assumption that the existence of thought is preceded

33
by a level of physical reality (Foster, 2000). Idealism, often correlated with the belief systems of religion,
presuppose the existence of a creator, an idea which gave shape to existing structures and the
developments that ensued therefrom. Materialism as a philosophy has three main pillars, based on Roy
Bhaskar (1992):

1) Ontological materialism, asserting the unilateral dependence of social upon biological (and more
generally physical) being and the emergence of the former from the latter;
2) Epistemological materialism, asserting the independent existence and transfactual – that is, causal
and lawlike – activity of at least some of the objects of scientific thought;
3) Practical materialism, asserting the constitutive role of human transformative agency in the
reproduction and transformation of social forms.

Marx was in his work mainly focused on practical materialism because in his view, man’s position to
nature has always been “practical from the outset, that is, relations established by action” (Marx, 1975,
p. 190). However, in his more general view on materialism he embraced both ontological materialism
and epistemological materialism, calling them essential in the pursuit of science (Foster, 2000).

In Marx’s view, historical materialism guided societies throughout history and the struggle of the
exploited against the exploiter permeates this history. Although Marx criticises capitalism, he does not
do so blindly; he realizes that this mode of production is a result of the struggle of workers against certain
divisions of labour. Capitalism as a mode of production is the result of a long historical process
characterized by struggle, and different degrees of freedom for the worker and private ownership (which
are inherently connected). It is thus important to state that this stage of production was preceded by other
stages, namely: primitive communism, slavery and feudalism. Each of these stages, including capitalism,
is characterized by a specific division of labour (and a specific relationship with nature that results from
this). Primitive communal production was characterized by common ownership of the means of
production. The concept of class is alien to this society and did thus not yet exist. This changed with the
introduction of slavery, where common ownership of the means of production was transformed into
private ownership. Not only did this entail the creation of classes, but society was transformed from a
situation where the means of production were socially owned, to a situation where those social elements,
the citizens themselves, became the means of production. Human labour was thus commodified and

34
traded as such. The slave owner not only owns the labour, he owns the slave as well. This changed with
the transition into the feudal system, where the feudal lord owns the means of production, but does not
fully own the worker. He can buy or sell the serf, whose labour he controls, but cannot kill him or
completely treat him like a commodity, as is the case in the slave system. However, private ownership
is further developed and exploitation is nearly as severe as it was under slavery, be it slightly mitigated.
The last step is then a transition into the capitalist mode of production, where the worker is completely
free (see above), both from being treated as a commodity – the capitalist producer cannot kill or trade in
workers – and from the means of production, as private ownership is developed even further. An
inherent contradiction in the capitalist mode of production is the social character of the production
process, but the private ownership of the means of production. There is thus an irreconcilable
contradiction between the character of the productive forces and the relations of production, which are
translated into a class struggle between the exploited and the exploiter. This led Marx to assume a final
stage, Socialism, where these contradictions are lifted and the means of production become socially
owned (Commission of the Central Committee of the C.P.S.U, 1939).

Illustration Practical Materialism

Practical materialism and the reasons for focusing on labour by Marx and Engels become more
apparent when analysing the case made by Engels in his unfinished manuscript “The Part played by
Labour in the Transition from Ape to Man” from 1876. Basing his ideas on Darwin’s theory of
evolution, Engels conducted research into the development of the hand in humans and compared it
with primates genetically similar to humans (eg. chimpanzees).

A first conclusion to be drawn is that the adaptation of feet for developing an erect posture has aided
in freeing hands for labour and thus is an essential evolutionary period in human history related to the
metabolic relationship between man and nature which forms part of Marx’s analysis of capitalism
(see below). Darwin described this process as the law of correlation of growth, defining it as the law
that states that “the specialised forms of separate parts of an organic being are always bound up with
certain forms of other parts that apparently have no connection with them”.

35
A second and more important conclusion drawn from his enquiry was not the causes for the
development of the hand itself, but the consequences it had for the growth of the organism which this
hand forms part of. The enhanced capacity to take part in labour activity with the development of the
hand, necessarily brought humans closer together by increasing cases of mutual support and joint
activity, thus creating the need for the development of language to communicate.

Labour and (consequently) speech were the two most essential stimuli for the development of the
brain which gradually distinguished ape from man. And with this development went the development
of its most immediate instruments – the senses. As hearing developed gradually with speech, the
development of the brain as a whole is accompanied by a gradual refinement of the senses. And the
sense of touch, which “the ape hardly possesses in its crudest initial form, has been developed only
side by side with the development of the human hand itself, through the medium of labour.”

Engels finishes his work with a prophet-like conclusion, seemingly referring to the looming dangers
of climate change which demonstrates the scope of his research and the validity of Marxian theory
for analysing the ecological question:

“Let us not, however, flatter ourselves overmuch on account of our human victories over
nature. For each such victory nature takes its revenge on us. Each victory, it is true, in the
first place brings about the results we expected, but in the second and third places it has
quite different, unforeseen effects which only too often cancel the first. The people who, in
Mesopotamia, Greece, Asia Minor and elsewhere, destroyed the forests to obtain cultivable
land, never dreamed that by removing along with the forests the collecting centres and
reservoirs of moisture they were laying the basis for the present forlorn state of those
countries. When the Italians of the Alps used up the pine forests on the southern slopes, so
carefully cherished on the northern slopes, they had no inkling that by doing so they were
cutting at the roots of the dairy industry in their region; they had still less inkling that they
were thereby depriving their mountain springs of water for the greater part of the year, and
making it possible for them to pour still more furious torrents on the plains during the rainy

36
seasons. Those who spread the potato in Europe were not aware that with these farinaceous
tubers they were at the same time spreading scrofula. Thus at every step we are reminded
that we by no means rule over nature like a conqueror over a foreign people, like someone
standing outside nature – but that we, with flesh, blood and brain, belong to nature, and
exist in its midst, and that all our mastery of it consists in the fact that we have the advantage
over all other creatures of being able to learn its laws and apply them correctly.” (Engels,
1934)

5.2.3 Metabolism as an Ecological and Societal Concept

Marx believed that the relation of man to nature is an ever evolving process mediated by labour, which
defines man’s position towards the environment that it forms part of. He called this interaction
Stoffwechsel, or metabolism. This concept of metabolism was the key foundation to Marx’s analysis
and he based his theories on this assumption. In his work Capital, Marx writes:

Labour is, first of all, a process between man and nature, a process by which man, through his
own actions, mediates, regulates and controls the metabolism between himself and nature. […]
He sets in motion the natural forces which belong to his own body, his arms, legs, head and
hands, in order to appropriate the materials of nature in a form adapted to his own needs.
Through this movement he acts upon external nature and changes it, and in this way he
simultaneously changes his own nature. […] The labour process […] is the universal
condition for the metabolic interaction (Stoffwechsel) between man and nature, the everlasting
nature-imposed condition of human existence, and it is therefore independent of every form of
that existence, or rather it is common to all forms of society in which human beings live. (Marx,
1976a, pp. 283–290)

The concept of Stoffwechsel, metabolism, did not originate in Marx’s own writing. Conceptualized in
the 1840s, it emerged with the physicist Theodorr Schwann introducing the notion of cellular
metabolism in 1839. It refers to a biochemical process in which an organism or cell draws upon the

37
materials from its environment and converts it into the building blocks of growth. This concept is now
widely used by leading ecologists like Eugene Odum, who employ it to incorporate all biological levels,
starting from a single-cell organism and ending with a whole ecosystem. Marx, who studied under
Schwann, was extremely impressed by this concept and applied it to society together with Engels (Foster,
2000). Fischer-Kowalski states that “the concept of metabolism in biology has valuable features: it
refers to a highly complex self-organizing process that the organism seeks to maintain in widely varying
environments. This metabolism requires certain material inputs from the environment, and it returns
these materials to the environment in a different form” (Fischer-Kowalski, 1998, p. 64); for Marx, the
force mediating this process for society was labour and its development within historically specific social
formations (Foster, 2000). Most social scientists at that time were very interested in advances in biology
and natural sciences and criticized the separation and categorization between social sciences on one hand,
and exact or natural sciences on the other. The objective of Marx and Engels here was to establish a
mutual, interdependent relationship between man and nature, going beyond the widespread notion of
“man utilizing nature” (Fischer-Kowalski, 1998).

The idea of a separation in this metabolic relation is what Marx called the metabolic rift, a material
estrangement of mankind within capitalist society from the natural conditions which form the basis of
its existence (Foster, 2000). Marx defined this separation or estrangement as the metabolic rift, primarily
between mankind and the soil, but also between mankind and labour – which coincides with the concept
of alienation, a concept central to Marx’s analysis of capitalism. In Capital Vol. 3, he states that:

Large landed property reduces the agricultural population to an ever decreasing minimum
and confronts it with an ever growing industrial population crammed together in large towns;
in this way it produces conditions that provoke an irreparable rift in the interdependent process
of the social metabolism, a metabolism prescribed by the natural laws of life itself. The result
of this is a squandering of the vitality of the soil, which is carried by trade far beyond the bounds
of a single country. […] Large-scale industry and industrially pursued large-scale agriculture
have the same effect. If they are originally distinguished by the fact that the former lays waste
and ruins the labour-power and thus the natural power of man, whereas the latter does the
same to the natural power of the soil, they link up in the later course of development, since the

38
industrial system applied to agriculture also enervates the workers there, while industry and
trade for their part provide agriculture with the means of exhausting the soil. (Marx, 1991, pp.
949–950)

To assume that large-scale capitalist society created such a rift in the metabolism, was to argue that the
fundamental conditions of sustainability had been violated. This sustainability issue relates not only to
the soil, but also to the separation of town and country (read: urban and rural areas), which through its
antagonism disrupts the social metabolism. Marx problematized the consequences of increased
urbanization to the social metabolism and the effect it had on the recycling of nutrients. In Vol. 3 of
Capital, he notes that “there is a colossal wastage in the capitalist economy in proportion to their actual
use. In London, for example, they can do nothing better with the excrement produced by 4 million people
than pollute the Thames with it, at monstrous expense” (Marx, 1991, p. 195). Engels observed the same
issue as Marx in his work The Housing Question, claiming that “in London alone a greater quantity of
manure than is produced by the whole kingdom of Saxony is poured away every day into the sea with
an expenditure of enormous sums, and […] colossal works are necessary in order to prevent this
manure from poisoning the whole of London” (Engels, 1935, p. 95). It is no coincidence that Marx and
Engels both formulated such a specific analysis about the question of human waste in urban areas. In
fact, the development of the concept of metabolic rift was based on the work of the German agricultural
chemist Julius von Liebig, who made thorough analyses about the circulation of soil nutrients and its
relation to animal metabolism (Foster, 2000).

The idea that Marx had an intricate understanding of the concept of sustainability becomes clear in the
stance he makes against capitalist modes of production and the incompatibility it carries with ecological
sustainability:

[…] the way that the cultivation of particular crops depends on fluctuations in market prices
and the constant changes in cultivation with these price fluctuations - the entire spirit of
capitalist production, which is oriented towards the most immediate monetary profit - stands
in contradiction to agriculture, which has to concern itself with the whole gamut of permanent
conditions of life required by the chain of human generations. (Marx, 1991, p. 754)

39
The present-day notion of sustainability as captured in the definition from the Brundtland Commission
as “development which meets the needs of the present without compromising the ability of future
generations to meet their own needs” (World Commission on Environment, 1987) shows great
similarities to the emphasis by Marx on the need to maintain “the permanent conditions of life required
by the chain of human generations”. In a later section of Capital Vol. 3, he states that “an entire society,
a nation, or all simultaneously existing societies taken together, are not the owners of the earth. They
are simply its possessors, its beneficiaries, and have to bequeath it in an improved state to succeeding
generations, as boni patres familias” (Marx, 1991, p. 911), boni patres familias loosely translated to
“good heads of households”.

Illustration Marx and the Depletion of Resources

Marx himself was quite worried about deforestation and viewed this as the historical result of the
exploitative relation to nature that had characterized not only capitalism, but all civilizations up to that
point (Foster, 2000). In Capital Vol. 2, he wrote that “the development of civilization and industry in
general has always shown itself so active in the destruction of forests that everything that has been
done for their conservation and production is completely insignificant in comparison” (Marx, 1992,
p. 322).

In a letter to Engels he tackles the issue of climate change and deforestation related to agricultural
cultivation. Aside from this paragraph being further proof of his interest in the natural sciences, Marx
tackles the issue of climate change directly, although it should be clear that this should be seen in the
context of 19th century scholarship, where knowledge about issues regarding climate change was
rather limited. In the letter, he writes: “Very interesting is the book by Fraas (1847): Klima und
Pflanzenwelt in der Zeit, eine Geschichte beider, namely as proving that climate and flora change in
historical times. He is a Darwinist before Darwin, and admits even the species developing in
historical times. But he is at the same time agronomist. He claims that with cultivation – depending
on its degree – the 'moisture' so beloved by the peasants gets lost (hence also the plants migrate from
south to north), and finally steppe formation occurs. The first effect of cultivation is useful, but finally
devastating through deforestation, etc.” (Marx & Engels, 1987b, pp. 558–559).

40
Engels himself was aware of the dangers of resource depletion and excessive energy use. In a letter
to Marx he analyses a work by Podolinsky, who would end up introducing the second law of entropy
into economics which would ultimately become one of the foundations of ecological economics.
Although he criticises Podolinsky for his initial economic conclusions, he readily adopts the physical
conclusions that Podolinsky draws, stating that “his real discovery is that human labour is capable
of retaining solar energy on the earth's surface and harnessing it for a longer period than would
otherwise have been the case.” (Marx & Engels, 1992, p. 410). However, he takes the ecological
question a step further:

What Podolinski has completely forgotten is that the working individual is not only a
stabiliser of the present but also, and to a far greater extent, a squanderer of past, solar heat.
As to what we have done in the way of squandering our reserves of energy, our coal, ore,
forests, etc., you are better informed than I am. From this point of view, hunting and fishing
may also be seen not as stabilisers of fresh solar heat but as exhausters and even incipient
squanderers of the solar energy that has accumulated from the past. (Marx & Engels, 1992,
p. 411).

5.2.4 Commodities and the Creation of Value

The widespread belief that Marx’s labour theory of value excludes or minimizes the role of nature in the
production process is not only erroneous, it limits the possibilities of broadening Marxian theory with
ecological perspectives. In separating use-value from exchange-value in commodities, Marx offers a
broader understanding of the role of nature in the production of these commodities. In his analysis,
exchange-value is the monetary value that is represented in the market and that is determined by the
labour time spent producing a certain commodity. In an economy where social production occurs in
private enterprises, the acceptance of any labour is determined ex post by the exchange-values that
certain products command on the market (Burkett, 2014).

41
The difference with use-value lies in the fact that use-value is intrinsic to the commodity itself, while
exchange-value is the value it represents in trade. Marx defines use-values as:

[…] a use-value […] is conditioned by the physical properties of the commodity, and has no
existence apart from the latter. […] Use-values are only realized [verwirklicht] in use or in
consumption. They constitute the material content of wealth, or whatever its social form may
be. In the form of society to be considered here they are also the material bearers [Trager] of
[…] exchange-value. (Marx, 1976a, p. 126)

Use-value, also called utility (different from the neoclassical concept of utility, which is subjectively
defined by the consumer), varies above all in quality; while exchange-value varies only in quantity and
“does not contain an atom of use-value” (Marx, 1976a, p. 128). This exchange-value has another
consequence, namely the abstraction of labour. As we disregard use-value and make an abstraction of
this, labour is also abstracted. The table, house or yarn ceases to be any useful thing, as all of its sensuous
characteristics are extinguished; and so the labour needed to produce them is not anymore the labour of
the joiner, the mason or the spinner, or any kind of productive labour, but leads instead to the
disappearance of specific forms of labour. Labour cannot anymore be distinguished and is reduced to
the same kind of labour, “human labour in the abstract” (Marx, 1976a, p. 128). Products are now a
crystallization of human labour-power in social form, which determines its value. Or, as Marx states in
Value, Price & Profit (1976b):

A commodity has a value, because it is a crystallisation of social labour. The greatness of its
value, or its relative value, depends upon the greater or less amount of that social substance
contained in it; that is to say, on the relative mass of labour necessary for its production. (Marx,
1976b, p. 31)

It is necessary to mention that labour is here regarded as homogeneous labour, or the expenditure of one
uniform labour-power. In Capital Vol. 1, he writes:

The total labour-power of society, which is manifested in the values of the world of
commodities, counts here as one homogeneous mass of human labour-power, although
composed of innumerable individual units of labour-power. Each of these units is the same as

42
any other, to the extent that it has the character of a socially average unit of labour-power and
acts as such; i.e. only needs, in order to produce a commodity, the labour time which is
necessary on an average, or in other words is socially necessary. (Marx, 1976a, p. 129)

The regulation of wealth production by exchange-values here presupposes that the acquisition of use-
values requires purchasing of commodities with money received through commodity sales. An
economy based on commodities thus excludes individuals from reproducing themselves independently
from the market nexus. The key point to this process of social production featuring commodified
conditions of production is that it inherently requires a social separation from the human producers of
these commodities from these conditions. Labourers can attain necessary use-values only by selling their
labour power to the capitalists controlling the necessary production conditions of these values. This
regulation of the division of labour by monetary exchange-values representing socially necessary labour
time (see above) thus becomes the dominant form of production to such an extent as it is able to maintain
the dominant social relation of production between capital and free labourers (free is here referred to as
free from the means of production, not as the wider, philosophical concept of freedom) (Burkett, 2014).
Or, as Marx states in Capital Vol. 1:

Free workers, in the double sense that they neither form part of the means of production
themselves, as would be the case with slaves, serfs, etc., nor do they own the means of
production, as would be the case with self-employed peasant proprietors. The free workers are
therefore free from, unencumbered by, any means of production of their own. With the
polarization of the commodity-market into these two classes, the fundamental conditions of
capitalist production are present. The capital-relation presupposes a complete separation
between the workers and the ownership of the conditions for the realization of their labour.
(Marx, 1976a, p. 874)

This separation has repercussions for the conception of society’s relation to nature. Burkett (2014) gives
three aspects of significance of Marx’s approach:

1) Marx states that exchange-value is a form of value and not the reverse. This is aimed at the
creation of value, which Marx insists arises only in production and not in the realm of exchange.

43
This is important because it seems to have been missed or forgotten by many of his ecological
critics, who wish to ascribe value to nature (and not just use-value).

2) The subordination of exchange-value and use-value as particular forms of value leads to the
increasing domination of production for profitable sales (M-C-M’, where M stands for money
and C for commodity) instead of for productive use (in which any monetary exchange occurs
to satisfy the acquisition of alternative use-values, as contained in C-M-C’). Marx here states
that:

[…] the circulation M-C-M both the money and the commodity function only as different
modes of existence of value itself […]. It is constantly changing from one form into the other,
without becoming lost in this movement […]. If we pin down the specific forms of appearance
assumed in turn by self-valorizing value in the course of its life, we reach the following
elucidation: capital is money, capital is commodities. In truth, however, value is here the
subject of a process in which, while constantly assuming the form in turn of money and
commodities, it changes its own magnitude, throws off surplus-value from itself considered as
original value, and thus valorizes itself independently. (Marx, 1976a, p. 255)

This is important as insofar the concept of value encapsulates capitalism’s inherent antagonism
with nature (see below), any environmental policies in which value remains the active factor
are unlikely to alleviate ecological crises, as it is unable to alleviate the inherent class conflict
between capital and labour. This principally relates to private or government-imposed
monetary rents on natural resources.

3) The last, yet most important observation considers the detrimental effect of subordination of
exchange-value and use-value under value, representing a social abstraction of use-value
(which is the material need-satisfying character of production). In Marx’s view, this ultimately
leads to the separation of wealth creation from its natural and material basis. He states in Capital
Vol. 1 that “material wealth, the world of use values, exclusively consists of natural materials
modified by labour, hence appropriated solely through labour, and the social form of this
wealth, exchange value, is nothing but a particular social form of the objectified labour

44
contained in the use values” (Marx, 1988, p. 40). Stated differently, the contradictory
relationship between exchange-value and use-value that arises from a commodity economy
inherently infers a secondary contradiction between wealth’s specifically capitalist production
form and its natural basis and substance. While nature contributes to the production of use-
values, capitalism represents wealth in a purely quantitative, socio-formal abstraction: labour
time in general. The free appropriation of nature through capital, which occurs when nature
contributes to the production of use-value but does not add to value production, manifests this
contradiction “insofar as it is enabled by the system’s valuation of nature according to the
social labour time necessary for its appropriation in commodity production, not according to
the real contribution of nature to wealth or human need satisfaction” (Burkett, 2014, p. 82).

Marx himself was aware of the main contradiction in this theory, namely that if exchange-value is
determined by the labour-time contained in a commodity, this gives no explanation to the exchange-
value for products which contain no labour-time; in other words, how the exchange-value of natural
forces arises. His solution was the concept of capitalist rents, which though extremely complex in its
details, is quite straightforward in its basic premise. The rent of land is treated as a redistribution of
surplus value, or as Engels put it in his letter to Schmidt, it “represents the surplus profit, in excess of
the general rate, arising from a monopolisation of a natural force” (Marx & Engels, 2010, p. 464). Or,
as Marx stated it in Capital Vol. 3:

Wherever natural forces can be monopolized and give the industrialist who makes use of them
a surplus profit, whether a waterfall, a rich mine, fishing grounds or a well-situated building
site, the person indicated as the owner of these natural objects, by virtue of his title to a portion
of the earth, seizes this surplus profit from the functioning capital in the form of rent. […] One
section of society here demands a tribute from the other for the very right to live on the earth,
just as landed property in general involves the right of the proprietors to exploit the earth's
surface, the bowels of the earth, the air and thereby the maintenance and development of life.
(Marx, 1991, pp. 908–909)

It should be noted that the real underlying tension is the one between use-value on one hand, and
exchange-value on the other. The monetary claim on natural conditions enables the commanding of a

45
positive market price, or exchange-value – defined as the product of social labour time – but without the
utilization of human labour; in other words, the barrier to capitalist competition as posed by landed
property has a material basis in the irreplaceable character of natural wealth (Burkett, 2014).

Illustration Examples of Systemic Contradictions Arising from Marx’s Value Theory

The contradiction between use-value and exchange-value, or the qualitative and material substance
of a commodity versus its quantitative and homogenous form, culminates in the necessity for money.
Indeed, Marx identified money as a necessary form for solving this contradiction (or at the very least,
masking it). In determining money as a representative for and carrier of value, an abstraction occurs
from the natural and human substances of wealth. He wrote in his Grundrisse that “as a value, the
commodity is an equivalent; as an equivalent, all its natural properties are extinguished; it no longer
takes up a special, qualitative relationship towards the other commodities” (Marx, 1973, p. 141).
Money therefore abstracts not only from labour, but also from natural and ecological diversity, insofar
these are not manifested in the quantity of social labour time required to appropriate and productively
utilize natural conditions (Burkett, 2014). This leads to a phenomenon that Rachel Carson calls a
“shotgun approach to nature”. She stated in her 1962 book that our attitude towards plants

is a singularly narrow one. If we see any immediate utility in a plant we foster it. If for any
reason we find its presence undesirable or merely a matter of indifference, we may condemn
it to destruction forthwith. […] Many are marked for destruction merely because, according
to our narrow view, they happen to be in the wrong place at the wrong time. (Carson, 1962,
pp. 63–64)

The abstraction of value from nature that arises from a commodity economy inevitably leads to a
homogenization of social and ecological diversity and thus entails an inherent contradiction that
shakes the foundation of this production system. The abstraction of value from nature is not merely
an influence in the development of capitalist society; its social and material power accumulates as the
combined productivity of labour and nature is subsumed under nature (Burkett, 2014).

46
Another aspect that arises from the qualitative homogeneity of value is the infinite divisibility
problem. As stated by Marx, use-value is (mainly) defined by its quality, while exchange-value is
defined by its quantity. This entails that value and money do not only abstract from qualitative
environmental distinctions and relationships, they also valuate nature as to quantify it and thus enable
its artificial division or fragmentation (Burkett, 2014). This aspect too, did not go unnoticed by Marx.
He mentions in his Grundrisse that “as a value, every commodity is equally divisible; in its natural
existence this is not the case. As a value it remains the same no matter how many metamorphoses
and forms of existence it goes through” (Marx, 1973, p. 141). Marxian geographer David Harvey
interpreted this division in his article ‘The nature of Environment: the Dialectics of Social and
Environmental Change’ (1993), where he analyses the dubious relationship of exchange-value
towards natural conditions. He very eloquently puts forth a strong critique concerning this relationship
and claims that

money prices attach to particular things and presuppose exchangeable entities with respect
to which private property rights can be established or inferred. This means that we conceive
of entities as if they can be taken out of any ecosystem of which they are a part. We presume
to value the fish, for example, independently of the water in which they swim. The money
value of a whole ecosystem can be arrived at, according to this logic, only by adding up the
sum of its parts, which are construed in an atomistic relation to the whole. This way of
pursuing monetary valuations tends to break down, however, when we view the
environment as being constructed organically, ecosystemically or dialectically […] rather
than as a Cartesian machine with replaceable parts. Indeed, pursuit of monetary valuations
commits us to a thoroughly Cartesian-Newtonian-Lockeian and in some respects 'anti-
ecological' ontology of how the natural world is constituted. (Harvey, 1993, p. 6)

47
5.3 Ecological Economics
Contrary to classical Marxian theory, which was developed mainly by Marx and Engels, ecological
economics is a field which is characterised by internal discussions and fierce debate. There are, however,
certain key concepts that distinguish this economic field from others. In what follows the foundations of
this theory considered to be relevant to the topic at hand will be outlined.

5.3.1 On the Character of Economics

Ecological economics (EE) is a discipline that arised out of a conception of sustainability that challenged
the dominant perception of it. While many academic fields still use the traditional perspective of
sustainability, namely that in order for a society to be considered sustainable, it needs to conform to the
three basic pillars that each carry the weight of sustainability. A society is truly considered sustainable,
in this view, when it satisfies the economic, social and environmental conditions that are needed to
sustain its existence indefinitely (represented as × in Figure 5.1a). The economist René Passet was the
first one to challenge this conception and the field of EE is based on his representations in the book
L’économique et le vivant (1979).

Figure 5.1 A representation of the traditional view on sustainability (a) and the view
proposed by ecological economics (b)

Biosphere
Environmental

×
Human-social sphere
Social

Economic sphere
Economic
(a) (b)

Source: own image based on Martínez-Alier & Muradian (2015)

48
The main assumption in the conception of reality by the field of EE is that the economic sphere is
preceded by the environmental and social sphere. In other words, there was a nature before human
society, and a society before the development of the economic market as it is known today. This concept,
seemingly influenced by historical materialism as proposed by Marx, has severe repercussions for the
perception of economics. Economics is not a sphere that exists outside of the ecosystem and produces
externalities that damage it, but instead is a sphere that is embedded in environmental and social relations.
The study of the market should thus come after the study of ecology and social institutions. Not the
internalities but the externalities come first (Martínez-Alier & Muradian, 2015).

A central aspect to this is the inclusion of the laws of thermodynamics into economics. Ecological
economics see the economic system as an open system, being defined as a system with entry and exit
of materials and energy – closed systems are those with entry and exit of energy but not of materials,
isolated systems are those without entry or exit of either materials or energy – and argument their claims
using the first and second law of thermodynamics. Although there is debate concerning the introduction
of the thermodynamic laws, many ecological economists have stated a desire to include an ecological
foundation based on natural sciences in economics.

The first law of thermodynamics is quite straightforward, their main premise consists of the fact that
nothing can be created from or returned to nothing; which implies that material and energy cannot be
destroyed, but rather interchanged. In burning wood, energy (in the form of heat) is created out of a
material. This process also produces CO2 and other emission gases and is telling about the waste
problem human society faces, as plastic can in most cases not completely be recycled but is either burned
(releasing toxic fumes into the atmosphere) or ends up on a landmass. Raw material inputs are extracted
from the environment, and to the environment they return in the form of waste (Ayres, 1998).

It is the second law of thermodynamics that is more relevant to ecological economics, however, and that
has sparked most of the debate in academia. This law states that the universe generally moves from a
state of low entropy to a state of high entropy, it never decreases. Entropy is the function of the state of
a particular system and is very difficult to define, even for the most trained physicists. Nicholas
Georgescu-Roegen, one of the pioneers of ecological economics, describes it as follows in his chapter

49
The Entropy Law and the Economy Problem (1994) of the book Valuing the Earth: Economics, Ecology,
Ethics by Daly & Townsend:

Energy exists in two qualitative states – available or free energy, over which man has almost
complete command, and unavailable or bound energy, which man cannot possibly use. […]
When a piece of coal is burned, its chemical energy is neither decreased nor increased. But
the initial free energy has become so dissipated in the form of heat, smoke and ashes that man
can no longer use it. It has been degraded into bound energy. […] Bound energy is […]
chaotically dissipated energy. […] This is why entropy is also defined as a measure of disorder.
It fits the fact that a coper sheet represents a lower entropy than the copper ore from which it
was produced. (Georgescu-Roegen, 1994)

This is reflected in our economic system through the use of resources and its environmental
repercussions. Humans have historically depended on two sources of energy; either stocks of
concentrated minerals or the solar flow of radiant energy. The latter used to be more important, as
traditional societies used a limited amount of fossil minerals and maximised the use of solar energy to
maximize agricultural production. The Industrial Revolution, which enhanced the drive for capital
accumulation but also the vast technological leaps and the increased spending of the State in military –
which quickly increased during the two World Wars – altered the economic system. The role of fossil
minerals like oil, gas and petrochemicals became more important and were introduced in agriculture
with industrial machinery and methods replacing the conventional techniques, and chemical fertilizers
replacing organic ones. This shift and transition to modernity changed all social-ecological interactions
or, as previously mentioned, the social metabolism. The biomass- and solar-driven local economy was
replaced by a petrochemical international economy with a much higher dependence on concentrated
minerals (Splash, 2017). What this means for the conception of economic flows is shown in Figure 5.2.

50
Figure 5.2 The Economic System in Ecological Economics

Matter and
high entropy High entropy
energy energy (mainly
low-grade
Nature/Economy Boundary thermal
energy)
Circular flow model
Output
markets
Firm Households
s
Input markets

Low entropy
energy (mainly Matter and low
solar energy) entropy energy
(natural resources)
Earth System Boundary

Source: own image based on Daly & Farley (2011)

It is through this conception of economics as part of a bigger ecological system that the neoclassical
perspective on economics as a standalone system is challenged. Ecological economics attempt to point
out that a given stock is by definition finite and that a society built on the depletion of a non-renewable
stock will collapse. Only if these finite resources can be replenished or substituted can disaster be
avoided. The Earth is here regarded as a virtually closed system in terms of matter, which moves through
the biogeochemical cycles, but an open system in terms of energy. The matter and energy that are
expelled out of the economic system are qualitatively lower – high entropy, unconcentrated – than the
inputs it takes in. This matter and energy, defined as ‘waste’ by human society, must end up in the air,
water or land. Matter and energy are in the neoclassical view not only transformed in form, but in
concept: they are externalities that can be regulated by the market. Or, as Splash (2017, p. 9) states it,
pollution is “an inevitable part of the economic process, not an avoidable externality that disappears if

51
the prices are ‘right’, and it inevitably increases with economic growth because that growth is
dependent upon material and energy throughput”.

As Burkett (2006) correctly points out, this relationship is mediated by the use of technology. The
entropy law corresponded originally to mechanistic systems, not to systems which were purposefully
driven. This means that when analysing the impact of entropy on the economic system, which could be
seen as a purposefully driven system, the ability of technology to mediate the effects of entropy need to
be taken into account. However, it would be foolish to do away with the theory altogether. Sir Arthur
Stanley Eddington stated the importance of the second law of thermodynamics as follows:

The law that entropy always increases holds, I think, the supreme position among the laws of
nature. […] If your theory is found to be against the second law of thermodynamics, I can give
you no hope; there is nothing for it but to collapse in deepest humiliation. (Eddington, 1929, p.
74)

The inherent unsustainability of the current system thus leads, according to ecological economics, to the
necessity of de-growth in order to reach a steady-state economy. This means that the developed countries
would not strive for economic growth, but instead try to control it, or even completely abandon the
paradigm of growth. This has faced tremendous criticism and it is easy to understand why: there are
many unanswered questions considering the construction of such an economy and what it would mean
to the current system. What does this mean for interest rates? For capital accumulation in general? Is it
compatible with capitalism? Ecological economics has not yet reached the degree of maturity as a
discipline to have an answer to these questions, which will prove to be crucial and divisive in the
evolution of the academic field (Trainer, 2012).

5.3.2 The Incommensurability of Values

The theory of incommensurability of values is based on the belief that humans face inherent needs which
are not automatically substitutable by other goods. There is a sort of sacredness to the need for water to
sustain human life, for example (Martínez-Alier & Muradian, 2015). In the same way, endosomatic
energy (the use of energy that a human consumes inside the body, also expressed as food energy,
expressed in kcal) cannot readily be substituted by exosomatic energy (the use of energy in broader

52
society, eg. to cook the food). Industrial societies use considerably more exosomatic energy than
agricultural societies, but in both cases exosomatic energy only increases once the requirements for
endosomatic energy are fulfilled (Fischer-Kowalski & Haberl, 2007).

Stressing the existence of lexicographic preferences in economics (here carrying the meaning of a
hierarchical order of preferences), ecological economists assume a limitation on the substitutability of
certain goods, there is thus an incommensurability of values. Ecological economics rejects the premise
of categorizing unique value types under a single numeric denominator. Or, as Funtowicz & Ravetz
(1994) state:

The issue is not whether it is only the marketplace that can determine value, for economists
have long debated other means of valuation; our concern is with the assumption that in any
dialogue, all valuations or ‘numeraires’ should be reducible to a single one-dimensional
standard. (Funtowicz & Ravetz, 1994, p. 199)

In the same article, Funtowicz & Ravetz do admit that certain ecological economists have pleaded for a
development of monetary valuation schemes for ecosystems, which is expressed from a more pragmatic
angle to protect the environment; but this seems to be more of a coping strategy in the current economic
climate than a valid alternative to the current economic system. Indeed, Funtowicz & Ravetz admit that
these proposals are mostly asserted to secure compensation for damage that has already been done.
Martínez-Alier & Muradian (2015) come to the same conclusion, arguing that money valuation is
perhaps most appropriate when trying to hold third parties accountable in civil litigations for past
environmental liabilities. Concerning the use of money valuation as a standalone system for combating
environmental degradation, they state that:

[…] the insistence on money valuation clearly makes less visible the biological and ecological
importance of Nature, and also livelihood and cultural values. The beauty and sacredness of
mountains such as the Niyamgiri Hills in Odisha might seem negligible when compared to the
very large money value of their bauxite deposits. The mountains are better defended outside
money valuation […]. (Martínez-Alier & Muradian, 2015, p. 10)

53
Plurality of values is inherently bound to rational decision-making, according to O’Neill (2017). Taking
into account the value of social-ecological systems for many communities worldwide, it is important to
realize that it is difficult, if not near impossible, to translate a livelihood into a monetary value. Indeed,
as O’Neill asserts, monetary values merely capture the intensity of preferences but not the soundness of
the reasons for them. This has consequences for policymaking, as he states that “rational public choice
requires deliberative institutions that allow citizens to transform preferences through reasoned dialogue,
rather than measuring through money given preferences for the purposes of aggregating them to arrive
at an ‘optimal’ outcome” (O’Neill, 2017, p. 234).

Ecological economics thus challenges the idea that rational decision-making demands value
commensurability. Starting from the assumption of a plurality of needs, they conclude that certain goods
cannot be substituted by others and that an incorporation of this plurality is thus inherent to the study of
economics and rational decision-making.

5.3.3 Social Institutions and the Economy

Another crucial aspect of ecological economics involves not the ecological, but rather the social nature
of economics. For ecological economists, economics is embedded in a social system which it forms part
of, and it cannot be understood without taking this social system into account. The difference with
neoclassical economics is here embodied by the car engine metaphor, as demonstrated by Martínez-
Alier & Muradian (2015). According to the neoclassical economists, the road system is of no importance
in the design and engineering of a car engine. Regardless of which rules apply, a car engine will be made
in the same way and with the same materials. The design of car engines ultimately determines
transportation as we know it.

Ecological economists reject this mechanical approach to economics. The economy has no meaning
without its social and political context and needs thus to be interpreted in its socio-political environment.
Engineering economics, following the metaphor, can only be justified as a small part of specialists
operating in a larger team of specialists bringing together the necessary social sciences. Regarding the
construction of car engines as a standalone concept implies that there is no conceptualisation of the
vehicle they are supposed to drive or what is required in terms of the social system that would make

54
such vehicles operative. The institutional design and social structure of a transportation system, and the
resulting human behaviour in that system, are much more important than the technical and engineering
design (Splash, 2017). The creation of car engines then only acquires meaning once it is embedded
within a larger notion of transportation. The same counts for economics and social sciences.

Ecological economists claim that preferences and decision-making strategies are a consequence of
social norms and structures. This is inherently tied to the conceptualization of the plurality of values (see
above) and carries an inherent institutional dimension with it. It therefore should not come as a surprise
that institutional economics has been an influential source of ideas for ecological economics. It has
turned to various sophisticated models to determine human behaviour and the role of institutional actors
in certain environmental outcomes (Paavola & Adger, 2005).

The interdisciplinary nature of ecological economics should be clear by now, as it not only draws ideas
from ecology, but also from institutional economics and social psychology. The question now becomes:
what commonalities are there to be drawn with Marxian ecology?

5.4 Towards a ME-EE Nexus


It is clear that a lot of similarities exist between Marxian Ecology and Ecological Economics. This is
somewhat surprising as these academic fields are separated by approximately one century, but does
contain logic when assuming that ecology as a systems theory today has been strongly influenced by
Marxian theory. The first author to introduce the second law of thermodynamics, Podolinsky, was very
much inspired by Marx’s writing and even engaged in correspondence with Marx and Engels.
Materialism was the foundation of the analysis for most original ecologists. This influence has led
Bhandari (2015) to describe Marxian philosophy and economics as the “original systems theory” and
Foster (2015) to state that “ecology as we know it today thus represents the triumph of a materialist
systems theory”. The concept of materialism permeates the science of ecology and seems as much a
part of ecological economics as it is of Marxian ecology. The concept of a social metabolism, which is
crucial to ecological economics and resulted from a materialist approach to science, was first coined by
Marx.

55
The question why this materialist aspect was lost in 20th century Marxism is related to the split between
Soviet Marxism and Western Marxism. Soviet Marxism was shaped by thinkers who continued to see
“complex, historical, interconnected views of development, associated with dialectical reasoning, as
essential to the understanding of nature and science” (Foster, 2015, p. 4), while in Western Marxism
and the Frankfurt School in particular, Marxian economics were separated from their natural basis and
ecological critiques formulated by neo-Marxists lacked a thorough interdisciplinary analysis of ecology
and instead rested on vague ethical criticism. The tragedy of this separation exists in the fact that the
scientific, materialist foundation of Marxism went lost in both cases: in the Soviet Union, Stalinist purges
of the scientific community ensured that it took on a more politically repressive role, with the
technological optimism that characterized this period reinforcing this rigidity; while in the West the
abstraction of Marxian ecology from its scientific base influenced Green philosophy, where a radical
anti-technological stance and a neglect of natural-scientific developments started to appear. The Green
philosophy then evolved with new insights of ecology but left behind its long-lost materialist roots
(Foster, 2015).

The biggest criticism from ecological economists, and Martínez-Alier in particular (the most renowned
historian of ecological economics), towards Marx and Engels is uttered against their position towards
the second law of thermodynamics, which plays a crucial role in the theory of ecological economics.
Indeed, in 2007 he criticized Engels’ “unwillingness to accept that the First and Second Laws of
thermodynamics could apply together” (Martínez-Alier, 2007, p. 275), Bensaïd said the same in his
work Marx For Our Times, stating that Engels “adhered to the first principle (conservation of energy),
while rejecting the second (its progressive dissipation)” (Bensaïd, 2002, p. 331) and Frenay even went
so far as to state that Marx and Engels “proceeded down a road that pointedly ignored the 2nd Law
(Engels misunderstood it, thinking it contradicted the 1st) and environmental considerations in general”
(Frenay, 2006, p. 364).

This critique towards Engels (and, by implication, Marx) is based on a letter that Engels wrote to Marx
and four lines in Engels’ introduction of The Dialectics of Nature (Marx & Engels, 1987a) and stem
from a considerable lack of insight into the works of Marx and Engels, as will be shown below. In a

56
letter to Marx from 1869 (Marx & Engels, 1988, p. 246) – the full passage is here displayed to avoid
misrepresentation – Engels writes that

In Germany the conversion of the natural forces, for instance, heat into mechanical energy,
etc., has given rise to a very absurd theory, which incidentally follows with a certain
inevitability from Laplace's old hypothesis, but is now displayed, as it were, with mathematical
proofs: that the world is becoming steadily colder, that the temperature in the universe is
levelling down and that, in the end, a moment will come when all life will be impossible and
the entire world will consist of frozen spheres rotating round one another. I am simply waiting
for the moment when the clerics seize upon this theory as the last word in materialism. It is
impossible to imagine anything more stupid. Since, according to this theory, in the existing
world, more heat must always be converted into other energy than can be obtained by
converting other energy into heat, so the original hot state, out of which things have cooled, is
obviously inexplicable, even contradictory, and thus presumes a god. Newton's first impulse is
thus converted into a first heating. Nevertheless, the theory is regarded as the finest and highest
perfection of materialism; these gentlemen prefer to construct a world that begins in nonsense
and ends in nonsense, instead of regarding these nonsensical consequences as proof that what
they call natural law is, to date, only half-known to them. But this theory is all the dreadful rage
in Germany.

In this passage the bold text was added purposefully to display the lines that are frequently used to prove
Engels’ rejection of the second law of thermodynamics. The common misconception arises out of an
interpretation error: Engels’ criticism is here not levelled at the second law of thermodynamics itself, but
at the two controversial hypotheses that were commonly extrapolated from this law: the steady cooling
down of the earth and the heat death of the universe (Foster & Burkett, 2008). The fact that this is directed
towards cosmology is indicated by his criticism of Laplace’s old hypothesis and the implicit meaning
of world, where he refers not only to the earth, but the universe in general. This is also the case in
criticism which states that Engels “thought (together with other contemporary authors) that ways would
be found to re-use the heat irradiated into space” (Martínez-Alier, 2007, p. 276). This is directed
towards a passage in Dialectics of Nature, where Engels states that “the question is only finally solved

57
when it has been shown how the heat radiated into universal space becomes utilisable again” (Marx &
Engels, 1987a, p. 562), addressing loss of motion in physics. It is exactly because this discussion is
situated in physics that Martínez-Alier erroneously opens Engels up for ridicule. A term such as
utilizable energy (which Martínez-Alier wrongfully reconstructs as re-useable, which conceals its
original meaning) as employed in physics refers to physical forces independent of human action, it is
seen as the potential of a system to do work. Nowhere in his work does Engels suggest that there might
exist a possibility to put the irradiated energy into space back into use by humans (Foster & Burkett,
2008).

Other scholars criticise Engels’ approach to Podolinsky, as he supposedly sharply criticised the
introduction of environmentalism into economics. This is not based on a misrepresentation of Engels’
work, but on a limited knowledge about economics. It is correct that in his letter to Marx, Engels states
that “All the economic conclusions he draws […] are wrong” and that “to express economic conditions
in terms of physical measures is, in my view, a sheer impossibility”, stating that Podolinsky “has
confused the physical with the economic” (Marx & Engels, 1992, pp. 411–412). To assume that Engels
here does away with the second law of entropy is ridiculous, as recent publications have demonstrated
that “Podolinsky’s perfect-human-machine model of ecological economics was fundamentally flawed
from the standpoint of thermodynamics itself” and that “Podolinsky’s energy calculations in agriculture
were crude and incomplete (as Engels noted), leaving out the energy associated with fertilizers, fossil
fuels, and even direct sunlight” (Foster & Burkett, 2008, p. 4), something that Marx, Engels and other
materialist thinkers at that time were well aware of.

The misrepresentation of Marxism by scholars in the 20th and 21st century and the attack by ecological
economics on classical Marxism based on the presumed rejection of the second law of thermodynamics
by Engels has tragically reduced the possibility of constructing a unified theoretical framework to tackle
the ecological question. After analysing the differences and, more importantly, similarities between both
academic fields, there are some aspects where both academic fields can benefit from collaborating.

A first interesting commonality concerns the value issue. The incommensurability of value is something
that is touched upon by both Marxism and ecological economics. There is a strong rejection of
monetizing environmental values in both academic fields. The concept of capitalist rents, as laid out by

58
Marx, states that exchange-values can be attributed to valueless and scarce but monopolizable natural
conditions (Burkett, 2014). In his Grundrisse, he very eloquently states that “because money is the
general equivalent, the general power of purchasing, everything can be bought, everything may be
transformed into money. […] With that the individual is posited, as such, as lord of all things. There are
no absolute values, since, for money, value as such is relative. There is nothing inalienable, since
everything alienable for money. There is no higher or holier, since everything appropriable by money”
(Marx, 1973, pp. 838–839). This seems to be in line with the plurality of values approach of ecological
economics (see above).

A second commonality concerns the criticism uttered against neoclassical economics. In his 2017
chapter Social Ecological Economics, Splash claimed that modern economics have become “divorced
from actual and empirical economic systems and their operations” and that “they make
recommendations for society on the basis of a discipline that has no theory of society” (Splash, 2017, p.
5). Although Marx has often been regarded as a mechanistic, almost Promethean thinker, nothing could
be further from the truth. While pointing out the flaws of capitalism and placing them in a historical
perspective by focusing on the material-productive conditions of society, Marx never lost sight of the
necessary relationship of these conditions to natural history (Foster, 2000). In this regard, contributions
from Marxian theory could help avoid the pitfall of denying the structural role of capital accumulation
and its consequences for human-Nature relationships (Douai, 2017).

The third commonality where both disciplines could find common ground is related to the concept of
unequal ecological exchange. Studies of unequal exchange “seek to measure, in biophysical terms, the
ecological disadvantages that have been systematically imposed on the periphery. In short, in the same
way as unequal economic exchange theory postulated the exchange of more labour for less, unequal
ecological exchange theory is based on the exchange of more ecological “units” for less” (Douai, 2017,
p. 64). Although this topic is still very debated within ecological economics, there have been efforts to
align this theory with Marx’s theory of the metabolic rift. This is interesting for our topic not only
because of the structural power relations that are present in international trade and the position that
agroecological producers occupy within those power relations, but also because of the unequal

59
ecological exchange that occurs between these producers and conventional producers in the Pampa-
region (see below).

In what follows, the ME-EE nexus will be applied to the findings from the interviews in the Pampa-
region. The different topics that arose from the interviews will be touched upon and analysed to see
which lessons can be drawn for the development of the agroecological paradigm. The findings are
divided into four categories depending on different dimensions: ecological, economic, social and
political aspects are all touched upon. The objective is twofold in this analysis, namely to 1) determine
how the agroecological paradigm as such is comparable and reconcilable with the ME-EE nexus; and
2) what that means for the development of this paradigm. This analysis will also entail why, according
to actors within the agroecological paradigm, the conventional agricultural model does not adequately
respond to the demands in society, which are ecological, economic, social and political.

60
6 Results and Discussion

6.1 Ecological Analysis


6.1.1 Soil

To start the analysis of agroecology with a discussion about the ecological aspects of it and, more
specifically, about soil, seems fitting when considering the emphasis that Marx put on soil science and
the interest and admiration he fostered for scientists like Justus von Liebig. The issue of soil came up
often in the interviews, which should come as no surprise as soil is the foundation of agriculture. A very
popular saying amongst agroecological producers in the Pampa-region is “suelo sano, alimentos sanos”,
which can be loosely translated into “a healthy soil renders healthy products”.

The Pampa-region has, despite being the most fertile and productive region in Argentina, caused
concern about soil quality in the last decades. Several reports (Buschiazzo et al., 2009; Galantini & Suñer,
2008; Sartorato, 2006) have indicated that soil quality in the humid Pampa has been decreasing and that
this is very likely a result of exhaustion through a combination of application of fertilizers, pesticides
and the use of monocultures. In interviews with local farmers and academics, most compared the status
of the soil to “corcho” (cork) because of the inability of the soil of absorbing water. With the Pampas
being the region where farmers mainly focus on soy monocultures, many agroecological farmers
claimed that a lack of crop rotation and a practice of overburdening the soil lies at the foundation of this
inability. Floodings of the Pampean regions are (partially) a consequence of this that have disastrous
effects on ecosystems and production. Soy is a crop that does not consume much water and this also
contributes to the floodings that seemingly get worse with climate change (Carta, 2017).

The alarming rate of soil deterioration does not go unnoticed by the Argentinean authorities. In a 2011
report, the Ministry of Agriculture wrote that (translated from Spanish) “the specialization system with
high productivity [referring to cereal monocultures] has negatively affected the soil and environment
due to their use of contaminating inputs – fertilizers, pesticides, fossil fuels, etc. – and an increasing
propensity to erosion in a lot of marginalised areas has manifested itself” (IFAD, 2011, p. 87). They
further write of the alarming rate of deforestation in the Pampa-region and the eutrophication of lakes in
the region and the consequential disappearance of fish and an overall reduction in biodiversity as a result

61
of excessive concentrations of fertilizers and pesticides. Species of birds have disappeared from the
region and this has interrupted the trophic (food) chain, also leading to a reduction in biodiversity.
Referring to soil quality, they state that “the change in the use of the soil has potentially affected the
cycle of floodings and droughts” (IFAD, 2011, p. 87). The capacity of the ecosystems to soften climatic
cycles has been diminished through the reduction in pastures in favour of annual crops with little
application of organic matter (IFAD, 2011). There is thus something to say for the concern of
agroecological farmers in the region.

The effect of cultivation on natural forests in the Pampa-region shown in Figure 6.1.

Figure 6.1 Soil Use in the Pampa-region: annual cultivation (a), natural forests (b) and
natural pastures (c)

a b c

Source: IFAD (2011)

The Human-nature relationship that Marx posited at the foundations of materialist theory and the
metabolic rift that afterwards occurred could help interpreting this problem. The disregard of ecosystems
and climatic conditions by producers of soy monocultures was criticized by agroecological farmers and
an integrated crop rotation with biodynamic methods which consider surrounding ecosystems are a
possible alternative to the conventional cultivation methods. This could help mitigate the metabolic rift
posited as such by Marx. [2] even stated that “we need to improve the human-Nature metabolism”,
explicitly using the same wording as Marx used. Although no surveys were taken regarding soil quality,
all farmers that engaged in interviews stated that the soil quality drastically improved after applying
agroecological methods.

62
6.1.2 Biodiversity

Although it did not came up as many times as other pressing topics like soil quality or health benefits
related to agroecology, biodiversity and respect for Nature lie at the foundation of the choice to convert
to agroecology. Respondent [1] stressed the importance of research into functional biodiversity.
Biodiversity conservation does not and cannot begin and end with conserving separate areas where
biodiversity flourishes, while neglecting biodiversity in our socio-economic activities. Mankind is an
inherent and unseparable part of Nature and in order to make strides in mitigating the deterioration of
biodiversity that is occurring not only in Argentina, but in the whole world, it is necessary to start asking
what the position of mankind and agriculture is in the wider conception of biodiversity. Do we not form
part of that same biodiversity? If the answer to this question is yes, then logic obligates us to say that an
artificial separation between an anthropologic and an untouched world does not only make no sense,
but is also impossible to execute and an ontological and epistemological delusion.

Biodiversity is an instrument that permits limiting the amount of (chemical) inputs into agriculture. It is
also very different depending on location (as the word diversity in the word clearly denotes) and thus if
agriculture attempts at utilising biodiversity to limit the amount of external inputs, this inherently leads
to the conclusion that there needs to be a multitude of agricultures depending on surrounding ecosystems
and natural conditions. This leads us to an incompatibility with conventional cropping methods, which
does not take into account differing ecological conditions and instead prescribes the same medicine for
all places. An interesting story from [1] – it is here loosely translated from Spanish –demonstrates this
lack of diversity in the conventional approach to agriculture:

In my formation as an agronomist, I always thought my education was too simplistic, technical


and even boring. When we learned about maize cultivation in class, my first instinct was to ask
the professor “which type of maize are we analysing?”, to which he replied: “maize is maize”.
There is no room for alternatives in classical agronomics and there is a clear limitation to the
capacity of this discipline to analyse diversity and variability – which necessarily raises the
question how biodiversity can be taken into account.

63
This also applies to the importance of family farming. Especially in the Pampa-region, where land is
very expensive and concentrated in the hands of some big landowners who engage it in soy
monocultures, the question arises whether these landowners (who often live in big cities like Buenos
Aires or Rosario) really know what ecological challenges their region is up against. Family farmers live
on the land they work and are thus more aware of the ecosystems that surround his land. This leads to a
different type of efficiency, one that not only takes into account economic costs, but also ecological costs
(it should be clear that this is not an attempt to monetize these costs, which has been criticised above).
Family farmers know what their land needs because they live there.

The ME-EE nexus has a distinctive view on ecological conservation, criticising the extent to which
Nature is conserved in and through capitalism, which in and by itself subjects political, social and
ecological diversities to market dynamics (Büscher, Sullivan, Neves, Igoe, & Brockington, 2012).
Büscher state three main critiques on neoliberalism in their work Towards a Synthesized Critique of
Neoliberal Conservation (2012): 1) internal contradictions of capitalism (as touched upon here); 2)
appropriation and misrepresentation; and 3) the disciplining of dissent. The latter two will be relevant
later in the analysis. It was O’Connor (1998) who first defined a second contradiction of capitalism as
“the way that the combined power of capitalist production relations and productive forces self-destruct
by impairing or destroying rather than reproducing the conditions necessary to their own reproduction”
(O’Connor, 1998, p. 195). Although it is true that this subjugation of natural conditions has been cause
for debate within the school of ecological economics, more research is needed into the topic of economic
valuation and common ground exists for collaboration between these two academic fields.

6.2 Economic Analysis


6.2.1 Value Issue

The value issue, which has been touched upon extensively in the introduction to the theoretical
framework (see above), also plays a role in agroecological cultivation and raises two questions about
economic sustainability of the agroecological paradigm: 1) is agroecology economically rentable? In
other words, does a shift to agroecological production reduce income and is this reduction drastic; and
2) how is value attributed according to the agroecological paradigm?

64
The first question has been the cause of one of the biggest criticisms at the address of advocates for the
agroecological paradigm. It is also a question that, according to these advocates, starts from an erroneous
assumption: that neoclassical economics captures value in monetary terms, and that what exists outside
the boundaries of monetary value as captured by this discipline is by extention of this assumption not
valuable. This phenomenon was repeated by most respondents and is surprisingly similar to the
distinction that Marx made between use-values and exchange-values.

In purely economic terms, most farmers did say that they did not see a decrease in income, but no
surveys were taken and the sample population might not have been representative of the general
population. Still, as one farmer said, “if agroecology were not rentable, we would just go back to using
conventional methods. So yes, economic rentability is crucial, but agroecology is good business”. [1]
mentioned a fundamental negationism, referring to the strategy of many multinationals and academics
who use prejudice to discredit agroecology as a paradigm. He also denied the allegation that agroecology
would be less rentable and stated that “agroecology can be done in the right or the wrong way, just like
conventional agriculture; but there has yet to be a study published that proves that agroecology, across
the board, is not economically rentable or even less rentable than conventional agriculture”. While
there have been plenty of studies or reports concerning the productivity of agroecological farming
compared to conventional agriculture which indicate that it does not necessarily lead to lower yields or
income (Altieri, Funes-Monzote, & Petersen, 2012; Badgley et al., 2007; Christian Aid, 2011; Foresight,
2011; Funes-Monzote et al., 2009, 2011; Pretty, Morison, & Bragg, 2003; Rodríguez-Izquierdo et al.,
2017; Rosset & Altieri, 2017; SANE, 1998), these efforts have been rather unequally distributed and
have not been able to convince most academics of the paradigm. In Argentina, several experiments were
conducted in order to calculate the efficiency of sustainable agriculture. In Mendoza, Van den Bosch et
al. (2018) proved that organic grapes were profitable as an economic activity, but added that this is a
niche activity and has been gaining popularity through a rise in demand for organic wines. INTA (2017)
proved that although conventional methods score higher on economic indicators, agroecological
production methods score higher in all other sustainable indicators (social, nutritional and
environmental). They standardized the scores on a percentual scale, which can be observed in Figure
6.2. Indicators are total gross margin (MBT), gross margin/cost (MB/$), carbon balance (BC), nutrient
balance (BN), environmental impact through application of pesticides and fertilizers (EIQ) and

65
work/person in hours (HsH). 100% is here indicated as being the most efficient, 0% of being the least
efficient. However, as labour/person was here calculated as a proxy for social interaction, agroecology
scores higher on it than conventional agriculture. In terms of efficiency, conventional agriculture thus
scores higher, but this is not reflected in Figure 6.2. The most famous experiment was conducted by the
FAO (2016) in an agroecological project called La Aurora, situated in the Pampa-region in the province
Buenos Aires. Over a period of 11 years, not only the productivity of this project was analysed, but also
its porosity (capacity to absorb water), the potentially mineralizable nitrogen and soluble phosphorus.
La Aurora scored best across the board with the exception of productivity, which was comparable to
other farms in the region. However, it used much less inputs over this period and operated under a much
lower cost than its neighbouring farmers. The results of this are seen in Figure 6.3.

Figure 6.2 Standardized Scale of Comparative Analysis of Sustainability Between


Conventional and Agroecological Agriculture

Conventional Agroecological

100%
80%
60% Conventional Agroecological
40%
MBT 100% 54.30%
MB/$ 100% 38.20%
20% BC 50.57% 100%
0% BN 50.04% 100%
EIQ 9.31% 100%
HsH 40.68% 100%

Source: own image based on INTA (2017)

66
Figure 6.3 Comparison La Aurora and Farms in the Area, Average Production (a),
Environmental Indicators (b)

211.6
3350 3316 250
a b
3300
200
3250
3200 150

85

62.3
61.5
55.8
3150 3116 100

51.8
35
3100

11.7
50
3050
0
3000 Nitrogen Porosity Phosphoru
Average
(mg/Kg) (%) s (ppm)
Production
(Kg/ha) La Aurora 85 61.5 51.8
La Aurora 3116 Farms in the Area 35 55.8 11.7
Farms in Conserved Natural
3316 62.3 211.6
the Area Area

La Aurora Farms in the Area La Aurora Farms in the Area Conserved Natural Area

Source: own images based on Kiehr & Sarandón (2016)

Nonetheless, the academic field concerning the economic viability of agroecology is still divided and
studies display different results, with some reports also pointing out lower productivity levels, especially
in the first years of the transition. It thus rests us to wait for a systematization of the paradigm to make it
ready for implementation and systematic analysis. However, the high degree of diversity and variability
that characterizes agroecology – considering the assumption of a diversity of agricultures, depending on
the location and based on the surrounding ecosystems – makes any effort to do so particularly difficult.

The second question concerns the question of how value should be attributed according to advocates of
the paradigm. This is part of a big debate between agroecological academics and generally follows the
line of the debate that is also present in ecological economics: should a plea be made for monetary
valuation of natural conditions, or should the paradigm adhere to the plurality of values approach that
characterizes ecological economics? The pragmaticism here often clashes with the desire to maintain a
coherent ideological line, with [1] even stating that it is necessary to “recalculate hidden costs”. A quote
by [10] might offer more insight into this:

67
Rentability should not be defined only in its economic, but also in its social and ecological
aspects. Current agribusiness calculates rentability with methodologies that do not take into
account the soil quality and environment (negative externalities). Not using agrochemicals
implies a drastic reduction in costs. Not only economic costs, but social and environmental
costs as well. If the ecosystem services and the health benefits of producing agroecologically
are included, agroecology would be more rentable than conventional agriculture. This is why
further research and more data regarding this issue is needed.

Whether these calculations will assume a monetary character is not clear, but is being debated fiercely
within the agroecological field.

6.2.2 Local Production and Consumption

Another objective which is key to agroecology is the shortening of the value chain. One of the
cornerstones of achieving this is the local production and consumption of produce. [10] states that “we
should orient ourselves towards an alternative commercialization of produce. Through fairs and public
events but also regional markets, producer and consumer could come into contact with each other,
resulting in a product of higher quality and a better price for both”. The interdisciplinary importance of
this objective thus becomes quite clear: not only is local production and consumption important to
mitigate climate change or stimulate healthier consumption, it also opens up the possibility for
communal life to find its place in consumption and even for this to result in cheaper prices through
drastically reducing the commercial chain.

An interesting Marxian viewpoint that helps out with interpreting this objective is what he calls the
separation of town and country. For Marx, “the division of labour inside a nation leads at first to the
separation of industrial and commercial from agricultural labour, and hence to the separation of town
and country and to the conflict of their interest” (Marx, 1998, p. 38). In the Greek polis the focus of
society was more on town while in feudal society the focus was on country, but it is in capitalism that
this antagonism becomes fully developed (Foster, 2000), leading to “the most important division of
material and mental labour” (Marx, 1998, p. 72). To overcome this separation and to abolish this
contradiction is “one of the first conditions of communal life, a condition which again depends on a mass

68
of material premises and which cannot be fulfilled by the mere will, as anyone can see at the first glance.
(These conditions have still to be set forth)” (Marx, 1998, p. 72). This implies local production and
consumption and more importantly, an integrated system of material and cultural exchange, which
agroecology puts a lot of emphasis on.

There is an ecological side to this too. Marx argued that international trade in food made the problem of
alienation of the constituent elements of the soil much worse. In England, exhaustion from the soil
reached problematic levels to the point that at a certain moment English field were manured with
Peruvian guano, which was imported weekly – leading Marx to conclude that capitalist large-scale
agriculture prevents truly rational application of the new science of soil management (Marx, 1976a, p.
348). Although soil management and the application of chemical fertilizers has partially resolved this
problem of international trade, soil exhaustion is a big problem which the application of these fertilizers
are partially responsible for and it has brought with it a new problem: the effect on biodiversity.

Organizations which subscribe to the agroecological paradigm, like La Via Campesina (which is the
biggest international organization defending the rights of small-scale farmholders) have claimed that
they are inspired by the division between town and country as put forth by Marx and Engels (Via
Campesina, 2017). It is thus not far-fetched to imagine agroecology as a paradigm developing further
according to Marxian theory and to draw similarities between both fields concerning local food
production and consumption.

6.3 Social Analysis


One of the most remarkable things about the paradigm of agroecology as such is its focus on social
aspects. Gender equality, horizontality and a conception of radical democracy are all key to the
development of the paradigm. The question is then where this comes from and what separates this
paradigm from conventional agriculture as such? [1] explains this social aspect as follows:

The importance of social aspects in agroecology comes from the notion of an agro-ecosystem
as a system altered by human beings. This means that one cannot forget about the human
aspect in agriculture. Systems need social sustainability as well. It is fundamental. What

69
agroecology is trying to do is reconstruct what has been destroyed by the conventional
paradigm. But that is difficult to implement in practice.

This not only reminds of the phrase of Marx concerning labour, that “labour is, first of all, a process
between man and nature, a process by which man, through his own actions, mediates, regulates and
controls the metabolism between himself and nature”, which has been disrupted by the metabolic rift
and thus refers to the social metabolism as Marx stated; but it also refers to the perception of
sustainability in the field of ecological economics, where social sustainability encapsulates economic
sustainability instead of those two existing separately.

Likewise, an agroecological producer is embedded in his/her social environment and this has strong
consequences for the interactions that this producer engages in. [2] stated that “it is important to create
awareness with customers”, [6] that “the difference with our produce is that it is healthy and that we
know who we sell this to”, and [4] that “there is a side to agroecology which values social aspects in a
different way than conventional extractivist agriculture through the way it attributes value”. This last
quote goes back to the division of use-values and exchange-values and shows how this distinction has
fundamental consequences. Not only does agroecology encapsulate economic values – which are
important still, it is important not to idealize agroecological farmers beyond limitations, if it would not
be rentable then most would not engage in agroecology – but also social and ecological values.

In interviews with people, health often came up as an important factor in their decision to get involved
in agroecology. Some farmers did this out of past health experiences dealing with diseases in the family,
some because of general experiences they had with the application of chemical pesticides, “we once fed
water with chemical pesticides to our horse by accident. It died a bit later. This is what is awaiting us,
the only difference being that we are dying slowly. The horse was better off” [2], and some because of
their conviction that these pesticides are harmful to both humans and Nature. In any case, it becomes
clear that there is an educational side to agroecology linked to creating awareness, which separates it
from conventional agriculture and which contributes to the social identity of the field.

Another social aspect considers the issue of poverty. On the website of Pro Huerta, a branch of INTA
which works locally and helps smallholder farmers develop their own orchards, one of the main
motivations is the combating of poverty. The first sentence in their 2011 annual report states that “Pro

70
Huerta is directed to the population who are facing poverty issues” (Pro Huerta, 2011) and this is
reflected in the interviews that were conducted with agroecological farmers. [7] stated that “urban
agriculture allows for relatively quick results and ensures that those who are poor and need solutions
more urgently are helped in an effective way”, while [8] stated that “the work in a collective, without a
boss, dignifies the work of the farmer and teaches him to work the land. This is crucial, in our perspective,
to combat poverty in the region”. Agroecology has an important social foundation, with some projects
even having their own form of social security systems. The focus on poverty thus should not be a
surprise as it forms part of the identity of agroecology as a paradigm.

Gender equality is one of the most prominent objectives of the agroecological movement and this serves
as a constant reminder of the position of women in agriculture. A big reason why agriculture is
underperforming in some developing countries relates to the fact that women do not have the access to
resources or opportunities to be more productive (FAO, 2018a). In their 2018 report, the FAO stated
that many of the inefficiencies leading to losses or wastes in the food value chain could be mitigated by
a more gender equal distribution of access to resources. This means that regions or countries with high
degrees of gender inequality will also expect to see higher losses in potential efficiencies over the whole
food value chain (FAO, 2018b). The agroecological movement strongly supports the feminist
movement in Argentina (Ni Una Menos), of which the large amount of posters in the research center of
INTA served as proof. Not only the position of women in society is important, the actual interaction
between men and women is taken into account in many agroecological projects in order to generate a
more inclusive atmosphere. It is not uncommon for members of agroecological projects to use the
female pronoun to denote a group of (mixed gender) people, for which in Spanish normally the male
pronoun is used. Inclusiveness is crucial to working together and to make sure everybody has access to
the resources they need in order to be productive. Agroecology as a paradigm has understood this very
well.

A last social aspect concerns the horizontality of the organization. Most agroecological projects have a
sense of radical democracy in the way their structures are organized. They are organized in assemblies
(for bigger projects), where the farmers come together and discuss the difficulties they have. Decisions
for the project are made with consensus and they sell their produce together. For smaller projects they

71
discuss responsibilities between themselves and this thus constitutes a very horizontal decision-making
process. [2] states that “Our project consists of 17 families and our decisions are made in an assembly.
It is important to be organized well so the work is distributed evenly between everyone”, while the
smaller project [6] states that “we have no bosses here. We hold small meetings when important
decisions need to be made and our responsabilities rotate to make it fair for everyone”. Agroecological
projects often consist of horizontal forms of cooperation, as this case shows. The sale of produce in
cooperation with other farmers in the region requires a lot of coordination, but also reaps rewards as
sales are done on a bigger scale. There is something to be said for the democratization of economic
decision-making in agriculture. In order to do this, however, it is important to have control over the
means of production, which in this case is land. The separation of farmers from the land and the conflict
with landowners that ensues from this separation is lined out further in the analysis of political conflicts.

6.4 Political and Conflict Analysis


6.4.1 Producer-State Interactions and Perspectives

The state plays an important role in mitigating and responding to certain conflicts that might appear.
When the state foregoes not only this responsibility, but expected responsibilities which it might not
engage in but is expected to, conflicts occur. These are lined out in this section.

A first clear conflict occurs between conventional and agroecological producers. Due to the surface of
agricultural land in the Pampa-region, a common practice is the aerial fumigation of crops with
glyphosate. The consequences of this chemical treatment of (predominantly) soy monocultures have
been cause for concern in the medical community of the Pampa-region. Verzeñassi, a medical
researcher in the National University of Rosario found that after analysing 90,001 people, cancer and
hyperthyroidism instances had quadrupled in the regions where these aerial fumigations were utilized
(Verzeñassi, 2014). In their 2017 article, Arancibia et al. (2017) stated (here translated from Spanish)
that “while hegemonic academic sectors actively promote the expansion of this technology [referring
to GMOs], the complaints about the incidence of grave health problems in rural populations exposed
to agrochemicals, which this model uses extensively, have multiplied” (Arancibia et al., 2017, p. 105).
Not only cancer and hyperthyroidism cases increased, a study by Paganelli et al. (2010) showed through
microbiological analysis of xenopus and chicken embryos that glyphosate-based herbicides interfere

72
with key molecular mechanisms regulating early development, leading to congenital malformations
(birth defects). Benítez-Leite et al. (2009) found a statistically significant correlation between living close
to fumigated areas and birth defects. And the problem might be worse than first thought, as traces of
DDT have been found on Antarctica, where nothing is produced, which means that pesticides can travel
farther than previously assumed (IDEP Salud, 2014). Indeed, a report published by medical researcher
Kaczewer in 2002 stated that (translated from Spanish) “today we can affirm with absolute certainty
that every child on the planet is exposed to pesticides from conception, during the gestation and even
during lactation, regardless of the place of birth” (Kaczewer, 2002, p. 1).

Agroecological producers are aware of this danger and this is a main cause for their conflicts with
neighbouring producers. Members of the project [6] stated they had to leave their fields several days a
month to escape the fumigations that go on close to them and [2] asked himself “why do we invest this
much into agroecology if next to us they fumigate that much? The authorities should be controlling this,
it is their responsibility to look after the health of the population”. The main conflict thus arises from the
fumigations that take place closeby (especially in places where parcels are located closely to each other,
like the peri-urban area of La Plata or Rosario) and the fact that the government does not take enough
action to regulate these fumigations.

Another key conflict concerns access to land. Oxfam stated in its 2016 report that the largest 1% of farms
in Argentina occupied 35.93% of all cropland, and that small farms (defined as smaller than 500 ha)
make up 83% of all farms, but only possess 13% of all cropland (Oxfam America, 2016). The Ministry
of Agriculture stated that 2% of farms controlled about 50% of all farmland and that 57% of all farms
only controlled 3%. The Pampa-region is however, the region in Argentina where agrarian plots are
most equally divided. The smallest producers, who cultivate a surface smaller than 100 ha, make up 38%
of the producers but only own 3% of the land, while the largest producers (surface larger than 2500 ha)
make up only 3% of the producers but control 39% of the land. Since 1988, the inequality has grown
however, with around 54,000 producers disappearing in the 90s alone. This is a process that still
continues (IFAD, 2011). From the interviews, it becomes clear that the biggest problems occur in the
peri-urban areas, where land is very expensive. [4] claimed that “access to land is fundamental. This
region [La Plata] is peri-urban and is very expensive. A big part of the workers here are immigrants

73
from Bolivia and they rent land here. The fact that their land is so expensive inhibits them from taking
too much risks”, [2] stated that “the dependence on the land is real. Agroecological producers who have
been working the land for years, improving the soil through biodynamic practices, are expelled from
the land when the price increases, which happens quite often”, and [8] claimed that “there is no such
thing as agroecology without taking into account the issue of land. A farmer can apply agroecological
methods to the land, but if that land is not his, it is possible that he will have to look for a new parcel of
land after a certain period of time”. Although the Pampa-region is the most equally distributed region
compared to other regions in Argentina, there are big differences within the region. Urban and peri-
urban regions are the most expensive, while many rural areas in the Pampa-region have more affordable
land. However, the access to public facilities and the lack of paved roads makes it difficult for family
farms to settle in these areas.

Most farmers that were interviewed were of the opinion that the government should intervene in the
relatively unequal land distribution problem: “urban land is very expensive, but the government should
intervene” [7], “we’re asking the government to help us with the problem of landless farmers” [2] and
“without access to land, there is no agroecology, because it is difficult to plan the cultivation of crops if
landowners give out leases for one or two years and the real estate competition drives the prices up.
The access should be facilitated by fiscal plans given out to individual farmers to buy the land” [10].
The government does not enough to facilitate access to land and the biggest victims are the poor, landless
farmers which often are immigrants from neighbouring countries (mostly from Bolivia).

Access to land plays an important role in Marxian theory. As stated above, in Capital Vol. 1 he states
that “the capital-relation presupposes a complete separation between the workers and the ownership of
the conditions for the realization of their labour” (Marx, 1976a, p. 874), forcing them to become wage
labourers. This separation was the result of what Marx called primitive accumulation. Accumulation in
a capitalist system presupposes surplus-value, surplus-value presupposes capitalist production and
capitalist production presupposes the availability of considerable masses of capital and labour force in
the hands of commodity producers (Marx, 1976a). This situation thus presupposes an initial, original
accumulation, which is not the result of the capitalist production process but its point of departure. In
Capital, he states that “the expropriation of the agricultural producer, of the peasant, from the soil is the

74
basis of the whole process. The history of this expropriation assumes different aspects in different
countries, and runs through its various phases in different orders of succession, and at different
historical epochs” (Marx, 1976a, p. 876). This accumulation was not a result of gradual historical
processes, but of violent expropriation, resulting in separating the producer from his/her means of
production of subsistence, as stated above. The case of colonialism which ensued in the expropriation
of indigenous inhabitants from their means of production and turning them into slaves for large
landowners in Latin America is the prime example of this. The slaves disappeared, the landowners did
not.

David Harvey, a Marxian geographer here introduced the term accumulation by dispossession, stating
that this accumulation did not take place in a series of events which made capitalism possible and which
is rendered useless after that condition has been fulfilled, but rather that capitalism is a system that
reproduces this primitive accumulation by absorbing elements into the system that were not present
there before. It reproduces expropriation of the means of production and privatization of the commons.
In his work The New Imperialism, he states that

The escalating depletion of the global environmental commons (land, air, water) and
proliferating habitat degradations that preclude anything but capital-intensive modes of
agricultural production have likewise resulted from the wholesale commodification of nature
in all its forms. […] What accumulation by dispossession does is to release a set of assets
(including labour power) at very low (and in some instances zero) cost. Overaccumulated
capital can seize hold of such assets and immediately turn them to profitable use. In the case
of primitive accumulation as Marx described it, this entailed taking land, say, enclosing it, and
expelling a resident population to create a landless proletariat, and then releasing the land
into the privatized mainstream of capital accumulation. (Harvey, 2003, pp. 148–149)

Note that this accumulation does not necessarily mean that a resident population is physically expelled
from its land, but that this also happens through economic means. Akhram-Lodi describes the loss of
land here as the expropriation of producers from their assets through the normal, everyday workings of
imperfect markets (Akram-Lodhi, 2012). The accumulated (or privatized) goods by capitalism undergo
a necessary transformation involving wide-ranging structural, institutional, and legal changes. Examples

75
of this are the massive privatizations by foreign multinationals in the Soviet-Union after its collapse or
the turn towards state-orchestrated capitalism in China. Or, the issue of land that many farmers are faced
with in the Pampa-region. This aspect of private accumulation (or, accumulation by dispossession) has
as a result that the character of labour differs between large-scale, conventional entrepreneurship and
small-scale family farming. The division of labour differs as labourers who have no access to land and
are thus separated from the means of production, consequently being forced to sell their labour to either
larger agricultural companies or to companies in urban areas. Family farms who do have access to land
are owners of the means of production and thus are involved in a different division of labour. This
division becomes under pressure, however, as speculation intensifies (especially in peri-urban areas) and
land prices soar. The increasing concentration of land in the Pampa-region in the last 20 years (see above)
is a great example of this. Small-scale farmers are under pressure and access to land is crucial for them
to maintain their productive capacities. Instead, multinational companies are buying up land in the region
and this complicates the ability of small-scale agroecological farms to survive.

The relation of these farmers to the government is quite complex and intricate. Small-scale farmers are
organized in syndicates and social movements, of which the Unión de Trabajadores de la Tierra (UTT)
and the Movimiento de Trabajadores Excluidos (MTE) are great examples. The UTT is especially
interesting because of their short history but rapid growth on one hand, and their interactions with the
government on the other. Originally founded to put pressure on governing bodies to implement
protectory measures for small-scale farmers, the union has grown to more than 15,000 members in only
eight years. It is a pressing group through both political and extra-political means. They hold close
contacts with the municipalities, with SENASA, the National Institute of Agrarian Technology (INTA)
and with the ministry of development; but they also engage in what is known as the verdurazos, the
organization of a vegetable market on central squares in Buenos Aires where produce is sold at often
three or four times less than the regular shelf price. These markets are organized irregularly, sometimes
once a week, sometimes once a month. But they are the main instrument the UTT uses to put pressure
on conventional producers and more so on the government to implement protectory laws for small-scale
farmers. Despite the contacts that the UTT has with governmental bodies, the police have generally
reacted to these verdurazos by oppression rather than appeasement. On the 15th of February, one day
after the interview with the UTT representative in La Plata [3], the police brutally cracked down on a

76
verdurazo that took place on Plaza Constitución in Buenos Aires. The municipality stated that the
vendors had not solicited permission to organize the verdurazo and police responded by removing the
vendors by force from the square. A photograph of an old woman picking up vegetables while flanked
by police officers went viral after this incident.

This event brings us to the second critique on neoliberalism by Büscher et al. (2012): disciplining dissent.
In their article, they state that in a society dominated by the neoliberal paradigm, “dissenting voices are
either denied or put in a particular light” and that “dissenting voices are […] rather marginal or often
met with the disciplining force of denial or disregard” (Büscher et al., 2012, p. 22). It should be
mentioned that this disciplining does not always occur through openly violent actions, as some
neoliberal outlets do stimulate dissenting voices as a sort of catharsis without promoting these voices to
be converted into real change (Büscher et al., 2012).

The perspective towards the government and their actions also differs between the diverse actors that
are active in the agroecological paradigm. Workers from INTA had their funding cut by the Macri
administration and were forced to move their
offices to an old University experimental site,
and some of them stated that when applying for
research funding they could not use the word
agroecology in their proposal, as the
government deems it “too political”. One of the
pioneers of urban agriculture in Rosario [7]
stated that the government “refuses to
collaborate with these kind of experiences”,
while [10] stated that there is a “lack of political Source: Ávila (2019)

support to make the agroecological transition


happen”. However, [9] stated that “people who criticize this government seem to forget that it was the
previous administration who did nothing but promote the introduction of GMOs”, while [2] mentioned
that “the current administration put all support for agroecology on hold” but that “politicians in general
come and go without much regard for the producers”.

77
There are thus conflicting perspectives on the role of the government, but it does seem that
agroecological actors do not feel represented by the current government. The consensus is that the
government should do more to facilitate a transition to sustainable agriculture but has failed to do so.
This is what Lenin tackles in his work State and Revolution, stating that governments are not
independent bodies of legislation, but rather the product of class contradictions. The state arises when
people transition from a certain mode of production to another, the transition from primitive
communism into capitalism – through slavery and feudalism – has been accompanied by an evolution
of the state itself. This state has been dominated by the exploiting class throughout history, which have
created bodies of power in order to safeguard the privileges that these modes of production grant them
(police, prisons, etc.). These bodies and institutions are characterized by certain degrees of power and
violence. This leads Lenin to conclude that the state is thus not an independent and neutral body, standing
above the class contradiction that characterizes the modes of productions. The state cannot be readily
used by all classes in society but is instead a reflection of the ruling class, whose interests it defends
because it has been structurally and historically designed to do so (Lenin, 2014). The police repression
of the verdurazos is a great example of this, as many respondents expressed their conviction that the
state should be protecting its citizens (also through healthy food consumption) and that the current
government does not go far enough to do so.

6.4.2 Agroecology as a Movement

It is impossible to separate agroecology from its social foundations, which is one of the aspects that
separate it from organic agriculture as a paradigm. Agroecology itself has its foundations in social
thought and movements which emerged in opposition to early agricultural industrialization and which
has developed dialectically against this industrialization (Sevilla Guzmán & Woodgate, 2013). To
approach agroecology only in agronomic terms would mean that a significant aspect of it, which
contributes to its ontological reality, is ignored.

In their 2013 paper, Sevilla Guzmán & Woodgate state that

If the science of agroecology is separated from the agrarian social thought and movements
with which it has grown up, we would argue that its transformative potential will be lost and

78
agroecology will become just another instrumental discipline in the continuing saga of
capitalism’s struggle to overcome its own internal contradictions. (Sevilla Guzmán &
Woodgate, 2013, pp. 42–43)

This aspect of agroecology becomes clear in the case of Malvinas Argentinas. This village of
aproximately 8000 inhabitants was the center of agroecological resistance against Monsanto between
2012 and 2016. After Monsanto announced its plans to construct a new plant in Malvinas Argentinas,
the citizens started to organise in an Assembly (Asamblea Lucha por la Vida) and blocked access to the
construction site. Monsanto decided to give up after four years and started deconstructing the site in
2016. [8] stated that

We started organising through small assemblies and through contact with our neighbours. But
that was not all, several social organisations from outside got involved and helped us out with
our struggle. We all knew what Monsanto was here for. So after an event that we organised,
Spring Without Monsanto, we decided to take the plant. That is when the encampment started.

The fight against Monsanto was not only waged on the construction site, as Monsanto entered into an
agreement with the University of Cordoba at the same time the plant was constructed. [8], at that time a
student, stated that “we organized assemblies through student organizations and with several professors,
until the dean organized a public debate to discuss the possibility of entering into an agreement with
Monsanto. We won that debate, and with that the agreement was annulled”. Agroecology as a
cultivation method is not just a scientific method, but exists within a social and political structure and
receives its meaning in the dialectic struggle with this structure. It is thus impossible to understand this
paradigm without understanding its social foundations. This implies an inherent indivisibility between
agroecology as a science, movement or practice (Sevilla Guzmán & Woodgate, 2013).

6.5 Capitalism and the Development of Agroecology


The most controversial topic in the conducted interviews concerned the question whether agroecology
as a paradigm, with its focus on horizontal decision-making processes, local production and
consumption, and the focus on indigenous crops and an increased biodiversity, is compatible with
globalized capitalism. This is where agroecology as a paradigm has not yet given a concise answer and

79
it seems that a lot of debate concerning this topic is still ongoing. There are two main ideological camps
within the paradigm regarding this question: a pragmatic, moderate camp and a more radical camp.

The more pragmatic actors in the agroecological paradigm tend to focus on the concept of “buen vivir”,
a term that refers to “good living”, where emotional and social aspects are valued more than they are
today. This approach is quite similar to Amartya Sen’s capability approach, where issues like health,
freedom and education are included factors in economics. These agroecological actors seem convinced
that agroecology will never reach its full potential under capitalism, but that it might help limit its
excesses. As [1] stated that “we should think about other forms of living, with less emphasis in economic
growth and consumption, but on a way of life which is more focused on well-being”. They tend to agree
that capitalism has its flaws, but that there is no other convincing economic paradigm out there. A way
of focusing on well-being in this case involves valueing social aspects, which are not commonly
captured by monetary valuation systems. In other words, they want to reduce pure economic growth (as
proposed by ecological economics) by including social and ecological use-values and giving them an
exchange-value (as described in Marxian theory).

The more radical camp does not believe it is possible for agroecology to develop within the capitalist
paradigm and argues that the two are inherently impossible to combine. This camp criticizes the
tendency of capitalism to privatize common goods and to disregard social and cultural aspects of society
which are present in agroecology. In their eyes, the development of capitalism in agriculture is what
agroecology struggles against and thus goes firmly against the social aspect of this paradigm. [4] stated
that:

Capitalism is not compatible with agroecology, but meanwhile we have to keep struggling. In
an ideal situation I would tell you that through redistribution of money capitalism would be
compatible with agroecology, or I would tell you that globalised capitalism does not
discriminate against the poorer countries. But we cannot wait for capitalism to get demolished
first, there are many ways in which we can and should fight to counteract it.

It is clear that these two camps have in common that they think capitalism has its flaws, but the
conception of agroecology within this system is what divides them into opposing sides. The acceptance
of capitalism as a paradigm for agroecology to develop within both offers perspective and risks, with

80
the biggest risk, according to most respondents, being coopted by the system. This is why the perspective
towards capitalism is that important.

The dialectics of agroecology with capitalism passes through three stages: 1) negation; 2) confrontation;
and 3) co-optation. This is not a step-for-step process but a collection of processes that happen at the
same time. Negation refers to the denial of agroecology as a valid alternative. Confrontation refers to the
active combating of this paradigm. This occurs when agroecology has a large foundation of actors who
actively manifest their support for the paradigm. The third step is co-optation. This happens when the
movement loses its “revolutionary” (in terms of struggling against the structures that constrain it)
potential and becomes integrated into the system. A perfect example of this is organic agriculture. It is
deemed to be a cultivation method which takes into account the environment, while in 2017 52.9% of
all organic produce from Argentina was exported to the United States (SENASA, 2018). One of the
biggest fears of agroecological actors in the Pampa-region is “becoming the next organic agriculture”
[4]. One of the ways in which SOCLA attempts to circumvent this challenge is to design a certification
method of agroecological events, where they will support certain events through certification if they
deem these events to represent real agroecology.

This relates to the third critique on neoliberalism by Büscher et al. (2012): appropriation and
misrepresentation. In their critique, the authors refer to neoliberal conservation of the environment,
stating that “[…] neoliberal conservation intertwines propaganda, marketing, and governmentality to
open up new conservation spaces for capitalist expansion while providing the marketed appearance
that this trajectory bears no contradiction with ecological integrity or social equity” (Büscher et al.,
2012, p. 18). The appeared contradiction of conservation on one hand, creating areas of pristine
wilderness (which is an illusion) while promoting capitalist expansion in other areas of the world not
only neglects the human-Nature metabolism as stated by Marx, it exacerbates its divisions and the
resulting separation of town and country. Indeed, a recent study by Hallman et al. (2017) showed a 75%
decrease in insect biomass in protected areas over a period of 27 years. This mechanism is repeated
when applying the same logic to the co-optation of agroecological or organic agriculture and its
introduction into the market. On one hand, the introduction of organic agriculture in the global market
promotes capitalist expansion, while the introduction of regulations and certification promotes the idea

81
of harmony with the environment, which is openly propagated. The focus on human-Nature
relationships is moved to the background and it often ceases to play an active role in the cultivation of
organic crops. Indeed, several studies have shown that organic agriculture often poses a threat to its
surrounding ecosystems and that in some cases, it is worse for the environment than conventional
agriculture (Clark & Tilman, 2017; Gabriel, Sait, Kunin, & Benton, 2013; Landquist, Nordborg, &
Hornborg, 2016; Ramankutty & Rhemtulla, 2012; Ritchie, 2017).

The co-optation into capitalism is a valid concern, but here another question arises: by giving exchange-
values to social or ecological use-values, or in other words, by giving a monetary value to social or
ecological aspects, does this not facilitate an integration into the capitalist framework? Here we conclude
with the same remark that the theoretical framework (ME-EE) has extensively touched upon:

[…] the insistence on money valuation clearly makes less visible the biological and ecological
importance of Nature, and also livelihood and cultural values. The beauty and sacredness of
mountains such as the Niyamgiri Hills in Odisha might seem negligible when compared to the
very large money value of their bauxite deposits. The mountains are better defended outside
money valuation. (Martínez-Alier & Muradian, 2015, p. 10)

Though actors of the agroecological paradigm seem to be aware of its biggest challenges, caution is
warranted so that they not dig their own grave in focusing on the monetary valuation of social and
ecological aspects of their environment.

Finally, the question arises how agroecology is supposed to develop within the confines of globalised
capitalism. The main answer that respondents gave is that it is important to focus on education and
research. It is of crucial importance to keep evolving as a discipline and the formation of scholars
contributes to this. Also, the objective of scholars in agroecology is not just to exist next to the
conventional paradigm, but to replace it. As [1] states:

Do we need to create faculties of agroecology or teach agroecology in agronomy? It should


be the latter, we need a paradigm shift. Agroecology will not coexist next to conventional
agriculture as a niche, as a variation in the market. What is called agronomy today, will be

82
called agroecology in the future. The problem is the model. It is important to solve the problems,
not the symptoms. The solution is to focus on the productive structure.

This quote brings us to another Marxian concept that might help understand the conflict that agroecology
faces with capitalism: dual power by Lenin (1974). In his April Theses, he states that “the highly
remarkable feature of our revolution is that it has brought about a dual power. This fact must be grasped
first and foremost: unless it is understood, we cannot advance” (Lenin, 1974, p. 38), referring to the
surge of a new government alongside the existing one. A government which represented another class
in society, that of the proletariat and the soldiers. He refers to this dual power as a transitional phase,
because “there can be no dual power in a state” (Lenin, 1974, p. 146). This referred to the Soviets that
grew during the different steps of the Russian Revolution in 1917, as one dominant paradigm of power
was replaced by another. Although there is certainly no proletarian revolution occurring in Argentina,
the framework of dual power can be applied to the struggle between the conventional and the
agroecological paradigm. The concept of long-lasting patterns of a fragmented state and a continuous
tension between two polar opposites as expressed by dual power can be helpful to think of the
Argentinean context. A great example of this is the question of agrarian reform in Latin America, which
was implemented in several countries during the 1960s and 1970s, but lost their legitimacy in the 1980s
and 1990s as neoliberal approaches gained support. Agrarian reform was largely neglected, except to
the extent that they could be organized through the market (Enríquez, 2013). Indeed, it was Enríquez
who cleverly pointed to the relevance of dual power in the context of agroecology, although erroneously
contributing the concept to Trotsky instead of Lenin, stating (about the Venezuelan case) that

Here is where Trotsky’s (1957) concept of dual power is insightful. He spoke of certain parts
of government continuing to be in the hands of the previously predominant class, at the same
time as the newly powerful class gains influence in others. (Enríquez, 2013, p. 633)

Even within INTA, there are branches which are deeply invested in agroecology, while other branches
are involved and even cooperate with multinationals like Monsanto. Both paradigms are involved in a
struggle in different institutions, which by themselves or through branches which they consist of,
propagate and aid the development of one of both paradigms. This struggle will most likely continue
until one of both replaces the other because, as stated above, there can be no dual power in a state. This

83
might occur through co-optation, where agroecology is absorbed by the dominant capitalist framework
and undergoes a transformation like organic agriculture went through, or through the replacement of the
conventional agricultural model by agroecology. It seems like agroecological actors are aware of this
struggle, as “agroecology will not coexist next to conventional agriculture as a niche, as a variation in
the market” [1] (see above). It is unclear whether the replacement of conventional agriculture will lead
to co-optation by the capitalist framework anyway and if this is even avoidable. This question will
determine whether agroecology is truly compatible with a capitalist framework, or not.

84
7 Conclusion
The quantitative analysis revealed that the transition to organic agriculture, and by extension
agroecological agriculture, strongly depends on the organic price premium on one hand, and on farm
size on the other. A slight percentual increase in the organic premium has strong effects on the decision
to make the transition and this is also reflected in the evolution of the amount of certified organic
producers in the last twenty years. When analysing how these numbers have evolved in the Pampa-
region, a very volatile pattern shows up, which tends to confirm the findings of the dynamic
programming model. The difficulties of transitioning to sustainable agriculture are perfectly
demonstrated by this.

Still, the quantitative model is not sufficient by itself to grasp the complex social, ecological and political
realities of transitioning to sustainable agriculture. Marxian ecology was combined with ecological
economics to provide a theoretical framework in which to analyse agroecology as a paradigm. Marxian
ecology was chosen for its analysis of not only the capitalist production system, but of man’s position in
and relation to nature. Ecological economics was then chosen to build upon this foundation and analyse
the inconsistencies that characterize current neoclassical economics. Although these two academic
schools differ in some ways, through collaboration they have the potential to give a more consistent
social, economic and ecological critique of agricultural production systems.

After analysing the interviews with agroecological actors in the Pampean region in Argentina, it became
clear that the development of the paradigm is tilted against the social, ecological and political issues of
capitalism in Argentina, but that there is a difference in the conception of this paradigm towards the
dominant capitalist framework that it is struggling against and that strong debate exists whether the
paradigm can continue to develop itself within the confines of this framework. Both camps admit that it
is important to replace the conventional model, but it is unclear if this will go accompanied with co-
optation by the capitalist framework, or if this paradigm will combat the excesses that ensue from it.

More research into this topic is crucial and most agroecological actors are aware of this. Some of the
most important issues that might be of interest for future studies concern: 1) the development of shorter
value chains and how they make agroecological farming more profitable; 2) an attempt at defining

85
agroecology and a deeper comparative look into the trajectories of both organic agriculture and
agroecology; 3) a qualitative assessment about the decision-making processes of agroecological farmers;
4) a standardized study about the productivity and profitability of agroecology; and 5) further
development of integrating the development of the agroecological paradigm into the wider structural
(socio-economic and political) constraints. In this study, the last issue was of importance; but due to the
limited scope of the study and the focus on just one area, many questions remain unanswered.

Agroecology as a discipline exists in three dimensions: as a movement, a science and a practice. The
development of the paradigm against the conventional cultivation method thus depends on the success
of these three dimensions to evolve together. As a movement, it draws from its struggles against
agricultural modernization and the political conflicts that it is engaged in; as a science, it draws from the
formation of scholars and the academic debates that ensue from it; and as a practice it is inspired by both
traditional knowledge systems and scientific advances, as long as they are in harmony with the
surrounding ecosystems. This is what separates agroecology from other production systems and the
possibility to retain its identity will be crucial in determining whether they can replace the conventional
agricultural production methods.

86
87
8 Reference List
Akram-Lodhi, A. H. (2012). Contextualising land grabbing: contemporary land deals, the global
subsistence crisis and the world food system. Canadian Journal of Development Studies/Revue
Canadienne d’études Du Développement, 33(2), 119–142.
https://doi.org/10.1080/02255189.2012.690726

Altieri, M. A., Funes-Monzote, F. R., & Petersen, P. (2012). Agroecologically efficient agricultural
systems for smallholder farmers: Contributions to food sovereignty. Agronomy for Sustainable
Development, 32(1), 1–13. https://doi.org/10.1007/s13593-011-0065-6

Altieri, M. A., Nicholls, C. I., Henao, A., & Lana, M. A. (2015). Agroecology and the design of climate
change-resilient farming systems. Agronomy for Sustainable Development, 35(3), 869–890.
https://doi.org/10.1007/s13593-015-0285-2

Altvater, E. (2004). Is There an Ecological Marxism? In J. D. Schmidt (Ed.), Development studies and
political ecology in a North South perspective (pp. 2–25). DIR. Retrieved from
http://vbn.aau.dk/en/publications/development-studies-and-political-ecology-in-a-north-south-
perspective(0f5d54f0-2bc8-11db-b3d2-000ea68e967b).html

Altvater, E. (2007). A Marxist Ecological Economics. Monthly Review, 58(8), 55. Retrieved from
http://canterbury.summon.serialssolutions.com/2.0.0/link/0/eLvHCXMw3V07T8MwELYQE4
ihvMtL2VGQE9uJPTBUqKhLFyizdbETgZRWqARU_j3nJm7SCP4Aa5w48p3vu8_23ZkQF
t_RsIcJBfpdYK44V44-AxHRGOS5CVhQXNW1nztJr5uLNttn_0HxI5d_s0LtuaRjj2w-
__ijS0bRnqvX8rt3PjAu57C8HZXVF_jQXb8rkPZ2

Amujoyegbe, B., Opabode, J. T., & Olayinka, A. (2007). Effect of organic and inorganic fertilizer on
yield and chlorophyll content of maize (Zea mays L.) and sorghum Sorghum bicolour (L.)
Moench). African Journal of Biotechnology, 6(16), 1869–1873.
https://doi.org/10.5897/AJB2007.000-2278

Arancibia, F., Bocles, I., Massarini, A., & Verzeñassi, D. (2017). Tensiones entre los saberes
académicos y los movimientos sociales en las problemáticas ambientales. Metatheoria – Revista

88
de Filosofía e Historia de La Ciencia, 8(2), 105–123. Retrieved from
http://www.argenbio.org/index.php?action=cultivos&opt=5

Avila, B. (2019). Suicidio. Retrieved June 2, 2019, from http://socompa.info/opinion/suicidio/

Ayres, R. (1998). Eco-thermodynamics: Economics and the Second Law. Ecological Economics, 28,
189–209.

Badgley, C., Moghtader, J., Quintero, E., Zakem, E., Chappell, M. J., Avilés-Vázquez, K., … Perfecto,
I. (2007). Organic agriculture and the global food supply. Renewable Agriculture and Food
Systems, 22(2), 86–108. https://doi.org/10.1017/S1742170507001640

Bartolomé, J. M. G., Casado, G. G., de Molina, M. G., & Guzmán, E. S. (2001). Introducción a la
agroecología como desarrollo rural sostenible. Reis, (95), 213. https://doi.org/10.2307/40184357

Bellamy, A. S., & Ioris, A. A. R. (2017). Addressing the knowledge gaps in agroecology and identifying
guiding principles for transforming conventional agri-food systems. Sustainability (Switzerland),
9(3), 1–17. https://doi.org/10.3390/su9030330

Benítez-Leite, S., Macchil, M. L., & Acosta, M. (2009). Malformaciones congénitas asociadas a
agrotóxicos. Revista de La Sociedad Boliviana de Pediatría, 48(3), 204–217. Retrieved from
http://www.scielo.org.bo/pdf/rbp/v48n3/a13.pdf

Bensaïd, D. (2002). Marx For Our Times. London: Verso.

Bhandari, R. (2015). Marxian Economics: The Oldest Systems Theory Is New Again (or Always?).
Retrieved May 23, 2019, from https://www.ineteconomics.org/perspectives/blog/marxian-
economics-the-oldest-systems-theory-is-new-again-or-always

Bottomore, T. (1992). A Dictionary of Marxist Thought. Wiley. Retrieved from


https://books.google.co.kr/books?id=q4QwNP_K1pYC

Burkett, P. (2006). Marxism and Ecological Economics. Sante Publique (Vol. 28). Leiden: Koninklijke
Brill NV. https://doi.org/10.1017/CBO9781107415324.004

Burkett, P. (2014). Marx and Nature. A Red and Green Perspective. Chicago: Haymarket Books.

89
Büscher, B., Sullivan, S., Neves, K., Igoe, J., & Brockington, D. (2012). Towards a Synthesized Critique
of Neoliberal Biodiversity Conservation. Capitalism Nature Socialism, 23(2), 4–30.
https://doi.org/10.1080/10455752.2012.674149

Buschiazzo, D. E., Panebianco, J., Guevara, G., Rojas, J., Zurita, J., Bran, D., … Hurtado, P. (2009).
INCIDENCIA POTENCIAL DE LA EROSIÓN EÓLICA SOBRE LA DEGRADACIÓN
DEL SUELO Y LA CALIDAD DEL AIRE EN DISTINTAS REGIONES DE LA
ARGENTINA. Ciencia Suelo, 27(2), 255–260. Retrieved from
https://www.suelos.org.ar/publicaciones/vol_27n2/Buschiazzo et al 255-260.pdf

Cabrera, A. (1976). Regiónes Fitogeográficas Argentinas. In Enciclopedia Argentina de Agricultura y


Jardinería. Buenos Aires: Acme.

Carson, R. (1962). Silent Spring. Boston: Houghton Mifflin.

Carta, H. (2017). Lo que pasa en la pampa húmeda no sólo es responsabilidad del clima. Retrieved May
26, 2019, from https://www.lanacion.com.ar/economia/campo/lo-que-pasa-en-la-pampa-
humeda-no-solo-es-responsabilidad-del-clima-nid2063322

Christian Aid. (2011). HEALTHY HARVESTS : The benefits of Sustainable Agriculture in Africa and
Asia. Africa, (September).

Clark, M., & Tilman, D. (2017). Comparative analysis of environmental impacts of agricultural
production systems, agricultural input efficiency, and food choice. Environmental Research
Letters, 12(6), 064016. https://doi.org/10.1088/1748-9326/aa6cd5

Commission of the Central Committee of the C.P.S.U. (1939). History of the Communist Party of the
Soviet Union. New York: International Publishers.

Daly, H., & Farley, J. (2011). Ecological Economics. Washington: Island Press.

De Ponti, T., Rijk, B., & Van Ittersum, M. K. (2012). The crop yield gap between organic and
conventional agriculture. Agricultural Systems, 108, 1–9.
https://doi.org/10.1016/j.agsy.2011.12.004

90
Delbridge, T. A., & King, R. P. (2016). Transitioning to Organic Crop Production: A Dynamic
Programming Approach. Journal of Agricultural and Resource Economics, 41(3), 481–498.
Retrieved from http://www.waeaonline.org/UserFiles/file/JARESeptember20168Delbridge481-
498.pdf

Douai, A. (2017). Marxist Ecology and Ecological Economics. In C. Splash (Ed.), Routledge Handbook
of Ecological Economics (pp. 57–66). New York: Routledge.

Eddington, A. S. (1929). The Nature of the Physical World. Cambridge: Cambridge University Press.

Edmonds, A. W., & Kennedy, T. D. (2017). An Applied Guide to Research Designs Quantitative,
Qualitative, and Mixed Methods (Second). Thousand Oaks: SAGE Publications, Inc.

Engels, F. (1934). The part played by labour in the transition from ape to man. Moscow: Progress
Publishers. Retrieved from https://books.google.co.kr/books?id=GudvAAAAIAAJ

Engels, F. (1935). The Housing Question. International Publishers. Retrieved from


https://books.google.co.kr/books?id=dwsvAAAAMAAJ

Enríquez, L. J. (2013). The paradoxes of Latin America’s “Pink Tide”: Venezuela and the project of
agrarian reform. Journal of Peasant Studies, 40(4), 611–638.
https://doi.org/10.1080/03066150.2012.746959

European Commission. (2019). Organic farming in the EU. A fast growing sector. Brussels. Retrieved
from http://ec.europa.eu/agriculture/markets-and-prices/market-briefs/index_en.htm

FAO. (n.d.). Production_Crops_E_All_Data. Rome.

FAO. (2001). Los Mercados Mundiales de Frutas y Verduras Orgánicas. Rome. Retrieved from
http://www.fao.org/3/y1669s/y1669s00.htm#Contents

FAO. (2003). Trade Reforms and Food Security. Conceptualizing the Linkages. Rome. Retrieved from
http://www.fao.org/3/a-y4671e.pdf

FAO. (2016). The State of Food and Agriculture: Climate Change, Agriculture and Food Security.
Rome. Retrieved from www.fao.org/publications

91
FAO. (2018a). Empowering rural women, empowering agriculture. Rome. Retrieved from
http://www.fao.org/3/CA2678EN/ca2678en.PDF

FAO. (2018b). Gender and food loss in sustainable food value chains: a guiding note. Rome. Retrieved
from http://www.fao.org/3/i8620en/I8620EN.pdf

Fischer-Kowalski, M. (1998). Society’s Metabolism. Journal of Industrial Ecology, 2(1), 61–78.


https://doi.org/10.1162/jiec.1998.2.1.61

Fischer-Kowalski, M., & Haberl, H. (2007). Socioecological Transitions and Global Change:
Trajectories of Social Metabolism and Land Use. Cheltenham: Edward Elgar Publishing.
Retrieved from https://books.google.be/books?id=2NfFI60MMx8C

Foresight. (2011). The Future of Food and Farming : Challenges and choices for global sustainability.
London.

Foster, J. B. (2000). Marx’s Ecology: Materialism and Nature. Monthly Review Press. Retrieved from
https://books.google.co.kr/books?id=5t4VCgAAQBAJ

Foster, J. B. (2015). Marxism and Ecology: Common Fonts of a Great Transition. Retrieved from
https://www.greattransition.org/images/GTI_publications/Foster-Marxism-and-Ecology.pdf

Foster, J. B., & Burkett, P. (2008). Classical Marxism and the Second Law of Thermodynamics.
Organization & Environment, 21(1), 3–37. https://doi.org/10.1177/1086026607313580

Frenay, R. (2006). Pulse: The Coming Age of Systems and Machines Inspired by Living Things. New
York: Farrar, Straus and Giroux.

Funes-Monzote, F. R., Martin, G. J., Suarez, J., Blanco, D., Reyes, F., Cepero, L., … Valle, Y. del.
(2011). Evaluacion inicial de sistemas integrados para la produccion de alimentos y energia en
Cuba. Pastos y Forrajes, 34(4), 445–462.

Funes-Monzote, F. R., Monzote, M., Lantinga, E. A., Ter Braak, C. J. F., Sanchez, J. E., & Van Keulen,
H. (2009). Agro-ecological indicators (AEIs) for dairy and mixed farming systems classification:
Identifying alternatives for the cuban livestock sector. Journal of Sustainable Agriculture, 33(4),

92
435–460. https://doi.org/10.1080/10440040902835118

Funtowicz, S. 0, & Ravetz, J. R. (1994). The worth of a songbird: ecological economics as a post-
normal science. Retrieved from http://www.nusap.net/downloads/funtowiczandravetz1994.pdf

Gabriel, D., Sait, S. M., Kunin, W. E., & Benton, T. G. (2013). Food production vs. biodiversity:
comparing organic and conventional agriculture. Journal of Applied Ecology, 50(2), 355–364.
https://doi.org/10.1111/1365-2664.12035

Galantini, J. A., & Suñer, L. (2008). Las fracciones orgánicas del suelo: análisis en los suelos de la
Argentina. Agriscientia, 25(1), 41–55. Retrieved from
https://revistas.unc.edu.ar/index.php/agris/article/viewFile/2740/2168

Georgescu-Roegen, N. (1994). The Entropy Law and the Economy Problem. In Valuing the Earth:
Economics, Ecology, Ethics (pp. 75–88). Cambridge: MIT Press.

Gonzalez de Molina, M. (2013). Agroecology and politics. how to get sustainability? about the
Necessity for a political agroecology. Agroecology and Sustainable Food Systems, 37(1), 45–59.
https://doi.org/10.1080/10440046.2012.705810

Hallmann, C. A., Sorg, M., Jongejans, E., Siepel, H., Hofland, N., Schwan, H., … de Kroon, H. (2017).
More than 75 percent decline over 27 years in total flying insect biomass in protected areas. PLOS
ONE, 12(10), e0185809. https://doi.org/10.1371/journal.pone.0185809

Harribey, J.-M. (2008). Ecological Marxism or Marxian Political Ecology? In Critical Companion to
Contemporary Marxism (pp. 189–207). Leiden: Koninklijke Brill NV.

Harvey, D. (1993). The Nature of the Environment. The Dialectics of Social and Environmental Change.
Socialist Register, 29, 1–51.

Harvey, D. (2003). The New Imperialism. Oxford: Oxford University Press. Retrieved from
http://eatonak.org/IPE501/downloads/files/New Imperialism.pdf

Herrera, F. F., Domené-Painenao, O., & Cruces, J. M. (2017). Agroecology and Sustainable Food
Systems The history of agroecology in Venezuela: a complex and multifocal process The history

93
of agroecology in Venezuela: a complex and multifocal process.
https://doi.org/10.1080/21683565.2017.1285842

IDEP Salud. (2014). En Argentina se cuadruplicaron los casos de cáncer en las zonas donde se cultiva
soja. Retrieved May 29, 2019, from http://idepsalud.org/en-argentina-se-cuadruplicaron-los-
casos-de-cancer-en-las-zonas-donde-se-cultiva-soja/

IFAD. (2011). La Problemática de la Tierra en Argentina. Rome. Retrieved from www.ifad.org

Ikerd, J. E. (1993). Two related but distinctly different concepts. Small Farm Today (USA), 10(1), 30–
31.

INTA. (2017). Toma de decisiones según criterios : comparación entre agrícultura tradicional y un
módulo agroecológico. Marcos Juarez.

Islam, S. (2013). RETAIL PRICE DIFFERENTIAL BETWEEN ORGANIC AND


CONVENTIONAL FOODS. In Proceedings of ASBBS (Vol. 20). Retrieved from
http://asbbs.org/files/ASBBS2013V1/PDF/I/IslamS(P537-545).pdf

Jansen, K. (2015). The Debate on Food Sovereignty Theory: Agrarian Capitalism, Dispossession and
Agroecology. The Journal of Peasant Studies, 42(1), 213–232.
https://doi.org/10.1080/03066150.2014.945166

Kaczewer, J. (2002). Uso de Agroquímicos en las fumigaciones periurbanas y su efecto nocivo sobre la
salud humana. Buenos Aires.

Kiehr, J. E., & Sarandón, S. J. (2016). Produccion Agroecologica de cereales y carne Bovina en un
Establecimiento Agropecuario Extensivo (650 Has) en el sudeste de la Provincia de Buenos Aires
de la Republica Argentina. El caso de “La Aurora”, una experiencia de 25 años. Buenos Aires.
Retrieved from http://www.fao.org/3/a-be861s.pdf

Kirchmann, H., & Torvaldsson, G. (2000). Challenging targets for future agriculture. European Journal
of Agronomy, 12, 145–161.

Kumar, A., Saini, J., & Chopra, P. (2018). The Comparative Cost and Profit Analysis of Organic and

94
Non-organic Farming Practices in the Mid Himalayan Region. International Journal of
Agriculture, Environment and Biotechnology Citation: IJAEB, 11(4), 633–639.
https://doi.org/10.30954/0974-1712.08.2018.4

La Via Campesina. (1996). The Right to Produce and Access to Land. In World Food Summit
Proceedings (p. 3). Rome. Retrieved from
http://www.acordinternational.org/silo/files/decfoodsov1996.pdf

Lampkin, N. H., & Padel, S. (1994). The Economics of Organic Farming: An International Perspective.
CAB International. Retrieved from https://books.google.be/books?id=joa5AAAAIAAJ

Landquist, B., Nordborg, M., & Hornborg, S. (2016). Litteraturstudie av miljöpåverkan från
konventionellt och ekologiskt producerade livsmedel. Retrieved from
https://www.livsmedelsverket.se/globalassets/publikationsdatabas/rapporter/2016/miljopaverka
n-fran-konventionellt-och-ekologiskt-producerade-livsmedel-nr-2-
2016.pdf?_t_id=1B2M2Y8AsgTpgAmY7PhCfg==&_t_q=ekologiskt&_t_tags=language:sv,si
teid:67f9c486-281d-47

Lenin, V. (1974). Collected Works Volume 24. Moscow: Progress Publishers.

Lenin, V. (2014). State and Revolution. Chicago: Haymarket Books.

Lotter, D. W. (2003). Organic Agriculture. Journal of Sustainable Agriculture, 21(4), 59–128.


https://doi.org/10.1300/J064v21n04

Lucretius. (1969). De Rerum Natura (On the Nature of Things).

MacCormack, H. (1995). Sustainable Agriculture versus Organic Farming. In “What is Sustainable


Agriculture?” Planting the Future: Developing an Agriculture that Sustains Land and
Community. Iowa City: Iowa State University Press.

Madison, M. G. (1997). “Potatoes Made of Oil”: Eugene and Howard Odum and the Origins and Limits
of American Agroecology. Environment and History, 3(2), 209–238.
https://doi.org/10.2307/20723041

95
Martínez-Alier, J. (2007). Social Metabolism and Environmental Conflicts. The Social Register, 43,
273–293.

Martínez-Alier, J., & Muradian, R. (2015). Handbook of Ecological Economics. Cheltenham: Edward
Elgar Publishing.

Marx, K. (1973). Grundrisse: Foundations of the Critique of Political Economy. New York: Vintage
Books.

Marx, K. (1975). Texts on method. Barnes & Noble Books. Retrieved from
https://books.google.co.kr/books?id=88YgAQAAIAAJ

Marx, K. (1976a). Capital: A Critique of Political Economy. Penguin Books Limited. Retrieved from
https://books.google.co.kr/books?id=A7GfucogfCAC

Marx, K. (1976b). Value, Price & Profit. New York: International Publishers.

Marx, K. (1988). Collected Works Volume 30. Karl Marx: 1861-63. New York: International Publishers.

Marx, K. (1991). Capital Vol. 3. Penguin Books Limited. Retrieved from


https://books.google.co.kr/books?id=1MnEAAAAIAAJ

Marx, K. (1992). Capital Vol. 2. Penguin Books Limited.

Marx, K. (1998). The German Ideology. New York: Prometheus Books.

Marx, K., & Engels, F. (1987a). Collected Works Volume 25. Engels. International Publishers.

Marx, K., & Engels, F. (1987b). Collected Works Volume 42. Marx and Engels 1864-1868.
International Publishers.

Marx, K., & Engels, F. (1988). Collected Works Volume 43. Letters 1868-1870. International Publishers.

Marx, K., & Engels, F. (1992). Collected Works Volume 46. Letters 1880-1883. International Publishers.

Marx, K., & Engels, F. (2010). Marx & Engels Collected Works. Volume 50. Letters 1892-95 (Vol. 50).
Lawrence & Wishart.

96
Ministerio de Producción y Trabajo. (n.d.). Estimaciones Agricolas. Buenos Aires.

Miranda, M. J., & Fackler, P. L. (2004). Applied Computational Economics and Finance. MIT Press.
Retrieved from https://books.google.co.kr/books?id=KUYPome1SxwC

National Ministry of Agriculture and Rural Development, C. (2004). La Competitividad de las Cadenas
Agroproductivas en Colombia. Bogotá: IICA Biblioteca Venezuela. Retrieved from
https://books.google.be/books?id=-niU32tEHs0C

National Treasury of Argentina. (2017). Origen provincial de las exportaciones. Buenos Aires.
Retrieved from www.indec.gob.ar/calendario.asp

O’Connor, J. R. (1998). Natural Causes: Essays in Ecological Marxism. Guilford Publications.


Retrieved from https://books.google.be/books?id=4_cYGQ2BafUC

O’Neill, J. (2017). Pluralism and Incommensurability. In C. Splash (Ed.), Routledge Handbook of


Ecological Economics (pp. 227–236). New York: Routledge.

Oxfam America. (2016). Unearthed: Land, power and inequality in Latin America. Boston. Retrieved
from https://www-cdn.oxfam.org/s3fs-public/file_attachments/bp-land-power-inequality-latin-
america-301116-en.pdf

Paavola, J., & Adger, W. N. (2005). Institutional ecological economics.


https://doi.org/10.1016/j.ecolecon.2004.09.017

Paganelli, A., Gnazzo, V., Acosta, H., López, S. L., & Carrasco, A. E. (2010). Glyphosate-Based
Herbicides Produce Teratogenic Effects on Vertebrates by Impairing Retinoic Acid Signaling.
Chemical Research in Toxicology, 23(10), 1586–1595. https://doi.org/10.1021/tx1001749

Pas, T. C. M., & Rees, R. M. (2014). Analysis of Differences in Productivity, Profitability and Soil
Fertility Between Organic and Conventional Cropping Systems in the Tropics and Sub-tropics.
Journal of Integrative Agriculture, 13(10), 2299–2310. https://doi.org/10.1016/s2095-
3119(14)60786-3

Passet, R. (1979). L’Économique et le Vivant. Geneva: Payot. Retrieved from

97
https://books.google.be/books?id=4RFOAQAAIAAJ

Patel, R. (2009). Food sovereignty. Journal of Peasant Studies, 36(3), 663–673.

Paz, J. V. (2011). The Cuban Agrarian Revolution: Achievements and challenges. Estudos Avançados,
25(72). Retrieved from http://www.scielo.br/pdf/ea/v25n72/en_a07v25n72.pdf

Pickett, J. A. (2013). Food security: intensifcation of agriculture is essential, for which current tools must
be defended and new sustainable technologies invented. Food Energy & Security, 2, 167–173.

Ponisio, L. C., M’Gonigle, L. K., Mace, K. C., Palomino, J., de Valpine, P., & Kremen, C. (2014).
Diversification practices reduce organic to conventional yield gap. Proceedings of the Royal
Society B: Biological Sciences, 282(1799), 20141396–20141396.
https://doi.org/10.1098/rspb.2014.1396

Pretty, J. N., Morison, J. I., & Bragg, R. E. (2003). Reducing food poverty by increasing agriculture
sustainability in developing countries, 8809(February 2018). https://doi.org/10.1016/S0167-
8809(02)00087-7

Pro Huerta. (2011). Plan Operativo Anual. Buenos Aires. Retrieved from
https://inta.gob.ar/sites/default/files/script-tmp-poa_2011.pdf

Ramankutty, N., & Rhemtulla, J. (2012). Can intensive farming save nature? Frontiers in Ecology and
the Environment, 10(9), 455–455. https://doi.org/10.1890/1540-9295-10.9.455

Reganold, J. P., & Wachter, J. M. (2016). Organic agriculture in the twenty-first century. Nature Plants,
2, 1–8. https://doi.org/10.1038/nplants.2015.221

Rigby, D., & Cáceres, D. (2001). Organic farming and the sustainability of agricultural systems.
Agricultural Systems, 68(1), 21–40. https://doi.org/10.1016/S0308-521X(00)00060-3

Ritchie, H. (2017). Is organic really better for the environment than conventional agriculture? Retrieved
from https://ourworldindata.org/is-organic-agriculture-better-for-the-environment

Rodríguez-Izquierdo, L., Rodríguez-Jiménez, S. L., Macías-Figueroa, O. L., Benavides-Martell, B.,


Amaya-Martínez, O., Perdomo-Pujol, R., … Miyares-Rodríguez, Y. (2017). Evaluation of food

98
and energy production in animal husbandry farms of Matanzas province, Cuba. Pastos y Forrajes,
40(3), 208–214. Retrieved from http://scielo.sld.cu/pdf/pyf/v40n3/en_pyf08317.pdf

Rodríguez, E., Gentile, N., Lupín, B., & Garrido, L. (2002). El Mercado Interno de Alimentos
Organicos: Perfil de los Consumidores Argentinos. Buenos Aires. Retrieved from
http://nulan.mdp.edu.ar/1010/1/00154.pdf

Rosset, P. M., & Altieri, M. A. (2017). Agroecology: science and politics. Rugby: Practical Action
Publishing.

Salcedo, S., & Guzmán, L. (2014). La agricultura familiar en cifras. Agricultura familiar en América
Latina y el Caribe. Recomendaciones de Política. Retrieved from
http://www.fao.org/docrep/019/i3788s/i3788s.pdf

SANE. (1998). Farmers, NGOs and Lighthouses: Learnings from Three Years of Training, Networking
and Field Activities. Berkeley.

Sartorato, C. R. (2006). Resumos do I Congresso Brasileiro de Agroecologia. Revista Brasileira de


Agroecologia, 1(1), 241–245.

SENASA. (2018). Situación de la Producción Organica en Argentina durante el año 2017. Buenos
Aires.

Serme, I., Ouattara, K., Ouattara, D., Ouedraogo, S., Youl, S., & Wortmann, C. (2018). Sorghum Grain
Yield Under Different Rates of Mineral and Organic Fertilizer Application in the South-Sudan
Zone of Burkina Faso BT - Improving the Profitability, Sustainability and Efficiency of Nutrients
Through Site Specific Fertilizer Recommendations . In A. Bationo, D. Ngaradoum, S. Youl, F.
Lompo, & J. O. Fening (Eds.) (pp. 235–248). Cham: Springer International Publishing.
https://doi.org/10.1007/978-3-319-58792-9_14

Seufert, V., Ramankutty, N., & Foley, J. A. (2012). Comparing the yields of organic and conventional
agriculture. Nature, 485. https://doi.org/10.1038/nature11069

Sevilla Guzmán, E., & Woodgate, G. (1997). Sustainable rural development: From industrial agriculture

99
to agroecology. In The international handbook of environmental sociology (pp. 83–100).
Cheltenham: Edward Elgar Publishing.

Sevilla Guzmán, E., & Woodgate, G. (2013). Agroecology: Foundations in agrarian social thought and
sociological theory. Agroecology and Sustainable Food Systems, 37(1), 32–44.
https://doi.org/10.1080/10440046.2012.695763

Shuaibu, Y. M., Bala, R. A., Kawure, S., & Shuaibu, Z. (2018). Effect of organic and inorganic fertilizer
on the growth and yield of sorghum (Sorghum bicolor (L.) Moench) in Bauchi state, Nigeria.
GSC Biological and Pharmaceutical Sciences, 2(1), 025–031.
https://doi.org/10.30574/gscbps.2018.2.1.0053

SOS Geografía. (n.d.). Economías Regionales. Retrieved June 3, 2019, from


https://argeo.jimdo.com/argentina/economías-regionales/

Splash, C. (2017). Social Ecological Economics. In C. Splash (Ed.), Routledge Handbook of Ecological
Economics (pp. 3–16). New York: Routledge.

Tanuro, D. (2010). Marxism, energy, and ecology: The moment of truth. Capitalism, Nature, Socialism,
21(4), 89–101. https://doi.org/10.1080/10455752.2010.524451

Trainer, T. (2012). De-growth: Do you realise what it means? Futures, 44, 590–599.
https://doi.org/10.1016/j.futures.2012.03.020

Trewavas, A. (2001). Urban myths of organic farming. Nature, 410, 409–410.

Van den Bosch, M. E., Abraham, L. I., & Alturria, L. V. (2018). Producción orgánica de uva en
Mendoza , Argentina : tipos de productores , caracterización técnica y económica Organic grape
production in Mendoza , Argentina : types of winegrowers , technical and economic
characterization. Cuyonomics, 1(2), 103–119. Retrieved from
http://revistas.uncu.edu.ar/ojs/index.php/cuyonomics

Verzeñassi, D. (2014). El modelo agrosojero y su impacto en nuestras vidas. In Agroindustria, Salud y


Soberanía (pp. 31–48). Buenos Aires: El Colectivo.

100
Via Campesina. (2017). Agroecology, a way of life, struggle, and resistance against capitalism.
Retrieved May 28, 2019, from https://viacampesina.org/en/agroecology-way-life-struggle-
resistance-capitalism/

Williams, S., Toulmin, C., & Pretty, J. (2011). Sustainable Intensification. Routledge.
https://doi.org/10.4324/9781849776844

World Commission on Environment. (1987). Our Common Future. Oxford. Retrieved from
http://www.un-documents.net/our-common-future.pdf

101
9 Appendices
Appendix A Interview 1
Introduction

1) What is your name?


2) How long have you been engaged in this project?
3) What does agroecological production mean to you and why did you decide to start with it?

Social analysis

4) How many people form part of this undertaking?


5) How is it organized? Who makes the decisions?
6) How do they organize labour? Who works? Would you consider there to be any gender gap
in the labour division?

Economic analysis

7) Is your produce for self-consumption or for market sale?


a. Which places do you sell your produce?
b. Why do you sell it there?
c. Do you feel that there are many alternatives?
8) Is the project economically feasible for you? Why (not)?

Ecological/Environmental analysis

9) Have you noticed a change in soil fertility since you made the transition?
10) Do you take into account the biodiversity of animals and crops in the region? How do you do
this?

Political conflict analysis

11) Are there many conventional producers in the area? Does it interrupt your own work?
12) Which inputs do you apply to the land and how do you acquire them?
13) Are you owner of the land?

102
14) What are the costs for you to access this land? Do you consider them to be excessive?
15) Do you have many contacts with other producers or organizations?

The further development of agroecology

16) Where does the agroecological knowledge come from?


17) Are there any ways the government assists you in your undertaking?
18) What are the main challenges to agroecology?
19) Do you consider yourself to be part of a certain ideological camp?
20) What are the most important principles of agroecology to you?
21) Is there anything else you would like to add, but which was not mentioned in this interview?

Appendix B Interview 2
Introduction

1) What is your name?


2) For how long have you dedicated yourself to the study of agroecology?
3) What made you study it?

Agroecology as a science

4) What are the areas where agroecology today still needs to investigate?
5) How does agroecology develop institutionally and do you feel that universities are adopting its
principles more readily?
6) Is there a connection between the research centers and the practice? Between theory and praxis?
How does this connection manifest itself?
7) What would you say are the biggest challenges to agroecology as a science today?
8) How do you see the difference between food security and food sovereignty And what does this
difference imply?

Agroecology as a practice

9) What can agroecology mean for mitigating climate change?

103
10) Most agroecological projects have an inherent social aspect to them. Where does this impulse
come from? And why is it not that prominent in conventional agriculture?
11) A strong critique towards agroecology is that it is not economically rentable and that its
productivity is lower. How would you respond to these critiques?
12) The history of Latin America is characterized by struggle fro land. What role does access to
land play in the development of the agroecological paradigm?

Agroecology as a movement

13) What kind of influence does the conventional agricultural paradigm have, and how do conflicts
between these paradigms manifest themselves?
14) Considering the emphasis that agroecology puts on horizontal decision-making, local
production and consumption, and a shorter supply chain, do you believe it is compatible with
globalized capitalism as a paradigm? If not, how could it further develop?
15) How can the political institutions facilitate a transition towards agroecology?
16) Do you feel there is much legal protection for smallholder farmers?

104

Anda mungkin juga menyukai