(Advances in Agronomy 116) Donald L. Sparks (Eds.) - Advances in Agronomy 116 (2012, Academic Press, Elsevier)
(Advances in Agronomy 116) Donald L. Sparks (Eds.) - Advances in Agronomy 116 (2012, Academic Press, Elsevier)
Advisory Board
PAUL M. BERTSCH RONALD L. PHILLIPS
University of Kentucky University of Minnesota
Permissions may be sought directly from Elsevier’s Science & Technology Rights
Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333;
email: [email protected]. Alternatively you can submit your request online by
visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting
Obtaining permission to use Elsevier material
Notice
No responsibility is assumed by the publisher for any injury and/or damage to persons or
property as a matter of products liability, negligence or otherwise, or from any use or
operation of any methods, products, instructions or ideas contained in the material
herein. Because of rapid advances in the medical sciences, in particular, independent
verification of diagnoses and drug dosages should be made
ISBN: 978-0-12-394277-7
ISSN: 0065-2113 (series)
Numbers in Parentheses indicate the pages on which the authors’ contributions begin.
K. J. Boote (41)
Agronomy Department, University of Florida, Gainesville, Florida, USA
Jean-Pierre Caliman (71)
PT SMART Research Institute (SMARTRI), Pekanbaru, Riau, Indonesia
Qing Chen (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
Xinping Chen (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
François Colin (71)
Montpellier-SupAgro, UMR-LISAH (Laboratory on Interactions between Soil,
Agrosystem and Hydrosystem), Montpellier cedex, France
Irina Comte (71)
Department of Natural Resource Sciences, Macdonald Campus of McGill Univer-
sity, Ste-Anne-de-Bellevue, Quebec, Canada, and CIRAD (International Coopera-
tion Centre in Agronomic Research for Development), Montpellier cedex, France
Zhenling Cui (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
Mingsheng Fan (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
Steven J. Fonte (123)
International Center for Tropical Agriculture (CIAT), Cali, Colombia
Olivier Grünberger (71)
IRD (Institut de Recherche pour le Développement), UMR-LISAH, Montpellier
cedex, France
Rongfeng Jiang (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
Xiaotang Ju (1)
Department of Plant Nutrition, China Agricultural University, Beijing, PR China
ix
x Contributors
DONALD L. SPARKS
Newark, Delaware, USA
xiii
C H A P T E R O N E
Contents
1. Introduction 2
2. Principles of INM 6
2.1. Optimizing nutrient inputs and taking all possible sources of
nutrients into consideration 9
2.2. Dynamically matching soil nutrient supply with crop
requirement spatially and temporally 11
2.3. Effectively reducing N losses in intensive managed Chinese
cropping systems 12
2.4. Taking all possible yield increase measures into consideration 15
3. Technology and Demonstration of INM in Different Cropping
Systems 16
3.1. INM for intensive wheat and maize system 18
3.2. INM for paddy rice 21
3.3. INM for vegetable systems 23
3.4. INM for orchards 26
4. Large-Scale Dissemination of INM 29
5. Summary and Conclusions 31
Acknowledgments 32
References 32
Abstract
While the concept of sustainability as a goal has become widely accepted, the
dominant agricultural paradigm still considers high yield and reduced environ-
mental impact being in conflict with one another. During the past 49years
(1961–2009), the 3.4-fold increase in Chinese agricultural food production can
1
2 Fusuo Zhang et al.
Abbreviations
1. Introduction
The Green Revolution helped to create the world’s “Miracle in China,”
with 9% of the world’s arable land feeding 22% of the world population. In the
past 49years (1961–2009), cereal grain yields have increased 3.5-fold from 1.2
to 5.4tha1, while total grain production has increased 3.4-fold from 110 to
483 million ton (MT) (FAO, 2011). In 1998, grain, meat, and egg production
per capita in China exceeded the world average. The increased demand in
Chinese grain production has affected the global food supply and the natural
Nutrient Management in China 3
resource bases required for nutrient production (fossil fuels, mineral sources of
P and K) and has attained world recognition.
However, this 3.4-fold increase in Chinese agricultural food production
during the past 49years can be partly attributed to a 48-fold increase in
chemical fertilizers from 1 to 49MT, including a 37-fold increase in N
fertilizer application and a 91-fold increase in P fertilizer use, and a 442-fold
increase in the area of irrigated croplands (Fig. 1). Total consumption of
chemical fertilizers worldwide increased by 3.9-fold from 32 to 164MT,
indicating that 36% of the global increase (132MT) came from China
during the past 49years. In the past 10years (2000–2009), 54% of the global
increase in chemical fertilizer consumption (27MT) was contributed
by China, including 11MT fertilizer N (54% of the global increase), 2.5
MT fertilizer P (52% of the global increase), and 1.1MT fertilizer K (58% of
the global increase) (Figs. 1 and 2A,B).
Cereal yields in the past 10years have continued to increase with no
proportional increases in fertilizer use in many developed countries or
regions such as Western Europe (rainfed cereal systems), North America
(rainfed and irrigated corn), and Japan and South Korea (irrigated rice)
(Dobermann and Cassman, 2005). For example, in the past 10years, chem-
ical fertilizer consumption in the United States increased by only 0.04MT
with 0.23% of total fertilizer consumption in 2009 and decreased by 0.32
MT in Western Europe (Fig. 2A). By contrast, the application rate of
600 80
Grain production
500 Total fertilizer Fertilizer consumption (MT)
N fertilizer
Grain production (MT)
60
400 P fertilizer
K fertilizer
300 40
200
20
100
0 0
1960 1970 1980 1990 2000 2010
Year
Figure 1 The trend of grain production and chemical fertilizer inputs (N, P, and K
fertilizers) in China from 1961 to 2009. The P and K fertilizers are calculated by P2O5
and K2O, respectively. Fertilizer consumption is defined as the difference between
fertilizer production and exports. Source: FAO (2011) and IFA (2011).
4 Fusuo Zhang et al.
200 80
A
160
120
40
80
20
40
0 0
1960 1970 1980 1990 2000 2010
Year
600
B
500
Fertilizer rate (kg ha )
-1
400
300
200
100
0
1960 1970 1980 1990 2000 2010
Year
Figure 2 Trend of total chemical fertilizer consumption (A) and fertilizer rate per
hectare (B) for global scale, China, United States, and Western European. Source: IFA
(2011).
2. Principles of INM
The overall principle of INM is to maximize biological potential for
improving crop productivity and resources use efficiency through root
zone/rhizosphere management. Plant roots take up nutrients from soils
via the rhizosphere, a narrow zone of the soil that is directly influenced by
root growth, root secretions, and associated soil microorganisms. In crop-
ping systems, a rhizosphere continuum in the root zone can be formed due
to root/rhizosphere interactions among individual plants. The rhizosphere
is the important interface where interactions among plants, soils, and micro-
organisms occur and is a “bottleneck” controlling nutrient transformations,
availability, and flow from soils to plants. Therefore, the chemical and
biological processes occurring in the rhizosphere determine the mobiliza-
tion and acquisition of soil nutrients together with microbial dynamics, and
also control NUE by crops, and thus profoundly influence cropping system
productivity and sustainability (Zhang et al., 2004, 2010).
As plant growth proceeds, the roots can respond to and/or sense changes
in soil nutrient availability including nutrient supply intensity and composi-
tion. These responses involve a series of adaptive alterations in root mor-
phology and root physiology. P-deficient plants can commonly increase
their root/shoot ratio, root branching, root elongation, root topsoil forag-
ing, and formation of cluster roots and root hairs (Lynch and Brown, 2008;
Shen et al., 2011b; Vance, 2008). Mycorrhizal associations can also enhance
the spatial availability of P, extending the nutrient absorptive surface by
formation of mycorrhizal hyphae (Marschner, 1995). On the other hand,
root-induced chemical and biological changes in the rhizosphere affect the
bioavailability of soil P, mainly involving rhizosphere acidification, carboxy-
late exudation, secretion of phosphatases or phytases, and Pi transporter
expression (Neumann and Römheld, 2002; Zhang et al., 2010). It has been
reported that P deficiency increases the formation of cluster roots by white
lupin (Lupinus albus L.; Shen et al., 2005; Wang et al., 2007), axial root length
and total root length, and larger amounts of lateral roots and more root hair
formation in maize (Zea mays L.) or Arabidopsis (Bates and Lynch, 1996;
Linkohr et al., 2002; Liu et al., 2004b; Schachtman et al., 1998; Schenk and
Barber, 1979). In crop species, Liu et al. (2004b) found that efficient use of
P in calcareous soil by maize is related to its large root system, with a greater
ability to acidify the rhizosphere, and a positive response of acid phosphatase
production and excretion in low P conditions. High P acquisition efficiency
by modifying root morphology and root physiology in terms of rhizosphere
biological and chemical processes is important for achieving high crop
yields with savings in nutrient inputs. The nutrient supply intensity or
concentrations in the rhizosphere/root zone in cropping systems can be
optimized to a critical level through nutrient management to maximize the
biological potential for efficient use of soil P by plants.
Nitrogen fertilization is the most common practice for the regulation of
root growth in field conditions. Maize roots respond to N supply in two
ways. First, in uniform N supply systems, N deficiency increases maize root
length, resulting in longer axial roots (primary, seminal, and nodal roots;
Tian et al., 2006; Wang et al., 2003). This helps the roots to explore a larger
8 Fusuo Zhang et al.
soil volume and thus increases spatial N availability. However, root elonga-
tion can be inhibited if the N supply is too high. In maize, for example, the
optimum nitrate level for root length seems to be around 5mmolL1 (Tian
et al., 2008). Second, root growth can be stimulated when plant roots
experience nutrient-rich patches, particularly when the patches are rich in
N and P (Drew, 1975; Hodge, 2004). When a maize plant is suffering from
N deficiency and part of the root mass is supplied with nitrate locally, the
growth of lateral roots in the supplied area is enhanced (Granato and Raper,
1989; Guo et al., 2005; Sattelmacher and Thoms, 1995). This helps plants to
compete with other plant species and/or microbes for limited N resources
(Hodge, 2004). It is suggested that NO 3 plays a key role as a nutrient signal
in regulating root proliferation (Zhang and Forde, 1998). Localized P
application effectively enhances crop growth and P use efficiency. More-
over, manipulating and managing nutrient supply intensity and composition
in the local fertilization zone can greatly strengthen root growth and
nutrient uptake through modifying rhizosphere processes and enlarging
the root absorbing surface. A field experiment showed that localized appli-
cation of P with addition of ammonium significantly enhanced P uptake
and crop growth through stimulating root proliferation and rhizosphere
acidification ( Jing et al., 2010). The leaf expansion rate was 20–50% higher,
the total root length 23–30% greater, and the plant growth rate 18–77%
greater with a localized supply of P plus ammonium compared with broad-
casting of these nutrients. Localized application of P combined with addi-
tion of ammonium significantly decreased rhizosphere pH in the fertilized
zone compared with the bulk soil ( Jing et al., 2010). The results suggest that
modifying rhizosphere processes in the field may be an effective manage-
ment strategy for increasing NUE and plant growth.
Rhizosphere management emphasizes maximizing the efficiency of
root/rhizosphere processes in nutrient acquisition and use by crops rather
than simply depending on excessive fertilizer inputs, which involves reg-
ulating the root system, rhizosphere acidification, carboxylate exudation,
microbial associations with plants, rhizosphere interactions in terms of
intercropping and rotation (Li et al., 2007), localized application of nutri-
ents, use of efficient crop genotypes and synchronizing rhizosphere nutri-
ent supply with crop demand. Rhizosphere management has been shown
to be an effective approach for increasing NUE and crop productivity
through “small causes with big effects” for sustainable agricultural produc-
tion (Zhang et al., 2010). Based on a better understanding of rhizosphere
processes, the key steps of INM are (1) optimizing nutrient inputs and
taking all possible sources of nutrients into consideration, (2) dynamically
matching soil nutrient supply with crop requirement spatially and tempo-
rally, (3) effectively reducing N losses in intensively managed Chinese
cropping systems, and (4) taking all possible yield increase measures into
consideration (Fig. 4).
Nutrient Management in China 9
result, high crop yields can be readily obtained in arable soils without
application of N fertilizers (Tong et al., 2004). This large indigenous N supply
aggravates N surpluses and increases the potential for N losses from agro-
ecosystems unless it is considered to be a component part of the plant-
available N when constructing an integrated N management plan.
The large indigenous N supply is attributed to high soil nitrate-N accu-
mulation and environmental N supply. Soil nitrate-N (NO 3 -N) accumula-
tion in the top 90 or 100cm of the soil was above 200kgNha1 under
conventional N practice in intensive wheat–maize systems (Cui et al.,
2008a,d; Liu et al., 2003); this residual NO 3 -N supply can reach 1173 and
613kgNha1, respectively, in greenhouse vegetable and orchard systems in
North China ( Ju et al., 2004, 2006). Ju et al. (2006) observed that residual soil
nitrate-N after winter wheat harvest was 275kgNha–1 in the top 90cm of the
soil profile and 213kgNha–1 in the 90–180cm soil depth increment. Liu et al.
(2004a) reported an average residual soil nitrate-N content of 314kgNha–1 in
the top 2m of the soil profile and 145kgNha–1 at 2–4m soil depth based on
on-farm soil tests in annual winter wheat production systems in Beijing
suburbs on the surroundings of NCP (n¼93). Wheat grain yields on the
NCP showed no response to applied N when initial nitrate-N before
sowing in the top 90cm soil layer exceeded 200kgNha–1, but residual
nitrate-N content after harvest and N losses significantly increased (Cui
et al., 2008a). High nitrate-N accumulation in the soil profile is like a “time
bomb” that could explode at any moment and will finally be lost to the
environment through either denitrification or leaching under high N appli-
cation rates ( Ju et al., 2009; Zhao et al., 2006). Ju et al. (2002) observed
residual nitrate-N from the first growing season (wheat) to leach from the
0 to 1m soil profile during the second growing season (maize) on the NCP
due to high rainfall in summer.
The total amount of N available from the environment in China in the
2000s has more than doubled compared with the 1980s because of the rapid
continuing increases in both oxidized and reduced N emissions (Liu and
Zhang, 2009; Liu et al., 2010, 2011). According to Fig. 5, N inputs from
atmospheric N deposition and irrigation water in China were up to 33 and
12kgNha1 yr1 in the 2000s but only 14 and 4kgNha1 yr1 in the 1980s.
Similar rapid increases in environmental sources of N from atmospheric
deposition and irrigation water were reported on the NCP and in the Taihu
Lake region ( Ju et al., 2009). Nitrogen input from deposition plus irrigation
was only 30kgNha1 yr1 on the NCP and in Taihu Lake region in the
1980s, and this value has increased to 99 and 89kgNha1 yr1, respectively,
in these two intensive agricultural regions in the 2000s ( Ju et al., 2009).
Evidence from the NCP indicates that dry N deposition (50–60kgNha1
yr1) (Shen et al., 2009, 2011a) is likely to be a major contributor to the
environmental N in this semihumid/semiarid region compared with wet
N deposition (30kgNha1 yr1) (Liu et al., 2006b; Zhang et al., 2008b).
Nutrient Management in China 11
35
1980s
30
2000s
N input (kg N ha-1 yr -1) 25
20
15
10
0
N deposition N from irrigation
Further studies confirm that about 50% of this air-borne N input (52kg
Nha1 yr1) can be utilized by current crops, according to the ITNI-system
which is based on a 15N dilution approach (He et al., 2007a, 2010).
meet the crop N requirement, establish growth, and promote the develop-
ment of healthy roots during the early periods of crop growth (Cui et al.,
2008a,d). Nitrogen application in excess of crop N demand during this
period increases the potential for nitrate-N leaching and results in excessive
crop growth that is more susceptible to disease and lodging. Therefore,
more N fertilizer (around 60–70% of the total N fertilizer input) should be
applied during the most rapid growth stages of the crop to achieve synchro-
nization between the N supply and crop demand (Cui et al., 2008b,c).
Restricted by old knowledge and habits, farmers often apply large
amounts of N fertilizers before planting or at the early growth stages as a
conventional management practice in crop production. For example, the N
supply before planting is usually about 50% of the total amount given (Chen,
2003; Cui et al., 2008c,d). Many rice farmers in China apply 50–90% of the N
as a basal dressing and an early top-dressing within the first 10 days after
transplanting to reduced transplanting shock and stimulate early tillering
(Zhang et al., 2011). Clearly, this large amount of N fertilizer at the early
growth stages has resulted in poor synchronization between soil N supply and
crop demand, leading to high soil inorganic N concentrations before the
occurrence of rapid crop N uptake (Chen et al., 2006; Tilman et al., 2002).
Although the average N application rate in China is excessive for the
maximum crop N requirement, some low-income farmers or those in
remote areas apply N inadequately. If we simply use 150–200kgNha–1 as
a reasonable amount of N fertilizer for the main wheat and maize growing
regions of China (n¼10,000), only one-third of farmers would be in this
range, one-third would be applying too much (N rate>200kgNha1), and
one-third would be applying too little (N rate<150kgNha1) (Wang,
2007). On a regional scale, higher crop yields are likely to be achieved
through a combination of increased N application in regions with a low N
input and improved NUE in regions where N fertilizer use is already high.
50
45 NH3 volatilization
40 Leaching
Denitrification
35
N losses (%)
30
25
20
15
10
5
0
Rice Wheat-South Wheat-North Maize
Taihu region North China Plain
Crops and regions
Table 1 Resource characteristics of nutrient and the technologies of INM for different
nutrients
N demand
N target value
Fertilizer+ N
from
environment,
N supply
etc.
Nutrient bioavailability
Planting
Nutrient spatial
availability
Nutrient temporal and spatial variability
Nitrogen target value for different grain yield for Chinese maize production
Figure 7 Model of in-season root-zone N management and N target value for differ-
ent grain yields for Chinese maize production. Date source: Cui et al., 2008b.
20 Fusuo Zhang et al.
and 5%, respectively. As expected, the ONR treatment had higher N use
efficiency and reduced apparent N loss compared with FNP (P<0.05).
Compared to FNP, apparent N losses were reduced by 73% from 159 to
43kgNha1 and 43% from 151 to 86kgNha1 in the wheat and maize
growing seasons, respectively (Fig. 8).
Clearly, in this study ONR using in-season root-zone N management
strategy was close to optimal, sustained high yields, and increased profits,
while minimizing the potential environmental impacts of N fertilization.
This superiority of in-season root-zone N management strategy can be
attributed to three major factors: (1) efficient utilization of all N sources
from the soil, environment, and fertilizer; (2) achievement of synchroniza-
tion between N supply and crop demand; and (3) addressing site-specific N
management. In China, small-scale farming has intensified variations in soil
N supply. In the research region, indigenous N supplies (average N uptake
in 0 N control) varied from 69 (site 14) to 191kgNha1 (site 5), with a
coefficient of variation of 24%. The large disparity among experimental sites
reflects the net variation in soil N transformations (mineralization, immobi-
lization, denitrification, and volatilization), possible N transport (leaching
and runoff), and crop uptake. Clearly, one rate does not fit all circumstances.
Figure 8 Performance of INM for wheat and maize in the North China Plain. (A)
INM can reduce N fertilizer by 30% while increasing grain yield of 16% compared with
farmers’ practice in Binzhou, Shandong province. (B) Maize lodging for FNP in
Xiaoxian, Anhui Province. (C) Mean N fertilizer application rate, apparent N losses,
AEN and PFPN of INM compared with FNP. Modified from Cui et al. (2008b,c).
Nutrient Management in China 21
Figure 9 Performance of INM for rice in China. (A) Experimental bases (red dots) for
INM demonstration of rice in China; (B) INM reduced N fertilizer by 38% while
increasing grain yield of 11% compared with FNP; (C) rice yield, N rate and AEN for
INM and FNP for paddy rice in Chinese main rice production since 2001.
250 120
100
200
80
150
60
100
40
50 20
N target (kg N ha-1)
0 0
100
200
300
I II III
Figure 12 Performance of INM in China. (A) 158 experimental bases (red dots) for
INM dissemination in China; (B) increased yield, reduced N fertilizer rate, and
increased net income of INM for different crops including wheat, maize, rice, vegeta-
ble, fruit, rape, and cotton, compared farmers’ practice. D yield, N fertilizer, and net
income mean the different yields, N fertilizer rate, and net income between INM and
farmer’s practice.
30 Fusuo Zhang et al.
Figure 13 Trends in fertilizer consumption for cereal crop and partial factor produc-
tivity for fertilizer (PFP) in China from 1981 to 2008. The PFP was defined the ratio of
crop yield per unit of applied chemical fertilizer. Modified from Zhang et al. (2011).
Nutrient Management in China 31
ACKNOWLEDGMENTS
We thank the National Basic Research Program of China (973 Program: 2009CB118606),
the Special Fund for the Agriculture Profession (201103003). This research was supported
by the National Natural Science Foundation of China (30890130) and the Innovative
Group Grant of the National Science Foundation of China (31121062) for financial
support. We will give special thanks to Prof. P. Christie in Queen’s University Belfast,
UK, and Prof. C. Tang in La Trobe University, Bundoora, Australia, for their comments
and linguistic revisions.
REFERENCES
Alloway, B. J. (2008). Micronutrients and crop production: An introduction.
In “Micronutrient Deficiencies in Global Crop Production” (B. J. Alloway, et al., Eds.),
pp. 1–39. Springer, The Netherlands.
Bates, T. R., and Lynch, J. P. (1996). Stimulation of root hair elongation in Arabidopsis
thaliana by low phosphorus availability. Plant Cell Environ. 19, 529–538.
Beman, J. M., Arrigo, K. R., and Matson, P. A. (2005). Agricultural runoff fuels large
phytoplankton blooms in vulnerable areas of the ocean. Nature 434, 211–214.
Berg, A. (1973). The relation of manganese to internal bark necrosis of apple. Science 74,
485–490.
Binford, G. D., Blackmer, A. M., and Cerrato, M. E. (1992). Relationships between corn
yields and soil nitrate in late spring. Agron. J. 84, 53–59.
Blankenau, K., Olfs, H. W., and Kuhlmann, H. (2000). Effect of microbial nitrogen
immobilization during the growth period on the availability of nitrogen fertilizer for
winter cereals. Biol. Fertil. Soils 32, 157–165.
Campbell, C. A., Myers, R. J. K., and Curtin, D. (1995). Managing nitrogen for sustainable
crop production. Fertil. Res. 42, 277–296.
Cao, Z. H., Huang, J. F., Zhang, C. S., and Li, A. F. (2004). Soil quality evolution after land
use change from paddy soil to vegetable land. Environ. Geochem. Health 26, 97–103.
Cassman, K. G. (1999). Ecological intensification of cereal production systems:
Yield potential, soil quality, and precision agriculture. Proc. Natl. Acad. Sci. USA 96,
5952–5959.
Cassman, K. G., Gines, G. C., Dizon, M. A., Samson, M. I., and Alcantara, J. M. (1996).
Nitrogen-use efficiency in tropical lowland rice systems: Contributions from indigenous
and applied nitrogen. Field Crop Res. 47, 1–12.
Cassman, K. G., Dobermann, A., and Walters, D. T. (2002). Agroecosystems, nitrogen use
efficiency, and nitrogen management. Ambio 31, 132–140.
Chai, Z. P., Jiang, P. A., and Wang, X. M. (2009). The investigation research of fertilization
about several main characteristic fruit trees in Xinjiang. Chin. Agric. Sci. Bull. 24, 231–
234 (in Chinese with English abstract).
Chen, X. P. (2003). Optimization of the N Fertilizer Management of a Winter Wheat/
Summer Maize Rotation System in the North China Plain.PhD Thesis, University of
Hohenheim, Stuttgart, Germany.
Chen, Q., Zhang, X., Zhang, H., Christie, P., Li, X., Horlacher, D., and Liebig, H. (2004).
Evaluation of current fertilizer practice and soil fertilizer in vegetable production in the
Beijing region. Nutr. Cycl. Agroecosyst. 69, 51–58.
Chen, X. P., Zhang, F. S., Römheld, V., Horlacher, D., Schulz, R., Böning-Zilkens, M.,
Wang, P., and Claupein, W. (2006). Synchronizing N supply from soil and fertilizer
and N demand of winter wheat by an improved Nmin method. Nutr. Cycl. Agroecosyst. 74,
91–98.
Nutrient Management in China 33
Chen, X. P., Zhang, F. S., Cui, Z. L., Li, F., and Li, J. L. (2010). Optimizing soil nitrogen
supply in the root zone to improve maize nitrogen management. Soil Sci. Soc. Am. J. 74,
1367–1373.
Chen, X. P., Cui, Z. L., Vitousek, P. M., Cassman, K. G., Matson, P. A., Bai, J. S.,
Meng, Q. F., Hou, P., Yue, S. C., Römheld, V., and Zhang, F. S. (2011). Integrated
soil-crop system management for food security. Proc. Natl. Acad. Sci. USA 108,
6399–6404.
China Agricultural Yearbook. (2011). Editorial committee of China Agricultural Yearbook,
Chinese Statistical Publishing House, Beijing.(in Chinese).
Chinese Ministry of Environmental Protection (2010). http://www.gov.cn/jrzg/2010-02/
10/content_1532174.htm..
Cong, L. M., Jiao, X. Y., Wang, L. G., Wang, J. S., and Dong, E. W. (2011). Present
situation of fertilization in solar-greenhouse in Quwo county and suggestions. J. Shanxi
Agric. Sci. 39(4), 342–345. (in Chinese with English abstract).
Cui, Z. L., Chen, X. P., Miao, Y. X., Li, F., Zhang, F. S., Li, J. L., Ye, Y. L., Yang, Z. P.,
Zhang, Q., and Liu, C. S. (2008a). On-farm evaluation of winter wheat yield response to
residual soil nitrate-N in North China Plain. Agron. J. 100, 1527–1534.
Cui, Z. L., Chen, X. P., Miao, Y. X., Zhang, F. S., Sun, Q. P., and Schroder, J. (2008b).
On-farm evaluation of the improved soil Nmin-based nitrogen management for summer
maize in North China Plain. Agron. J. 100, 517–525.
Cui, Z. L., Zhang, F. S., Chen, X. P., Miao, Y. X., Li, J. L., Shi, L. W., Xu, J. F., Ye, Y. L.,
Liu, C. S., Yang, Z. P., Zhang, Q., Huang, S. M., et al. (2008c). On-farm evaluation of an
in-season nitrogen management strategy based on soil Nmin test. Field Crop Res. 105, 48–55.
Cui, Z. L., Zhang, F. S., Miao, Y. X., Sun, Q. P., Li, F., Chen, X. P., Li, J. L., Ye, Y. L.,
Yang, Z. P., Zhang, Q., and Liu, C. S. (2008d). Soil nitrate-N levels required for high
yield maize production in the North China Plain. Nutr. Cycl. Agroecosyst. 82, 187–196.
Cui, Z. L., Zhanf, F. S., Mi, G. H., Chen, F. J., Li, F., Chen, X. P., Li, J. L., and Shi, L. F.
(2009). Interaction between genotypic difference and nitrogen management strategy in
determining nitrogen use efficiency of summer maize. Plant Soil 317, 267–276.
Cui, Z. L., Chen, X. P., and Zhang, F. S. (2010a). Current nitrogen management status and
measures to improve the intensive wheat–maize system in China. Ambio 39, 376–384.
Cui, Z. L., Zhang, F. S., Chen, X. P., Dou, Z. X., and Li, J. L. (2010b). In-season nitrogen
management strategy for winter wheat: Maximizing yields, minimizing environmental
impact in an over-fertilization context. Field Crop Res. 116, 140–146.
Cui, Z. L., Zhang, F. S., Chen, X. P., Li, F., and Tong, Y. P. (2011). Using in-season
nitrogen management and wheat cultivars to improve nitrogen use efficiency. Soil Sci.
Soc. Am. J. 75, 967–983.
Dinnes, D. L., Karlen, D. L., Jaynes, D. B., Kaspar, T. C., Hatfield, J. L., Colvin, T. S., and
Cambardella, C. A. (2002). Nitrogen management strategies to reduce nitrate leaching
in tile-drained midwestern soils. Agron. J. 94, 153–171.
Dobermann, A. (2007). Nutrient use efficiency-measurement and management. In “IFA
International Workshop on Fertilizer Best Management Practices,” International Fertil-
izer Industry Association, Paris, 7–9 March 2007, Brussels, Belgium.
Dobermann, A., and Cassman, K. G. (2005). Cereal area and nitrogen use efficiency are
drivers of future nitrogen fertilizer consumption. Sci. China C Life Sci. 48, 745–758.
Dobermann, A., and Fairhurst, T. (2000). Rice: Nutrient Disorders & Nutrient Manage-
ment. PPI, PPIC and IRRI, Singapore.
Drew, M. C. (1975). Comparison of the effects of a localized supply of phosphate, nitrate,
ammonium and potassium on the growth of the seminal root system, and the shoot, in
barley. New Phytol. 75, 479–490.
Drinkwater, L. E., and Snapp, S. S. (2007). Nutrients in agroecosystems: Rethinking the
management paradigm. Adv. Agron. 92, 163–186.
34 Fusuo Zhang et al.
Du, L. F., Zhang, W. L., and Li, Z. H. (2006). Soil quality with various planting patterns
in Yangtze River Delta Area. J. Agro Environ. Sci. 25, 95–99. (in Chinese with English
abstract).
Editorial Committee of China Agriculture Yearbook (ECCAY). (2006). China Agriculture
Yearbook. China Agriculture Press, Beijing.
Engels, T., and Kuhlmann, H. (1993). Effect of the rate of N fertilizer on apparent net
mineralisation of N during and after cultivation of cereal and sugar beet crops.
Z. Pflanzenernähr. Bodenkd. 156, 149–154.
Fahey, T. J., Williams, C. J., and Rooney-Varga, J. N. (1999). Nitrogen deposition in and
around an intensive agricultural district in central New York. J. Environ. Qual. 28,
1585–1600.
Fan, M., Lu, S., Jiang, R., Liu, X., Zeng, X., Goulding, K., and Zhang, F. (2007). Nitrogen
input, 15 N balance and mineral N dynamics in a rice-wheat rotation in southwest China.
Nutr. Cycl. Agroecosyst. 79, 255–265.
Fan, M. S., Cui, Z. L., Chen, X. P., Jiang, R. F., and Zhang, F. S. (2008). Integrated nutrient
management for improving crop yields and nutrient utilization efficiencies in China.
J. Soil Water Conserv. 63, 126–128.
Fan, M. S., Lu, S. H., Jiang, R. F., Liu, X. J., and Zhang, F. S. (2009). Triangular
transplanting pattern and split nitrogen fertilizer application increase rice yield and
nitrogen fertilizer recovery. Agron. J. 101, 1421–1425.
Fan, H. Z., Chen, Y. X., Lv, S. H., and Feng, W. Q. (2010a). Current status and evaluation
of nutrient management of turnip in Sichuan province. Soils 42(5), 828–832. (in Chinese
with English abstract).
Fan, M. S., Christie, P., Zhang, W., and Zhang, F. S. (2010b). Crop production, fertilizer
use and soil quality in China. In “Advances in Soil Science Food Security and Soil
Quality” (R. Lal, B. A. Stewart, et al., Eds.), pp. 87–108. CRC Press, Taylor & Francis
Group, Boca Raton, FL.
FAO. (2011). FAOSTAT Database—Agriculture Production. Food and Agriculture Orga-
nization of the United Nations, Rome.
Ferguson, I. B., and Watkins, C. B. (1989). Bitter pit in apple fruit. Hortic. Rev. 11, 289–355.
Galloway, J. N., and Cowling, E. B. (2002). Reactive nitrogen and the world: 200 years of
change. Ambio 31, 64–71.
Gao, Y. M., Xu, J., and Gao, S. Q. (2006). Effects of silicon application on apple internal
bark necrosis induced by high content of manganese. Plant Nutr. Fertil. Sci. 12(4), 571–
577. (in Chinese with English abstract).
Gao, B., Li, J. L., Chen, Q., Liu, Q. H., and Wang, J. (2009). Study on nitrogen
management and irrigation methods of greenhouse tomato. Sci. Agric. Sin. 42(6),
2034–2042. (in Chinese with English abstract).
Gao, B., Ju, X. T., Zhang, Q., Christie, P., and Zhang, F. S. (2011). New estimates of direct
N2O emissions from Chinese croplands from 1980 to 2007 using localized emission
factors. Biogeosci. Discuss. 8, 6971–7006.
Ge, X. Y. (2009). Assessment of NPK Fertilizer Consumption and Demand in Vegetable
System in China. Master Thesis, China Agricultural University, Beijing, China (in
Chinese with English abstract).
Goulding, K. W. T., Bailey, N. J., Bradbury, N. J., Hargreaves, P., Howe, M.,
Murphy, D. V., Poulton, P. R., and Willison, T. (1998). Nitrogen deposition and its
contribution to nitrogen cycling and associated soil processes. New Phytol. 139, 49–58.
Granato, T. C., and Raper, C. D. (1989). Proliferation of maize (Zea mays L.) roots in
response to localized supply of nitrate. J. Exp. Bot. 40, 263–275.
Guo, R. Y. (2007). Studies on Nitrogen Control in Root-Zone and Summer Catch Crop
Planting for Reducing N Loss in Greenhouse Cucumber Cropping System. PhD Thesis,
China Agricultural University, Beijing, China. (in Chinese with English abstract).
Nutrient Management in China 35
Guo, Y. F., Mi, G. H., Chen, F. J., and Zhang, F. (2005). Effect of NO 3 supply on lateral
root growth in maize plants. J. Plant Physiol. Mol. Biol. 31, 90–96. (in Chinese with
English abstract).
Guo, R. Y., Li, X. L., Christie, P., Chen, Q., and Zhang, F. S. (2008). Seasonal temperatures
have more influence than nitrogen fertilizer rates on cucumber yield and nitrogen uptake
in a double cropping system. Environ. Pollut. 151, 433–451.
Guo, J. H., Liu, X. J., Zhang, Y., Shen, J. L., Han, W. X., Zhang, W. F., Christie, P.,
Goulding, K. W. T., Vitousek, P. M., and Zhang, F. S. (2010). Significant acidification
in major Chinese croplands. Science 327, 1008–1010.
He, F. F., Xiao, W. L., Li, J. L., Chen, Q., Jiang, R. F., and Zhang, F. S. (2006). Integrated
nitrogen management in greenhouse tomato production. Plant Nutr. Fertil. Sci. 12(3),
394–399. (in Chinese with English abstract).
He, C. E., Liu, X. J., Fangmeier, A., and Zhang, F. S. (2007a). Quantifying the total airborne
nitrogen input into agroecosystems in the North China Plain. Agric. Ecosyst. Environ. 121,
395–400.
He, F. F., Chen, Q., Jiang, R. F., Chen, X. P., and Zhang, F. S. (2007b). Yield and nitrogen
balance of greenhouse tomato (Lycopersicum esculentum Mill.) with conventional and
site-specific management practice in Northern China. Nutr. Cycl. Agroecosyst. 77, 1–14.
He, F. F., Jiang, R. F., Chen, Q., Zhang, F. S., and Su, F. (2009). Nitrous oxide emissions
from an intensively managed greenhouse vegetable cropping system in Northern China.
Environ. Pollut. 157, 1666–1672.
He, C. E., Wang, X., Liu, X. J., Fangmeier, A., Christie, P., and Zhang, F. S. (2010).
Nitrogen deposition and its contribution to nutrient inputs to intensively managed
agricultural ecosystems. Ecol. Appl. 20, 80–90.
Hodge, A. (2004). The plastic plant: Root responses to heterogeneous supplies of nutrients.
New Phytol. 162, 9–24.
Hofman, G. (1999). Nutrient management legislation in European countries. NUMALEC
Report. Concerted action, Fair6-CT98-4215.
Huang, Y., and Sun, W. J. (2006). Changes in topsoil organic carbon of croplands in
mainland China over the last two decades. Chin. Sci. Bull. 51, 1785–1803.
IFA (2011). IFADATA—International Fertilizer Industry Association..
Jiang, Y. M., Zhang, H. Y., and Zhang, F. S. (2007). The Theory and Practice of Integrated
Nutrient Management for Fruit Trees in North China. China Agricultural University Press,
Beijing. (in Chinese).
Jiao, X. Y., Wang, L. G., Zhang, D. L., Zhang, J. S., and Dong, E. W. (2010). Present
situation of fertilizer application, its problems and suggestions concerning vegetable
production under conditions of solar-greenhouse. J. Shanxi Agric. Sci. 38(4), 37–41. (in
Chinese with English abstract).
Jing, J. Y., Rui, Y. K., Zhang, F. S., Rengel, Z., and Shen, J. B. (2010). Localized application
of phosphorus and ammonium improves growth of maize seedlings by stimulating root
proliferation and rhizosphere acidification. Field Crop Res. 119, 355–364.
Ju, X. T., Liu, X. J., Zou, G. Y., Wang, Z. H., and Zhang, F. S. (2002). Evaluation of
pathway of nitrogen loss in winter wheat and summer maize rotation system. Agric. Sci.
Chin. 1, 1224–12316. (in Chinese with English abstract).
Ju, X. T., Liu, X. J., Zhang, F. S., and Roelcke, M. (2004). Nitrogen fertilization, soil nitrate
accumulation, and policy recommendations in several agricultural regions of China.
Ambio 33, 300–305.
Ju, X. T., Kou, C. L., Zhang, F. S., and Christie, P. (2006). Nitrogen balance and
groundwater nitrate contamination: Comparison among three intensive cropping sys-
tems on the North China Plain. Environ. Pollut. 143, 117–125.
Ju, X., Xing, G., Chen, X., Zhang, S., Zhang, L., Liu, X., Cui, Z., Yin, B., Christie, P.,
Zhu, Z., and Zhang, F. (2009). Reducing environmental risk by improving N manage-
ment in intensive Chinese agricultural systems. Proc. Natl. Acad. Sci. USA 106, 3041–3046.
36 Fusuo Zhang et al.
Ju, X., Lu, X., Gao, Z., Chen, X., Su, F., Kogge, M., Römheld, V., Christie, P., and
Zhang, F. (2011). Processes and factors controlling N2O production in an intensively
managed low carbon calcareous soil under sub-humid monsoon conditions. Environ.
Pollut. 159, 1007–1016.
Le, C., Zha, Y., Li, Y., Sun, D., Lu, H., and Yin, B. (2010). Eutrophication of lake waters in
China: Cost, causes, and control. Environ. Manage. 45, 662–668.
Li, R. (2000). Efficiency and Regulation of Fertilizer Nitrogen in High-Yield Farmland: A
Case Study on Rice and Wheat Double Maturing System Agriculture Area of Tai Lake
for Deducing to Jiangsu Province. PhD Thesis, China Agricultural University, Beijing,
China. (in Chinese with English abstract).
Li, G. T., Li, B. G., and Chen, D. L. (2002). Method for measurement of ammonia
volatilization from large area field by Bowen Ratio System. J. Chin. Agric. Univ. 6,
56–62. (in Chinese with English abstract).
Li, L., Li, S. M., Sun, J. H., Zhou, L. L., Bao, X. G., Zhang, H. G., and Zhang, F. S. (2007).
Diversity enhances agricultural productivity via rhizosphere phosphorus facilitation on
phosphorus-deficient soils. Proc. Natl. Acad. Sci. USA 104, 11192–11196.
Li, Z. D., Wang, Y. H., and Li, L. Y. (2009). Orchard soil acidification status in Jiaodong
peninsula of Shandong province and control techniques. Phosphate Compound Fertil. 24
(6), 80–81. (in Chinese with English abstract).
Li, H., Zhang, W., Zhang, F., Du, F., and Li, L. (2010). Chemical fertilizer use and
efficiency change of main grain crops in China. Plant Nutr. Fertil. Sci. 16(5),
1136–1143. (in Chinese with English abstract).
Li, H., Huang, G., Meng, Q., Ma, L., Yuan, L., Wang, F., Zhang, W., Cui, Z., Shen, J.,
Chen, X., Jiang, R., and Zhang, F. (2011). Integrated soil and plant phosphorus
management for crop and environment in China. A review. Plant Soil 349, 157–167.
10.1007/s11104-011-0909-5.
Liang, J. (2011). Study on the Situation of Nitrogen Input and Output and the Potential of
Fertilizer Saving in Vegetable System in China. Master Thesis, China Agricultural
University, Beijing, China. (in Chinese with English abstract).
Linkohr, B. I., Williamson, L. C., Fitter, A. H., and Leyser, H. M. O. (2002). Nitrate and
phosphate availability and distribution have different effects on root system architecture
of Arabidopsis. Plant J. 29, 751–760.
Liu, X. J., and Zhang, F. S. (2009). Nutrient from environment and its effects in nutrient
management of ecosystems—A case study on atmospheric nitrogen deposition. Arid Zone
Res. 3, 306–311. (in Chinese with English abstract).
Liu, H. J., Ju, X. T., and Tong, Y. A. (2002). The status and problems of fertilization of main
fruit trees in Shaanxi Province. Agric. Res. Arid Areas 20(1), 38–44. (in Chinese with
English abstract).
Liu, X., Ju, X., Zhang, F., Pan, J., and Christie, P. (2003). Nitrogen dynamics and budgets in
a winter wheat–maize cropping system in the North China Plain. Field Crop Res. 83,
111–124.
Liu, H. B., Li, Z. H., and Zhang, Y. G. (2004a). Characteristics of nitrate distribution and
accumulation in soil profiles under main agro-land use types in Beijing. Sci. Agric. Sin. 37,
692–698. (in Chinese with English abstract).
Liu, Y., Mi, G., Chen, F., Zhang, J., and Zhang, F. (2004b). Rhizosphere effect and root
growth of two maize (Zea mays L.) genotypes with contrasting P efficiency at low P
availability. Plant Sci. 167, 217–223.
Liu, J. L., Liao, W. H., Zhang, Z. H., and Sun, J. S. (2006a). The change of soil nutrition and
the status of distribution in the apple orchard in the south and central part of Hebei
province. Acta Hortic. Sin. 33, 705–708. (in Chinese with English abstract).
Liu, X. J., Ju, X. T., Zhang, Y., Kopsch, J., and Zhang, F. S. (2006b). Nitrogen deposition in
agroecosystems in the Beijing area. Agric. Ecosyst. Environ. 113, 370–377.
Nutrient Management in China 37
Liu, X. J., Song, L., He, C. E., and Zhang, F. S. (2010). Nitrogen deposition as an important
nutrient from the environment and its impact on ecosystems in China. J. Arid Land 2,
137–143. (in Chinese with English abstract).
Liu, X. J., Duan, L., Mo, J. M., Du, E. Z., Shen, J. L., Lu, X. K., Zhang, Y., Zhou, X. B.,
He, C. E., and Zhang, F. S. (2011). Nitrogen deposition and its ecological impact in
China: An overview. Environ. Pollut. 159, 2251–2264.
Lu, S. C., Chen, Q., and Zhang, F. S. (2008). Analysis of nitrogen input and soil nitrogen
load in orchards of Hebei province. Plant Nutr. Fertil. Sci. 14(5), 858–865. (in Chinese
with English abstract).
Lynch, J. P., and Brown, K. M. (2008). Root strategies for phosphorus acquisition. In “The
Ecophysiology of Plant-Phosphorus Interactions” (P. J. White and J. P. Hammond,
Eds.), pp. 83–116. Springer, Dordrecht, The Netherlands.
Ma, W. J. (2010). The Growth, Nutrient Uptake and Accumulation in Different Cash
Crops. PhD Thesis, Northwest A & F University, Yangling, Shaanxi, China (in Chinese
with English abstract)..
Ma, W. Q., Mao, D. R., and Zhang, F. S. (1999). Current status and evaluation of crop
fertilization in Shandong province. Chin. J. Soil Sci. 30, 217–220. (in Chinese with
English abstract).
Magdoff, F. R., Ross, D., and Amadon, J. (1984). A soil test for nitrogen availability to corn.
Soil Sci. Soc. Am. J. 48, 1301–1304.
Marschner, H. (1995). Mineral Nutrition in Higher Plants. Academic Press, London.
Matson, P. A., Parton, W. J., Power, A. G., and Swift, M. J. (1997). Agricultural intensifi-
cation and ecosystem properties. Science 277, 504–509.
Matson, P. A., Naylor, R., and Ortiz-Monasterio, I. (1998). Integration of environmental,
agronomic, and economic aspects of fertilizer management. Science 280, 112–115.
Meisinger, J. J., Bandel, V. W., Angle, J. S., O’Keefe, B. E., and Reynolds, C. M. (1992).
Presidedress soil nitrate test evaluation in Maryland. Soil Sci. Soc. Am. J. 56, 1527–1532.
Neumann, G., and Römheld, V. (2002). Root-induced changes in the availability of
nutrients in the rhizosphere. In “Plant Roots, The Hidden Half” (Y. Waisel, A. Eshel,
U. Kafkafi, et al., Eds.), pp. 617–649. Marcel Dekker, Inc., New York.
Peng, L. H. (2006). Application of Real-Time Testing of Nitrate Technique in Nitrogen
Management of Cucumber in Greenhouse. PhD Thesis, China Agricultural University,
Beijing, China. (in Chinese with English abstract).
Peng, S., Buresh, R., Huang, J., Zhong, X., Zou, Y., Yang, J., Wang, G., Liu, Y., Hu, R.,
and Tang, Q. (2010). Improving nitrogen fertilization in rice by site-specific N manage-
ment. A review. Agron. Sustain. Dev. 30, 649–656.
Raun, W. R., and Johnson, G. V. (1999). Improving nitrogen use efficiency for cereal
production. Agron. J. 91, 357–363.
Raun, W. R., Solie, J. B., Johnson, G. V., Stone, M. L., Mullen, R. W., Freeman, K. W.,
Thomason, W. E., and Lukina, E. V. (2002). Improving nitrogen use efficiency in cereal
grain production with optical sensing and variable rate application. Agron. J. 94, 815–820.
Ren, H. Q. (2007). Characteristics of Nitrogen Input and in Peach Orchards and Effects on
Nitrate Content of Groudwater in the Farms of Pinggu Arable Land. Master Thesis,
China Agricultural University, Beijing, China (In Chinese with English abstract).
Ren, T., Christie, P., Wang, J. G., Chen, Q., and Zhang, F. S. (2010). Root zone soil
nitrogen management to maintain high tomato yields and minimum nitrogen losses to
the environment. Sci. Hortic. 125, 25–33.
Robertson, G. P., and Vitousek, P. M. (2009). Nitrogen in agriculture: Balancing the cost of
an essential resource. Annu. Rev. Environ. Resour. 34, 97–125.
Roelcke, M., Han, Y., Schleefk, K. H., Zhu, J. G., Liu, G., Cai, Z. C., and Richter, J.
(2004). Recent trends and recommendations for nitrogen fertilization in intensive
agriculture in eastern China. Pedosphere 14, 449–460.
38 Fusuo Zhang et al.
Roth, G. W., and Fox, R. H. (1990). Soil nitrate accumulation following nitrogen fertilized
corn in Pennsylvania. J. Environ. Qual. 9, 243–248.
Sattelmacher, B., and Thoms, K. (1995). Morphology and physiology of the seminal root
system of young maize (Zea mays L.) plants as influenced by a locally restricted nitrate
supply. Z. Pflanz. Bodenkunde. 158, 493–497.
Schachtman, D. P., Reid, R. J., and Ayling, S. M. (1998). Phosphorus uptake by plants:
From soil to cell. Plant Physiol. 116, 447–453.
Schenk, M. K., and Barber, S. A. (1979). Root characteristics of corn genotypes as related to
P uptake. Agron. J. 71, 921–924.
Schleef, K. H., and Kleihanss, W. (1994). Mineral Balance in Agriculture in EU. Institute of
Farm Economics, Federal Agricultural Research Centre, Brounchweig, Germany.
Schmitt, M. A., and Randall, G. W. (1994). Developing a soil nitrogen test for improved
recommendations for corn. J. Prod. Agric. 7, 328–334.
Shen, J., Li, H., Neumann, G., and Zhang, F. (2005). Nutrient uptake, cluster root
formation and exudation of protons and citrate in Lupinus albus as affected by localized
supply of phosphorus in a split-root system. Plant Sci. 168, 837–845.
Shen, J. L., Tang, A. H., Liu, X. J., Fangmeier, A., Goulding, K. W. T., and Zhang, F. S.
(2009). High concentrations and dry deposition of reactive N species at two sites in the
North China Plain. Environ. Pollut. 157, 3106–3113.
Shen, J., Liu, X., Fangmeier, A., Goulding, K., and Zhang, F. (2011a). Atmospheric
ammonia and particulate ammonium from agricultural sources in the North China
Plain. Atmos. Environ. 45, 5033–5041.
Shen, J., Yuan, L., Zhang, J., Li, H., Bai, Z., Chen, X., Zhang, W., and Zhang, F. (2011b).
Phosphorus dynamics: From soil to plant. Plant Physiol. 156, 997–1005.
Shi, W. M., Yao, J., and Yan, F. (2008). Vegetable cultivation under greenhouse conditions
leads to rapid accumulation of nutrients, acidification and salinity of soils and ground-
water contamination in South-Eastern China. Nutr. Cycl. Agroecosyst. 83, 73–84.
Shorrocks, V. M., and Nicholson, D. D. (1980). The influence of boron deficiency on fruit
quality. Acta Hortic. (ISHS) 92, 103–110.
Sogbedji, J. M., Van Es, H. M., and Yang, C. L. (2000). Nitrate leaching and nitrogen
budget as affected by maize nitrogen rate and soil type. J. Environ. Qual. 29, 1813–1820.
Song, X. Z., Zhao, C. X., Wang, X. L., and Li, J. (2009). Study of nitrate leaching nitrogen fate
under intensive vegetable production pattern in northern China. C. R. Biol. 332, 385–392.
Tang, L. L., Chen, Q., Li, X. L., Chen, Y. Z., and Ding, G. G. (2005). Studies on the target
value of nitrogen supply for greenhouse tomato growth during autumn-winter season.
Plant Nutr. Fertil. Sci. 11(2), 230–235. (in Chinese with English abstract).
Tian, Q. Y., Chen, F. J., Zhang, F. S., and Mi, G. H. (2006). Genotypic difference nitrogen
acquisition ability in maize plants is related to the coordination leaf and root growth.
J. Plant Nutr. 29, 317–330.
Tian, Q. Y., Chen, F. J., Liu, J. X., Zhang, F. S., and Mi, G. H. (2008). Inhibition of maize
root growth by high nitrate supply is correlated to reduced IAA levels in roots. J. Plant
Physiol. 165, 942–951.
Tilman, D. (1999). Global environmental impacts of agricultural expansion: The need for
sustainable and efficient practices. Proc. Natl. Acad. Sci. USA 96, 5995–6000.
Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R., and Polasky, S. (2002). Agricultural
sustainability and intensive production practices. Nature 418, 671–678.
Tong, T. A., Emteryd, O., and Zhang, S. L. (2004). Evaluation of over-application of
nitrogen fertilizer in China’s Shanxi province. Sci. Agric. Sin. 37, 1239–1244. (in Chinese
with English abstract).
Vance, C. P. (2008). Plants without arbuscular mycorrhizae. In “The Ecophysiology of
Plant-Phosphorus Interactions” (P. J. White and J. P. Hammond, Eds.), pp. 117–142.
Springer, Dordrecht, The Netherlands.
Nutrient Management in China 39
Vitousek, P. M., Naylor, R., Crews, T., David, M. B., Drinkwater, L. E., Holland, E.,
Johnes, P. J., Katzenberger, J., Martinelli, L. A., Matson, P. A., Nziguheba, G.,
Ojima, D., et al. (2009). Nutrient imbalances in agricultural development. Science 324,
1519–1520.
Wang, J. Q. (2007). Analysis and Evaluation of Yield Increase of Fertilization and Nutrient
Utilization Efficiency for Major Cereal Crops in China. PhD Thesis, China Agricultural
University, Beijing, China (in Chinese with English abstract).
Wang, J. G. (2011). Management of Degraded Vegetable Soils in Greenhouse. China
Agricultural University Press, Beijing, China (in Chinese).
Wang, X., Cao, Y., Zhang, F., and Chen, X. (1995). Application of building-up and
maintenance approach in agriculture. Plant Nutr. Fertil. Sci. 1, 59–63. (in Chinese with
English abstract).
Wang, L. J., Jiang, W. B., He, Q. F., and Fan, H. B. (2001). Studies on the relationship of
development of bitter pit in apple fruits with the contents of calcium and magnesium and
the activities of antioxidant enzymes. Acta Hortic. Sin. 28(3), 200–205. (in Chinese with
English abstract).
Wang, Y., Mi, G. H., Chen, F. J., and Zhang, F. (2003). Genotypic differences uptake by
maize inbred lines its relation to root morphology. Acta. Ecol. Sin. 23, 297–302.
Wang, B. L., Shen, J., Zhang, W. H., Zhang, F. S., and Neumann, G. (2007). Citrate
exudation from white lupin induced by phosphorus deficiency differs from that induced
by aluminum. New Phytol. 176, 581–589.
Wang, J. Y., Liu, Q. H., and Liu, J. L. (2010a). Analysis on the characteristic and cause of
orchard soil acidification in the area of Shandong peninsula. Chin. Agric. Sci. Bull. 26(16),
164–169. (in Chinese with English abstract).
Wang, X. Y., Hang, B., and Liu, C. L. (2010b). Distribution of calcium in bagged apple fruit
and relationship between antioxidant enzyme activity and bitter pit. Agric. Sci. Technol. 11
(1), 82–85. (in Chinese with English abstract).
Wehrmann, J. V., and Scharpf, H. C. (1979). Mineral nitrogen in soil as an indicator for
nitrogen fertilizer requirements (Nmin-method). Plant Soil 52(1), 109–126.
Xu, S. Y., Zhang, F. S., and Wang, H. (2008). Effects of environmental factors on internal
bark necrosis of apple trees. J. Fruit Sci. 25(1), 73–77. (in Chinese with English abstract).
Ye, Y. L., Yang, S. Q., Liu, S. L., and Wang, W. L. (2008). Study on vegetable production,
fertilizer application, soil chemical and physical property variance in suburb of Zhengz-
hou city. Henan Sci. 26(1), 51–55. (in Chinese with English abstract).
Yoshida, S. (1981). Fundamentals of Rice Crop Science. International Rice Research
Institute, Los Baños, Philippines. 269 pp.
Zhang, X. M. (2010). Effect of Fertilization on the Yield and the Apparent N Balance for
Greenhouse Tomato in Shouguang. Master Thesis, China Agricultural University, Beij-
ing, China. (in Chinese with English abstract).
Zhang, H. M., and Forde, B. G. (1998). An Arabidopsis MADS box gene that controls
nutrient-induced changes in root architecture. Science 279, 407–409.
Zhang, S. L., Cai, G. X., Wang, X. Z., Xu, Y. H., Zhu, Z. L., and Freney, J. R. (1992). Loss
of urea-nitrogen applied to maize grown on a calcareous fluvo-aquic soil in North China
Plain. Pedosphere 2, 171–178.
Zhang, F. S., Shen, J. B., Li, L., and Liu, X. (2004). An overview of rhizosphere processes
related with plant nutrition in major cropping systems in China. Plant Soil 260, 89–99.
Zhang, X. S., Liao, H., Chen, Q., Christie, P., Li, X. L., and Zhang, F. S. (2007). Response
of tomato on calcareous soils to different seedbed phosphorus application rates. Pedosphere
17, 70–76.
Zhang, F. S., Wang, J. Q., Zhang, W. F., Cui, Z. L., Ma, W. Q., Chen, X. P., and Jiang, R. F.
(2008a). Nutrient use efficiencies of major cereal crops in China and measures for
improvement. Acta Pedolog. Sin. 45, 915–924. (in Chinese with English abstract).
40 Fusuo Zhang et al.
Zhang, Y., Liu, X. J., Fangmeier, A., Goulding, K. T. W., and Zhang, F. S. (2008b).
Nitrogen inputs and isotopes in precipitation in the North China Plain. Atmos. Environ.
42, 1436–1448.
Zhang, F. S., Chen, X. P., and Chen, Q. (2009). The Fertilization Guideline of Major Crop
in China. China Agricultural University Press, Beijing. (in Chinese).
Zhang, F., Shen, J., Zhang, J., Zuo, Y., Li, L., and Chen, X. (2010). Rhizosphere processes
and management for improving nutrient use efficiency and crop productivity: Implica-
tions for China. Adv. Agron. 107, 1–32.
Zhang, F. S., Cui, Z. L., Fan, M. S., Zhang, W. F., Chen, X. P., and Jiang, Q. F. (2011).
Integrated soil-crop system management: Reducing environmental risk while increasing
crop productivity and improving nutrient use efficiency in China. J. Environ. Qual. 40,
1–7.
Zhao, X., and Xing, G. (2009). Variation in the relationship between nitrification and
acidification of subtropical soils as affected by the addition of urea or ammonium sulfate.
Soil Biol. Biochem. 41, 2584–2587.
Zhao, R., Chen, X., Zhang, F., Zhang, H., Schroder, J., and Römheld, V. (2006).
Fertilization and nitrogen balance in a wheat–maize rotation system in North China.
Agron. J. 98, 938–945.
Zhao, T. S., Yu, L. C., and Jiao, R. (2007). Study on the relationship between calcium
nutrition and bitter pit in bagged apples. J. Fruit Sci. 24(5), 649–652. (in Chinese with
English abstract).
Zheng, X. H., Han, S. H., Huang, Y., Wang, Y. S., and Wang, M. X. (2004). Re-
quantifying the emission factors based on field measurements and estimating the direct
N2O emission from Chinese croplands. Glob. Biogeochem. Cycl. 18, 1–19. GB2018.
Zhou, J. P. (2009). The problems of manure fertilization of orchards and correct methods.
Bull. Agric. Sci. Technol. 3, 151–152. (in Chinese with English abstract).
Zhu, Z. L. (1998). The status, problems and countermeasures of nitrogen fertilizer applica-
tion in China (in Chinese). In “Fertilizer Issues of Sustainable Agriculture Development
in China” (Q. K. Li, et al., Eds.), pp. 28–51. Jiangsu Science and Technology Publishing,
Nanjing, Jiangsu.
Zhu, J. H., Li, X. L., Christie, P., and Li, J. L. (2005). Environmental implications of low
nitrogen use efficiency in excessively fertilized hot pepper (Capsicum frutescens L.)
cropping systems. Agr Ecosyst Environ 111, 70–80.
C H A P T E R T W O
Contents
1. Introduction 42
2. Vegetative Development 43
2.1. Germination and emergence 43
2.2. Leaf appearance and leaf number 44
3. Canopy Expansion and Growth Processes 45
3.1. Leaf thickness 45
3.2. Leaf area and stem elongation 46
3.3. Leaf senescence 47
3.4. Stomatal conductance and transpiration 47
3.5. Photosynthesis 48
3.6. Net assimilation and growth rates 49
4. Reproductive Development and Growth 50
4.1. Appearance of flowers, pegs, and pods 50
4.2. Rate of flower production 51
4.3. Pollen production and viability and fruit-set 53
4.4. Number of pegs, pods, and seeds 54
4.5. Pod and seed growth rates and their size 55
5. Total Dry Matter, Pod, and Seed Yield 56
6. Harvest Index and Shelling Percentage 58
6.1. Harvest index 58
6.2. Shelling percentage 59
7. Root Growth and Root-to-Shoot Ratio 59
7.1. Root growth 59
7.2. Root-to-shoot ratio 60
* International Crops Research Institute for the Semi-Arid Tropics (ICRISAT), Patancheru, Andhra Pradesh,
India
{
Agronomy Department, University of Florida, Gainesville, Florida, USA
41
42 Uttam Kumar et al.
Abstract
Global warming is changing climate in terms of increased frequency of extreme
weather events as well as increased air temperature and vapor pressure deficit
of air and spatial and temporal change in rainfall. In spite of the beneficial effect
of increased atmospheric CO2 concentration, climate change will adversely
impact the production and productivity of groundnut grown in subtropical and
tropical regions of the world. This chapter reviews the current state of knowl-
edge on effects of climate change factors on the growth and development of
groundnut. This review identifies research gaps and suggests upgrades to
groundnut models, such as the CROPGRO-Groundnut model, which is being
used as a tool to assess impacts of climate change on groundnut crop. This
review revealed that the direct and indirect effects of most climate change
factors on plant growth and development processes are well understood and
already incorporated in the CROPGRO-Groundnut model. Extreme events asso-
ciated with climate change may sometimes cause water-logging, extreme soil
water deficiency, or extreme humidity conditions, and these effects could be
better addressed in the models.
1. Introduction
The Fourth Assessment report of the Inter-Governmental Panel on
Climate Change (IPCC, 2007) has reconfirmed that the atmospheric con-
centrations of carbon dioxide, methane, and nitrous oxide greenhouse gases
(GHGs) have increased markedly since 1750. The global increases in CO2
concentrations are due primarily to fossil-fuel use and land-use change,
while those of methane and nitrous oxide are primarily due to agriculture.
The IPCC has also shown that these increases in GHGs have resulted in
warming of the climate system by 0.74 C over the past 100years, and the
projected increase in temperature by 2100 is about 1.8–4.0 C. For the
South Asia region, the IPCC has projected 0.5–1.2 C rise in temperature
by 2020, 0.88–3.16 C by 2050, and 1.56–5.44 C by 2080 depending upon
the scenario of future development. Overall, the temperature increases are
Effect of Climate Change Factors on Processes of Crop Growth 43
likely to be much higher in winter season than in rainy season. With climate
change, more frequent hot days, heat waves, and warm spells are expected
to increase. These increases in the temperatures are likely to result in both
spatial and temporal variations in rainfall. Overall, there will be an increase
in rainfall especially in the tropical regions. The pattern of precipitation is
already changing and will become more erratic and intense with warming of
the globe. Because of the increase in temperatures, vapor pressure deficit of
the air will increase in spite of the increase in humidity with the increase in
rainfall. For the A1B SRES scenario, the expected increase in CO2 concen-
tration will be 420ppm by 2020, 530ppm by 2050, and 650ppm by 2080 as
estimated by the SPAM model (IPCC, 2001).
These changes in climatic factors (CO2, temperature, vapor pressure defi-
cit, and rainfall) will alter plant growth and development processes and most
likely have negative impact on crop productivity, especially in the semiarid
tropical regions, where the current temperatures are already high and close to
the upper limits beyond which the plant processes will be adversely affected.
Therefore, in spite of some expected benefits of the increased CO2 concentra-
tion on some crops, global warming poses a potential threat to agricultural
production and productivity throughout the world. Increased incidence of
weeds, pests, and plant diseases with climate change may cause even greater
economic losses to agricultural production. It is projected that even a small rise
in temperature (1–2 C) at lower latitudes, especially in the seasonally dry
tropical regions (IPCC, 2007), would decrease crop productivity.
Groundnut (Arachis hypogaea L.) is one of the major oilseed and food
crops grown in subtropical and tropical regions of the world. It is grown in
different rainfall and temperature regimes on a variety of soils. Being a C3
crop, higher temperatures and other climatic factors may affect its produc-
tivity and to some extent its distribution. This chapter attempts to review
the current state of knowledge of climate factor effects on growth and
development response of groundnut and revisits the need to fine-tune the
CROPGRO and other groundnut models to determine the impacts and
adaptation of groundnut to the climate change in future.
2. Vegetative Development
2.1. Germination and emergence
After groundnut seeds are sown, germination and emergence are primarily
determined by the temperature and soil moisture in the seeding zone.
The processes of germination and emergence have a minimum threshold
value, optimum range, and maximum threshold value for both temperature
and soil moisture contents. At minimum threshold values of temperature
(base temperature) and soil moisture content, the processes of germination are
44 Uttam Kumar et al.
not initiated. At the optimum range of temperature and soil moisture both,
germination and emergence takes place at a maximum rate. Between their
minimum threshold and lower optimum values, the rates of germination and
emergence increase with the increase in temperature and soil moisture.
Above their optimum range, these processes are progressively slowed down
until they completely stop at their respective maximum threshold values
(damaging thresholds). For example, Awal and Ikeda (2002) and Prasad
et al. (2006) reported that base temperature for germination of groundnut is
approximately 10 C and the optimum temperature (OT) for emergence is
between 25 and 30 C. Mohamed et al. (1988) and Angus et al. (1981)
reported base temperatures ranging from 8 to 13 C for groundnut seed
germination. These differences in base temperature suggest genotypic differ-
ence among cultivars studied. In terms of soil temperature, the optimum
mean soil temperature for seed germination is between 29 and 30 C
(Mohamed et al., 1988) and for root growth it is close to 30 C (Suzuki,
1966). Leong and Ong (1983) also reported that in two cooler (wet) soil
temperatures (19 and 22 C) less than 50% emergence of groundnut seedling
took place, while at warmer temperatures (25, 28, and 31 C) the percentage
of emergence varied from 70% to 80%. Seedling emergence started within
5 days after sowing (DAS) in warm temperatures but in 10 DAS at 19 C.
treatment, the maximum leaf area index (LAI) reached 7.5 in ambient CO2
and 8.0 in elevated CO2 at the end of the season. Under drought conditions,
elevated CO2 had a highly significant effect on canopy development. Plants
achieved a maximum LAI of 3 in ambient CO2 and 4.3 in elevated CO2.
Later when the drought conditions intensified, LAI declined to 1.9 in the
ambient CO2 and 3.0 in the elevated CO2. Groundnut plants grown under
elevated CO2 in drought conditions maintained less negative leaf water
potential than the plants grown in ambient CO2, which helped in main-
taining the turgor potential for growth and expansion of leaves. These
results showed that elevated CO2 benefits the crop growth under both
water limiting and nonlimiting conditions; however, the relative benefits
are more under water limiting conditions (something that model simula-
tions also show).
3.5. Photosynthesis
Talwar et al. (1999) recorded higher net photosynthetic rate in three
groundnut genotypes grown at 35/30 C as compared to those grown at
25/25 C at 30 and 60 DAS. They also observed genotypic differences in net
photosynthesis at both temperatures. In crops like groundnut (C3 crops),
Rubisco is not saturated by the current concentration of CO2 in the
atmosphere. So an increase in CO2 concentration will improve the balance
of CO2 and O2 at Rubisco site, thus improving the CO2-exchange rate
(CER) of the plant by providing more substrate for photosynthesis. Prasad
et al. (2003) reported that doubling of ambient CO2 concentration (350 vs.
700mmolmol1) enhanced leaf photosynthesis of groundnut by 27% across
a range of daytime temperatures (32–44 C), but they found no CO2 by
temperature interaction on leaf photosynthesis. On the other hand, some
researchers have suggested that optimum growth temperature for several
plants may rise significantly with increasing concentration of atmospheric
CO2 (Berry and Bjorkman, 1980; McMurtrie and Wang, 1993; McMurtrie
et al., 1992; Stuhlfauth and Fock, 1990). Long (1991) calculated from well-
established plant physiological principles that most C3 plants should increase
their OT for growth by approximately 5 C with 300ppm increase in
CO2 concentration. Thus, photosynthetic rates are expected to rise with
Effect of Climate Change Factors on Processes of Crop Growth 49
800mmolmol1 CO2, indicates that the HT regime chosen in this study was
still in the OT range for biomass production of groundnut. Rao (1999)
reported increased dry weight of shoot in elevated CO2 (660 vs. 300ppm)
even at 40 C.
Prasad et al. (2003) reported increase in total dry matter production of
groundnut with increase in CO2 between temperatures of 32/22 and
40/30 C. Further increase in temperature to 44/34 C decreased total dry
matter under both ambient (350mmolmol1) and elevated CO2 (700mmol
mol1). As the temperature increased from 32/22 to 44/34 C, pod yield
decreased by 89% and 87% under ambient and elevated CO2, respectively.
With the same increase in temperature, the seed yield decreased by 90% and
88% under ambient and elevated CO2, respectively. Temperature and CO2
effect on total dry matter, pod, and seed yields was statistically significant;
however, the interaction between temperature and CO2 for all yields was
not significant. On average, total dry matter yield increased by 36%, and
both pod and seed yields increased by 30% under elevated CO2 across all the
temperature regimes. The study showed that when the groundnut crop is
exposed to HTs throughout the full season, total dry matter production is
reduced at temperatures above 40/30 C (35 C), whereas the pod and seed
yields are adversely affected above temperatures of 32/22 C (27 C). These
results differ from the Cox (1979) study results that OT for dry matter
production ranges from 25 to 30 C with a mean of 27.5 C, whereas the
pod and seed yields start declining above 24 C. The study of Cox (1979)
used pot-grown plants at lower light intensity.
Clifford et al. (1993) reported that, in well-irrigated conditions,
elevated CO2 (700ppm) increased above-ground dry matter accumulation
by an average of 16% over the ambient CO2 concentration (350ppm).
Droughted plants grown at elevated CO2 produced more than double the
dry matter of plants grown at ambient CO2. Average increase in pod yield
with elevated CO2 was 25%, from 2.73 to 3.42tha1 in well-irrigated
plots, with a sixfold increase from 0.22 to 1.34tha1 in the droughted
treatment. The reason for such differential response to CO2 in two
moisture regimes was discussed earlier as a result of CO2-induced water
conservation in the section on stomatal conductance and photosynthesis.
Timing and intensity of water stress can enhance or reduce yield of
groundnut. Rao et al. (1985) reported that when groundnut plants received
12–15% less water during vegetative growth (or up to start of pegging) pod
yields increased by 12–19% compared to the fully irrigated control. Earlier
work at ICRISAT (ICRISAT Annual Report, 1981); Ong (1984) showed
similar increase in pod yield under mild water stress during vegetative phase
of groundnut. In the Rao et al. (1985) study when plants were stressed from
start of flowering to start of seed growth, total biomass and pod yield were
reduced as much as 50% and 77%, respectively. Greatest reduction in kernel
yield occurred when stress was imposed during the seed-filling phase.
58 Uttam Kumar et al.
As fruit initiation continues even after the start of kernel growth, soil water
deficits during pod-filling stage reduce both the initiation and the develop-
ment of pods (Boote et al., 1976; Matlock et al., 1961; Ono et al., 1974;
Pallas et al., 1979; Underwood et al., 1971).
weight was 3.28–3.67gplant1 when applied from pod to harvest and 2.79–
3.41gplant1 when applied from seed filling to harvest stage. Similarly, Rao
(1999) using open top chamber observed increase in dry weight of root with
CO2 enrichment from 330 to 660ppm at 35 to 40 C. Pilumwong et al.
(2007) in a rhizotron study observed that when CO2 concentration was
increased from 400 to 800mmolmol1 the fibrous root dry weight of ground-
nut plants increased at 25/15 C but decreased at 35/25 C. Clifford et al.
(1993) using closed environment glasshouse observed that under ambient
(350ppm) and elevated (700ppm) CO2 the dry root weights were 180.2 and
177.3gm2 in the irrigated treatment and 274.0 and 274.7gm2 in the
drought treatment in the respective CO2 concentrations. This indicates that
root dry weight was unaffected by CO2 at a given moisture regime but was
increased by drought. These differences in root weight response to CO2 in
different studies may be attributed to the differences in the crop growth
facility used for experimentation.
9. Concluding Comments
Groundnut (A. hypogaea L.) is one of the major oilseed and food crops of
the subtropical and tropical regions of the world. It is grown in different rainfall
and temperature regimes on a variety of soils. Depending upon the location on
the globe, climate change may benefit or adversely affect the productivity of
Effect of Climate Change Factors on Processes of Crop Growth 65
this crop. This chapter has reviewed the current state of knowledge on effects
of climate change factors, such as extremes of air and soil temperatures, relative
humidity, water availability and their interactions with photoperiod, light
intensity, and increased atmospheric CO2 concentration, on the growth and
development of groundnut. This review identified research gaps and needs to
generate information to upgrade the CROPGRO-Groundnut model. This
review revealed that the direct and indirect effects of most climate change
factors on plant growth and development processes are well understood and
already incorporated in the model. Extreme events associated with climate
change such as water-logging, extreme soil water deficiency or extreme
humidity conditions will affect the productivity of the crop. Low-light inten-
sity affects flowering and high air and soil temperatures affect root growth and
root-to-shoot ratio. The effects of these factors on groundnut crop growth and
development need to be sufficiently understood before these are suitably
incorporated in the model to enhance its capability for better assessment
of climate change impacts and to develop adaptation strategies to cope up
with climate change in different agro-climates. Direct comparison of model
simulations against experimental data reported in some studies listed in this
review would be useful.
ACKNOWLEDGMENT
We are grateful to ICRISAT for providing financial support through the USAID
linkage fund.
REFERENCES
Ahmed, F. E., Hall, A. E., and DeMason, D. A. (1992). Heat injury during f loral develop-
ment in cowpea (Vigna unguiculata, Fabaceae). Am. J. Bot. 79, 784–791.
Angus, J. F., Cunningham, R. B., Moncur, M. W., and MacKenzeie, D. H. (1981). Phasic
development in field crops. I. Thermal response in the seedling phase. Field Crop Res. 3,
365–378.
Awal, M. A., and Ikeda, T. (2002). Effects of changes in soil temperature on seedling
emergence and phenological development in field-grown stands of peanut (Arachis
hypogaea). Environ. Exp. Bot. 47, 101–113.
Bagnall, D. J., and King, R. W. (1991a). Response of peanut (Arachis hypogaea) to tempera-
ture, photoperiod and irradiance 1. Effect on flowering. Field Crop Res. 26, 263–277.
Bagnall, D. J., and King, R. W. (1991b). Response of peanut (Arachis hypogaea) to tempera-
ture, photoperiod, and irradiance. 2. Effect on peg and pod development. Field Crop Res.
26, 279–293.
Bell, M. J., Bagnall, D. J., and Harch, G. (1991). Effect of photoperiod on reproductive
development of peanut (Arachis hypogaea L.) in a cool subtropical environment. II.
Temperature interactions. Aust. J. Agric. Res. 42, 1151–1161.
Berry, J., and Bjorkman, O. (1980). Photosynthetic response and adaptation to temperature
in higher plants. Annu. Rev. Plant Physiol. 31, 491–543.
66 Uttam Kumar et al.
Bolhuis, G. G., and De Groot, W. (1959). Observations on the effect of varying temperatures
on the flowering and fruit set in three varieties of groundnut. Neth. J. Agric. Sci. 7, 317–326.
Boote, K. J., and Pickering, N. B. (1994). Modeling photosynthesis of row crop canopies.
HortScience 29, 1423–1434.
Boote, K. J., Varnell, R. J., and Duncan, W. G. (1976). Relationships of size, osmotic
concentration, and sugar concentration of peanut pods to soil water. Soil Crop Sci. Soc.
Fla. Proc. 35, 47–50.
Boote, K. J., Jones, J. W., Mishoe, J. W., and Wilkerson, G. G. (1986). Modeling growth
and yield of groundnut. In “Agrometeorology of Groundnut: Proceedings of an Interna-
tional Symposium, 21–26 Aug 1985, ICRISAT Sahelian Center, Niamey, Niger,”
pp. 243–254. ICRISAT, Patancheru, A.P., India.
Boote, K. J., Jones, J. W., and Singh, P. (1991). Modeling growth and yield of groundnut—
State of the art. In “Groundnut—A Global Perspective: Proceedings of an International
Workshop, 25–29 Nov 1991,” pp. 331–343. ICRISAT Center, India.
Boote, K. J., Jones, J. W., Hoogenboom, G., and Pickering, N. B. (1998). The CROPGRO
model for grain legumes. In “Understanding Options for Agricultural Production”
(G. Y. Tsuji, G. Hoogenboom, and P. K. Thornton, Eds.), pp. 99–128. Kluwer
Academic Publishers, Dordrecht.
Boote, K. J., Allen, L. H. Jr., Vara Prasad, P. V., and Jones, J. W. (2010). Testing effects of
climate change in crop models. In “Handbook of Climate Change and Agroecosystems”
(D. Hillel and C. Rosenzweig, Eds.), pp. 109–129. Imperial College Press, London UK.
Chen, J. J., and Sung, J. M. (1990). Gas exchange rate and yield response of Virginia-type
peanut to carbon dioxide enrichment. Crop Sci. 30, 1085–1089.
Clifford, S. C., Stronach, I. M., Mohamed, A. D., Azam-Ali, S. N., and Crout, N. M. J. (1993).
The effects of elevated atmospheric carbon dioxide and water stress on light interception,
dry matter production and yield in stands of groundnut (Arachis hypogaea L.). J. Exp. Bot. 44,
1763–1770.
Clifford, S. C., Black, C. R., Roberts, J. A., Stronach, I. M., Singleton-Jones, P. R.,
Mohamed, A. D., and Azam-Ali, S. N. (1995). The effect of elevated atmospheric
CO2 and drought on stomatal frequency in groundnut (Arachis hypogaea L.). J. Exp.
Bot. 46, 847–852.
Cox, F. R. (1979). Effect of temperature treatment on peanut vegetative and fruit growth.
Peanut Sci. 6, 114–117.
Craufurd, P. Q., Prasad, P. V. V., and Summerfield, R. J. (2002). Dry matter production and
rate of change of harvest index at high temperature in peanut. Crop Sci. 42, 146–151.
Dreyer, J., Duncan, W. G., and McClaud, D. E. (1981). Fruit temperature growth and yield
of peanut. Crop Sci. 21, 686–688.
Flohr, Marie-Luise, Williams, J. H., and Lenz, F. (1990). The effect of photoperiod on
the reproductive development of a photoperiod sensitive groundnut (Arachis hypogea L.)
cv. NC Ac 17090. Exp. Agric. 26, 397–406.
Fortanier, E. J. (1957). Control of flowering in Arachis hypogaea L. PhD. Thesis. Mededelin-
gen van de Landouwhoogexhool te Wageningen, The Netherlands.
Golombek, S. D., and Johansen, C. (1997). Effect of soil temperature on vegetative and
reproductive growth and development in three Spanish genotype of peanut (Arachis
hypogaea L.). Peanut Sci. 24, 67–72.
Hardy, R. W. F., and Havelka, U. D. (1977). Possible routes to increase the conversion of
solar energy to food and feed by grain legumes and cereal grains (crop production): CO2
and N fixation, foliar fertilization, and assimilate partitioning. In “Biological Solar Energy
Conversion” (A. Mitsui, et al., Eds.), pp. 299–322. Academic Press, New York.
Hertog, L. D., Stulen, I., and Lambers, H. (1993). Assimilation, respiration and allocation of
carbon in Plantago major as affected by atmospheric CO2 Levels—A case study. In “CO2
and Biosphere” ( J. Rozema, et al., Eds.), pp. 369–378. Kluwer Acadamic Publishers,
Belgium.
Effect of Climate Change Factors on Processes of Crop Growth 67
Ong, C. K. (1984). The influence of temperature and water deficits on the partitioning of
dry matter in groundnut (Arachis hypogaea L.). J. Exp. Bot. 35, 746–755.
Ono, Y. (1979). Flowering and fruiting of peanut plants. Jpn. Agric. Res. Q. 13, 226–229.
Ono, Y., Nakayama, K., and Kubota, M. (1974). Effects of soil temperature and soil
moisture in podding zone on pod development of peanut plants. Proc. Crop Sci. Soc.
Jpn. 43, 247–251.
Pallas, J. E. Jr., Stansell, J. R., and Koske, T. J. (1979). Effect of drought on florunner
peanuts. Agron. J. 71, 853–858.
Pilumwong, J., Senthonga, C., Srichuwongb, S., and Ingram, K. T. (2007). Effects of
temperature and elevated CO2 on shoot and root growth of peanut (Arachis hypogaea L.)
grown in controlled environment chambers. Sci. Asia 33, 79–87.
Prasad, P. V. V., Craufurd, P. Q., and Summerfield, R. J. (1999a). Sensitivity of peanut to
timing of heat stress during reproductive development. Crop Sci. 39, 1352–1357.
Prasad, P. V. V., Craufurd, P. Q., and Summerfield, R. J. (1999b). Fruit number in relation
to pollen production and viability in groundnut exposed to short episodes of heat stress.
Ann. Bot. 84, 381–386.
Prasad, P. V. V., Craufurd, P. Q., Summerfield, R. J., and Wheeler, T. R. (2000a). Effects of
short episodes of heat stress on flower production and fruit-set of groundnut (Arachis
hypogaea L.). J. Exp. Bot. 51, 777–784.
Prasad, P. V. V., Craufurd, P. Q., and Summerfield, R. J. (2000b). Effect of high air and soil
temperature on dry matter production, pod yield and yield components of groundnut.
Plant Soil 222, 231–239.
Prasad, P. V. V., Boote, K. J., Allen, L. H. Jr., and Thomas, J. M. G. (2003). Super-optimal
temperatures are detrimental to peanut (Arachis hypogaea L.) reproductive processes and
yield at both ambient and elevated carbon dioxide. Glob. Chang. Biol. 9, 1775–1787.
Prasad, P. V. V., Boote, K. J., Thomas, J. M. G., Allen, L. H. Jr., and Gorbet, D. W. (2006).
Influence of soil temperature on seedling emergence and early growth of peanut cultivars
in field conditions. J. Agron. Crop Sci. 192, 168–177.
Pressman, E., Peet, M. M., and Pharr, M. (2002). The effect of heat stress on tomato pollen
characteristic is associated with changes in carbohydrate concentration in the developing
anthers. Ann. Bot. 90, 631–636.
Rao, K. V. (1999). The combined effect of elevated CO2 levels and temperature on growth
characteristics of groundnut (Arachis hypogaea L.). Indian J. Plant Physiol. 4, 297–301.
Rao, R. C. N., Singh, S., Sivakumar, M. V. K., Srivastava, K. L., and Williams, J. H. (1985).
Effect of water deficit at different growth of peanut. I. yield responses. Agron. J. 77, 782–786.
Singh, P., Boote, K. J., and Virmani, S. M. (1994a). Evaluation of the groundnut model
PNUTGRO for crop response to plant population and row-spacing. Field Crop Res. 39,
163–170.
Singh, P., Boote, K. J., Yogeswara Rao, A., Iruthayaraj, M. R., Sheikh, A. M.,
Hundal, S. S., Narang, R. S., and Singh, Phool (1994b). Evaluation of the groundnut
model PNUTGRO for crop response to water availability, sowing dates and seasons.
Field Crop Res. 39, 147–162.
Stronach, I. M., Clifford, S. C., Mohamed, A. D., Singleton-Jones, P. R., Azam-Ali, S. N., and
Crout, N. M. J. (1994). The effect of elevated carbon dioxide, temperature and soil moisture
on the water use of stands of groundnut (Arachis hypogeae L.). J. Exp. Bot. 45, 1633–1638.
Stuhlfauth, T., and Fock, H. P. (1990). Effect of whole season CO2 enrichment on the
cultivation of a medicinal plant, Digitalis lanata. J. Agron. Crop Sci. 164, 168–173.
Suzuki, M. (1966). Studies on thermoperiodicity of crops. II. The effects of soil temperature
on fructification of peanuts. Chiba Univ. Tech. Bull. 13, 95–101.
Suzuki, K., Takeda, H., Tsukaguchi, T., and Egawa, Y. (2001). Ultrastructural study of
degeneration of tapetum in anther of snap bean (Phaseolus vulgaris L.) under heat-stress.
Sex. Plant Reprod. 13, 293–299.
Effect of Climate Change Factors on Processes of Crop Growth 69
Talwar, H. S. (1997). Physiological Basis for Heat Tolerance During Flowering and Pod
Setting Stages in Groundnut (Arachis hypogaea L.). JIRCAS Visiting Fellowship Report
1996–97 JIRCAS, Okinawa.
Talwar, H. S., Takeda, H., Yashima, S., and Senboku, T. (1999). Growth and photosyn-
thetic responses of groundnut genotypes to high temperature. Crop Sci. 39(2), 460–466.
Underwood, C. V., Taylor, H. M., and Hoveland, C. S. (1971). Soil physical factors
affecting peanut pod development. Agron. J. 63, 953–954.
Weiss, E. A. (2000). Oilseed Crops. Blackwell Science, London.
Williams, J. H., and Boote, K. J. (1995). Physiology and modelling—Predicting the unpre-
dictable legume. In “Advances in Peanut Science” (H. E. Pattee and H. T. Stalker, Eds.),
pp. 301–335. APRES, Stillwater, Oklahoma.
Williams, J. H., Wilson, J. H., and Bate, G. C. (1975). The growth of groundnuts (Arachis
hypogaea L. cv. Makulu Red) at three altitudes. Rhodosian J. Agric. Res. 13, 33–43.
Wood, I. M. W. (1968). The effect of temperature at early flowering on the growth and
development of peanuts (Arachis hypogaea). Aust. J. Agric. Res. 19, 241–251.
Xi, X. Y. (1991). Development and structure of pollen and embryo sac in peanut (Arachis
hypogaea L.). Bot. Gaz. 152, 164–172.
Yamagata, M., Kouchi, H., and Yoneyama, T. (1987). Partitioning and utilization of photo-
synthate produced at different growth stages after anthesis in soybean (Glycine max L.
Merr.): Analysis by long term 13C-labelling experiments. J. Exp. Bot. 38, 1247–1259.
C H A P T E R T H R E E
Contents
1. Introduction 72
2. Expansion of Oil Palm Cultivation in Indonesia and
Environmental Stakes 74
2.1. Expansion of oil palm cultivation 74
2.2. Environmental stakes 75
3. Oil Palm Cultivation 81
3.1. Climate and soil conditions 81
3.2. Production systems: Industrial versus smallholder plantations 82
3.3. Land clearing and site preparation 82
3.4. Water and soil management 83
3.5. Nutrient-demand assessment 85
3.6. Fertilizer management 90
3.7. Synthesis 92
4. Hydrological Processes and Associated Nutrient Transfers in Oil
Palm Plantations 93
4.1. Precipitation in Indonesia 94
4.2. Interception 94
4.3. Evapotranspiration 95
71
72 Irina Comte et al.
Abstract
Rapid expansion of oil palm (Elaeis guineensis Jacq.) cultivation in Southeast Asia
raises environmental concerns about deforestation and greenhouse gas emis-
sions. However, less attention was paid to the possible perturbation of hydrolog-
ical functions and water quality degradation. This work aimed to review (i) the
agricultural practices commonly used in oil palm plantations, which potentially
impact hydrological processes and water quality and (ii) the hydrological changes
and associated nutrient fluxes from plantations. Although many experimental
trials provide clear recommendations for water and fertilizer management, we
found that few studies investigated the agricultural practices actually followed by
planters. Our review of hydrological studies in oil palm plantations showed that
the main hydrological changes occurred during the first years after land clearing
and seemed to dissipate with plant growth, as low nutrient losses were generally
reported from plantations. However, most of those studies were carried out at the
plot scale and often focus on one hydrological process at a single plantation age.
So, there is insufficient information to evaluate the spatiotemporal fluctuations in
nutrient losses throughout the entire lifespan of a plantation. Furthermore, few
studies provided an integrated view at the watershed scale of the agricultural
practices and hydrological processes that contribute to nutrient losses from oil
palm plantations and the consequences for surface and groundwater quality.
Future research efforts need to understand and assess the potential of oil palm
plantations to change hydrological functions and related nutrient fluxes, consid-
ering agricultural practices and assessing water quality at the watershed scale.
1. Introduction
Oil palm (Elaeis guineensis) is one of the most rapidly expanding crops
in the tropics. Since the early 1980s, the global land area under oil palm
production has more than tripled, reaching almost 15 million ha in 2009 and
accounting for almost 10% of the world’s permanent crop land (FAOSTATS,
2011; Sheil et al., 2009) Most of this increase has taken place in Southeast Asia.
Together, Malaysia and Indonesia account for almost 85% of the 46.5 million
tons of crude oil palm produced in the world, Indonesia being the top
producer since 2007 (Oil World, 2011; USDA, 2007). The area covered by
smallholder plantations in Indonesia increased nearly 1000-fold between 1979
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 73
and 2008, reaching almost 3 million ha, representing 39% of Indonesian oil
palm plantations currently, the remaining 4.5 million ha being large private
(52%) and government-owned (8%) plantations (IMA, 2010).
Although oil palm cultivation is a strong driver of economic development
in Indonesia, providing jobs and incomes to millions of people (USDA, 2007),
it is strongly denigrated for its environmental impacts. Many media and NGOs
accuse oil palm plantation development in Southeast Asia of triggering defor-
estation, loss of biodiversity, peatland degradation, and high greenhouse gas
(GHG) emissions (Greenpeace, 2011; WWF, 2011). In the scientific commu-
nity, there is controversy about the positive and negative aspects of the
expanding oil palm cultivation and potential environmental risks, which has
been discussed at length in the scientific literature (Basiron, 2007; Lamade and
Bouillet, 2005; Nantha and Tisdell, 2009; Sheil et al., 2009). The development
of oil palm plantations, which frequently cover tens of square kilometers in
Southeast Asia, involves land clearing, road and drainage network construc-
tion, and sometimes earthworks such as terracing on undulating areas. The use
of agrochemicals, such as fertilizers and pesticides, might represent a potential
risk for the sustainability of aquatic ecosystem and hydrological functions
when agricultural practices are not optimized. In particular, oil palm growers
usually apply large amounts of commercial fertilizer and thus are among the
largest consumers of mineral fertilizers in Southeast Asia (Härdter and
Fairhurst, 2003). However, hydrological processes within oil palm plantations
are still not fully understood, and few studies have examined the impacts of
agricultural practices on terrestrial hydrological functions and water quality in
nearby aquatic ecosystems (Ah Tung et al., 2009), although aspects that impact on
water quality are by far the largest component of an environmental risk register accounting
for nearly 50% of all entries in oil palm plantations (Lord and Clay, 2006).
This review aims to document the current state-of-knowledge of agricul-
tural practices in oil palm plantations that potentially impact hydrological
functions and water quality in surface waters, with a focus on nutrient loading
of surface waterways, and to highlight research gaps in the understanding of
these processes. This work focuses on the situation in Indonesia, with examples
from other oil palm producing countries in the humid tropics as appropriate.
First, the expansion of oil palm cultivation in Indonesia, relevant environmen-
tal issues, and polemics will be presented. Next, typical agricultural practices in
industrial and smallholder oil palm plantations will be discussed, focusing on
nutrient, soil, and water management. Finally, the last section gives the state-of-
the-art knowledge of hydrological changes and associated nutrient fluxes from
oil palm plantations compared to tropical rainforests, which were the dominant
natural ecosystem prior to oil palm plantation establishment. Relevant pro-
cesses in the hydrological cycle and their magnitude and relevance in oil palm
plantations will be explained in this section, but we do not provide an in-depth
discussion of hydrological processes in rainforests, as a number of reviews were
already published on this topic (Bruijnzeel, 1991, 2004; Elsenbeer, 2001).
74 Irina Comte et al.
Figure 1 Map of Indonesian palm oil production in 2007/2008 (from USDA, 2009).
tons of CPO by 2020 (IMA, 2010; Rist et al., 2010). According to USDA
(2007), the availability of land in Indonesia, coupled with other factors—
high seed sales, record energy prices, and high vegetable oil prices—ensures
that Indonesia will continue to lead the world in palm oil production for
years to come. However, few developments generate as much controversy
as the rapid expansion of oil palm in developing countries such as Indonesia
(Koh and Wilcove, 2008; Nantha and Tisdell, 2009). Negative conse-
quences reported by environmental groups include deforestation, loss of
biodiversity, peatland degradation, GHG emissions, and water pollution.
harbors numerous endemic or rare species (Koh and Ghazoul, 2008; WRI,
2002). Many sources are claiming that virgin tropical forests are being
cleared for oil palm plantations, leading to natural habitat loss for many
endangered species and biodiversity reduction. For instance, it was reported
that Sumatran orangutans (Pongo abelii) and Bornean orangutans (Pongo
pygmaeus) face extinction due to plantation expansion (Nantha and
Tisdell, 2009; Nellemann et al., 2007; Tan et al., 2009). Herds of elephants,
tigers, and rhinos are reported to be critically threatened due to this
expansion (Danielsen et al., 2008; WRI, 2002). Studies in oil palm frontier
areas on the island of Sumatra concluded that oil palm plantations result
in a significant reduction in biodiversity if plantations replace natural forests,
secondary forests, agroforests, or even degraded forests and scrubby
unplanted areas (Gillison and Liswanti, 1999; Sheil et al., 2009). However,
others mentioned that the expansion of oil palm plantations is only one of
the factors contributing to herd displacement and local extinction, as other
anthropogenic activities like illegal logging, forest fires, and illegal hunting
are also problematic for large mammals (Nellemann et al., 2007; Tan et al.,
2009). According to Koh and Ghazoul (2008), at least 56% of the oil palm
expansion in Indonesia during the period 1990–2005 occurred at the
expense of primary, secondary, or plantation forests, and 44% was on
cropland area. Deforestation in Southeast Asia cannot be attributed solely
to oil palm production. According to the World Rainforest Movement
(WRM, 2002), the immediate causes of rainforest destruction in Southeast
Asian countries are logging by commercial companies, shifting agriculture,
monoculture plantations (e.g., rubber in Thailand), cattle ranching, fuel-
wood harvesting, hydroelectric dams, mining and oil exploitation, and
colonization schemes.
of solid oil wastes, palm fiber, and shells. Moreover, 10 million tons of palm
oil mill effluent (POME), a polluted mix of crushed shells, water, and fat
residues, was produced and often returned without treatment to natural
watercourses downstream from the mill (Lord and Clay, 2006), leading
to the degradation of the aquatic ecosystems (Briggs et al., 2007; Sheil et al.,
2009). The POME is an acidic colloidal suspension characterized by high
concentration of suspended solids, with a biological oxygen demand (BOD)
of 25,000ppm and a chemical oxygen demand (COD) of 60,000ppm
( Jacquemard, 1995; Olaleye and Adedeji, 2005). According to Olaleye
and Adedeji (2005), riparian rivers and streams receiving untreated mill
effluent are expected to be heavily polluted. Fortunately, technologies
have been developed to treat and reclaim drinking water from the POME
(Ahmad et al., 2006; Rupani et al., 2010; Singh et al., 2010; Yi Jing et al.,
2010), and they need to be widely implemented to protect downstream
waters. Sewage from the worker population in the plantation is another
waste byproduct that is expected to elevate COD, BOD, and fecal coliform
levels in waterways.
greatly reduce yields (Basiron, 2007). The oil palm is cultivated predomi-
nantly on tropical soils in the orders Ultisol, Oxisol, and Inceptisol. These
soils are highly acidic with low buffering capacities (Ng, 2002) as a conse-
quence of cation leaching (Caliman et al., 1987). Once the pH drops below
5.5, aluminum and manganese compounds start to dissolve, which may
cause root deterioration (Godbold et al., 1988). However, oil palm is
adapted to acidic conditions (Omoti et al., 1983), and with appropriate
management, oil palm plantations can also be productive on “problem soils”
such as acid sulfate soils, deep peat and acidic high aluminum soils, where
few other crops are successful (Corley and Tinker, 2003).
popular agronomic handbooks for oil palm cultivation are Corley and
Tinker (2003), Jacquemard (1995), Rankine and Fairhurst (1998a,b,c), and
Hartley (1988).
Once the site is selected, establishment of the oil palm plantation starts
with land clearing. At present, mechanical methods are used in all major oil
palm-growing countries with chainsaws, winches, and bulldozers. Then,
the felled vegetation is either burnt or allowed to rot. The issue of whether
or not to burn the felled vegetation has remained a subject of controversy
for years (Corley and Tinker, 2003) due to environmental impacts of
burning: smoke, haze, and large nutrient losses through volatilization and
ash carried away (Mackensen et al., 1996). Since the massive fires in
Kalimantan and Sumatra in 1997, the Indonesian government prohibited
burning (Corley and Tinker, 2003). However, according to some media,
laws to limit agricultural burning are poorly enforced and burning still
continues (Mongabay Editorial, 2006).
The general layout of a plantation is decided by the topography, the
drainage, the position of the mill, and the distance to transport fresh fruit
bunches (FFB) to the nearest road. Hartley (1988) recommends a gap of 320
m between roads, giving a density of 33m roads ha1, that is, 3% of the land
area. In hilly areas, the road density should increase, with distances between
roads not exceeding 200m, due to the difficulty of transporting FFB across
platforms and terraces. With terraces, the distance should be 125m, the
density being 80m roads ha1. A typical road layout in hilly estates has to be
arranged in relation to the drainage lines and streams. If the roads can run
parallel to streams, this reduces the number of bridges needed and helps to
avoid crossing over swampy areas (Corley and Tinker, 2003). Steps for the
establishment and exploitation of an oil palm plantation on a typical private
estate are summarized in Fig. 2.
Field establishment
Field lining and holing
Final culling
Transplant seedlings
Fertilizer application
Pest control
General field upkeep
Harvesting
Transportation of fresh fruit bunches to mills
Replanting Abandonment
Removal of old palms Evacuation of plantation staff and workers
Figure 2 Typical oil palm plantation development activities (adapted from ECD,
2000).
(APOC, 2011; Goh and Chew, 1995). To achieve this, a good outlet with
sufficient capacity to discharge excess water is needed (Corley and Tinker,
2003). Another possibility is to install controlled drainage systems that retain
subsurface water in the drains prior to dry periods. Generally, the drainage
system consists of an interconnected network of collection and main drains of
varying dimensions depending on the hydrological and rainfall characteristics
of the area (Othman et al., 2010). The slope, intensity, and dimension of
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 85
targeted yields (Tarmizi and Mohd, 2006); (ii) to determine input levels to
achieve an economically optimum production (income vs. costs of produc-
tion) (Breure, 2003; Caliman, 2001; Caliman et al., 1994; Goh and Härdter,
2003) including recommended types, rates, and timing of fertilizer applica-
tions to minimize nutrient losses (Goh and Chew, 1995); and also (iii) to
develop agricultural practices aiming at minimizing chemical fertilizers
inputs such as the establishment of a LCC during the immature stage
(Agamuthu and Broughton, 1985), recycling of pruned fronds and male
inflorescences (Ng and Thamboo, 1967), mulching of EFB (Caliman et al.,
2001; Chiew and Rahman, 2002; Loong et al., 1987), and POME spread-
ing (Rupani et al., 2010; Wood et al., 1979). These studies led to an
abundant literature on recommendations for optimum nutrient manage-
ment in oil palm plantations (Corley and Tinker, 2003; Fairhust et al.,
2005; Kee and Goh, 2006).
Fertilizer management in oil palm plantations is based on nutrient
balance principle, which estimates the total demand of the palm and
matches it with the nutrient supply in the oil palm plantation and from
supplemental fertilizers (Goh et al., 1999). Nutrient demand can be divided
in two categories: nutrient uptake and nutrient losses from soil through
processes such as runoff, leaching, and gaseous emissions. Within nutrient
uptake, we distinguish nutrient stocks exported by harvesting (FFB) and
nutrients immobilized in the palm biomass (for growth) (Goh, 2004; Goh
and Härdter, 2003). Some authors also calculate nutrients recycled from
pruned fronds and male inflorescences because they are usually returned to
the soil (Tarmizi and Mohd, 2006). The nutrient requirements of oil palm
vary widely, depending on the target yield, genetic potential of the planting
material used, and numerous environmental factors such as tree spacing,
palm age, soil fertility, groundcover conditions, and climate (Fairhurst and
Mutert, 1999; Goh and Härdter, 2003; Tarmizi and Mohd, 2006). Table 1
compares the total nutrient stocks in standing biomass of oil palm plantation
and tropical forest, illustrating that 1ha of forest generally contains more
nutrients in plant biomass than a plantation. Table 2 shows nutrient uptake
and allocation for production, immobilization in palm biomass, and recy-
cling, based on data from trials in Southeast Asia. Generally, a larger
proportion of nutrient uptake is needed for FFB than for immobilization.
For example, Ng et al. (1999) reported, for a target yield of 25tha1 year1,
the annual nutrient uptake in FFB was 2.3-fold greater than nutrients
immobilized in new biomass.
The soil nutrient supply in an oil palm plantation comes from dissolved
nutrients in atmospheric deposition, including precipitation, nutrients
recycled from pruned fronds and male inflorescences when these are
returned to the soil, nutrients leached by rainfall from the leaf canopy (leaf
wash), nutrient returns from LCC, available nutrients present in the soil, and
fertilizer applications (Goh, 2004; Goh and Härdter, 2003; Goh et al., 1999;
Table 1 Standing stock biomass in oil palm plantation and topical forest (adapted from Henson, 1999)
Production
Harvested fruit bunches 24 72.5 12.1 93.2 20.7 – Ng and Thamboo (1967),
Ng et al. (1968)
25 73.2 11.6 93.4 20.8 19.5 Ng et al. (1968)
30 97.6 10.0 105.4 18.2 – Tarmizi and Mohd (2006)
30 99.1 15.6 129.3 33.3 – Ng et al. (1999)
Immobilized
Trunk 30 42.4 4.1 121.6 10.2 – Ng et al. (1999)
Roots – 16.6 1.1 2.8 0.42 – Corley et al. (1971)
Trunk & roots 30 18.5 2.4 61.9 3.8 – Tarmizi and Mohd (2006)
Immobilized in new biomass 25 40.0 3.1 55.7 11.5 13.8 Ng et al. (1968)
Recycled
Pruned fronds and male inflorescences 24 78.4 11.3 102.1 28.1 – Ng and Thamboo (1967),
Ng et al. (1968)
FFB, Fresh Fruit Bunch.
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 89
Outer
Inner cycle
Replanting
cycle Shell, fibre
cycle Increment
120 kg EFB to land 39 kg
(25 years) Inflorescences,
63 kg
rainfall leaching,
litter fall, POME Loss on
Felled pruned fronds to land combustion
Trunk & 30 kg
palms & roots leaves 290 kg Pruned fronds
1900 kg 1800 kg & cover crop Boiler ash Fertilizer
80 kg 210 kg
Ground level
Figure 3 Total demands and sinks for potassium in an oil palm plantation with 30ton
FFB yield (from Corley and Tinker, 2003). EFB, empty fruit bunch; FFB, fresh fruit
bunch; POME, palm oil mill effluent.
Tarmizi and Mohd, 2006). As an example, the potassium fluxes in oil palm
plantation are illustrated in Fig. 3.
Some research groups developed complex models, based on the nutrient
balance principle, to assess nutrient requirements, taking account of the
commodity price and fertilizer cost. Corley and Tinker (2003) reported on
three of these models, which are summarized below:
1. The Applied Agriculture Research group in Malaysia developed a linked
group of models for oil palm management, which include a model to
assess site-specific yield potential (ASYP), a model for predicting the
fertilizer requirements, the integrated site-specific fertilizer recommen-
dation system (INFERS), and expert systems for determining best
month for fertilizer application, and the timing and allocation of differ-
ent fertilizers (Kee et al., 1994; Kok et al., 2000).
2. Leaf analysis is the most common diagnostic tool to determine the nutri-
tional status of oil palm and estimate the appropriate fertilizer rates because
of significant relationship between leaf nutrient concentration and FFB
yield (Foster and Chang, 1997; Goh, 2004). It is largely used by the
International Cooperation Centre in Agronomic Research for
90 Irina Comte et al.
(1996), field palms yielding 30t FFBha1 in West Sumatra return to the
soil about 10tha1 year1 dry matter from pruned fronds, which contain
125kg N, 10kg P, 147kg K, and 15kg Mg. Also, both EFB and POME
contain substantial amounts of nutrients and organic matter that can replen-
ish the soil fertility and help meet the nutrient requirements for oil palm.
According to Taillez (1998), mulched EFB can reduce the need for chemi-
cal fertilizers by more than 50% in immature stands and by 5% in mature
stands. Application of 40–60t EFBha1 year1 or 750m3 POMEha1 year1
is recommended to add organic matter and improve soil fertility on poor
inland soils (Goh et al., 1999).
3.7. Synthesis
Industrial and smallholder planters do not have the same resources to ensure
optimum nutrient management. Specifically, independent smallholders do
not benefit from techniques such as leaf diagnosis and soil analysis (Fairhurst
and Mutert, 1999; Pushparajah, 1994) to assess the nutrient requirements of
their plantations. In many smallholdings, low-productivity palms are
planted unevenly without terracing, and fertilizer use is inadequate and
unbalanced (urea applied alone) (FAO, 2005; Webb et al., 2011; Zen
et al., 2005). Some smallholders observe and copy the industrial plantations
in recycling pruned fronds and male inflorescences, but not all. Moreover,
they do not benefit from organic inputs of EFB and POME, which are
available only in industrial or governmental estates where the mills are
located. Unfortunately, few data are available on actual fertilization practices
in most oil palm plantations in Indonesia. Some authors provide estimates of
average yields of palm oil from the different management groups, which are
generally higher for private plantations than smallholdings, although the
FAO estimated similar yields for these two groups (Table 4). However,
these estimates should be interpreted with caution as they do not account
for the high variability for cultural practices among planters. Although some
Table 4 Average yields of palm oil production for the different production systems in
Indonesia (tha1)
Industrial
Smallholder Government Private Source
Evapotranspiration Rainfall
Evaporation Interception
Stemflow
Water in
biomass
Overland flow
Infiltration
Stormflow Streamflow
Deep Capillary
drainage rise
Groundwater Baseflow
Figure 4 The hydrological cycle in an oil palm plantation. Boxes indicate storage
pools, arrows indicate fluxes.
4.2. Interception
Studies of hydrological processes in tropical rainforests indicate the impor-
tance of interception by the leaves and branches (canopy) of plants. In his
review, Bruijnzeel (1990) concluded that forest interception was between
4.5% and 22% of the rainfall incident on the canopy, with an average value
of 13%. Compared to natural forest vegetation, a lower proportion of
rainfall is expected to be intercepted by palms as a result of lower leaf area
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 95
index, even in mature oil palm plantations (Henson, 1999) due to clearing
of most understory vegetation. Dufrêne (1989) reported that less than
5% rainfall was intercepted following a 30-mm precipitation event and
lower values were recorded at higher rainfall intensity. However, Squire
(1984) found that interception was 17–22% of precipitation in oil palm
plantation, the amount varying with palm age, erectness of canopy, and
rainfall intensity.
In tropical forests, only a small amount of rain falls directly on the ground
or into water bodies, as most of the rainfall will reach the soil via throughfall
and stemflow, which are responsible for transferring nutrients from the
canopy to the soil. Some Malaysian studies found that throughfall trans-
ferred 70–78% of rainfall in mature oil palm plantations (Kee et al., 2000). In
Papua New Guinea where rainfall is higher, Banabas et al. (2008) found that
83% of the rain reached the ground as throughfall. These values are the same
order of magnitude as those reported by Bruijnzeel (1990) for lowland
rainforest: 77–93% of incident rainfall (on average 85%) was transported
by throughfall. We are not aware of published data on throughfall, stem-
flow, and rainfall interception by immature oil palm when the canopy has
not yet closed.
4.3. Evapotranspiration
A number of evapotranspiration (ET) studies were carried out in tropical
forests (Noguchi et al., 2004; Tanaka et al., 2008; Zulkifli et al., 1998). In
three catchments in Selangor, Malaysia with more than 94% forest cover,
Low and Goh (1972) found ET accounted for half of the 2162–2482mm
annual rainfall when estimated as the difference between measured rainfall
and water outflow from the catchment. In a study in the Sungai Tekam
Experimental Basin, Malaysia (DID, 1989) that received average annual
rainfall of 1878mm for the period 1977/1978 to 1985/1986, the actual ET
of a forested catchment averaged 1500mmyear1 (range: 1374–1583mm
year1), which represents 99% of the potential ET, on average 1515mm
year1 (range: 1449–1567mmyear1) based on the Penman & Thornwaite
method of estimation. In his review, Bruijnzeel (2004) reported ET typi-
cally ranged from 1000 to 1800mmyear1 in lowland and hill dipterocarp
forests in Peninsular Malaysia.
There have been few ET studies in oil palm plantations (Yusop et al.,
2008). Micrometeorological measurements in mature oil palm on a Malaysian
coastal site showed that ET accounted for 83% of rainfall and was close to
potential ET calculated by the Penman equation (Henson, 1999). A study of
oil palm on volcanic soils in Papua New Guinea gave an estimated ET of
1334–1362mmyear1; Banabas et al. (2008) noted that water deficit was
unlikely to limit transpiration in this area, which had average annual rainfall
between 2398 and 3657mm. Radersma and de Ridder (1996) estimated oil
96 Irina Comte et al.
increases soil moisture storage, which resulted in soils reaching field capacity
earlier during rainfall events. Infiltration rates will also be reduced immedi-
ately following deforestation when soils are compacted by heavy machinery.
However, DID (1989) showed that infiltration rates return to predistur-
bance levels following the establishment of crops due to improvement in
soil structure under vegetative ground cover.
According to Banabas (2007), soils under mature oil palm typically have
high infiltrability (80–8500mmh1, depending on soil texture). However,
the infiltrability of the soil is highly variable due to the ordered structure of
vegetation in oil palm plantations. A study in West Sumatra on 10-year-old
oil palms demonstrated significant spatial variability when comparing water
infiltration rates in soil beneath the palm circle, harvest path, and frond piles.
Infiltration rate increased in the order path<circle<frond pile (Fairhurst,
1996). In Papua New Guinea, Banabas et al. (2008) recorded similar results
(Table 5). They attributed the highest values in the frond pile zones to the
macroporosity-enhancing effect of the organic matter present in this zone.
They ascribed the lower values found in the weeded circle and harvest-path
zones to topsoil compaction from falling bunches in the weeded circle,
wheel and foot traffic in the harvest paths, and sparse understory vegetation.
Thus, apart from its high dependence on soil texture, infiltrability will
vary temporally and spatially in oil palm plantation. After a strong decrease
due to soil compaction after land clearing, infiltrability in oil palm plantation
may partly recover due to plant growth and organic matter addition to the
soil. Yet, infiltrability will remain low along roads and harvest pathways and
in weeded circles.
Table 5 Soil infiltrability (mmhr1) for major soil types and at specific locations in oil
palm plantations
et al., 1975). Depending on the amount of water draining out of the rooting
zone, leached solutes may simply accumulate in a deeper layer of the soil
profile or may reach the underlying groundwater (Ah Tung et al., 2009).
Table 6 summarizes results from Bruijnzeel (1991), including estimates of
annual runoff and nutrient losses in drainage water from tropical rainforests.
The author also reported that forest clear-cutting often increased nutrient
losses to streams. Brouwer and Riezebos (1998) compared nutrient leaching
in closed and logged tropical rainforest in Guyana. Logging clearly increased
leaching, which they ascribed to an increased percolation of water through
the soil (2800mm after 22months logging compared to 1800mm in closed
forest) and increased solute concentrations in the percolating water. They
found that Ca, K, and Mg concentrations were 2–10 times greater in the gaps
after logging than in closed forest, while the NO3 concentration was 10–20
times greater than closed forest. The major pulse of leaching occurred during
the first year after logging, and most solute concentrations in percolating soil
water remained higher up to 15months after logging. However, vegetative
regrowth reduced leaching losses as plants absorbed soluble nutrients and
immobilized them in standing stock biomass.
Large nutrient losses are expected in oil palm plantations (Goh et al.,
2003). According to Ng et al. (2003), most of the oil palm root biomass is
found within 1m of the soil, but the distribution of oil palm active roots
favors the nutrient uptake in the upper 30cm, which may increase the
potential risk of nutrient leaching. Omoti et al. (1983) measured amounts
of nutrients leached under immature (4years) and mature (22years) oil
palms, distinguishing between the loss of native nutrients and the loss of
applied nutrients (fertilizers). The losses of native nutrients from closed and
logged tropical forest on one hand and in immature and mature oil palm
plantations on the other hand are summarized in Table 7. To our knowl-
edge, there is no chronological study comparing the amounts of nutrient
leached under natural forest and oil palm plantation established on the same
site neither after forest clearing nor between forested and oil palm catch-
ments with similar climatic and soil conditions. This makes it difficult to
quantify the impact of oil palm plantation establishment on native nutrient
leaching. The most complete study to assess the impact on hydrological
processes of forest conversion to tree crop and oil palm plantations (DID,
1989) did not evaluate leaching, leaving a considerable gap in our knowl-
edge of this process in plantations at the catchment scale.
Despite the paucity of large-scale data on leaching processes, many plot-
scale studies investigated the percentage of applied fertilizers lost through
leaching in oil palm plantations, some of them comparing young and mature
oil palm stands (Chang and Zakaria, 1986; Foong et al., 1983; Maena et al.,
1979). For example, a field lysimeter study conducted on Munchong series
soil in Malaysia found higher fertilizer losses when the palm was 1–4years old
(17% for N and 10% for K), which declined to 2.1% of applied fertilizer N
Table 6 Catchment studies of annual rainfall, annual runoff, and nutrient losses in drainage water from South East Asian tropical forests (modified
from Bruijnzeel, 1991)
Nutrient losses
Annual Annual
(kgha1 year1)
Catchment rainfall runoff Q/P Sources quoted by
Location Type of forest Soil area (ha) P (mm) Q (mm) (%) Ca Mg K P N Bruijnzeel (1991)
Ulu Gombak, Partly disturbed Oxisol 31 2500 750 30 2.1 1.5 11.2 – – Kenworthy (1971)
Malaysia dipterocarp forest
Bt. Berembun, Undisturbed Deep Ultisols (2/3) 29.6 2005 225 11 5.8 3.6 8 – – Abdul Rahim and
Malaysia dipterocarp forest and Oxisols (1/3), Zulkifli (1986),
sandy clay (loam) Zulkifli (1989),
and Zulkifli et al.
(1989)
Watubelah, Plantation forest of Andesitic tuffs 18.7 4670 3590 77 29 30.5 22 0.7 10.6 Bruijnzeel (1983a,c,
Indonesia Agathis dammara underlain by 1984)
andesitic breccias
Kinta Valley, Lowland rainforest Limestone – 2845 1605 56 795 90 76 – – Crowther (1987a,b)
Malaysia (karst terrain)
Ei Creek, Papua Colline rainforest Basaltic volcanic 16.25 2700 1480 55 24.8 51 14.9 – – Turvey (1974)
agglomerates
Gua Anak Lowland rainforest Limestone – 2440 1255 51 764 45 20 – – Crowther (1987a,b)
Takun,
Malaysia
Table 7 Nutrient leaching losses in undisturbed forest, highly disturbed forest, and oil palm plantations
and 2.7% of fertilizer K when the palm was 5–14years old (Foong, 1993).
Higher nutrient losses through leaching from immature palm implies less
plant nutrient uptake, whereas older palms have more extensive root system
that can absorb applied and indigenous soil nutrients, a greater nutrient
demand and a higher transpiration rate that lowers water loss via leaching.
However, a field-scale study on Orlu and Algba series (Rhodic Paleudult)
soils in Nigeria showed no significant differences in nutrient leaching
from applied fertilizers between the immature (4years) and mature plantations
(22years) (Omoti et al., 1983). According to some authors, the adult stage
poses a high risk of nutrient losses because ground vegetation is sparse due to
poor light penetration through the closed oil palm canopy (Breure, 2003).
Moreover, the LCC dies off at canopy closure, releasing a large amount of
N from the decomposing legume biomass and increasing the risk of N loss via
leaching (Campiglia et al., 2010; Goh et al., 2003). According to Goh and
Chew (1995), leaching losses also depends on soil texture and greater losses
were recorded in sandier soils, as summarized in Table 8. In general, leached
P losses are low due to the relative immobility of P in acidic, weathered
tropical soils (Goh et al., 2003; Omoti et al., 1983).
A plot-scale study by Ah Tung et al. (2009) is the only one to our
knowledge that investigated leaching losses of inorganic N and K and
measured their concentrations in groundwater. They found leaching losses
of inorganic N represented between 1.0% and 1.6% of applied N fertilizer
and the K losses were between 2.4% and 5.3%, depending on fertilizer
application rates. The concentration of N and K in the soil solution
decreased with soil depth, which they explained by nutrient removal and
uptake by palm roots, resulting in lower nutrient concentrations in the soil
solution with depth. However, another explanation they did not mention
is that fertilizers are applied near the soil surface, so the topsoil layers have a
higher concentration of nutrients in soil solution; this naturally declines
with depth because there was no fertilizer injection deeper in the soil
profile. The measured concentrations of NH4-N, NO3-N, and K in
groundwater ranged from 0.23 to 2.7, 0.07 to 0.25, and 0.63 to 9.54mgl1,
respectively, which did not exceed the water quality standards set by the
World Health Organization (WHO, 2008). The authors did not specify
the distance of groundwater sampling wells from the palm circles where
the fertilizers were applied. However, they mentioned the possibility of
groundwater pollution when excessive N fertilizer was applied or if NO3-N
leached from the soil profile into groundwater in the intertree spaces between
palms. This supposition is supported by Schroth et al. (2000), who reported
pronounced spatial pattern of NO3-N concentrations within the plantation.
Low NO3-N concentrations were measured in the soil profile within 1m of
palm trees, indicating efficient absorption of mineral N by the palms, whereas
soil NO3-N concentrations increased with increasing distance from palm
trees. At 4m away from the trees, the vertical NO3-N concentration gave
Table 8 Nutrients leached as percentages of applied fertilizers in immature (1–4years) and mature (>4years) oil palm plantations
Annual
rainfall (P) Runoff Q/P Dissolved loss TSS outflow
Soil Study scale mm (Q) mm (%) Soil erosion (tha1 year1) (tha1 year1) Notes
Red-yellow Plot 210m 2862–3563 2.50 0.5 0.41gm2 – – Period 27/3 to 6/6/98;
podsol rainfall 495mm
2.50 0.8 1.35gm2 – – Period 28/12/97 to
21/2/98; rainfall 316mm
Gleyic Runoff plot 50–200m2; 1950 56.55 2.9 38kgha1 0.16 0.3 Excluding valley bottom
podsol catchment scale year1
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 105
weeded circle (where stem flow causes the highest local water inflows and the
infiltrability is quite low) and the harvest-path zone (where infiltrability is
lowest due to soil compaction) (Table 10).
While erosion is never excessive (i.e., greater than the rate of soil
formation) in forests, soil loss can be pronounced at particular stages in an
oil palm plantation. Many studies reported highest erosion rates immedi-
ately after land clearing, resulting from increased exposure of the soil surface
to erosion and surface runoff losses (Bruijnzeel, 1990, 2004; Douglas, 1999;
Goh et al., 2003). According to Clay (2004), bare soil resulting from road
construction and other infrastructure such as bridges, culverts, and drains
increases soil erosion in oil palm plantations. DID (1989) showed that
deforestation activities such as timber harvesting, construction of roads,
and preparation of land for crop planting account for as much as 91% of
all the sediment exported from the catchments. Results from erosion plots
on two soil types revealed five to seven times more erosion from deforested
land than forested lands during the first year after planting the LCC.
However, once the ground cover was established, erosion was greatly
reduced but not eliminated (DID, 1989). In mature plantations, erosion
still occurs from harvest paths, roads, and localized areas of steep elevation.
Clay (2004) reported that in Papua New Guinea, every 100m of road has
the potential to produce as much sediment as each hectare of oil palm, but
this is not unrelated as there are 50 linear meters of road for every hectare of
oil palm planted.
Some authors estimated contribution of rainfall to runoff and associated
nutrient losses, usually expressed as percentage of applied fertilizers, carrying
out plot-scale studies in oil palm plantations. Some of them computed
nutrient losses, via runoff and/or sediment transport to be large, accounting
for up to 10% of applied fertilizers (Maena et al., 1979). They observed
greater losses from surface runoff in the uncovered soil in the harvest path,
compared to the interrows, where pruned fronds provide soil cover
(Fairhurst, 1996; Goh et al., 2003; Maena et al., 1979). Others reported
low losses of nutrient via runoff in oil palm plantation (Banabas et al., 2008).
Results from key papers are summarized in Table 11. Moreover, runoff
losses of applied fertilizers also depend on the lag time between the applica-
tion and the subsequent rainfall. While Chew et al. (1995) showed that high
rainfall prior to fertilizer application resulted in substantial nutrient loss, Kee
and Chew (1996) found that the first rain event following fertilizer applica-
tion in a wet month gave N concentrations in runoff water of 89 and 135
mgkg1 for 65 and 130kgN ha1 rates, respectively, compared to 4mgN
kg1 in the unfertilized control plot. Thus, the amount of fertilizer nutrients
lost through runoff and sediment transport depends on the soil texture, the
age of the oil palms, the local topography and infiltrability, and the lag time
between fertilizer application and rainfall (Banabas et al., 2008). The con-
tinual compaction of harvest pathways and roads and the disappearance of
Table 10 Soil erosion and nutrient losses in surface runoff water from spatial components of an oil palm plantation on a Typic Hapludult in
Malaysia (after Goh et al., 1999; Maena et al., 1979)
Age of oil Annual Annual Annual nutrient losses, kgha1 year1 (% of applied fertilizers)
palm rainfall runoff
Location Soil plantation (mm) (% of rainfall) N P K Mg Ca Transport Source
Malaysia Orthoxic 11 1426 2.8–30.6% 9.93 (11.1%) 1.43 (2.8%) 10.40 (5.0%) 1.82 (5.6%) 4.04 (5.2%) In runoff Maena et al.
Tropudult 5.57 (6.2%) 3.63 (7.0%) 8.79 (0.0%) 21.10 (64.3%) 7.40 (9.4%) In eroded (1979)
sediment
Malaysia Typic 4.5–7.2 0.7–1.1 20.8–33.0 3.6–6.8 – In runoff Kee and
Paleudult (4.4–7.2%) (0.5–0.8%) (9.7–15.4%) (4.0–7.6%) – Chew
0.5–0.8 0.5–1.3 Trace 0.1 (0.1%) – In eroded (1996)
(0.5–0.9%) (0.3–0.9%) sediment
PNG Typic Mature 2398 6 (0–44 for 2.2 – – – – In runoff Banabas
Hapluland (135 374 individual et al.
palms events) (2008)
ha1)
PNG Typic Mature 3657 1.4 (0–8 for 0.3 – – – – In runoff
Udivitrand (135 682 individual
palms events)
ha1)
understory cover (including LCC) due to canopy closure may also contrib-
ute to soil and nutrient losses via runoff and erosion within a mature oil palm
plantation (Table 11).
Annual
rainfall EC (ms Turbidity DO TDS TSS K Ca Mg Na NH3 NO3 PO4 Cl SO4 SiO2
Land use Soil (mm) Notes pH cm1) (NTU) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) (mgl1) Source
Variously 2235 3.2– 14.3– 4.7–28.7 0.3–6.4 22.7– 1.2– 0.007– 0.7–2.9 0.0– 0.0–2.0 Gasim
vegetated 6.3 85.7 184 79.1 0.57 0.50 et al.
with 50% (2006)
forest and
15% oil
palm
Two- Araceneous 2348– Low 5.6 7.3– 0.34– 0.17– 0.32– 0.25– 0.03 0.08– 0.1 0.4–0.5 0.005 9.23– Yusop
forested series 3169 flow 7.5 0.38 0.19 0.35 0.28 0.1 9.24 et al.
catchments (derived Storm 5.1– 10.2– 0.83– 0.2– 0.37– 0.2– 0.04 0.23– 0.1 0.6 0.005 5.15– (2006)
(30ha) from flow 5.2 11.6 0.92 0.24 0.45 0.21 0.32 6.34
sandstone)
Forested Tropeptic 1878 6years 6.9 55.7 48.6 30.1 1.48 6.81 2.48 3.36 1.29 26.50 DID
control Harplothox average (1989)
catchment
(56ha)
Cleared Tropeptic 1878 6years 7.0 83.6 43.2 47.8 2.56 9.04 3.83 2.94 1.55 20.49 DID
catchment Harplothox average (1989)
for oil palm
(97ha)
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 111
Despite the potential risk of water pollution expected from oil palm
plantation activities, there have been very few studies at the watershed scale
to assess water quality in streams within a plantation at different develop-
ment stages (i.e., immature vs. mature palms).
4.7. Synthesis
There is an abundance of literature on hydrological processes in tropical
rainforest ecosystems, immediate and short-term impacts of rainforest distur-
bance (logging, clearing), as depicted in Fig. 5. It is generally accepted that
natural regrowth in tropical forests leads to a relatively fast return to previous
levels of soil infiltrability, streamflow, water budget, and soil nutrient stocks.
However, the impact of oil palm establishment on changes to hydrological
processes and associated nutrient losses and their evolution during oil palm
growth are much less investigated, documented, and understood.
Table 13 summarizes the evolution of hydrological processes and asso-
ciated nutrient losses occurring from forested land to mature oil palm planta-
tion stage, based on observed or expected outcomes and highlights research
gaps in the understanding of these processes. We compare consecutive stages
(cleared land vs. tropical forest; immature plantation vs. cleared land; mature
vs. immature plantation) and also compare both immature and mature stages
to forest, the original land cover. Expected trends for each stage are described
qualitatively because observations were often made for a specific plantation
age (without long-term monitoring to cover all stages) across areas with a
broad range of climatic and soil conditions. The impacts of forest clearing rely
on many studies reviewed by Bruijnzeel (1990, 1991) that generally reported
strong impacts of complete forest clearing. Information regarding the imma-
ture oil palm stage is based on the study by DID (1989), that represents, to our
knowledge, the only chronological study focused on the evolution of hydro-
logical process dynamics from forest to oil palm plantation establishment, at
both plot scale and watershed scale. Unfortunately, it was stopped before oil
palms reached maturity and did not investigate leaching process at the plot
scale. It concluded that after clearing, the growth of oil palm (with LCC)
tends to counteract the negative impacts of clearance, without always return-
ing to predisturbance levels. Runoff and erosion remain high in compacted
areas such as roads, harvest paths, and weeded circles. Due to the high
nutrient-uptake rate and large evaporative demand of the palms, low-nutrient
losses via leaching were generally reported in oil palm plantations in Southeast
Asia despite high-rainfall intensities.
Few studies compared the water and nutrient budgets between young
and mature oil palm stands, although leaching losses at the plot scale were
examined by Foong et al. (1983) and Omoti et al. (1983) and ET was
measured by Yusop et al. (2008). In mature plantations, data were available
from a number of plot scale focusing on single hydrological processes, such
112 Irina Comte et al.
Forest clearing
Increased soil
moisture Increased soil Increased
erosion nutrient
losses
Increased
sediment
loads
Reduced Increased
groundwater stormflow
recharge
5. Conclusion
Since the 1960s, research effort focused on plot-scale trials in South-
east Asia to provide agronomic recommendations for plantation managers
that would increase productivity and economic returns for the palm oil
industry. Growing awareness of environmental impacts from the rapidly
114 Irina Comte et al.
ACKNOWLEDGMENTS
This study was partly funded by the International Cooperation Centre in Agronomic
Research for Development (CIRAD) and the Natural Sciences and Engineering Research
Council of Canada (NSERC).
REFERENCES
Abdul Rahim, N., and Zulkifli, Y. (1987). Stream water quality of undisturbed forest catche-
ments in Peninsular Malaysia. In “Workshop on Impacts of man’s activities on tropical
upland forest ecosystems” (H. Yusuf, et al., Eds.), pp. 289–308. UPM, Serdang.
Abdul Rahim, N., and Harding, D. (1992). Effects of selective logging methods on water
yield and streamflow parameters in Peninsular Malaysia. J. Trop. For. Sci. 5, 130–154.
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 115
Agamuthu, P., and Broughton, W. J. (1985). Nutrient cycling within the developing oil
palm-legume ecosystem. Agric. Ecosyst. Environ. 13, 111–123.
Ah Tung, P. G., Yusoff, M. K., Majid, N. M., Joo, G. K., and Huang, G. H. (2009). Effect
of N and K fertilizers on nutrient leaching and groundwater quality under mature oil
palm in Sabah during the monsoon period. Am. J. Appl. Sci. 6, 1788–1799.
Ahmad, A. L., Chong, M. F., Bhatia, S., and Ismail, S. (2006). Drinking water reclama-
tion from palm oil mill effluent (POME) using membrane technology. Desalination
191, 35–44.
Aldrian, E., and Dwi Susanto, R. (2003). Identification of three dominant rainfall regions
within Indonesia and their relationship to sea surface temperature. Int. J. Climatol. 23,
1435–1452.
Anderson, J. M., and Spencer, T. (1991). Carbon, nutrient and water balances of tropical rain
forest ecosystems subject to disturbance. In “MAB digest 7,” p. 95. UNESCO, Paris.
Andriesse, J. P. (1988). Nature and management of tropical peat soils. FAO Soils Bulletin 59,
Rome.
APOC (American Palm Oil Council). Available at: http://www.americanpalmoil.com/
(consulted on March 7th, 2011).
Banabas, M. (2007). Study of nitrogen loss pathway in oil palm (Elaeis guineensis Jacq.)
growing agro-syatems on volcanic ash soils in Papua New Guinea. PhD thesis, Massey
University, Palmerston North.
Banabas, M., Turner, M. A., Scotter, D. R., and Nelson, P. N. (2008). Losses of nitrogen
fertiliser under oil palm in Papua New Guinea: 1. Water balance, and nitrogen in soil
solution and runoff. Aust. J. Soil Res. 46, 332–339.
Bangun, D. (2006). In “Indonesian palm oil industry”. Paper Presented at the National
Institute of Oilseed Products Annual Convention, Phoenix, Arizona.
Barnett, R. (2007). Top 10 Outcomes 2007. 2011 World Resources Institute, Washington,
DC, USA.
Basiron, Y. (2007). Palm oil production through sustainable plantations. Eur. J. Lipid Sci.
Tech. 109, 289–295.
Bidin, K., Douglas, I., and Greer, T. (1993). Dynamic response of subsurface water levels in a
zero-order tropical rainforest basin, Sabah, Malaysia. In “Hydrology of Warm Humid
Regions” ( J. S. Gladwell, Ed.), Vol. 216, pp. 491–496. IAHS Pub., Wallingford.
Bilotta, G. S., and Brazier, R. E. (2008). Understanding the influence of suspended solids on
water quality and aquatic biota. Water Res. 42, 2849–2861.
Bonnell, M. (2005). Runoff geenration in tropical forests. In “Forest-Water-People in the
Humid Tropics: Pas, Present and Future Hydrological Research for Integrated Land and
Water Management” (M. Bonnell and L. A. Bruinjnzeel, Eds.), pp. 314–406. Cambridge
University Press, Cambridge.
Breure, K. (2003). The search for yield in oil palm: Basic principles. In “Oil Palm:
Management for Large and Sustainable Yields” (T. Fairhurst and R. Hardter, Eds.),
pp. 59–98. Potash & Phosphate Institute/Potash Institute of Canada and International
Potash Institute, Singapore.
Briggs, A. O., Stanley, H. O., Adiukwu, P. U., and Ideriah, T. J. K. (2007). Impact of palm
oil (Elaeis guineensis Jacq; Banga) mill effluent on water quality of receiving Oloya lake in
Niger Delta, Nigeria. Res. J. Appl. Sci. 2, 842–845.
Brouwer, L. C., and Riezebos, H. T. (1998). Nutrient dynamics in intact and logged tropical
rain forest in Guyana. In “Soils of Tropical Forest Ecosystems. Characteristics, Ecology
and Management” (A. Schulte and D. Ruhiyat, Eds.), pp. 73–86. Springer-Verlag Berlin
Heidelberg, New York.
Bruijnzeel, L. A. (1983a). Hydrological and biogeochemical aspects of man-made forests in
South-central Java, Indonesia. PhD thesis, Free University, Amsterdam. p. 256.
116 Irina Comte et al.
Crowther, J. (1987a). Ecological observations in tropical karst terrain, West Malaysia. II.
Rainfall interception, litterfall and nutrient cycling. J. Biogeogr. 14, 145–155.
Crowther, J. (1987b). Ecological observations in tropical karst terrain. West Malaysia. III.
Dynamics of the vegetation-soil-bedrock system. J. Biogeogr. 14, 157–164.
Danielsen, F., Beukema, H., Burgess, N. D., Parish, F., Brühl, C. A., Donald, P. F.,
Mudiyarso, D., Phalan, B., Reijnders, L., Struebig, M., and Fitzherbert, E. B. (2008).
Biofuel plantations on forested lands: Double jeopardy for biodiversity and climate.
Conserv. Biol. 23, 348–358.
Darussamin, A. (2004). Impact assessment on oil palm development. In “The 2nd Routd-
table Meeting on Sustainable Palm Oil”, Jakarta.
DE (Down to Earth). (2005). Pesticide Use in Oil Palm Plantations. N 66. Available at:
http://dte.gn.apc.org/66pes.htm. (consulted on March 8th, 2011).
DID (Department of Irrigation and Drainage). (1989). In “Sungai Tekam Experimental
Basin Final Report, July 1977 to June 1986. Malaysian Ministry of Agriculture, Depart-
ment of Irrigation and Drainage, Kuala Lampur”.
Dolmat, M. T., Lim, K. H., Zakaria, Z. Z., and Hassan, A. H. (1987). Recent studies on the
effects of Land Application of Palm Oil Mill Effluent on oil palm and the environment.
In “Proceedings of the International Oil Palm/Palm Oil Conferences” (H. A. H. Halim,
P. S. Chew, B. J. Wood, and E. Pushparaja, Eds.), Vol. 2, pp. 596–604. PORIM and
Incorporated Society of Planters, Kuala Lampur.
Douglas, I. (1999). Hydrological investigations of forest disturbance and land cover impacts
in South-East Asia: A review. Philos. Trans. R. Soc. Lond. B Biol. Sci. 354, 1725–1738.
Dufrêne, E. (1989). Photosynthèse, consommation en eau et modélisation de la production
chez le palmier à huile (Elaeis guineensis Jacq.). PhD Thesis, Université de Paris XI, Orsay,
Paris, pp. 156.
Dufrêne, E., Ochs, R., and Saugier, B. (1990). Oil palm photosynthesis and productivity
linked to climatic factors. Oleagineux 45, 345–353.
ECD (Environment Conservation Department). (2000). Environmental Impact Assessment
(EIA) Guidelines on Oil Palm Plantation Development. Environmental Conservation
Department, Sabah, Malaysia. Available at: http://www.sabah.gov.my/jpas/programs/
ecd-cab/technical/OP211100.pdf. (consulted on Apr 13th, 2011).
Elsenbeer, H. (2001). Hydrologic flowpaths in tropical rainforest soilscapes: A review.
Hydrol. Processes 15, 1751–1759.
Fairhurst, T. (1996). Management of Nutrients for Efficient Use in Smallholder Oil Palm
Plantations. Wye College, London.
Fairhurst, T., and Mutert, E. (1999). Introduction to oil palm production. Better Crops Int.
13, 3–6.
Fairhust, T., Caliman, J.-P., Härdter, R., and Witt, C. (2005). Oil Palm: Nutrient Disorders
and Nutrient Managemen: diagnosis, causes, prevention, treatmentt. Potash & Phosphate
Institute/Potash & Phosphate Institute of Canada and International Potash Institute (PPI/
PPIC and IPI), Singapore.
FAO (Food and Agriculture Organization of the United Nations). (1990). Irrigation Water
Management: Irrigation methods, by C. Brouwer, K. Prins , M. Kay & M. Heibloem.
Irrigation water management Training manuals No. 5. Rome.
FAO (Food and Agriculture Organization of the United Nations). (2005). Fertilizer Use by
Crop in Indonesia, FAO, Natural Resources Management and Environment. Available
at: http://www.fao.org/docrep/008/y7063e/y7063e00.htm. (consulted on Feb 14th,
2011).
FAOSTATS, 2011. Available at: http://faostats.fao.org. (consulted on Feb 12th, 2011).
Foong, S. F. (1993). Potentail evaporation, potential yields and leaching losses of oil palm.
In “Proceeding of the 1991 PORIM International Palm Oil Conference” (Y. Basiron,
S. Jalani, K. C. Chang, S. C. Cheah, I. E. Henson, N. Kamarudin, K. Paranjothy,
118 Irina Comte et al.
N. Rajanaidu, and D. Tayeb, Eds.), pp. 105–119. Agriculture Palm Oil Research
Institute, Kuala Lampur.
Foong, S. F., Syed Sofi, S. O., and Tan, P. Y. (1983). A lysimetric simulation of leaching
losses from an oil palm field. In “Proceedings of the Seminar on Fertilizers in Malaysian
Agriculture” (C. P. Soon, K. B. Lian, W. S. W. Harun, and Z. Z. Zakaria, Eds.),
pp. 45–68. Malaysian Society of Soil Science, Kuala Lampur.
Foster, H. L. (1995). Experience with fertilizer recommendation systems for oil palm.
In “Proceedings of the 1993 PORIM International Palm Oil Congress—Agriculture”
(B. S. Jalani, et al., Eds.), pp. 313–328. Palm Oil Research Institute, Kuala Lampur.
Foster, H. L., and Chang, K. C. (1977). The diagnosis of the nutrient status of oil palms in
West Malaysia. In “International Development in Oil Palms” (D. A. Earp and
W. Newall, Eds.), pp. 290–312. Incorporated Society of Planters, Kuala Lampur.
Foster, H. L., Tarmizi Mohammed, A., and Zin, Z. Z. (1986). Fertilizer recommendations
for oil palm in Peninsular Malaysia. PORIM Techonol. 13, 1–25.
Foster, H. L., Tarmizi Mohammed, A., and Zakaria, Z. Z. (1988). Foliar diagnosis of
oil palm in Peninsular Malaysia. In “Proceedings of the 1987 International Oil Palm/
Palm Oil Conference—Agriculture” (H. A. H. Halim, P. S. Chew, B. J. Wood, and
E. Pushparajah, Eds.), pp. 244–261. PORIM and Incorporated Society of Planters, Kuala
Lampur.
Galdikas, B. M. (2009). Indonesian monsoons. Climate, Forest Ecology and Orangutans.
Available at: http://www.orangutan.org/archives/550. (consulted on Feb 14th, 2011).
Gasim, M. B., Toriman, M. E., Rahim, S. A., Islam, M. S., Chek, T. C., and Juahir, H.
(2006). Hydrology and water quality and land-use assessment of Tasik Chini’s feeder
rivers, Malaysia. Geografia 3, 1–16.
Germer, J., and Sauerborn, J. (2008). Estimation of the impact of oil palm plantation
establishment on greenhouse gas balance. Environ. Dev. Sustainability 10, 697–716.
Gillison, A., and Liswanti, N. (1999). Impact of Oil Palm Plantations on Biodiversity in
Jambi, Central Sumatra, Indonesia. CIFOR. Available at: http://www.asb.cgiar.org/
data/dataset/IDA1AMZB.htm. (consulted on March 9th, 2011).
Glastra, R., Wakker, E., and Richert, W. (2002). Oil Palm Plantations and Deforestation in
Indonesia. What Role Do Europe and Germany Play? WWF, Germany, Amsterdam,
Netherlands.
Godbold, D. L., Fritz, E., and Hüttermann, A. (1988). Aluminum toxicity and forest decline.
Proc. Natl. Acad. Sci. USA 85, 3888–3892.
Goh, K. J. (2004). Fertilizer recommendation systems for oil palm: Estimating the fertilizer
rates. In “Proceedings of MOSTA Best Practices Workshops: Agronomy and Crop
Management” (C. P. Soon and T. Y. Pau, Eds.), pp. 235–268. Malaysian Oil Scientists
and Technologies Association, Kuala Lampur.
Goh, K. J., and Chew, P. S. (1995). Managing soils for plantation tree crops. 1. General soil
management. In “Course on Soil Survey and Managing Tropical Soils” (S. Paramanathan,
Ed.), pp. 228–245. MSSS and PASS, Kuala Lampur.
Goh, K. J., and Härdter, R. (2003). General oil palm nutrition. In “Oil Palm: Management
for Large and Sustainable Yields” (T. Fairhurst and R. Hardter, Eds.), pp. 191–230. PPI/
PPIC and IPI, Singapore.
Goh, K. J., Teo, C. B., Chew, P. S., and Chiu, S. B. (1999). Fertiliser management in oil
palm: Agronomic principles and field practices. In “Fertiliser Management for Oil Palm
Plantations”, p. 44. ISP North-east Branch, Sandakan. Available at: http://www.aarsb.
com.my/AgroMgmt/OilPalm/FertMgmt/Principle/Fertiliser%20management%20in%
20oil%20palm–%20Agronomic%20Principles%20an%E2%80%A6.pdf (downloaded on
Jan 12 th 2011).
Goh, K. J., Härdter, R., and Fairhurst, T. (2003). Fertilizing for maximum return. In “Oil
Palm: Management for Large and Sustainable Yields” (T. Fairhurst and R. Hardter, Eds.),
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 119
pp. 279–306. Potash & Phosphate Institute/Potash & Phosphate Institute of Canada and
International Potash Institute (PPI/PPIC and IPI), Singapore.
Greenpeace. (2011). Deforestation for Palm Oil. Available at:, http://www.greenpeace.org/
usa/en/campaigns/forests/forests-worldwide/paradise-forests/our-work-in-paradise/
(consulted on April 4th, 2011).
Hansen, S. (2007). Feasibility study of performing an life cycle assessment on crude palm oil
production in Malaysia (9 pp). Int. J. LCA 12, 50–58.
Härdter, R., and Fairhurst, T. (2003). Introduction: Oil palm: Management for large and
sustainable yields. In “Oil Palm: Management for Large and Sustainable Yields”
(T. Fairhurst and R. Härdter, Eds.), pp. 1–12. PPI/PPIC IPI, Singapore.
Hartley, C. W. S. (1988). The Oil Palm Longman, London.
Henson, I. E. (1999). Comparative ecophysiology of oil palm and tropical rain forest. In “Oil
Palm and the Environment—A Malaysian Perspective” (G. Singh, L. Kim Huan,
T. Leng, and D. Lee Kow, Eds.), pp. 9–39. Malaysian oil palm grower’s concil, Kuala
Lampur.
Hooijer, A., Silvius, M., Wosten, H., and Page, S. (2006). Peat-CO2: Assessment of CO2
emissions from drained peatlands in South East Asia. Wetlands International. In “Delft
Hydraulics 36,” Delft, The Netherlands.
Hui, C. M. (2008). “Comparison of Rainfall Runoff Characteristics and Evapotranspiration
in Oil Palm Catchments”, p. 204. Universiti Teknologi Malaysia, Kuala Lampur,
Malaysia.
IEG (Independent Evaluation Group). (1993). Nucleus Estates and Smallholders Projects in
Indonesia the World Bank Group. Available at: http://lnweb90.worldbank.org/oed/
oeddoclib.nsf/DocUNIDViewForJavaSearch/95d104dd2107d21d852567f5005d8461?
OpenDocument&Click¼ (consulted on Jan 10th, 2011).
IMA (Indonesian Ministry of Agriculture). (2010). Area and Production by Category of
Producers: Palm Oil, Direktorat Jenderal Perkebunan. Kementerian Pertanian. Available
at: http://ditjenbun.deptan.go.id/index.php/direktori/3-isi/4-kelapa-sawit.html. (con-
sulted on April 13th, 2011).
Jacquemard, J. C. (1995). Le palmier à huile. Maisonneuve et Larose, Paris.
Johnson, D., Cole, D. W., and Gessel, S. P. (1975). Processes of nutrient transfer in a tropical
rainforest. Biotropica 7, 208–215.
Kee, K. K., and Chew, P. S. (1996). Nutrient losses through surface runoff and soil
erosion—Implications for improved fertilizer efficiency in mature oil palms.
In “Proceedings of the PORIM Internation Palm Oil Congress” (A. Ariffin,
M. B. Wahid, N. Rajanaidu, D. Tayeb, K. Paranjothy, S. C. Cheah, K. C. Chang,
and S. Ravigadevi, Eds.), pp. 153–169. Palm Oil Research Institute of Malaysia, Kuala
Lampur.
Kee, K. K., and Goh, K. J. (2006). Efficient fertiliser management for higher productivity
and sustainability in oil palm production. In “Proceedings of the International Planters
Conference on Higher Productivity and Efficient Practices for Sustainable Plantation
Agriculture”, Technical Papers, Vol. 1, pp. 157–182. Incorporated Society of Planters,
Kuala Lampur.
Kee, K. K., Goh, K. J., Chew, P. S., and Tey, S. H. (1994). An integrated site specific
fertilizer recommendation system (INFERS) for high productivity in mature oil palms.
In “Management for Enhanced Profitability in Plantations” (K. H. Chee, Ed.),
pp. 83–100. The Incorporated society of Planters, Kuala Lampur.
Kee, K. K., Goh, K. J., and Chew, P. S. (2000). Water cycling and balance in a mature oil
palm agroecosystem in Malaysia. In “Proceedings of the International Planters Confer-
ence” (E. Pushparajah, Ed.), pp. 153–169. The Incorporated society of Planters, Kuala
Lampur.
120 Irina Comte et al.
Kenworthy, J. B. (1971). Water and nutrient cycling in a tropical rain forest. In “The Water
Relations of Malesian Forests” (J. R. Flenley, Ed.), Miscellaneous Series No. 11, p. 11.
University of Hull.
Kemp, P., Sear, D., Collins, A., Naden, P., and Jones, I. (2011). The impacts of fine
sediment on riverine fish. Hydrol. Processes 25, 1800–1821. 10.1002/hyp.7940.
Khalid, H., Zin, Z. Z., and Anderson, J. M. (2000). Nutrient cycling in an oil palm
plantation: The effects of residue management practices during replanting on dry matter
and nutrient uptake of young palms. J. Oil Palm Res. 12, 29–37.
Koh, L. P., and Ghazoul, J. (2008). Biofuels, biodiversity, and people: Understanding the
conflicts and finding opportunities. Biol. Conserv. 141, 2450–2460.
Koh, L. P., and Wilcove, D. S. (2008). Is oil palm agriculture really destroying tropical
biodiversity?. Conserv. Lett. 2, 1–5.
Kok, T. F., Goh, K. J., Chew, P. S., Gan, H. H., Heng, Y. C., Tey, S. H., and Kee, K. K.
(2000). Advances in oil palm agronomic recommendations. In “Plantation Tree Crop in
the New Millennium: The Way Ahead” (E. Pushparajah, Ed.), pp. 215–232. The
Incorporated society of Planters, Kuala Lampur.
Lamade, E., and Bouillet, J.-P. (2005). Carbon storage and global change: The role of oil
palm. Oléaginaux Corps Gras Lipides 12, 154–160.
Ling, A. H. (1979). Some Lysimetric measurements of evapotranspiration of oil palm in
Central Peninsular Malaysia. In “Proceedings Symposium on Water in Malaysian agri-
culture”, pp. 89–99. Malaysian Soil Science Society, Kuala Lampur.
Loong, S. G., Mazeeb, M., and Letchumanan, A. (1987). Optimising the use of EFB mulch
on oil palms on two different soils, p. 24. PORIM International Palm Oil Development
Conference, Kuala Lampur.
Lord, S., and Clay, J. (2006). Environmental impacts of oil palm—Practical considerations in
defining sustainability for impacts on air, land and water. In “International Planters
Conference on Higher Productivity and Efficient Practices for Sustainable Agriculture,”
The Incorporated Society of Planters, Putrajaya.
Low, K. S., and Goh, G. C. (1972). The water balance of five catchments in Selangor.
J. Trop. Geogr. 35, 60–66.
Mackensen, J., Hölscher, D., Klinge, R., and Fölster, H. (1996). Nutrient transfer
to the atmosphere by burning of debris in eastern Amazonia. For. Ecol. Manage. 86,
121–128.
Maena, L. M., Thong, K. C., Ong, T. S., and Mokhtaruddin, A. M. (1979). Surface wash
under mature oil palm. In “Proceedings of the Symposium on Water Agriculture in
Malaysia” (E. Pushparajah, Ed.), pp. 203–216. Malaysian Society of Soil Science, Kuala
Lampur.
Malmer, A. (1996). Hydrological effects and nutrient losses of forest plantation establishment
on tropical rainforest land in Sabah, Malaysia. J. Hydrol. 174, 129–148.
Malmer, A., and Grip, H. (1994). Converting tropical rainforest to forest plantation in Sabah,
Malaysia. Part II. Effects on nutrient dynamics and net losses in streamwater. Hydrol.
Processes 8, 195–209.
Ministerial Decree on Agriculture. (2009). Peraturan Menteri Pertanian nomor: 14/Per-
mentan/PL.110/2/2009.
Mongabay Editorial. (2006). Forest Fires Result from Government Failure in Indonesia.
Available at: http://news.mongabay.com/2006/1015-indonesia.html. (consulted on
March 4th, 2011).
Mutert, E., Fairhurst, T. H., and von Uexküll, H. R. (1999). Agronomic management of oil
palms on deep peat. Better Crops Int. 13, 22–27.
Nantha, H. S., and Tisdell, C. (2009). The orangutan–oil palm conflict: Economic con-
straints and opportunities for conservation. Biodivers. Conserv. 18, 487–502.
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 121
Nellemann, C., Miles, L., Kaltenborn, B. P., Virtue, M., and Ahlenius, H. (2007). The Last
Stand of the Orangutan—State of Emergency: Illegal Logging, Fire and Palm Oil in
Indonesia’s National Parks. UNEP GRID-Arendal, UNEP-WCMC, UNESCO, Vol.
2011.
Ng, S. K. (2002). Nutrition and nutrient management of the oil palm. New thrust for the future
perspective. In “Potassium for Sustainable Crop Production. International Symposium on
Role of Potassium in India” (N. S. Pasricha and S. K. Bansal, Eds.), pp. 415–429. Potash
Research Institute of India, and International Potash Institute, New Delhi.
Ng, S. K., and Thamboo, S. (1967). Nutrient contents of oil palms in Malaysia. I. Nutrients
in reproductive tissue fruit bunches and male inflorescence. Malays. Agric. J. 46, 3–15.
Ng, S. K., Thamboo, S., and de Souza, P. (1968). Nutrient contents of oil palms in Malaysia.
II. Nutrients in reproductive tissue, fruit bunches and male inflorescence. Malays. Agric. J.
46, 332–391.
Ng, P. H. C., Chew, P. S., Goh, K. J., and Kee, K. K. (1999). Nutrient requirements and
sustainability in mature oil palms—An assessment. The Planter 75, 331–345.
Ng, S. K., von Uexküll, H., and Härdter, R. (2003). Botanical aspects of the oil palm
relevant to crop management. In “Oil Palm: Management for Large and Sustainable
Yields” (T. Fairhurst and R. Härdter, Eds.), pp. 13–26. Potash & Phosphate Institute/
Potash Institute of Canada and International Potash Institute, Singapore.
Noguchi, S., Nik, A., Yusop, Z., Tani, M., and Sammori, T. (1997a). Rainfall-runoff
responses and roles of soil moisture variations to the response in tropical rain forest,
bukit tarek, peninsular Malaysia. J. Forest Res. 2, 125–132.
Noguchi, S., Rahim Nik, A., Kasran, B., Tani, M., Sammori, T., and Morisada, K. (1997b).
Soil physical properties and preferential flow pathways in tropical rain forest, Bukit
Tarek, Peninsular Malaysia. J. Forest. Res. 2, 115–120.
Noguchi, S., Abdul Rahim, N., Shamsuddin, S. A., Tani, M., and Sammori, T. (2004).
Evapotranspiration estimates of a tropical rain forest, Bukit Tarek Experimental Water-
shed in Peninsular Malaysia, using short-time period water-budget method. J. Jpn. Soc.
Hydrol. Water Resour. 17, 427–444.
Noor, Y. R., Cahyo Wibisono, I. T., and Suryadiputra, I. N. (2007). Poverty alleviation
projects in peatland in Indonesia. In “Proceedings of the International Symposium and
Workshop on Tropical Peatland” (J. O. Rieley, C. J. Banks, and B. Ragkagukguk, Eds.),
pp. 177–186. Gadjah Mada University, Indonesia and University of Leicester, United
Kingdom, Yogyakarta, Indonesia, EU CARBOPEAT and RESTORPEAT Partnership.
Oil World, 2011. Available at: http://www.oilworld.biz/app.php?fid¼310&fpar¼YToy
OntzOjI6IklkIjtzOjQ6IjYzMTQiO3M6NDoicGNpZCI7czoyOiIxMiI7fQ%3D%3D
&isSSL¼0&aps¼0&blub¼26bc92353fe185973450e4160e44224d&ista¼apdxujfvn (con-
sulted on May 6th, 2011).
Okwute, L. O., and Isu, N. R. (2007). The environmental impact of palm oil mill effluent
(pome) on some physico-chemical parameters and total aerobic bioload of soil at a dump
site in Anyigba, Kogi State, Nigeria. Afr. J. Agric. Res. 2, 656–662.
Olaleye, V. F., and Adedeji, A. A. (2005). Planktonic and water quality of palm oil effluent
impacted river in Ondo state, Nigeria. Int. J. Zool. Res. 1, 15–20.
Omoti, U., Ataga, D., and Isenmila, A. (1983). Leaching losses of nutrients in oil palm
plantations determined by tension lysimeters. Plant Soil 73, 365–376.
Osuji, G. E., Okon, M. A., Chukwuma, M. C., and Nwarie, I. I. (2010). Infiltration
characteristics of soils under selected land use practices in Owerri, Southeastern Nigeria.
World J. Agric. Sci. 6, 322–326.
Othman, A., Mohammed, A. T., Harun, M. H., Darus, F. M., and Mos, H. (2010). Best
management practices for oil palm planting on peat: Optimum groundwater table.
Malaysian Palm Oil Board (MPOB) Information series 472.
122 Irina Comte et al.
Pushparajah, E. (1994). Leaf Analysis and Soil Testing for Plantation Tree Crops. Interna-
tional Board for Soil Research and Management (IBSRAM), Thailand. Available at:
http://www.agnet.org/library/eb/398/. (consulted on Jan 13th 2011).
Radersma, S., and deRidder, N. (1996). Computed evapotranspiration of annual and
perennial crops at different temporal and spatial scales using published parameter values.
Agric. Water Manag. 31, 17–34.
Rankine, I. R., and Fairhurst, T. (1998a). Field Handbook—Oil Palm Series. Nursery.
Vol. 1, Potash and Phosphate Institute, Singapore.
Rankine, I. R., and Fairhurst, T. (1998b). Field Handbook—Oil palm. Immature. Vol. 2,
Potash and Phosphate Institute, Singapore.
Rankine, I. R., and Fairhurst, T. (1998c). Field Handbook—Oil palm. Mature. Vol. 3,
Potash and Phosphate Institute, Singapore.
Rieley, J. O. (2007). Tropical peatland - The amazing dual ecosystem: coexistence and
mutual benefit. In “Proceedings of the international symposium and workshop on
tropical peatland” (J. O. Rieley, C. J. Banks, and B. Ragkagukguk, Eds.), pp. 1–14.
Gadjah Mada University, Indonesia and University of Leicester, United Kingdom,
Yogyakarta, Indonesia, EU CARBOPEAT and RESTORPEAT Partnership.
Rieley, J. O., and Page, S. O. (1997). Biodiversity and sustainibility of tropical peatlands.
In “Proceedings of the International Symposium on Biodiversity, Environmental Impor-
tance and Sustainability of Tropical peat and Peatlands. Palangka Raya, Central
Kalimantan, 4-8 September, 1995,” Samara Publishing Limited, Cardigan.
Rist, L., Feintrenie, L., and Levang, P. (2010). The livelihood impacts of oil palm: Small-
holders in Indonesia. Biodivers. Conserv. 19, 1009–1024.
Rupani, P. F., Singh, R. P., Ibrahim, M. H., and Esa, N. (2010). Review of current palm oil
mill effluent (POME) treatment methods: Vermicomposting as a sustainable practice.
World Appl. Sci. J. 11, 70–81.
Schindler, D. W., Armstrong, F. A. J., Holmgren, S. K., and Brunskil, G. J. (1971).
Eutrophication of lake 227, experimental lakes area, Northwester Ontario, by addition
of phosphate and nitrate. J. Fish. Res. Board Can. 28, 1763–1782.
Schrevel, A. (2008). Oil-Palm Estate Development in Southeast Asia: Consequences for Peat
Swamp Forests and Livelihoods in Indonesia. Available at: ftp://ftp.fao.org/docrep/fao/
011/i0314e/i0314e06.pdf. (consulted 10th Jan 2011).
Schroth, G., Rodrigues, M. R. L., and D’Angelo, S. A. (2000). Spatial patterns of nitrogen
mineralization, fertilizer distribution and roots explain nitrate leaching from mature
Amazonian oil palm plantation. Soil Use Manag. 16, 222–229.
Sheil, D., Casson, A., Maijaard, E., van Noordwijk, M., Gaskell, J., Sunderland-groves, J.,
Wertz, K., and Kanninen, M. (2009). The Impacts and Opportunities of Oil Palm in
Southeast Asia Center for International Forestry Research (CIFOR), Bogor.
Siegel, D. I., and Glaser, P. (2006). The hydrology of peatlands. In “Boreal Peatland
Ecosystems” (R. K. Wieder and D. H. Vitt, Eds.), pp. 289–311. Springer, Berlin
Heidelberg.
Singh, R. P., Ibrahim, M. H., Esa, N., and Iliyana, M. S. (2010). Addressing the threats
to biodiversity from oil-palm agriculture. Composting of waste from palm oil mill:
A sustainable waste management practice. Rev. Environ. Sci. Biotechnol. 9, 311–344.
Squire, G. R. (1984). Techniques in environmental physiology of oil palm. 2. Partitioning of
rainfall above ground. PORIM Bull. 9, 1–9.
Strack, M., and Waddington, J. M. (2007). Response of peatland carbon dioxide and
methane fluxes to a water table drawdown experiment. Global Biogeochem. Cycles 21,
GB1007.
Taillez, B. (1998). Oil Palm : A new crop. . . for what future? Oléagineux Corps gras Lipides 5,
106–109.
Tan, K. T., Lee, K. T., Mohamed, A. R., and Bhatia, S. (2009). Palm oil: Addressing issues
and towards sustainable development. Renew. Sustain. Energy Rev. 13, 420–427.
Agricultural Practices, Nutrient Fluxes and Water Quality in Oil Palm Plantations 123
Tanaka, N., Kume, T., Yoshifuji, N., Tanaka, K., Takizawa, H., Shiraki, K., Tantasirin, C.,
Tangtham, N., and Suzuki, M. (2008). A review of evapotranspiration estimates from
tropical forests in Thailand and adjacent regions. Agric. Forest. Meteorol. 148, 807–819.
Tarmizi, M. A., and Mohd, T. D. (2006). Nutrient demands of Tenera oil palm planted on
inlands soils of Malaysia. J. Oil Palm Res. 18, 204–209.
Turner, R. E., and Rabalais, N. N. (1994). Coastal eutrophication near the Mississippi river
delta. Nature 368, 619–621.
Turvey, N. D. (1974). Nutrient cycling under tropical rain forest in central Papua. Occa-
sional paper, 10. Department of Geography. University of Papua New Guinea, Port
Moresby, p. 96.
UN (United Nations). (2004). World Population to 2300. United Nations, Department
of Economic and Social Affairs, Population Division. Available at: http://www.un.org/
esa/population/publications/longrange2/WorldPop2300final.pdf. (consulted on March
9th, 2011).
USDA (United States Department of Agriculture). (2006). Keys to Soil Taxonomy.10th
edn.. United States Department of Agriculture and Natural Resources Conservation
Service, Washington, DC. Available at: http://terra-geog.lemig2.umontreal.ca/donnees/
geo2142/Keys%20taxonomy.pdf. (consulted on May 6th, 2011).
USDA (United States Department of Agriculture). (2007). Indonesian Palm Oil Production.
Available at: http://www.pecad.fas.usda.gov/highlights/2007/12/Indonesia_palmoil/.
(consulted on Jan 7th, 2011).
USDA (United States Department of Agriculture). (2009). Indonesia: Palm Oil Production
Growth to Continue. Available at: http://www.pecad.fas.usda.gov/highlights/2009/03/
Indonesia/. (consulted on March, 17th, 2011).
Vermeulen, S., and Goad, N. (2006). Towards better practices in smallholder palm oil
production. In “Natural Resource Issues Series”, 5, International Institute for Environ-
ment and Development (IIED), London.
von Uexküll, H. R. (2007). Oil Palm (Elaeis guineensis Jacq.). E. & S.E. Asia Program for the
Potash & Phosphate Institute/International Potash Institute, Singapore.
Wahid, M. B., Abdullah, S. N. A., and Henson, I. E. (2005). Oil palm—Achievement and
potential. Plant Prod. Sci. 8, 288–297.
Wakker, E. (1999). Forest Fires and the Expansion of Indonesia’s Oil-Palm Plantations.
WWF-Indonesia, Jakarta.
Webb, M. J., Nelson, P. N., Rogers, L. G., and Curry, G. N. (2011). Site-specific fertilizer
recommendations for oil palm smallholders using information from large plantations.
J. Plant Nutr. Soil Sci. 174, 311–320.
WEPA (Water Environment Partnership in Asia). (2011). State of Water Environmental
Issues. Available at: http://www.wepa-db.net/policies/state/indonesia/indonesia.htm.
(consulted on Feb 25th, 2011).
WHO (World Health Organization). (2008).Guidelines for drinking-water quality: Incor-
porating the 1st and 2nd addenda. 3rd edn. , Recommendations. 1, World Health
Organization, Geneva.
WI (Wetlands International). (2010). Input to World Bank Group Palm Oil Strategy
Consultations, World Bank Group Framework & IFC Strategy in the Palm Oil Sector.
Available at: http://www.ifc.org/ifcext/agriconsultation.nsf/Content/2E1FC2F0E9F1
D8FD8525772F0073CFB6?OpenDocument. (consulted on March 1st, 2011).
Wilkinson, G. E., and Aina, P. O. (1976). Infiltration of water into two Nigerian soils under
secondary forest and subsequent arable cropping. Geoderma 15, 51–59.
Wood, B. J., Pillai, K. R., and Rajaratnam, J. A. (1979). Palm oil mill effluent disposal on
land. Agric. Wastes 1, 103–127.
Wösten, J. H. M., Ismail, A. B., and van Wijk, A. L. M. (1997). Peat subsidence and its
practical implications: A case study in Malaysia. Geoderma 78, 25–36.
124 Irina Comte et al.
Wösten, J. H. M., Clymans, E., Page, S. E., Rieley, J. O., and Limin, S. H. (2008). Peat-
water interrelationships in a tropical peatland ecosystem in Southeast Asia. Catena 73,
212–224.
WRI (World Resource Institute). (2002). The State of the Forest: Indonesia. Available at:
http://www.wri.org/publication/state-of-the-forest-indonesia. (consulted on March
9th, 2011).
WRM (World Rainforest Movement. (2002). The Direct and Underlying Causes of Forest
Loss, WRM Briefings. Available at: http://www.wrm.org.uy/. (consulted on April 7th,
2011).
WWF (World Wild Fund). (2011). Orangutans Under Threat. How Unsustainable Palm Oil
Is Destroying the Orangutan’s Home. Available at: http://www.wwf.org.uk. (consulted
on March 9th, 2011).
Yi Jing, C., Mei Fong, C., and Chung Lim, L. A. W. (2010). Biological treatment of
anaerobically digested palm oil mill effluent (POME) using a Lab-Scale Sequencing Batch
Reactor (SBR). J. Environ. Manage. 91, 1738–1746.
Yusop, Z., Douglas, I., and Nik, A. R. (2006). Export of dissolved and undissolved nutrients
from forested catchments in Peninsular Malaysia. For. Ecol. Manage. 224, 26–44.
Yusop, Z., Chan, C. H., and Katimon, A. (2007). Runoff characteristics and application of
HEC-HMS for modelling stormflow hydrograph in an oil palm catchment. Water Sci.
Technol. 56, 41–48.
Yusop, Z., Hui, C. M., Garusu, G. J., and Katimon, A. (2008). Estimation of evapotranspi-
ration in oil palm catchment by short-time period water-budget method. Malays. J. Civil
Eng. 20, 160–174.
Zaidel’man, F. (2008). Protecting soils against degradation. Herald Russ. Acad. Sci. 78,
362–369.
Zen, Z., Barlow, C., and Gondowarsito, R. (2005). Oil Palm in Indonesian Socio-
Economic Improvement—A Review of Options. Department of Economics, Research
School of Pacific and Asian Studies, Australian National University, Canberra.
Zhang, X., Yu, X., Wu, S., and Liu, H. (2007). Effects of forest vegetation on runoff
and sediment transport of watershed in Loess area, west China. Front. Forest. China 2,
163–168.
Zulkifli, Y., Anhar, S., and Mohd Fauzi, Z. (1987). Effects of selective logging on physical
stream water quality in hill tropical rain forests. In “Paper Presented at Workshop on
Impact of Operations in Natural and Plantation Forests on Conservation of Soil and
Water Resources,” Universiti Pertanian Malaysia, Serdang, Malaysia. (23–26th June).
Zulkifli, Y., Abdul Rahim, N., Anhar, S., and Fauzi Zakaria, M. (1989). Rainfall chemistry
and nutrient loading in a Peninsular Malaysia forest site. J. Trop. For. Sci. 1, 201–214.
Zulkifli, Y., Abdul Rahim, N., and Baharuddin, K. (1998). Evapotranspiration loss from a
second growth forest of Peninsular Malaysia. In “International Conference of Hydrology
and Water resources of Humid Tropics,” Pangkor, Malaysia. (24–26th Nov).
C H A P T E R F O U R
Pathways to Agroecological
Intensification of Soil Fertility
Management by Smallholder Farmers
in the Andean Highlands
Steven J. Fonte,* Steven J. Vanek,† Pedro Oyarzun,‡
Soroush Parsa,* D. Carolina Quintero,* Idupulapati M. Rao,* and
Patrick Lavelle*,§
Contents
1. Introduction: Agricultural and Soil Fertility Issues in the High Andes 126
1.1. Cropping systems of the Andes 126
1.2. Biophysical limitations and risks 127
1.3. Socioeconomic and cultural setting of Andean agriculture 129
1.4. Current challenges and emerging threats 131
1.5. Ecologically based intensification in the Andean context 132
2. Examining Soil Fertility and Management Strategies in
Smallholder Systems 133
2.1. General concept of soil fertility 133
2.2. Approaches to examining soil fertility 133
3. Additional Considerations for Soil Fertility Interventions 162
3.1. Need to incorporate co-limiting crop growth factors 162
3.2. Integrating local needs and knowledge into soil
fertility research 162
4. Conclusions and Recommendations 163
4.1. Outlook for agroecological intensification in the
Andean context 163
4.2. Recommendations for future research and interventions 165
Acknowledgments 166
References 166
125
126 Steven J. Fonte et al.
Abstract
Small farmers in the high Andes (>2500m) of Bolivia, Ecuador, and Peru face
increasing threats to their livelihoods due to land degradation, climate change,
and overall decreases in agricultural productivity. The fragile nature of these
agroecosystems and limited capacity of resource-poor farmers in the region to
adopt the large-scale use of conventional fertilizer and pest control technolo-
gies suggest the need for agroecological intensification to restore soil function-
ing and ensure long-term sustainability in these systems. This review addresses
soil fertility decline from a management perspective and considers six basic
approaches to enhance nutrient cycling, crop nutrient acquisition, and long-
term productivity. A mass balance approach first defines basic boundaries for
nutrient cycling and suggests that erosion control and identification of alterna-
tive nutrient sources (e.g., peri-urban wastes, rock phosphate) are critical for
reversing negative nutrient budgets. Meanwhile, short-term nutrient dynamics
could benefit greatly from improved management of organic residues in combi-
nation with low-level inorganic fertilizer applications. There is also a need for
greater understanding of soil physiochemical properties throughout much of the
Andes and the impacts of management. Similarly, soil biological functioning is
critical for successful agroecological intensification and there is great potential
for both inoculative and management strategies to promote beneficial soil
communities. Crop breeding for smallholder environments should complement
strategies of agroecological intensification, taking advantage of high regional
agrobiodiversity and experience from stress breeding programs in other regions.
Finally, we suggest several means by which the spatial and temporal organization
of farms may be improved to enhance overall agroecosystem function.
A N Annual
B N Annual mean
W E
Precipitation (mm) W E
temperature (°C)
S S
0–261 –6–7
262–610 8–12
611–792 13–18
793–1010 19–22
1011–1357 23
1358–1627 24
1628–1875 25
1876–2276 26
2277–2786 27
2787–6081
Figure 1 Average annual precipitation (A) and temperature (B) for the Andean focus
region (Ecuador, Peru, and Bolivia). The maps were generated using the WorldClim
database (http://www.worldclim.org) following methods of Hijmans et al. (2005).
formed. For example, allophanic Andosols represent some of the most fertile
soils in the world, while non-allophanic Andosols, which dominate the
Ecuadorean highlands (Poulenard et al., 2001), can have severe problems of
phosphorus deficiency (due to high P-fixation), acidity, and aluminum tox-
icity (Dahlgren et al., 2004). Meanwhile, soils of the central Andes (southern
Peru and Bolivia; Fig. 2) are generally low in SOM and nutrient content,
coarser in texture, highly susceptible to erosion, and on average have a more
neutral pH (Bottner et al., 2006; Cárdenas et al., 2008; Valente and Oliver,
1993). Despite these generalities, it should be noted that targeted manage-
ment with manure and fallows on less vulnerable parts of farmed landscapes
can lead to localized areas of high fertility that support impressive productivity
(Garcı́a, 2011). Also, in contrast to the highly weathered lowland soils
commonly associated with tropical latitudes, many soils in the central Andes
are similar to Ustic soils found in temperate regions (e.g., southern Europe,
parts of North America), thus permitting some degree of knowledge transfer
on management from these better studied regions.
It is generally accepted that crops face the greatest threat from specialist
herbivores, both pests and pathogens, in their centers of origin ( Jennings
and Cock, 1977). The Andean region is no exception. For example,
potatoes in the Andes face the dual threats of potato late blight and several
species of potato tuber moths, weevils, and nematodes, each with the
Andean Soil Fertility 129
Andosols- N. Andes
Altiplano soils
Figure 2 Broad characterization of soil environments in the high Andes, divided into
four key regions: Andosols, hillslope soils with sedimentary or mixed parent materials,
and soils associated with the Alitplano. Boundaries are approximate and considerable
intergrading occurs among zones.
potential to cause total crop losses. In recent years, the severity of pest
problems in the Andes has increased significantly, presumably due to the
influences of agricultural intensification and climate change, among other
factors (Parsa, 2010). Despite abundant research, solutions to these problems
are lagging, particularly for neglected crops such as oca (Oxalis tuberosa),
which are facing increasing farmer abandonment due to heavy pest infesta-
tions (Hersh, 2000).
climate change) have created more demand for their use (Sarmiento et al.,
1993). Locally available inputs of organic matter have also become more
limiting, due to increasing demand for fuel and fodder, as well as lower
biomass production—driven by declining soil fertility and competing land
uses (Orsag, 2009; Swinton and Quiroz, 2003). A long tradition of agriculture
in the region, combined with the preservation of indigenous cultures and
knowledge, has endowed many Andean communities with a sophisticated
level of agroecological management (Sandor and Furbee, 1996; Winklerprins,
1999). However, due to rapidly shifting agricultural, climatic, and socioeco-
nomic contexts, new information and increased knowledge exchange are
needed to help farmers adapt to emerging challenges.
Access to agricultural inputs and external knowledge is especially
limited for rural communities located far from population centers. Due
to their isolation, imported products have a higher cost and technology
transfer (both externally and locally) is often slow, while the distance to
markets is so great that farmers have little incentive to invest in their land
beyond what is needed for home consumption and local trade (Swinton
and Quiroz, 2003). Poverty also represents a significant obstacle, as poor
farmers have neither the monetary resources to invest in new technologies
nor the economic luxury to assume the risk that is associated with these
activities. These issues are particularly relevant for the Altiplano, where
poverty levels greatly exceed the national averages for Bolivia and Peru,
and over 60% of the population lives in rural areas (Quiroz et al., 2003).
Labor shortages can also limit farmer investment in rural areas, as more
farming families seek to diversify their incomes through off-farm employ-
ment (Zimmerer, 1993).
In addition to resource access, other factors influence farm productivity
and farmers’ ability to adopt new technologies. For example, land tenure
can greatly influence farmer decisions, as farmers are generally reluctant to
invest in long-term agriculture improvements (i.e., soil conservation struc-
tures, proper management of SOM) on land that they do not own (Tenge
et al., 2004). Ongoing shifts in land management from communal to private
control represent another unique facet of Andean agriculture that effectively
results in smaller management units and intensification of these fields
(Cárdenas et al., 2008; Mayer, 1979). Despite the potential for more rapid
innovation to occur under individualized management, growing evidence
suggests that a lack of coordinated community decision-making about
agricultural practices can exacerbate problems of soil fertility, livestock
management, and pest regulation (Mayer, 1979; Orsag, 2009; Parsa,
2010). Local and cultural food preferences also affect farmers’ use of agri-
cultural inputs and practices. For example, many farmers choose not to
grow improved potato varieties or use agrochemicals for crops consumed at
home or sold locally, as they feel that traditional practices and varieties yield
superior taste and quality (Caycho-Ronco et al., 2009).
Andean Soil Fertility 131
soil nutrient depletion (van Groenigen et al., 2006). Decreased plant cover
and increased variability in precipitation in some areas will likely result in
greater erosion and subsequent nutrient loss. At the same time, higher
temperatures will likely accelerate losses of soil C due to decomposition of
SOM (Buytaert et al., 2010a; Davidson and Janssens, 2006) and can nega-
tively impact soil decomposer communities (Briones et al., 2009).
High microclimatic heterogeneity, complex topography, and low reso-
lution of global climate models generate substantial uncertainty regarding
future climate predictions in the Andes and limit the ability of farmers and
regional decision makers to anticipate and adapt to these changes (Buytaert
et al., 2010b; Valdivia et al., 2010). Alterations to climate along with
agricultural intensification may lead to increased pest pressure, as warmer
weather can favor pest reproduction and weaken plant resistance due to
increased water stress (Bale et al., 2002; Garret et al., 2006). Farmer surveys
and preliminary evidence suggest that climate change is already yielding
negative impacts on Andean croplands via enhanced water deficits, erosion,
and SOM loss (Aguilera, 2010) and emphasizes the need to develop new
agricultural strategies that improve ecological resilience and flexibility of
agroecosystems.
Community/regional/national scales
Market
Inorganic
export
fertilizers
Agricultural field
Organic N Internal
(labile and Inorganic N cycling
recalcitrant pools)
Crop N
removal
Losses
Gaseous Leaching
Erosion
N losses
Figure 3 Generalized nitrogen flow diagram for a typical smallholder field, incorpor-
ating spatial scales of nutrient sources. Rectangles represent the boundaries of the field
(inner solid rectangle), household or farm management unit (outer rectangle), and
boundary between the community scale and urban markets that supply inorganic
fertilizer and other external fertility materials and also are a destination of marketed
agricultural products and N export. Drawing shows that some N is cycled internal to
the field and most crops and residue nutrients are cycled internally within farms or lost
to erosion. Rangeland can represent an important fertility resource accessed via live-
stock grazing in extensive systems in the central Andes. Arrows represent approximate
size of loss pathways.
Andean Soil Fertility 135
critical throughout the Andes from both a nutrient balance perspective and
more general goals of agroecosystem sustainability.
The susceptibility of soils to erosion depends greatly not only on topog-
raphy, slope, climate, and soil type, but also to a large extent on management
factors such as vegetation cover, cropping system, tillage, and livestock
intensity (Coppus et al., 2003; Inbar and Llerena, 2000). Given the widely
acknowledged role of agricultural disturbance in exacerbating erosion
(Montgomery, 2007), a number of technologies have been put forth to
conserve soils in the Andes and elsewhere. For steeper hillside farms, stone
and earthen terraces represent one of the oldest approaches to soil conser-
vation in the region and have proven largely successful in combating erosion
(Goodman-Elgar, 2008; Sandor and Eash, 1995), yet these structures may
no longer represent a viable option due to high labor requirements and
altered political and socioeconomic conditions (Dehn, 1995; Posthumus
and De Graaff, 2005). Terraces formed by live barriers are attractive as they
require considerably lower initial investments, offer key byproducts (e.g.,
fodder, fuel, organic matter inputs), and can effectively control erosion
(Craswell et al., 1998; Sims et al., 1999). Sims et al. (1999) working in
Bolivia suggested that grasses (Phalaris sp.) offer the best erosion control in
such systems, as they grow rapidly and provide fodder for livestock. How-
ever, competition for nutrients and water, along with high spatial variability
within live barrier terraces, indicates that such technologies require addi-
tional study and/or modification (Dercon et al., 2006). Cover cropping
offers another viable means to control erosion on hillside soils, by protecting
the soil when and where crops are absent (Bunch, 2004; Sims et al., 1999).
The development of adapted cover crop technologies for erosion control in
the Andes is highly desirable, as cover crops can also contribute significantly
to soil nutrient and organic matter stores (Snapp and Silim, 2002; Wheeler
et al., 1999). Despite the great potential for plant-based technologies to
control erosion, maintenance costs (e.g., labor, seed) and knowledge
requirements can limit their adoption by small farmers (Bunch, 2004;
Posthumus et al., 2010; Snapp et al., 1998). Other erosion control strategies
focus on reduced soil disturbance (tillage) in order to decrease labor require-
ments and conserve soils. For example, the Wacho rozado system from
Ecuador (Sherwood et al., 1999) as well as the chiwa and chacmeo systems
from the Central Andes (Oswald et al., 2009) all represent unique indige-
nous potatoes planting systems that seek to minimize soil movement on
smallholder farms. Reduced tillage options have also been developed for
larger farms that employ animal traction or mechanized tillage (Mamani
et al., 2001; Quintero, 2009; Wall, 1999) and may be especially relevant for
hillside farms in the northern Andes where mechanical tillage is more
common and poses a greater threat. However, significant investment is
still needed to work with farmers in developing locally adapted technologies
and implements that minimize soil disturbance and provide viable
Andean Soil Fertility 137
IF
Nutrient availability/uptake
Combined OR +IF
OR
Plant demand
Figure 4 Theoretical curves for plant nutrient uptake an availability under different
fertility strategies (IF, Inorganic fertilizer; OR, Organic resources). Curves for plant
demand and organic matter decay represent what might be seen during a typical
growing season for a crop such as maize or potato. Twin peaks for IF represent two
fertilizations (at planting and at peak crop demand).
138 Steven J. Fonte et al.
Organic inputs have some challenging aspects for soil fertility management.
For example, compared to inorganic fertilizers, more must be transported
and applied due to their lower nutrient contents. Additionally, the mineral-
ization of plant nutrients from organic matter for plant uptake is often
harder to predict, as amount and timing of nutrient release depend on a
number of factors including organic matter quality, moisture, temperature,
and the soil decomposer community receiving organic additions (Lavelle
et al., 1993; Palm et al., 2001). The quality of organic resources (nutrient
content and ease of decomposition) also tends to be more heterogeneous
and can depend greatly on the source of material and timing of collection
or application. Thus, additional labor and knowledge may be required
to effectively manage organic resources, such that nutrient release is syn-
chronized with crop growth. Despite these potential drawbacks, organic
resources are the most important form of nutrient inputs for crop produc-
tion in the Andean highlands (Caycho-Ronco et al., 2009; Terrazas et al.,
1998) and offer several key benefits to farmers. First, organic resources are
typically less expensive and more readily available to farmers than inorganic
nutrient sources, especially in rural areas. Additionally, Andean smallholders
have developed complex local knowledge that is well adapted to the use of
manures and fallow residues and that may be readily transferable to innova-
tions in organic inputs. Organic resource application also contributes to
the maintenance of SOM in cropping systems (Fernandes et al., 1997), with
vital implications for soil structure, water storage and movement, nutrient
supply and retention, and the promotion of healthy soil biological commu-
nities (Craswell and Lefroy, 2001). Finally, they generally release nutrients
more slowly and over time, potentially decreasing the susceptibility of
nutrients to loss, as compared to synthetic fertilizers (Kramer et al., 2006).
Recognizing the inherent complexities of organic nutrient sources,
researchers have invested considerable effort toward classifying and processing
these materials in order to simplify management, provide a more predictable
release of nutrients, and to enrich materials for reducing bulk and accelerating
nutrient mineralization. For example, Palm et al. (2001) proposed a classifica-
tion system for plant residues and organic materials commonly found in
the humid tropics that is based on residue N content and the presence of
recalcitrant compounds (polyphenols and lignin). This system identifies four
organic matter quality classes with distinct recommendations of how best to
apply the residues for optimal nutrient release. Although a few studies have
examined the quality of organic resources available in the Andes (Coûteaux
et al., 2008; Machado et al., 2010; Mahboubi et al., 1997), this information is
limited and has not been adequately translated into targeted management
recommendations for the Andean agroecosystems.
Efforts for processing organic materials prior to application have become
commonplace throughout the world, as well as in the Andean highlands
(Felipe-Morales, 2002; Herbas, 2000). Composting encompasses a broad
Andean Soil Fertility 141
may not fully recover to the precultivation state for many years. Liming
(e.g., application of calcium carbonate) is perhaps the most commonly
prescribed means to increase the pH in acid soils, but required inputs can
be substantial and the costs prohibitively high for most smallholder farms.
Additions of ash and organic matter have also been shown to reduce acidity
(Aguilera, 2010; Garcı́a et al., 2010) and may represent a more affordable
alternative for Andean farmers in some circumstances. Organic matter
inputs have also been shown to increase CEC (Garcı́a et al., 2010; Oorts
et al., 2003) and can significantly improve other physical and chemical soil
properties such as aggregation, water-holding capacity, and soil buffering
capacity (Craswell and Lefroy, 2001; Kong et al., 2005). Biochar offers
many of the same benefits of organic matter, but with potentially more
immediate and longer lasting impacts. For example, biochar has been shown
to increase soil pH, CEC, water-holding capacity (Chan et al., 2007; Jha
et al., 2010; Karhu et al., 2011) and alter pore size distribution with multiple
impacts for plant growth and soil biological communities (see section 2.2.4).
As for soil structure, research concerning the characterization and manage-
ment of soil chemical environments is notably lacking across much of the
Andes and merits further emphasis in future research endeavors.
of soil disturbance (Carney and Matson, 2005; Thies and Grossman, 2006).
Intensification on smallholder farms often leads to a decrease in the diversity
of residue inputs and an increase in the tillage intensity, thus reducing the
abundance and diversity of soil communities and the SOM which they
depend on (Postma-Blaauw et al., 2010). For example, in maize and wheat
cropping systems of the Mexican highlands, residue retention and reduced
tillage have been shown to foster greater overall bacterial populations,
including beneficial actinomycetes and fluorescent pseudomonad bacteria
(Govaerts et al., 2008). Improved residue management (greater quantity and
diversity of inputs) has also been shown to increase microbial biomass and
diversity when compared to a conventionally managed agriculture (with
low residue inputs) in California (Briar et al., 2011). Root derived C can also
contribute to SOM and soil aggregation and is a key resource for soil
microbes, in addition to the direct role of soluble root exudates from living
plants (Kong and Six, 2010; Puget and Drinkwater, 2001). A closer exami-
nation of different residue management strategies is particularly important
for the Andean region where residues are predominantly removed for
livestock, yet little is known about the consequences of this practice for
soil biological functioning. In addition to management, temporal changes
over the fallow cycle in the Andean region also drive microbial community
dynamics and can provide benchmarks for the functional microbial and
faunal diversity that managed fallows need to provide. For example, in the
Bolivian Altiplano, Sivila and Herve (1999) found changes in the microbial
communities along a chronosequence of fallowed fields, such that mycor-
rhizal spore counts increased with fallow age and SOM levels, but decreased
following rotations of non-mycorrhizal quinoa and with the removal of
Baccharis shrub vegetation.
The remarkable diversity and range of functions attributed to microbes
and fauna in the rhizosphere of crops are now widely acknowledged, growing
out of early knowledge of important plant symbionts such as rhizobia
and mycorrhizae, to learning that microbes mediate many of the processes
in soils that control plant growth (Murphy et al., 2007; Osorio-Vega, 2007;
Rodrı́guez et al., 2006). Symbioses between legumes and rhizobia/bradyrhi-
zobia, and associations between Frankia actinomycetes and trees such as the
agroforestry species Alnus spp., currently account for most of the biologically
fixed N on Earth (Freiberg et al., 1997). Legume species in Andean cropping
systems include endemic Andean lupine (Lupinus mutabilis) infected by Bra-
dyrhizobium lupini; vetch, pea, and fava bean infected by Rhizobium legumino-
sarum bv. vicieae; and alfalfa and various endemic and introduced clover species
(Trifolium spp.). Arbuscular mycorrhizae (AM) are another well-studied sym-
biont that infects the roots of most crop species. They expand the volume
of soil explored by the plant and increase access to P and micronutrients,
especially under dry soil conditions (Aroca and Ruiz-Lozano, 2009). In the
Andes, potatoes, maize, cereals, and legumes like Vicia spp. are hosts for AM.
146 Steven J. Fonte et al.
inoculants since they are strong rhizosphere competitors via their antibiosis
and predation of other fungi, thus addressing a key challenge for inoculants
(Schuster and Schmoll, 2010; Verma et al., 2007). Trials in Bolivia have
found improved health and vigor of onion transplants when inoculated with a
Trichoderma harzianum (Medrano Echalar and Ortuño, 2007). As for other
inoculant research to date, trials on Trichoderma have focused frequently on
greenhouse or horticultural production, whereas future efforts should evaluate
their effectiveness in intensified smallholder crop rotations, especially where
soil disease pressure is high and positive effects would be most expected.
Research on PGPM that improve crop access to P is of particular interest
throughout the region and worldwide (Barroso and Nahas, 2005; Chen et al.,
2006; Jorquera et al., 2008; Keneni et al., 2010; Oliveira et al., 2009) and may
prove especially useful in P-sorbing volcanic soils (Andosols) in the northern
Andes. In Peru, Oswald et al. (2010) demonstrated the P-solubilizing activity
of 44% of Azotobacter and 58% of Bacillus bacterial strains isolated from the
rhizosphere in a single potato field, while 28% of bacterial isolates from
quinoa and one-quarter of rhizobial strains from fava bean in an Altiplano
collecting area had P-solubilizing ability (Ortuño, 2010). Despite the appar-
ently common occurrence of P-solubilizing bacteria in the Andes, their
impact on crop growth is not always clear. While Faccini et al. (2007)
found that P-solubilizing bacteria allowed equivalent potato yields with half
the P fertilizer, other researchers in the northern Andes observed little impact
of P-solubilizing bacteria on crop growth (Nustez and Acevedo, 2005;
Rodriguez et al., 2010). Previous work in the region has largely focused on
solubilizers of inorganic P forms, yet the prevalence of manure use in the
Andes suggests that increased focus on mineralization of organic P by PGPM
might be productive. Interestingly, P mineralizing and solubilizing bacteria
may have utility in both low-P environments and also in intensified farming
environments where fixation of applied fertilizer or manure P occurs (Chabot
et al., 1996).
Plant growth-promoting bacteria with hormonal modes of action pro-
vide glimpses of fascinating mechanisms of plant microbe interaction that
may prove useful in field settings. For example, inoculation with ethylene-
degrading PGPM has been found to increase yield of pea, tomatoes, and
peppers under in-season drought, by removing the ethylene stress signal to
roots and allowing a quicker resumption of growth when soil moisture
conditions improve (Arshad et al., 2008; Belimov et al., 2009; Mayak et al.,
2004). These results suggest that some soil microbes play a mediating role in
plant signaling to respond to stresses (Glick et al., 2007) and are especially
relevant given the risk of in-season drought for smallholders in the Andes.
For the inoculant organisms reviewed here, domains of soil properties
and management where inoculation is warranted need to be better char-
acterized. For the example of AM inoculants, high P availability in soil can
make mycorrhizal inoculation redundant or even inefficient for crops.
150 Steven J. Fonte et al.
microsites that are favorable as bacterial habitat (Thies and Rillig, 2009). In
reviewing impacts of biochar on AM, Warnock et al. (2007) documented
neutral to very positive responses from char application on mycorrhizal
colonization. As possible mechanisms for AM–charcoal relations, they
advanced local alteration of soil physicochemical properties, indirect effects
via impacts on other soil microbes, plant–fungal signaling interference, and
protection from fungal grazers within charcoal. In addition to biochar,
zeolites, nanoporous aluminosilicates of varied chemical composition,
have the potential to increase the cation exchange and water-holding
capacity of soils (Ramesh et al., 2010). The combination of small pore size
and nutrient-holding capacity may make them especially active in micro-
bially mediated nutrient transformations that prolong the availability of P,
N, and other nutrients, especially via sorption of ammonium and other
cations (Flores Macias et al., 2007; Ippolito et al., 2011).
While functioning as a source of carbon and nutrients to support
decomposer food webs and improve soil physiochemical properties, com-
post and vermicomposts can also provide beneficial microbial inoculants
to soils that can competitively suppress soil-borne diseases and have similar
effects to inoculation with other rhizosphere microbes (Cardoza, 2011;
Litterick et al., 2004; Quilty and Cattle, 2011). In addition, composts and
other biomedia have been tested as vehicles to carry inoculant microbes into
the soil (Seneviratne et al., 2011; Singhai et al., 2011). Rock phosphate or
other minerals are sometimes mixed with the compost to foster their
solubilization (Lange, 1994). In effect, compost may be thought of as a
concentrated version of the effect desired in the whole soil for improving
crop access to nutrients or protecting crop roots from disease. Despite some
promising results with compost as inoculant promoters, their potential may
be limited in rural, extensive smallholder farms in the Andes, where labor to
produce compost may be at a premium and composting feedstocks have
competing uses. However, in intensive areas of smallholder production near
cities where waste feedstocks may be more available, composting as an
integrated source of beneficial microbes and available nutrients for crops
may hold greater promise.
The case of composting illustrates the need for inoculant technologies to
be accessible for smallholders in the Andes. Cost and local access must be
considered alongside technical optimization so that inoculant technologies
can achieve maximum impact for smallholders, and thus research on inocu-
lants from the region often includes analyses of economic and environmen-
tal costs and benefits. To avoid what Rosset and Altieri (1997) termed the
threat of import substitution in smallholder sustainable agriculture, researchers
have demonstrated promising results with local, on-farm methods for
inoculating crops with AM using roots of infected crops, often with supe-
rior results to commercial inoculants (Davies et al., 2005a, Douds et al.,
2007, Mäder et al., 2011). This approach is well suited to AM and Rhizobia,
152 Steven J. Fonte et al.
where propagules like spores are necessarily associated with roots and may
be more difficult with associative fungi and bacteria that have no intraradical
structures. Regardless, such field-to-field approaches for inoculation carry
risks for transmission of soil pathogens, which should be taken very seriously
given their potential for drastic impact in the Andes.
While the role of soil microbial communities has been well studied in
the Andes, the impacts of soil fauna in the region remain largely unknown,
despite their significant impact on numerous key soil processes in diverse
ecosystems around the globe. Soil fauna comprise a diverse group of
organisms and alter soil functioning in a variety of ways. Soil invertebrates
generally increase nutrient availability by (1) comminution of plant residues
and accelerating decomposition and mineralization processes and (2) feed-
ing on soil bacteria and fungi, thus releasing of nutrients bound within
microbial tissues (Lavelle and Spain, 2001). The maintenance of healthy soil
food webs (diverse and active soil fauna) has also been shown to control
plant pathogens and may have important implications for crop growth
(Blouin et al., 2005; Sanchez-Moreno and Ferris, 2007). Soil macrofauna,
particularly those representing ecosystem engineers (e.g., earthworms, ants,
termites), can drastically alter soil microbial communities via effects on soil
structure, water movement, nutrient dynamics, and SOM (Lavelle et al.,
1997). These activities can have varied implications for soil functioning but
have generally proven to be beneficial for plant growth (Brown et al., 1999;
Evans et al., 2011). Preliminary findings suggest that macrofauna can be
quite relevant for at least some agroecosystems in the Andes. For example,
Tonneijck and Jongmans (2008) examined biostructures and SOM distri-
bution in páramo soils of Ecuador and concluded that bioturbation by
earthworms was a fundamental process regulating the vertical distribution
and turnover of SOM in these soils. Findings of Morales and Sarmiento
(2002) suggest that soil disturbance associated with cropping reduces soil
macrofauna abundance and diversity in Venezuelan páramo. Although some
reports on soil fauna exist for other regions (Righi and van der Hammen,
1996; Zurita, 1997), information for highland Andean agroecosystems is
notably scarce. Given the global importance of soil fauna and their potential
to offer rapid and inexpensive indicators of soil health (Ferris et al., 2001;
Velasquez et al., 2007), improved understanding of soil faunal activity and
the factors affecting their abundance in agroecosystems remains a critical gap
for sustainable soil fertility management in the Andes.
and climatic stresses, notably low nutrient availability and drought stress,
both of which will reduce crop yields under the likely scenarios of climate
change and soil degradation in the region. Great environmental and social
heterogeneity in the Andes somewhat limits the broad applicability of crop
breeding as an approach to agroecological intensification in the region.
However, we suggest that careful consideration of this diversity and
attempts to understand it can provide important opportunities for improv-
ing crop productivity and yield stability through breeding at both local and
regional scales.
With a few exceptions highlighted below, there are substantial gaps in
breeding and selection of highland Andean crops (native and introduced)
for improved performance under nutrient or drought stress. At the same
time, the range of variation of nutrient acquisition efficiency or drought
resistance already present in existing native varieties is poorly documented.
This lack of focus on stress tolerance for crops in the region may be traced to
activities that compete for breeders’ attention. These include expenditures
in resources and expertise toward assembling and maintaining important
agrobiodiversity collections (e.g., potatoes, quinoa); emphasis on forage
quality, market quality characteristics, and disease or pest resistance that
are more easily selected for at experiment stations (Sciarascia Mugnozza
et al., 1987); and a breeder bias toward wealthier farmers who have greater
access to fertility inputs and irrigation. Nevertheless, substantial interest in
abiotic stress breeding as well as a lack of adequate technical preparation for
local staff to undertake novel breeding approaches was revealed by a 2006
survey of Bolivian plant breeding programs (Gabriel, 2006), thus suggesting
high potential for new targeted breeding programs. Local biodiversity banks
and contacts with the CGIAR center breeding resources offer tremendous
opportunity to characterize existing genotypes for adaptation to low nutri-
ent supply and drought stress, and select new genotypes in stress environ-
ments along gradients of soil fertility and/or degradation. It should also
be noted that the lack of attention to breeding of Andean crops mirrors the
worldwide “orphaning” of these plants in comparison to other crops (e.g.,
maize, soybean, rice).
With the exception of soils of the Altiplano that have been characterized as
C- and N-limited, Andean valleys and particularly Andosols in the northern
Andes suffer from P limitation (Cárdenas et al., 2008; Dahlgren et al., 2004;
Valente and Oliver, 1993). Further, the soil P challenge in the Andes is one of
availability rather than absolute supply, suggesting that improving P foraging
ability of crops is a plausible strategy in this region. While N limitation is also
important for nonlegume crops in Andean agroecosystems (prompting inter-
est in N use efficiency for crops such as potatoes; Errebhi et al., 1999; Zebarth
et al., 2008), we suggest that N deficiencies are more easily addressed through
management (e.g., enhanced incorporation of legumes, see Sections 2.2.1 and
2.2.6) and merit less focus here.
154 Steven J. Fonte et al.
enhance biological control of soil pests (Degenhardt et al., 2009). The positive
agroecological feedbacks from P foraging ability and other abiotic stress traits
in smallholder systems should be recognized. For example, Henry et al.(2009)
demonstrated that far from mining the soil for additional P, bean and soybean
varieties that scavenged unavailable P covered soil better in a smallholder
hillside context, reducing erosion P losses and creating a better P balance than
P-inefficient varieties.
Unpredictable weather extremes, including frost and in-season drought,
have always threatened crop productivity and food security in the Andes.
However, increases in average temperature at high elevations due to climate
change are predicted to exceed the global average (Bradley et al., 2006), so
that evapotranspiration may increasingly outstrip the water-supplying ability
of soil and both in-season and terminal drought stress may increase. There is
already evidence that farmers perceive increased drought risk and are adapting
their crop mixtures to climate change, for example, by increasing the propor-
tion of early-maturing barley relative to other crops on the Bolivian Altiplano
to avoid crop failure (A. Bonifacio, personal communication). Crop breeding
for drought resistance and yield stability to confront climate change is already
a worldwide effort, and much can be gained from linking Andean crop
breeders to these global efforts (Khan et al., 2010). It should also be noted
that plant traits targeted for agroecological intensification and climate change
are also likely to be appropriate for possible alterations in the soils of Andean
agroecosystems (discussed in Section 2.2.1), as most soil consequences of
climate change are likely to occur through alterations to plant cover.
Mechanisms of drought resistance in crops encompass drought escape
or earliness, dehydration avoidance (conservation of water in the plant, or
ways to access more water in the soil profile), and drought tolerance (Khan
et al., 2010). Drought resistance includes plant physiological mechanisms
that improve yields under drought stress, such as deep rooting and osmotic
adjustment of plant cells that allow photosynthesis and other metabolic
processes to support plant productivity even under drought-stressed condi-
tions (Blum, 2005; Chaves et al., 2003). Breeders and crop physiologists have
elucidated these and other traits conferring drought resistance at a physiologi-
cal level (Cattivelli et al., 2008), and are beginning to manipulate these traits
with marker-assisted breeding and other molecular approaches (Araus et al.,
2008; Miklas et al., 2006). Breeding methods for physiological drought
resistance apply across crop species and could be useful for breeding efforts
in the Andean region. Nevertheless, it should be recognized that drought
tolerance mechanisms like osmotic adjustment have had limited success in
explaining yield stability under drought, in comparison to avoidance of
drought stress via plant size, earliness, or root architecture for water capture.
A number of studies in the region and elsewhere have examined differ-
ences in drought resistance among genotypes of Andean crops. For legumes,
Siddique et al. (2001) tested a wide array of Mediterranean-type legumes
156 Steven J. Fonte et al.
and concluded that earliness of flowering and pod formation (drought escape)
was the dominant trait increasing yield under low rainfall. Studies on fava
bean suggest that drought resistance and better recovery from drought are
more due to stomatal closure and water conservation within the plant, rather
than osmotic adjustment (Amede et al., 1999; Khan et al., 2010; Sau and
Minguez, 2000). Analogous to root morphology traits used in screening for P
foraging ability, Khan et al. (2007) suggested a number of physiological
parameters of fava bean that are promising as physiological screening criteria
for drought resistance. Drought physiology has not been greatly studied in L.
mutabilis (see Carvalho et al., 2004), but Australian research on a number of
old-world lupines showed that yield under drought stress was related to
earliness and the speed of pod and bean filling rather than osmotic adjustment
of cells (French and Buirchell, 2005; Palta et al., 2007). Others have argued
that maximal water capture in the soil profile (e.g., deep rootedness observed
for L. mutabilis) and effective use of that water in plant productivity are more
important under drought stress than a sole focus on photosynthesis per unit of
water (Blum, 2009). Meanwhile, Tourneux et al. (2003a,b) examined six
potato varieties under drought stress at tuberization in Bolivia and suggested
that yield differences were not correlated with physiological parameters, but
rather were due to reduced leaf area as well as the inability to translocate
carbohydrates to tubers in sensitive varieties.
Regardless of physiological mechanisms of drought tolerance or avoid-
ance, approaches to drought stress breeding for rainfed farming systems such
as those in the Andes have recently emphasized the importance of selecting
environments for breeding. Selection of varieties in environments with a
particular drought stress, including the timing of the stress in the life cycle of
the plant, is important to success in breeding for drought tolerance (Ceccarelli
et al., 1991). For example, potatoes are highly sensitive to drought at tuber-
ization (DallaCosta et al., 1997); breeding efforts for drought tolerance in
maize focus on flowering and grain-filling stages (Banziger et al., 2006); and
fava beans are most sensitive to drought at the pod set and early pod-filling
phase (Khan et al., 2010). Ceccarelli et al. (1991) argued that the unpredictable
stresses found in many smallholder farm environments mean that a suite of
drought resistance traits must be selected for as a whole, a critique that has
fostered an evolutionary strategy of breeding with farmer participation that is
of interest to breeding efforts in the Andes (Ceccarelli et al., 2010). Never-
theless, it still makes sense to target the most vulnerable phases of crop growth
for a given stress as a first priority in stress breeding, to foster mechanistic
understanding of stress resistance and also since varieties that resist drought
stress during these vulnerable phases will likely fare better in the average year
when water becomes limiting.
Breeders working to select for stress tolerance in multi-environment trials
have developed expertise in the efficient design and sophisticated interpre-
tation of these trials, which often present challenges in teasing apart
Andean Soil Fertility 157
A promising area for improved spatial integration at the plot scale is the
implementation of polycultures, or intercropping. Combining multiple species
within the same field may complicate management in some cases but can
increase overall yield through increased species complementarity of resource
use (Fornara and Tilman, 2008). For example, Li et al. (2007) found a maize–
fava bean intercropping systems in China to overyield (produce more than
either crop would in monoculture) due to differing rooting depth and timing
of nutrient uptake (reduced competition) as well as improved availability of P
resulting from organic acid production by fava bean roots (facilitation). In this
same cropping system, increased competition for available N by maize forced
fava beans to fix more N from atmospheric sources, thus augmenting overall N
inputs to the system via N-fixation (Li et al., 2009). Just as for crop rotation,
increased spatial diversification in Andean agriculture can also help alleviate
pest issues, but through increases in predation/parasitism of pests and decreased
density of resources (Gianoli et al., 2006; Poveda et al., 2008). Although
offering clear potential for cropping situations, such benefits of intercropping
have been demonstrated in multiple ecosystem types and seem to be especially
well suited for grassland systems (Fornara and Tilman, 2008; Oberson et al.,
1999). Such practices could perhaps be most easily implemented during non-
cropping stages of sectoral fallow systems, as management requirements are
minimal and risks to farmers relatively low. However, there is also great
potential for more intensified crop–pasture rotations. The idea of improved
fallows and pastures (where species composition is intentionally managed) for
the Andes has been suggested previously (Barrios et al., 2005; Garcı́a, 2011;
Sarmiento et al., 2001), but principally with a focus on leguminous plants.
Others have suggested mixtures of legumes and grasses as an option to maxi-
mize N-fixation and productivity for fodder production (Bartl et al., 2009;
Bentley et al., 2007). Such mixtures could incorporate exotic plants that are
well adapted to local climates (e.g., cold tolerant, able to set seed before arrival
of the dry season; Wheeler et al., 1999), as well as native species that could help
reduce farmer investment and ensure survival (Bartl et al., 2009; Nezomba
et al., 2010). Further, improved fallows (with optimal varieties and species
mixtures) combined with strategic fertilization (particularly P additions) could
greatly enhance N-fixation, fodder production, and SOM stabilization, ulti-
mately accelerating the recovery of soil functioning and nutrient reserves.
Another means of enhancing spatial integration at the plot and farm level is
through agroforestry practices. Although potential for tree growth in some
parts of the Andes (i.e., very dry regions and/or high elevations) may be low,
evidence suggests that for large parts of the Andean highlands trees may have
been much more common than at present (Chepstow-Lusty and Winfield,
2000; Hensen, 2002), and that potential for agroforestry technologies may be
higher than often acknowledged (Cotler and Maass, 1999; Mahboubi et al.,
1997; Reynel and Felipe-Morales, 1990). For example, Mahboubi et al.
(1997) compared tree survival of candidate species in the Bolivian Altiplano
160 Steven J. Fonte et al.
and found several species (e.g., the native Buddleja coriacea) to tolerate condi-
tions well at elevations as high as 4200m, with a greater number of tree species
suitable for lower elevations. They also suggested that residues obtained from
these trees (all except Eucalyptus sp.) could provide valuable nutrients to
enhance crop growth. Comparisons of contour row species near Cocha-
bamba, Bolivia, suggest that woody species can grow well and also provide
effective erosion control (Sims et al., 1999). Additionally, the incorporation of
legumes (particularly woody species) into grass contour barriers has been
suggested as an effective means of contributing to soil fertility and rapid erosion
control, while reducing plant resource competition (Mutegi et al., 2008; Sims
and Rodriguez, 2001). Along with contour hedgerows, alternative options for
integrating woody species into agroecosystems include fodder banks, wood
lots, cut and carry systems, and/or windbreaks at field margins (Reynel and
Felipe-Morales, 1990; Young, 1997), as well as trees intermixed in the field, a
highly successful strategy for lower elevation hillsides used in the Quesungual
agroforestry system of Honduras (Fonte et al., 2010; Hellin et al., 1999).
Ultimately, the incorporation of trees into the landscape offers great potential
to provide farmers with supplementary fuel, fodder, and soil fertility amend-
ments (Barrios et al., 2005), while contributing key ecosystem services (e.g.,
erosion control, climate change mitigation, improved water dynamics) and
enhancing regional biodiversity (Antle et al., 2007; Otero and Onaindiam,
2009; Smukler et al., 2010). It should be noted, however, that proposed
agroforestry technologies must seek to minimize the potential negative
impacts of trees, such as increased competition with crops for light, water,
and nutrients; greater labor requirement; and promotion of crop pests (e.g.,
birds). Based on the relative paucity of materials encountered in our literature
search, research to date has yet to fully explore the full potential of tree-based
options in the Andes. Although the flexibility of agroforestry practices in the
Andean highlands is perhaps more limited than their lowland counterparts,
large potential exists and future research needs to address these possibilities.
In addressing spatial organization of farms, soil heterogeneity frequently
represents a major challenge for soil fertility management. Andean agroeco-
systems often have extreme and diverse topography which can generate steep
and somewhat predictable fertility gradients over the landscape (Buytaert
et al., 2007). However, fertility gradients are also generated by anthropogenic
factors at the farm scale, with improved nutrient status typically found closer
to the household (Tittonell et al., 2005; Vanek, 2010). In farming systems of
the east African highlands, Vanlauwe et al. (2006) suggested that fields further
from the farmer homes (outfields) should receive greater nutrients invest-
ment, as they are likely show a greater response to fertilizer additions.
However, in more mountainous regions such as the Andes, outfields may
represent fields with steeper slopes and/or rockier sites, where soils do not
have the same yield potential as flatter infield sites and increased nutrient
applications might be at greater risk to loss, via erosion (Vanek, 2010). In
either situation, there may be distinct soil fertility management strategies that
Andean Soil Fertility 161
reflect differential labor inputs associated with far versus near farm plots. For
example, in the Andes it may be better to apply manure and compost
generated near the home to infield plots, so as to minimize transport labor.
At the same time more distant plots, especially those prone to erosion, are
perhaps better candidates for cover crops and/or improved fallows combined
with strategic fertilizer applications in order to generate vegetative cover and
organic matter inputs on-site, and thus avoid transport. Such fertility gradients
are equally important at smaller scales as well and perhaps more manageable.
For example, Dercon et al. (2006) demonstrated that erosion within contour
hedgerows in Ecuador generated gradients of SOM, P, texture, such that
fertility was lowest on the upper portion of slopes and highest at the bottom,
where live barriers slowed runoff leading to soil deposition. Crop yield within
terraces varied significantly, with up to fourfold differences for wheat growing
at the top versus the bottom of a single terrace. They conclude that erosion
control is critical in slowing the formation of such gradients. However, it may
also be necessary to improve organic matter management (via manure or plant
residue additions) on the upslope portions of the terrace to restore SOM and
the capacity of these soils to provide and retain nutrients.
Organization at larger scales is also important for optimizing the
efficiency of farm management and ensuring a range of ecosystem func-
tions at the farm and landscape level. Although climate, topography, and
soil type largely determine where different crop and livestock elements are
placed, there is likely to be some room to better organize farms and
landscapes for increasing farm efficiency and the production of ecosystem
services and reducing labor. For example, active landscape management in
the Andes has been observed whereby farmers strategically disperse their
crop fields as means of distributing risks and improving overall yield
stability (Goland, 1993). At the same time, the physical arrangement of
crops and other farm attributes can greatly influence less obvious agroe-
cosystem functions in a variety of ways. The maintenance of biodiversity
and pollination services probably offers the clearest case for the role of
landscape organization in the provision of agroecosystem services. Polli-
nators are critical for the production of numerous crops around the globe
(Klein et al., 2007), including a number of important Andean crops (e.g.,
onions, fava beans, Andean lupine), and are greatly influenced by land-
scape structure and composition (Ricketts et al., 2008). Pest control agents
are also impacted by the landscape (Bianchi et al., 2006), but often in
counterintuitive ways (Parsa et al., 2011). Perhaps more relevant for soil
fertility, landscape diversity can influence soil fauna biodiversity and
distribution. For example, patches of native vegetation may function as a
reservoir or refuge of earthworms for adjacent agricultural plots, thus
maintaining their diversity, abundance, and activity on farmed plots
(Marichal, 2011). The composition of management attributes at the farm
level can also impact key ecosystem services such as C storage, N loss, and
erosion control (Robertson and Swinton, 2005; Smukler et al., 2010).
162 Steven J. Fonte et al.
ACKNOWLEDGMENTS
We thank a number of people who provided assistance and advice at various stages in the
development of this chapter including Sady Garcia, Carlos Meneses, Maria Scurrah, Julio
Alegre, Ruddy Meneses, Karl Zimmerer, Noel Ortuño, Marcela Quintero, Edward
Guevara, Oscar Ortiz, Javier Aguilera, Alejandro Bonifacio, Claire Nicklin, Carlos Perez,
and many others. This work was made possible by support from the McKnight Foundation
Collaborative Crop Research Program.
REFERENCES
Abadı́n, J., González-Prieto, S. J., Sarmiento, L., Villar, M. C., and Carballas, T. (2002).
Successional dynamics of soil characteristics in a long fallow agricultural system of the
high tropical Andes. Soil Biol. Biochem. 34, 1739–1748.
Andean Soil Fertility 167
Abreu, Z., Llambı́, L. D., and Sarmiento, L. (2009). Sensitivity of soil restoration indicators
during Páramo succession in the high tropical Andes: Chronosequence and permanent
plot approaches. Restor. Ecol. 17, 619–628.
Aguilera, J. (2010). Impacts of Soil Management Practices on Soil Fertility in Potato-Based
Cropping Systems in the Bolivian Andean Highlands. University of Missouri, Columbia.
Alegre, J. C., Felipe-Morales, C., and LaTorre, B. (1990). Soil erosion studied in Peru. J. Soil
Water Conserv. 45, 417–420.
Altieri, M. (2004). Linking ecologists and traditional farmers in the search for sustainable
agriculture. Front. Ecol. Environ. 2, 35–42.
Alyokhin, A., Porter, G., Groden, E., and Drummond, F. (2005). Colorado potato beetle
response to soil amendments: A case in support of the mineral balance hypothesis? Agric.
Ecosyst. Environ. 109, 234–244.
Amede, T., Kittlitz, E. V., and Schubert, S. (1999). Differential drought responses of faba
bean (Vicia faba L.) inbred lines. J. Agron. Crop Sci. 183, 35–45.
Antle, J. M., Stoorvogel, J. J., and Valdivia, R. O. (2007). Assessing the economic impacts of
agricultural carbon sequestration: Terraces and agroforestry in the Peruvian Andes. Agric.
Ecosyst. Environ. 122, 435–445.
Ao, J., Fu, J., Tian, J., Yan, X., and Liao, H. (2010). Genetic variability for root morph-
architecture traits and root growth dynamics as related to phosphorus efficiency in
soybean. Funct. Plant Biol. 37, 304–312.
Arandia, W., Ortuño, N., Gutierrez, E., and Cáceres, A. (2007). Evaluación en invernadero
de la respuesta del cultivo de cebolla (Allium cepa) de la variedad rosada criolla al uso
combinado con aciduantes y micorrizas (m.a.). Facultad de Agronomı́a. Universidad
Mayor San Simon, Cochabamba, Bolivia.
Araus, J. L., Slafer, G. A., Royo, C., and Serret, M. D. (2008). Breeding for yield potential
and stress adaptation in cereals. Crit. Rev. Plant Sci. 27, 377–412.
Aroca, R., and Ruiz-Lozano, J. M. (2009). Induction of plant tolerance to semi-arid environ-
ments by beneficial soil microorganisms—A review. Sustain. Agric. Rev. 2, 121–135.
Arshad, M., Shaharoona, B., and Mahmood, T. (2008). Inoculation with Pseudomonas spp.
containing ACC-deaminase partially eliminates the effects of drought stress on growth,
yield, and ripening of pea (Pisum sativum L.). Pedosphere 18, 611–620.
Ashby, J., Beltran, A. J., Guerrero, M., and Ramos, H. F. (1997). Improving the acceptabil-
ity to farmer of soil conservation practices. J. Soil Water Conserv. 51, 309–312.
Bale, J. S., Masters, G. J., Hodkinson, I. D., Awmack, C., Bezemer, T. M., Brown, V. K.,
Butterfield, J., Buse, A., Coulson, J. C., Farrar, J., Good, J. E. G., Harrington, R., et al.
(2002). Herbivory in global climate change research: Direct effects of rising temperature
on insect herbivores. Glob. Change Biol. 8, 1–16.
Banziger, M., Setimela, P., Hodson, D., and Vivek, B. (2006). Breeding for improved
abiotic stress tolerance in maize adapted to southern Africa. Agric. Water Manage. 80,
212–224.
Barba, R., Coca, G., Tuma, N., and Pijnenborg, J. (2000). Pre-selección de cepas de (Brady)
rhizobium para leguminosas de grano en Bolivia. In “Compendio de trabajos presentados
por el proyecto rizobiologı́a (Cochabamba) en eventos y publicaciones de otras institu-
ciones” (R. Meneses, Ed.), pp. 47–49. Proyecto Rizobiologı́a Bolivia (CIAT-CIF-
PNLG-CIFP-UAW/DHV), Cochabamba, Bolivia.
Barrios, E., and Trejo, M. T. (2003). Implications of local soil knowledge for integrated soil
management in Latin America. Geoderma 111, 217–231.
Barrios, E., Cobo, J. G., Rao, I. M., Thomas, R. J., Amezquita, E., Jimenez, J. J., and
Rondon, M. A. (2005). Fallow management for soil fertility recovery in tropical Andean
agroecosystems in Colombia. Agric. Ecosyst. Environ. 110, 29–42.
Barroso, C. B., and Nahas, E. (2005). The status of soil phosphate fractions and the ability of
fungi to dissolve hardly soluble phosphates. Appl. Soil Ecol. 29, 73–83.
168 Steven J. Fonte et al.
Bartl, K., Gamarra, J., Gómez, C. A., Wettstein, H. R., Kreuzer, M., and Hess, H. D.
(2009). Agronomic performance and nutritive value of common and alternative grass and
legume species in the Peruvian highlands. Grass Forage Sci. 64, 109–121.
Bashan, Y., and Bashan, L. E.d. (2010). Chapter two—How the plant growth-promoting
bacterium Azospirillum promotes plant growth—A critical assessment. Adv. Agron. 108,
77–136.
Bayard, B., Curtis, M. J., and Dennis, S. (2006). The economics of adoption and manage-
ment of alley cropping in Haiti. J. Environ. Manage. 84, 62–70.
Bebbington, A. J., Carrasco, H., Peralbo, L., Ramon, G., Trujillo, J., and Torres, V. (1993).
Fragile lands, fragile organizations: Indian organizations and the politics of sustainability
in Ecuador. Trans. Inst. Br. Geogr. 18, 179–196.
Beebe, S., Lynch, J., Galwey, N., Tohme, J., and Ochoa, I. (1997). A geographical approach
to identify phosphorus-efficient genotypes among landraces and wild ancestors of com-
mon bean. Euphytica 95, 325–336.
Beebe, S. E., Rojas-Pierce, M., Yan, X., Blair, M. W., Pedraza, F., Munoz, F., Tohme, J.,
and Lynch, J. P. (2006). Quantitative trait loci for root architecture traits correlated with
phosphorus acquisition in common bean. Crop. Sci. 46, 413–423.
Beebe, S., Rao, I. M., Blair, M. W., and Butare, L. (2009). Breeding for abiotic stress tolerance
in common bean: Present and future challenges. SABRAO J. Breed. Genet. 41, 1–11.
Belimov, A. A., Dodd, I. C., Hontzeas, N., Theobald, J. C., Safronova, V. I., and Davies, W. J.
(2009). Rhizosphere bacteria containing 1-aminocyclopropane-1-carboxylate deaminase
increase yield of plants grown in drying soil via both local and systemic hormone signalling.
New Phytol. 181, 413–423.
Bell, L. W., Wade, L. J., and Ewing, M. A. (2010). Perennial wheat: A review of environmental
and agronomic prospects for development in Australia. Crop Pasture Sci. 61, 679–690.
Benizri, E., Baudoin, E., and Guckert, A. (2001). Root colonization by inoculated plant
growth-promoting rhizobacteria. Biocontrol Sci. Techn. 11, 557–574.
Bennet, A. E., and Bever, J. D. (2007). Mycorrhizal species differentially alter plant growth
and response to herbivory. Ecology 88, 210–218.
Bentley, J. W. (1989). What farmers don’t know can’t help them: The strengths and weak-
nesses of indigenous technical knowledge in Honduras. Agric. Human Values 6, 25–31.
Bentley, J., Velasco, C., Rodrı́guez, F., Oros, R., Botello, R., Webb, M., Devaux, A., and
Thiele, G. (2007). Unspoken demands for farm technology. Int. J. Agric. Sustain. 5, 70–84.
Bianchi, F. J. J. A., Booij, C. J. H., and Tscharntke, T. (2006). Sustainable pest regulation in
agricultural landscapes: A review on landscape composition, biodiversity and natural pest
control. Proc. R. Soc. B. Biol. Sci. B 273, 1715–1727.
Blouin, M., Zuily-Fodil, Y., Pham-Thi, A. T., Laffray, D., Reversat, G., Pando, A.,
Tondoh, J., and Lavelle, P. (2005). Belowground organism activities affect plant above-
ground phenotype, inducing plant tolerance to parasites. Ecol. Lett. 8, 202–208.
Blum, A. (2005). Drought resistance, water-use efficiency, and yield potential—Are they
compatible, dissonant, or mutually exclusive? Aust. J. Agr. Res. 56, 1159–1168.
Blum, A. (2009). Effective use of water (EUW) and not water-use efficiency (WUE) is the
target of crop yield improvement under drought stress. Field Crop. Res. 112, 119–123.
Boomsma, C. R., and Vyn, T. J. (2008). Maize drought tolerance: Potential improvements
through arbuscular mycorrhizal symbiosis? Field Crops Res. 108, 14–31.
Borie, B. F., Rubio, R., Morales, A., Curaqueo, G., and Cornejo, P. (2010). Arbuscular
mycorrhizae in agricultural and forest ecosystems in Chile. J. Soil Sci. Plant Nut. 10,
185–206.
Bossio, D., and Cassman, K. (1991). Traditional rainfed barley production in the Andean
highlands of ecuador: Soil nutrient limitations and other constraints. Mt. Res. Dev. 11,
115–126.
Bottner, P., Pansu, M., Sarmiento, L., Dominique Hervé, D., Ruben Callisaya-Bautista, R.,
and Metselaar, K. (2006). Factors controlling decomposition of soil organic matter in
Andean Soil Fertility 169
fallow systems of the high tropical Andes: A field simulation approach using 14C- and
15N-labelled plant material. Soil Biol. Biochem. 38, 2162–2177.
Bradley, R. S., Vuille, M., Diaz, H. F., and Vergara, W. (2006). Threats to water supply in
the Tropical Andes. Science 312, 1755–1756.
Briar, S. S., Fonte, S. J., Park, I., Six, J., Scow, K. M., and Ferris, H. (2011). The distribution
of nematodes and soil microbial communities across soil aggregate fractions and farm
management systems. Soil Biol. Biochem. 43, 905–914.
Briones, M. J. I., Ostle, N. J., McNamara, N. P., and Poskitt, J. (2009). Funtional shift
of grassland communities in response to soil warming. Soil Biol. Biochem. 41, 315–322.
Bronick, C. J., and Lal, R. (2005). Soil structure and management: A review. Geoderma 124,
3–22.
Brown, G. G., Pashanasi, B., Villenave, C., Patrón, J. C., Senapati, B. K., Giri, S., Barios, I.,
Lavelle, P., Blanchart, E., Blakemore, R. J., Spain, A. V., and Boyer, J. (1999). Effects of
earthworms on plant production in the tropics. In “Earthworm Management in Tropical
Agroecosystems” (P. Lavelle, L. Brussaard, and P. Hendrix, Eds.), pp. 87–147. CAB
International, Oxon.
Brush, S., Kesseli, R., Ortega, R., Cisneros, P., Zimmerer, K., and Quiros, C. (1995). Potato
diversity in the Andean center of crop domestication. Conserv. Biol. 19, 1189–1198.
Brussaard, L., Pulleman, M. M., Ouedraogo, E., Mando, A., and Six, J. (2006). Soil fauna
and soil function in the fabric of the food web. Pedobiologia 50, 447–462.
Bucheyeki, T. L., Shenkalwa, E. M., Mapunda, T. X., and Matata, L. W. (2008). On-farm
evaluation of promising groundnut varieties for adaptation and adoption in Tanzania. Afr.
J. Agric. Res. 3, 531–536.
Bunch, R. (2004). Adopción de abonos verdes y cultivos de cobertura. LEISA Revista de
Agroecologı´a 19(4), 11–13.
Buytaert, W., Deckers, J., Dercon, G., De BieÁvre, B., Poesen, J., and Govers, G. (2002).
Impact of land use changes on the hydrological properties of volcanic ash soils in South
Ecuador. Soil Use Manage. 18, 94–100.
Buytaert, W., Rolando Célleri, R., De Bièvre, B., Cisneros, F., Wyseure, G., Deckers, J.,
and Hofstede, R. (2006). Human impact on the hydrology of the Andean páramos. Earth
Sci. Rev. 79, 53–72.
Buytaert, W., Deckers, J., and Wyseure, G. (2007). Regional variability of volcanic ash soils in
south Ecuador: The relation with parent material, climate and land use. Catena 70, 143–154.
Buytaert, W., Cuesta-Camacho, F., and Tobón, C. (2010a). Potential impacts of climate
change on the environmental services of humid tropical alpine regions. Glob. Ecol.
Biogeogr. 20, 19–33.
Buytaert, W., Vuille, M., Dewulf, A., Urrutia, R., Karmalkar, A., and Célleri, R. (2010b).
Uncertainties in climate change projections and regional downscaling in the tropical Andes:
Implications for water resources management. Hydrol. Earth Syst. Sci. 14, 1247–1258.
Capowiez, Y., Cadoux, S., Bouchant, P., Ruy, S., Roger-Estrade, J., Richard, G., and
Boizard, H. (2009). The effect of tillage type and cropping system on earthworm
communities, macroporosity and water infiltration. Soil Till. Res. 105, 209–216.
Cárdenas, J., Choque, W., and Guzman, R. (2008). Fertilidad, uso, y manejo de suelos en la
zona del intersalar, departamentos de: Oruro y Potosı́. Fundación AUTAPO, programa
Quinoa sur, Oruro, Bolivia.
Cardoza, Y. J. (2011). Arabidopsis thaliana resistance to insects mediated by an earthworm-
produced organic soil amendment. Pest Manag. Sci. 67, 233–238.
Carney, K. M., and Matson, P. A. (2005). Plant communities, soil microorganisms, and soil
carbon cycling: Does altering the world belowground matter to ecosystem functioning?
Ecosystems 8, 928–940.
Carvalho, I. S., Ricardo, C. P., and Chaves, M. (2004). Quality and distribution of
assimilates within the whole plant of lupines (L-albus and L-mutabilis) influenced by
water stress. J. Agron. Crop Sci. 190, 205–210.
170 Steven J. Fonte et al.
Conklin, A. E., Erich, M. S., Liebman, M., Lambert, D., Gallandt, E. R., and
Halteman, W. A. (2002). Effects of red clover (Trifolium pratense) green manure and
compost soil amendments on wild mustard (Brassica kaber) growth and incidence of
disease. Plant Soil 238, 245–256.
Coppus, R., Imeson, A. C., and Sevink, J. (2003). Identification, distribution and characteristics
of erosion sensitive areas in three different Central Andean ecosystems. Catena 51, 315–328.
Córdoba, J. J., and Novoa, V. (1997). Problematica, experiencias y enfoque sobre la erosión,
manejo y conservación de suelos de ladera en Ecuador. In “Manejo Integral de Micro-
cuencas” (M. Tapia, Ed.), pp. 123–133. International Potato Center (CIP), Lima.
Cotler, H., and Maass, J. M. (1999). Tree management in the northwestern Andean
Cordillera of Peru. Mt. Res. Dev. 19, 153–160.
Coûteaux, M., Hervé, D., and Mita, V. (2008). Carbon and nitrogen dynamics of potato
residues and sheep dung in a two-year rotation cultivation in the Bolivian altiplano.
Commun. Soil Sci. Plant 39, 475–498.
Cox, T. S., Bender, M., Picone, C., Van Tassel, D. L., Holland, J. B., Brummer, E. C.,
Zoeller, B. E., Paterson, A. H., and Jackson, W. (2002). Breeding perennial grain crops.
Crit. Rev. Plant Sci. 21, 59–91.
Craswell, E. T., and Lefroy, R. D. B. (2001). The role and function of organic matter in
tropical soils. Nutr. Cycl. Agroecosyst. 61, 7–18.
Craswell, E. T., Sajjapongse, A., Howlett, D. B. J., and Dowling, A. J. (1998). Agroforestry
in the management of sloping lands in Asia and the Pacific. Agroforest. Syst. 38, 121–137.
Crookston, R. K., Kurle, J. E., Copeland, P. J., Ford, J. H., and Lueschen, W. E. (1991).
Rotational cropping sequence affects yield of corn and soybean. Agron. J. 83, 108–113.
Dahlgren, R. A., Saigusa, M., and Ugolini, F. C. (2004). The nature, property, and
management of volcanic soils. Adv. Agron. 82, 113–182.
DallaCosta, L., DelleVedove, G., Gianquinto, G., Giovanardi, R., and Peressotti, A. (1997).
Yield, water use efficiency and nitrogen uptake in potato: Influence of drought stress.
Potato Res. 40, 19–34.
Davidson, E. A., and Janssens, I. A. (2006). Temperature sensitivity of soil carbon decom-
position and feedbacks to climate change. Nature 440, 165–173.
Davies, F. T. Jr., Calderon, C. M., and Huaman, Z. (2005a). Influence of arbuscular
mycorrhizae indigenous to Peru and a flavonoid on growth, yield, and leaf elemental
concentration of ‘Yungay’ potatoes. HortScience 40, 381–385.
Davies, F. Jr., Calderon, C., Huaman, Z., and Gomez, R. (2005b). Influence of a flavonoid
(formononetin) on mycorrhizal activity and potato crop productivity in the highlands of
Peru. Sci. Hortic. (Amsterdam) 106, 318–329.
de Koning, G. H. J., Van de Kop, P. J., and Fresco, L. (1997). Estimates of sub-national
nutrient balances as sustainability indicators for agro-ecosystems in Ecuador. Agric.
Ecosyst. Environ. 65, 127–139.
de la Torres, C., and Burga, M. (1986). Andenes y camellones en el Perú andino: Historia,
presente y futuro. Consejo Nacional de Cencia y Tecnologı́a, Lima.
Degenhardt, J., Hiltpold, I., Koellner, T. G., Frey, M., Gierl, A., Gershenzon, J.,
Hibbard, B. E., Ellersieck, M. R., and Turlings, T. C. J. (2009). Restoring a maize
root signal that attracts insect-killing nematodes to control a major pest. Proc. Natl. Acad.
Sci. USA 106, 13213–13218.
Dehn, M. (1995). An evaluation of soil conservation techniques in the Ecuadorian Andes.
Mt. Res. Dev. 15, 175–182.
Dercon, G., Deckers, J., Poesen, J., Govers, G., Sánchez, H., Ramı́rez, M., Vanegas, R.,
Tacuri, E., and Loaiza, G. (2006). Spatial variability in crop response under contour
hedgerow systems in the Andes region of Ecuador. Soil Till. Res. 86, 15–26.
Deugd, M., Roling, N., and Smaling, E. M. A. (1998). A new praxeology for integrated
nutrient management, facilitating innovation with and by farmers. Agric. Ecosyst. Environ.
71, 269–283.
172 Steven J. Fonte et al.
Devaux, A., Vallejos, J., Hijmans, R., and Ramos, J. (1997). Respuesta agronómica de dos
variedades de papa (spp. tuberosum y andigena) a diferentes niveles de fertilización mineral.
Revista Latinoamericana de la Papa 9(10), 123–139.
Dick, R. P., Sandor, J. A., and Eash, N. S. (1994). Soil enzyme activities after 1500 years of
terrace agriculture in the Colca Valley of Peru. Agric. Ecosyst. Environ. 50, 123–131.
Douds, D. D., Nagahashi, G., Reider, C., and Hepperly, P. R. (2007). Inoculation with
arbuscular mycorrhizal fungi increases the yield of potatoes in a high P soil. Biol. Agric.
Hort. 25, 67–78.
Drinkwater, L. E., and Snapp, S. S. (2007). Nutrients in agroecosystems: Rethinking the
management paradigm. Adv. Agron. 92, 163–186.
Erenstein, O. (2003). Smallholder conservation farming in the tropics and sub-tropics:
A guide to the development and dissemination of mulching with crop residues and
cover crops. Agric. Ecosyst. Environ. 100, 17–37.
Errebhi, M., Rosen, C. J., Lauer, F. L., Martin, M. W., and Bamberg, J. B. (1999).
Evaluation of tuber-bearing Solanum species for nitrogen use efficiency and biomass
partitioning. Am. J. Potato Res. 76, 143–151.
Espinosa, J. (1991). Efecto residual de fosforo en andisoles. Revista Facultad de Agronomia
(Maracay) 17, 39–47.
Evans, J., Gregory, A., Dobrowolski, N., Morris, S. G., O’Connor, G. E., and Wallace, C.
(1996). Nodulation of field-grown Pisum sativum and Vicia faba: Competitiveness of
inoculant strains of Rhizobium leguminosarum bv. viciae determined by an indirect, com-
petitive ELISA method. Soil Biol. Biochem. 28, 247–255.
Evans, T. A., Dawes, T. Z., Ward, P. R., and Lo, N. (2011). Ants and termites increase crop
yield in a dry climate. Nat. Commun. 2, 1–7.
Faccini, G., Garzon, S., Martı́nez, M., and Varela, A. (2007). Evaluation of the effect of a dual
inoculum of phosphate-solubilizing bacteria and Azotobacter chroococcum, in crops of creole
potato (“papa criolla”) “yema de huevo” variety (Solanum phureja). In “First International
Meeting on Microbial Phosphate Solubilization” (E. V.a.C. Rodrı́guez-Barrueco, Ed.),
pp. 301–308. Dordrecht, The Netherlands.
Felipe-Morales, C. (2002). Manejo agroecologico del suelo en sistemas andinos. Agroeco-
logia. Ediciones Cientificas Sudamericanas, Buenos Aires.
Fernandes, E. C. M., Motavalli, P. P., Castilla, C., and Mukurumbira, L. (1997). Management
control of soil organic matter dynamics in tropical land-use systems. Geoderma 79, 49–67.
Ferris, H., Bongers, T., and de Goede, R. G. M. (2001). A framework for soil food
web diagnostics: Extension of the nematode faunal analysis concept. Appl. Soil Ecol. 18,
13–29.
Ferrufino, A., and Sanchez, M. (2006). Efecto de la inoculación con micorrizas en el
crecimiento de plántulas de tembe para palmito (Bactris gasipaes Kunth) en fase de vivero.
Rev. Agric. 37, 1–10.
Flores Macias, A., Galvis Spinola, A., Hernandez Mendoza, T. M., De Leon Gonzalez, F.,
and Payan Zelaya, F. (2007). Effect of zeolite (clinoptilolite and mordenite) amended
andosols on soil chemical environment and growth of oat. Interciencia 32, 692–696.
Fonte, S. J., Barrios, E., and Six, J. (2010). Earthworms, soil fertility and aggregate-
associated soil organic matter dynamics in the Quesungual agroforestry system.
Geoderma 155, 320–328.
Fornara, D. A., and Tilman, D. (2008). Plant functional composition influences rates of soil
carbon and nitrogen accumulation. J. Ecol. 96, 314–322.
Freiberg, C., Fellay, R., Balroch, A., Broughton, W. J., Rosenthal, A., and Peret, X.
(1997). Molecular basis of symbiosis between Rhizobium and legumes. Nature 387,
394–401.
French, R. J., and Buirchell, B. J. (2005). Lupin: The largest grain legume crop in Western
Australia, its adaptation and improvement through plant breeding. Aust. J. Agr. Res. 56,
1169–1180.
Andean Soil Fertility 173
Fuentes, M., Govaerts, B., De Leon, F., Hidalgo, C., Dendooven, L., Sayre, K. D., and
Etchevers, J. (2009). Fourteen years of applying zero and conventional tillage, crop
rotation and residue management systems and its effect on physical and chemical soil
quality. Eur. J. Agron. 30, 228–237.
Gabriel, J. (2006). Plant Breeding and Biotechnology Capacity Survey: BOLIVIA. Funda-
ción Proinpa, Cochabamba.
Garcı́a, S. (2011). Evaluating the Biophysical Resource Management Strategies of the Agro-
ecosystems in Farm Communities of the Mantaro Valley, Central Andes of Peru. p. 276.
Bioscience Engineering. Katholieke Universiteit Leuven, Belgium.
Garcı́a, S., Rodrı́guez, J., Vera, J., and Schrevens, E. (2010). Effect of compost application on
soil chemical and biological properties under potato crop in the Mantaro Valley—Peru.
In “Proceedings of the International Soil Science Congress, Samsun, Turkey,” pp. 211–217.
Garrett, K. A., Dendy, S. P., Frank, E. E., Rouse, M. N., and Travers, S. E. (2006). Climate
change effects on plant disease: Genomes to ecosystems. Annu. Rev. Phytopathol. 44,
489–509.
Geerts, S., Raes, D., Garcia, M., Del Castillo, C., and Buytaert, W. (2006). Agro-climatic
suitability mapping for crop production in the Bolivian Altiplano: A case study for
quinoa. Agric. Forest Meteorol. 139, 399–412.
Gianoli, E., Ramos, I., Alfaro-Tapia, A., Valdéz, Y., Echegaray, E. R., and Yábar, E. (2006).
Benefits of a maize-bean-weed mixed cropping system in Urubamba Valley, Peruvian
Andes. Int. J. Pest Manag. 52, 283–289.
Giller, K. E., Rowe, E. C., de Ridder, N., and van Keulen, H. (2006). Resource use dynamics
and interactions in the tropics: Scaling up in space and time. Agric. Syst. 88, 8–27.
Glick, B. R., Cheng, Z., Czarny, J., and Duan, J. (2007). Promotion of plant growth by
ACC deaminase-producing soil bacteria. Eur. J. Plant Pathol. 119, 329–339.
Goland, C. (1993). Field scattering as agricultural risk management: A case study from Cuyo
Cuyo, Department of Puno, Peru. Mt. Res. Dev. 13, 317–338.
Goodman-Elgar, M. (2008). Evaluating soil resilience in long-term cultivation: A study of
pre-Columbian terraces from the Paca Valley, Peru. J. Archaeol. Sci. 35, 3072–3086.
Govaerts, B., Mezzalama, M., Sayre, K. D., Crossa, J., Lichter, K., Troch, V., Vanherck, K.,
Corte, P.d., and Deckers, J. (2008). Long-term consequences of tillage, residue manage-
ment, and crop rotation on selected soil micro-flora groups in the subtropical highlands.
Appl. Soil Ecol. 38, 197–210.
Hahn, H., McManus, M. T., Warnstorff, K., Monahan, B. J., Young, C. A., Davies, E.,
Tapper, B. A., and Scott, B. (2008). Neotyphodium fungal endophytes confer physio-
logical protection to perennial ryegrass (Lolium perenne L.) subjected to a water deficit.
Environ. Exp. Bot. 63, 183–199.
Harden, C. (1988). Mesoscale estimation of soil erosion in the Rio Ambato drainage,
Ecuadorian Sierra. Mt. Res. Dev. 8, 331–341.
Harden, C. (1993). Land use, soil erosion, and reservoir sedimentation in an Andean
drainage basin in Ecuador. Mt. Res. Dev. 13, 177–184.
Harden, C. (1996). Interrelationships between land abandonment and land degradation: A
case from the ecuadorian andes. Mt. Res. Dev. 16, 274–280.
Harris, P. J. C., Allison, M., Smith, G., Kindness, H. M., and Kelley, J. (2001). The Potential
Use of Waste-Stream Products for Soil Amelioration in Peri-Urban Interface Agricul-
tural Production Systems. CAB International, Wallingford.
Hayat, R., Ali, S., Amara, U., Khalid, R., and Ahmed, I. (2010). Soil beneficial bacteria and
their role in plant growth promotion: A review. Ann. Microbiol. 60, 579–598.
Hellin, J., Welchez, L. A., and Cherrett, I. (1999). The Quezungual system: An indigenous
agroforestry system from western Honduras. Agroforest. Syst. 46, 229–237.
Henry, A., Kleinman, P. J. A., and Lynch, J. P. (2009). Phosphorus runoff from a phosphorus
deficient soil under common bean (Phaseolus vulgaris L.) and soybean (Glycine max L.)
genotypes with contrasting root architecture. Plant Soil 317, 1–16.
174 Steven J. Fonte et al.
Karak, T., and Bhattacharyya, P. (2011). Human urine as a source of alternative natural
fertilizer in agriculture: A flight of fancy or an achievable reality. Resour. Conserv. Recycl.
55, 400–408.
Karhu, K., Mattila, T., Bergström, I., and Regina, K. (2011). Biochar addition to agricultural
soil increased CH4 uptake and water holding capacity—Results from a short-term pilot
field study. Agric. Ecosyst. Environ. 140, 309–313.
Keneni, A., Assefa, F., and Prabu, P. C. (2010). Isolation of phosphate solubilizing bacteria
from the rhizosphere of faba bean of Ethiopia and their abilities on solubilizing insoluble
phosphates. J. Agric. Sci. Technol. 12, 79–89.
Kessler, C. A., and Stroosnijder, L. (2006). Land degradation assessment by farmers in
Bolivian mountain valleys. Land Degrad. Dev. 17, 235–248.
Khan, H. U. R., Link, W., Hocking, T. J., and Stoddard, F. L. (2007). Evaluation of
physiological traits for improving drought tolerance in faba bean (Vicia faba L.). Plant Soil
292, 205–217.
Khan, H. R., Paull, J. G., Siddique, K. H. M., and Stoddard, F. L. (2010). Faba bean
breeding for drought-affected environments: A physiological and agronomic perspective.
Field Crop. Res. 115, 279–286.
Kirkegaard, J. A., and Sarwar, M. (1998). Biofumigation potential of brassicas. I. Variation in
glucosinolate profiles of diverse field-grown brassicas. Plant Soil 201, 71–89.
Klein, A.-M., Vaissiere, B. E., Cane, J. H., Steffan-Dewenter, I., Cunningham, S. A.,
Kremen, C., and Tscharntke, T. (2007). Importance of pollinators in changing landscapes
for world crops. Proc. R. Soc. Lond. B. Biol. 274, 303–313.
Kochian, L. V., Piñeros, M. A., and Hoekenga, O. A. (2005). The physiology, genetics and
molecular biology of plant aluminum resistance and toxicity. Plant Soil 274, 175–195.
Kong, A. Y. Y., and Six, J. (2010). Tracing root vs. residue carbon into soils from
conventional and alternative cropping systems. Soil Sci. Soc. Am. J. 74, 1201–1210.
Kong, A. Y. Y., Six, J., Bryant, D. C., Denison, R. F., and van Kessel, C. (2005). The
relationship between carbon input, aggregation, and soil organic carbon stabilization in
sustainable cropping systems. Soil Sci. Soc. Am. J. 69, 1078–1085.
Kramer, S. B., Reganold, J. P., Glover, J. D., Bohannan, B. J. M., and Mooney, H. A.
(2006). Reduced nitrate leaching and enhanced denitrifier activity and efficiency in
organically fertilized soils. Proc. Natl. Acad. Sci. USA 103, 4522–4527.
Lagacherie, B., Bours, M., Giraud, J. J., and Sommer, G. (1983). Interaction between
Rhizobium lupini strains and species or cultivars of lupin (Lupinus albus, Lupinus luteus
and Lupinus mutabilis). Agronomie 3, 809–816.
Lange, F. R. (1994). Mejoramiento de la disponibilidad de fósforo por procesos biológicos de
la roca fosfórica. Facultad de Agronomı́a. Universidad Mayor San Simon, Cochabamba.
Lavelle, P., and Spain, A. V. (2001). Soil Ecology. Kluwer Academic Publishers, Dordretch,
Netherlands.
Lavelle, P., Blanchart, E., Martin, A., Martin, S., Spain, A., Toutain, F., Barois, I., and
Schaefer, R. (1993). A hierarchical model for the decomposition in terrestrial ecosystems:
Application to soils of the humid tropics. Biotropica 25, 130–150.
Lavelle, P., Bignell, D., Lepage, M., Wolters, V., Roger, P., Ineson, P., Heal, O. W., and
Dhillion, S. (1997). Soil function in a changing world: The role of invertebrate ecosystem
engineers. Eur. J. Soil Biol. 33, 159–193.
Lawler, J. J., Shafer, S. L., White, D., Kareiva, P., Maurer, E. P., Blaustein, A. R., and
Bartlein, P. J. (2009). Projected climate-induced faunal change in the Western Hemi-
sphere. Ecology 90, 588–597.
Leveau, J. H. J. (2007). The magic and menace of metagenomics: Prospects for the study of
plant growth-promoting rhizobacteria. Eur. J. Plant Pathol. 119, 279–300.
176 Steven J. Fonte et al.
Li, L., Li, S. M., Sun, J. H., Zhou, L. L., Bao, X. G., Zhang, H. G., and Zhang, F. S. (2007).
Diversity enhances agricultural productivity via rhizosphere phosphorus facilitation on
phosphorus-deficient soils. Proc. Natl. Acad. Sci. USA 104, 11192–11196.
Li, Y. Y., Yu, C. B., Cheng, X., Li, C. J., Sun, J. H., Zhang, F. S., Lambers, H., and Li, L.
(2009). Intercropping alleviates the inhibitory effect of N fertilization on nodulation and
symbiotic N-2 fixation of faba bean. Plant Soil 323, 295–308.
Liebman, M., and Dyck, E. (1993). Crop rotation and intercropping strategies for weed
management. Ecol. Appl. 3, 92–122.
Litterick, A. M., Wallace, P., Watson, C. A., and Wood, M. (2004). Compost extracts
in reducing pest and disease incidence and severity in sustainable temperate agricultural
and horticultural crop production—A review. Crit. Rev. Plant Sci. 23, 453–479.
Loon, L. C. (2007). Plant responses to plant growth-promoting rhizobacteria. Eur. J. Plant
Pathol. 119, 243–254.
Lorion, R. (2004). Rock Phosphate, Manure and Compost Use in Garlic and Potato Systems
in a High Intermontane Valley in Bolivia. Dept. of Crop and Soil Sciences, Washington
State University, Pullman, WA. p. 63.
Lorito, M., Woo, S. L., Harman, G. E., and Monte, E. (2010). Translational research on
Trichoderma: From ’omics to the field. Annu. Rev. Phytopathol. 48, 395–417.
Lynch, J. P. (2007). Roots of the second Green Revolution. Aust. J. Bot. 55, 493–512.
Machado, D., Sarmiento, L., and Gonzalez-Prieto, S. (2010). The use of organic substrates
with contrasting C/N ratio in the regulation of nitrogen use efficiency and losses in a
potato agroecosystem. Nutr. Cycl. Agroecosyst. 88, 411–427.
Mäder, P., Kaiser, F., Adholeya, A., Singh, R., Uppal, H. S., Sharma, A. K., Srivastava, R.,
Sahai, V., Aragno, M., and Wiemken, A. (2011). Inoculation of root microorganisms
for sustainable wheat–rice and wheat–black gram rotations in India. Soil Biol. Biochem. 43,
609–619.
Mahboubi, P., Gordon, A. M., Stoskopf, N., and Voroney, R. P. (1997). Agroforestry in the
Bolivian Altiplano: Evaluation of tree species and greenhouse growth of wheat on soils
treated with tree leaves. Agroforest. Syst. 37, 59–77.
Mamani, P., Botello, R., Condori, B., Moya, H., and Devaux, A. (2001). Efecto del tipo de
labranza con tracción animal en las caracterı́sticas fı́sicas del suelo, conservación de la
humedad y en el crecimiento y producción del cultivo de la papa. Revista Latinoamericana
de la Papa 12, 130–151.
Marichal, R. (2011). Impact de la déforestation sur les communautés de macrofaune et de
vers de terre en Amazonie—Relation avec les services écosystémiques. Université Pierre
et Marie Curie, Paris. p. 155.
Martinez-Viveros, O., Jorquera, M. A., Crowley, D. E., Gajardo, G., and Mora, M. L.
(2010). Mechanisms and practical considerations involved in plant growth promotion by
Rhizobacteria. J. Soil Sci. Plant Nutr. 10, 293–319.
Matson, P. A., Naylor, R., and Ortiz-Monasterio, I. (1998). Integration of environmental,
agronomic, and economic aspects of fertilizer management. Science 280, 112–115.
Mayak, S., Tirosh, T., and Glick, B. R. (2004). Plant growth-promoting bacteria that confer
resistance to water stress in tomatoes and peppers. Plant Sci. 166, 525–530.
Mayer, E. (1979). Land-Use in the Andes: Ecology and Agriculture in the Mantaro Valley of
Peru with Special Reference to Potatoes. International Potato Center, Lima, Peru.
Medrano Echalar, A. M., and Ortuño, N. (2007). Control de Damping off mediante la
aplicación e bioinsumos en almácigos de cebolla en el Valle Alto de Cochabamba,
Bolivia. Acta Nova 3, 660–679.
Mehari, A., Van Steenbergen, F., and Schultz, B. (2011). Modernization of spate irrigated
agriculture: A new approach. Irrig. Drain. 60, 163–173.
Meneses, R., Oller, V., and Waiijenberg, H. (2000). Inoculación y fertilización en el cultivo
de alfalfa en valles y alturas de Bolivia. In “Compendio de trabajos presentados por el
Andean Soil Fertility 177
Nustez, C. E., and Acevedo, J. C. (2005). Evaluating using Penicillium janthinellum Biourge
on the efficiency of phosphoric fertilisation of potato crops (Solanum tuberosum L. var.
Diacol Capiro). Agron. Colomb. 23, 290–298.
Nziguheba, G., Palm, C. A., Buresg, R. J., and Smithson, P. C. (1998). Soil phosphorus fractions
and adsorption as a affected by organic and inorganic sources. Plant Soil 198, 159–168.
Oberson, A., Friesen, D. K., Tiessen, H., Morel, C., and Stahel, W. (1999). Phosphorus
status and cycling in native savanna and improved pastures on an acid low-P Colombian
Oxisol. Nutr. Cycl. Agroecosyst. 55, 77–88.
Oberson, A., Bünemann, E. K., Friesen, D. K., Rao, I. M., Smithson, P. C., Turner, B. L.,
and Frossard, E. (2006). Improving phosphorus fertility in tropical soils through
biological interventions. “Biological Approaches to Sustainable Soil Systems” (Uphoff
et al., Eds.), pp. 531–546. CRC Press, Boca Raton.
Oliveira, C. A., Alves, V. M. C., Marriel, I. E., Gomes, E. A., Scotti, M. R.,
Carneiro, N. P., Guimaraes, C. T., Schaffert, R. E., and Sa, N. M. H. (2009). Phosphate
solubilizing microorganisms isolated from rhizosphere of maize cultivated in an oxisol of
the Brazilian Cerrado Biome. Soil Biol. Biochem. 41, 1782–1787.
Oorts, K., Vanlauwe, B., and Merckx, R. (2003). Cation exchange capacities of soil organic
matter fractions in a ferric lixisol with different organic matter inputs. Agric. Ecosyst.
Environ. 100, 161–171.
Orlove, B., and Godoy, R. (1986). Sectoral fallowing systems in the central Andes.
J. Enthnobiology 6, 169–204.
Orna, A. R. (2009). Evaluación del efecto de la aplicación de micorrizas en la producción de
tomate riñon (Solanum lycopersicon) bajo invernadero. Recursos Naturales. ESPOCH,
Riobamba, Ecuador.
Orsag, V. (2009). Degradación de suelos en el Altiplano Boliviano. Análisis—Instituto
Boliviano de Economı´a y Polı´tica Agraria 1, 27–30.
Ortuño, N., Navia, O., Medrano, A., Rojar, K., and Torrico, L. (2010). Desarrollo de
bioinsumos: Un aporte a la soberanı́a alimentaria de Bolivia. Rev. Agric. 47, 30–38.
Ortuño, N. (2010). Desarrollo de bioinsumos para la producción sosteniblle de hortalizas con
pequeños agricultores para una soberanı́a alimentariia en los Andes. CIAT, Colombia.
Osman, A. (1999). Soil Fertility Management in Cajamarca, Peru. Centro Internacional de la
Papa (CIP), Lima.
Osorio-Vega, N. W. (2007). Review of beneficial effects of rhizosphere bacteria on soil
nutrient availability and plant nutrient uptake. Revista de la Facultad Nacional de Agronomı´a
Medellin 60, 3621–3643.
Oswald, A. (2010). Soil Fertility Management and Crop Productivity of Potato Based
Cropping Systems of the Andean Highlands of Peru. Working Paper. Centro Interna-
cional de la Papa, pp. 1–29.
Oswald, A., Haan, S., Sanchez, J., and Ccanto, R. (2009). The complexity of simple tillage
systems. J. Agric. Sci. 147, 399–410.
Oswald, A., Calvo Velez, P., Zúñiga Dávila, D., and Arcos Pineda, J. (2010). Evaluating soil
rhizobacteria for their ability to enhance plant growth and tuber yield in potato. Ann.
Appl. Biol. 157, 259–271.
Otero, J., and Onaindiam, M. (2009). Landscape structure and live fences in Andes Colombian
agroecosystems: Upper basin of the Cane-Iguaque River. Int. J. Trop. Biol. 57, 1183–1192.
Oyarzun, P. J., Gerlagh, M., and Zadoks, J. C. (1998). Factors associated with soil receptivity
to some fungal root rot pathogens of peas. Appl. Soil Ecol. 10, 151–169.
Pacheco, V., Zelada, A., and Navarro, C. (1992). Recuperación de tierras en el Proyecto
Norte Chuquisaca. Proyecto Norte Chuquisaca, Sucre, Bolivia.
Palm, C. A., Gachengo, C. N., Delve, R. J., Cadisch, G., and Giller, K. E. (2001). Organic
inputs for soil fertility management in tropical agroecosystems: Application of an organic
resource database. Agric. Ecosyst. Environ. 83, 27–42.
Andean Soil Fertility 179
Palta, J. A., Turner, N. C., French, R. J., and Buirchell, B. J. (2007). Physiological responses
of lupin genotypes to terminal drought in a Mediterranean-type environment. Ann. Appl.
Biol. 150, 269–279.
Pansak, W., Dercon, G., Hilger, T., Kongkaew, T., and Cadisch, G. (2007). 13C isotopic
discrimination: A starting point for new insights in competition for nitrogen and water
under contour hedgerow systems in tropical mountainous regions. Plant Soil 298, 175–189.
Pansu, M., Sarmiento, L., Metselaar, K., and Bottner, P. (2007). Modelling the transforma-
tions and sequestration of soil organic matter in two contrasting ecosystems of the Andes.
Eur. J. Soil Sci. 58, 775–785.
Paredes, M. (2010). Peasants, Potatoes and Pesticides: Heterogeneity in the Context of
Agricultural Modernization in the Highland Andes of Ecuador. School of Social
Sciences. Wageningen Univiersity, Wageningen. p. 322.
Parsa, S. (2010). Native herbivore becomes key pest after dismantlement of a traditional
farming system. Am. Entomol. 56, 242–251.
Parsa, S., Ccanto, R., and Rosenheim, J. A. (2011). Resource concentration dilutes a key
pest in indigenous potato agriculture. Ecol. Appl. 21, 539–546.
Pearse, S. J., Veneklaas, E. J., Cawthray, G. R., Bolland, M. D. A., and Lambers, H. (2006).
Carboxylate release of wheat, canola and 11 grain legume species as affected by phospho-
rus status. Plant Soil 288, 127–139.
Perez, C., Nicklin, C., D’angles, O., Vanek, S. J., Sherwood, S., Halloy, S., Garrett, K., and
Forbes, G. (2010). Climate change in the high Andes: Implications and adaptation
strategies for small-scale farmers. Int. J. Environ. Cult. Econ. Soc. Sustain. 6, 71–88.
Pestalozzi, H. (2000). Sectoral fallow systems and the management of soil fertility: The rationality
of indigenous knowledge in the high Andes of Bolivia. Mt. Res. Dev. 20, 64–71.
Pimentel, D. (2006). Soil erosion: A food and environmental threat. Environ. Dev. Sustain. 8,
119–137.
Podwojewski, P., and Germain, N. (2005). Short-term effects of management on the soil
structure in a deep tilled hardened volcanic-ash soil (Cangahua) in Ecuador. Eur. J. Soil
Sci. 39, 39–51.
Posthumus, H., and De Graaff, J. (2005). Cost-benefit analysis of bench terraces. A case study
in Peru. Land Degrad. Dev. 16, 1–11.
Posthumus, H., Gardebroek, C., and Ruben, R. (2010). From participation to adoption:
Comparing the effectiveness of soil conservation programs in the Peruvian Andes. Land
Econ. 86, 645–667.
Postma-Blaauw, M., De Goede, R. G. M., Bloem, J., Faber, J. H., and Brussard, L. (2010).
Soil biota community structure and abundance under agricultural intensification and
extensification. Ecology 91, 460–473.
Poulenard, J., Podwojewski, P., Janeau, J.-L., and Collinet, J. (2001). Runoff and soil
erosion under rainfall simulation of Andisols from the Ecuadorian Páramo: Effect of
tillage and burning. Catena 45, 185–207.
Poveda, K., Gómez, M. I., and Martı́nez, E. (2008). Diversification practices: Their effect on
pest regulation and production. Rev. Colomb. Entomol. 34, 131–144.
Powlson, D. S., Gregory, P. J., Whalley, W. R., Quinton, J. N., Hopkins, D. W.,
Whitmore, A. P., Hirsch, P. R., and Goulding, K. W. T. (2011). Soil management in
relation to sustainable agriculture and ecosystem services. Food Policy 36, S72–S87.
Puget, P., and Drinkwater, L. E. (2001). Short-term dynamics of root- and shoot-derived
carbon from a leguminous green manure. Soil Sci. Soc. Am. J. 65, 771–779.
Quilty, J. R., and Cattle, S. R. (2011). Use and understanding of organic amendments in
Australian agriculture: A review. Aust. J. Soil Res. 49, 1–26.
Quintero, M. (2009). Effects of Conservation Tillage in Soil Carbon Sequestration and Net
Revenues of Potato-Based Rotations in the Colombian Andes. University of Florida,
Gainesville, FL. p. 103.
180 Steven J. Fonte et al.
Quiroz, R., León-Velarde, C., Valdivia, R., Zorogastúa, F. P., Baigorria, G., Barreda, C.,
Reinoso, J., Holle, M., and Li Pun, H. (2003). Making a difference to Andean livelihoods
through an integrated research approach. In “Research Towards Integrated Natural
Resources Management—Examples of Research Problems, Approaches and Partnerships
in Action in the CGIAR” (R. R. Harwood and A. H. Kassam, Eds.), pp. 111–122. FAO,
Rome.
Ramaekers, L., Remans, R., Rao, I. M., Blair, M. W., and Vanderleyden, J. (2010).
Strategies for improving phosphorus acquisition efficiency of crop plants. Field Crop.
Res. 117, 169–176.
Ramesh, K., Biswas, A. K., Somasundaram, J., and Rao, A. S. (2010). Nanoporous zeolites
in farming: Current status and issues ahead. Curr. Sci. India 99, 760–764.
Reed, S. C., Seastedt, T. R., Mann, C. M., Suding, K. N., Townsend, A. R., and
Cherwin, K. L. (2007). Phosphorus fertilization stimulates nitrogen fixation and increases
inorganic nitrogen concentrations in a restored prairie. Appl. Soil Ecol. 36, 238–242.
Reynel, C., and Felipe-Morales, C. (1990). Agroforesteria tradicional en los Andes del Perú:
un inventario de tecnologı́as y especies para la integración de la vegetación leñosa a la
agricultura. Proyecto FAO/Holanda/INFOR, Peru.
Ricketts, T. H., Regetz, J., Steffan-Dewenter, I., Cunningham, S. A., Kremen, C.,
Bogdanski, A., Gemmill-Herren, B., Greenleaf, S. S., Klein, A. M., Mayfield, M. M.,
Morandin, L. A., Ochieng, A., et al. (2008). Landscape effects on crop pollination
services: Are there general patterns? Ecol. Lett. 11, 499–515.
Righi, G., and van der Hammen, T. (1996). Distribución de especies de lombrices en las dos
vertientes de la cordillera central (transecto parque Los Nevados, Colombia). In “Estudios
de ecosistemas tropandinos 4” (T. Van der Hammen and A. G. dos Santos, Eds.),
pp. 475–483. Cramer (Brontraeger), Berlin-Stuttgart.
Robertson, G. P., and Swinton, S. M. (2005). Reconciling agricultural productivity
and environmental integrity: A grand challenge for agriculture. Front. Ecol. Environ. 3,
38–46.
Rodriguez, K. R., and Ortuño, N. (2007). Evaluación de micorrizas arbusculares en
interacción con abonos orgánicos como coadyuvantes del crecimiento en la producción
hortı́cola del Valle Alto de Cochabamba, Bolivia. Acta Nova 3, 697–720.
Rodriguez, E. A., Bolaños, B. M. M., and Menjivar Flores, J. C. (2010). Efecto de la
fertilización en la nutrición y rendimiento de ajı́ (Capsicum spp.) en el Valle del Cauca,
Colombia. Acta Agron. 59, 55–64.
Rodrı́guez, H., Fraga, R., Gonzalez, T., and Bashan, Y. (2006). Genetics of phosphate
solubilization and its potential applications for improving plant growth-promoting bac-
teria. Plant Soil 287, 15–21.
Romero-León, C. C. (2005). A Multi-Scale Approach for Erosion Assessment in the Andes.
Wageningen University, Wageningen, Netherlands. p. 162.
Rosset, P. M., and Altieri, M. A. (1997). Agroecology versus input substitution: A funda-
mental contradiction of sustainable agriculture. Soc. Natur. Resour. 10, 283–295.
Sanchez-Moreno, S., and Ferris, H. (2007). Suppressive service of the soil food web: Effects
of environmental management. Agric. Ecosyst. Environ. 119, 75–87.
Sandor, J., and Eash, N. (1995). Ancient agricultural soils in the Andes of southern Peru. Soil
Sci. Soc. Am. J. 59, 170–179.
Sandor, J., and Furbee, L. (1996). Indigenous knowledge and classification of soils in the
Andes of southern Peru. Soil Sci. Soc. Am. J. 60, 1502–1512.
Sarmiento, L. (2000). Water balance and soil loss under long fallow agriculture in the
Venezuelan Andes. Mt. Res. Dev. 20, 1–9.
Sarmiento, L., and Bottner, P. (2002). Carbon and nitrogen dynamics in two soils with
different fallow times in the high tropical Andes: Indications for fertility restoration. Appl.
Soil Ecol. 19, 78–79.
Andean Soil Fertility 181
Sarmiento, L., Monasterio, M., and Montilla, M. (1993). Ecological bases, sustainability, and
current trends in traditional agriculture in the Venezuelan high Andes. Mt. Res. Dev. 13,
167–176.
Sarmiento, L., Acea, M., Barrios, E., Bowen, W., Herrera, R., Llambı́, L., Ortuño, N.,
Sivila, R., and Varela, A. (2001). Un marco conceptual y metodológico para estudios de
fertilidad del suelo en los Andes tropicales. In “Memorias del IV Simposio Internacional
de Desarrollo Sustentable en los Andes: La estrategia para el siglo XXI”.
Sau, F., and Minguez, M. I. (2000). Adaptation of indeterminate faba beans to weather and
management under a Mediterranean climate. Field Crop. Res. 66, 81–99.
Schuster, A., and Schmoll, M. (2010). Biology and biotechnology of Trichoderma. Appl.
Microbiol. Biotnechnol. 87, 787–799.
Sciarascia Mugnozza, G. C., Avila, G., Claure, T., Rios, R., Pierola, L., and Crespo, M.
(1987). Investigaciones sobre el mejoramiento genético y cultural de trigo duro, girasol,
maı́z, frijol, lupino y haba en Bolivia. Instituto Italo-Latinoamericano Fundación
Pro-Bolivia, Centro de Investigaciones Fitoecogenéticos Pairumani, Rome.
Seneviratne, G., Jayasekara, A. P. D. A., De Silva, M. S. D. L., and Abeysekera, U. P. (2011).
Developed microbial biofilms can restore deteriorated conventional agricultural soils. Soil
Biol. Biochem. 43, 1059–1062.
Setimela, P., Chitalu, Z., Jonazi, J., Mambo, A., Hodson, D., and Banziger, M. (2005).
Environmental classification of maize-testing sites in the SADC region and its implica-
tion for collaborative maize breeding strategies in the subcontinent. Euphytica 145,
123–132.
Sharma, S. N. (2003). Effect of phosphate-solubilizing bacteria on efficiency of Mussoorie
rockphosphate in rice (Oryza sativa)-wheat (Triticum aestivum) cropping system. Indian J.
Agric. Sci. 73, 478–481.
Sherwood, S. G., Monar, C., and Suquillo, J. (1999). Wachu rozado: Vestigio del pasado,
oportunidad para el futuro. Centro Internacional de la Papa (CIP), Quito.
Siddique, K. H. M., Regan, K. L., Tennant, D., and Thomson, B. D. (2001). Water use and
water use efficiency of cool season grain legumes in low rainfall Mediterranean-type
environments. Eur. J. Agron. 15, 267–280.
Sims, B. G., and Rodriguez, F. (2001). Forage production and erosion control as a comple-
ment to hillside weed management. In “International Workshop in Integrated Manage-
ment for Sustainable Agriculture, Forestry and Fisheries” Centro Internacional de
Agricultura Tropical (CIAT) Cali, Colombia.
Sims, B., Rodrı́guez, F., Eid, M., and Espinoza, T. (1999). Biophysical aspects of vegetative
soil and water conservation pratices in the inter-Andean valles of Bolivia. Mt. Res. Dev.
19, 282–291.
Singhai, P. K., Sarma, B. K., and Srivastava, J. S. (2011). Biological management of
common scab of potato through Pseudomonas species and vermicompost. Biol. Control
57, 150–157.
Sivila de Cary, R., and Hervé, D. (1994). El estado microbiologico del suelo, indicador de
una restauracion de la fertilidad. In “Dinámicas del descanso de la tierra en los Andes”
(D. Hervé, Ed.), pp. 185–197. IBTA—OSTROM, La Paz, Bolivia.
Sivila, R., and Herve, D. (1999). Análisis de la microbiota en suelos cultivados del altiplano
central. In “Primer Congreso Boliviano de la Ciencia del Suelo Sociedad Boliviana de la
Ciencia del Suelo, La Paz” pp. 5–14.
Six, J., Conant, R. T., Paul, E. A., and Paustian, K. (2002). Stabilization mechanisms of soil
organic matter: Implications for C-saturation of soils. Plant Soil 241, 155–176.
Smaling, E. M. A., and Fresco, L. O. (1993). A decision-support model for monitoring
nutrient balances under agricultural land-use (NUTMON). Geoderma 60, 235–256.
Smaling, E. M. A., Stoorvegel, J. J., and Windmeijer, P. N. (1993). Calculating soil nutrient
balances in Africa at different scales: II District Scale. Fertil. Res. 35, 237–250.
182 Steven J. Fonte et al.
Smith, R. G., Gross, K. L., and Robertson, G. P. (2008). Effects of crop diversity on
agroecosystem function: Crop yield response. Ecosystems 11, 355–366.
Smukler, S. M., Sánchez-Moreno, S., Fonte, S. J., Ferris, H., Klonsky, K., O’Geen, A. T.,
Scow, K. M., Steenwerth, K. L., and Jackson, L. E. (2010). Biodiversity and multiple
ecosystem functions in an organic farmscape. Agric. Ecosyst. Environ. 139, 80–97.
Snapp, S. S., and Silim, S. N. (2002). Farmer preferences and legume intensification for low
nutrient environments. Plant Soil 245, 181–192.
Snapp, S. S., Mafongoya, P. L., and Waddington, S. (1998). Organic matter technologies for
integrated nutrient management in smallholder cropping systems of southern Africa.
Agric. Ecosyst. Environ. 71, 185–200.
Snapp, S., Swinton, S. M., Labarta, R., Mutch, D., Black, J. R., Leep, R., Nyiraneza, J., and
O’Neil, K. (2005). Evaluating cover crops for benefits, costs and performance within
cropping system niches. Agron. J. 97, 322–332.
Snoek, C., Vanderleyden, J., and Beebe, S. (2003). Strategies for genetic improvement
of Common Bean and Rhizobia: Towards efficient interactions. Plant Breed. Rev. 23,
21–72.
Stadel, C. (1991). Environmental stress and sustainable development in the tropical andes.
Mt. Res. Dev. 11, 213–223.
Stanish, C. (2007). Agricultural intensification in the Titicaca basin. In “Seeking a Richer
Harvest” (T. Thurston and C. Fisher, Eds.), pp. 125–139. Springer, New York.
Swinton, S., and Quiroz, R. (2003). Poverty and the deterioration of natural soil capital in
the Peruvian Altiplano. Environ. Dev. Sustain. 5, 477–490.
Takeda, M., and Knight, J. D. (2006). Enhanced solubilization of rock phosphate by
Penicillium bilaiae in pH-buffered solution culture. Can. J. Microbiol. 52, 1121–1129.
Tapia, M. (1994). Rotación de cultivos y su manejo en los Andes del Perú. In “Dinámicas del
descanso de la tierra en los Andes” (D. Hervé, Ed.), pp. 37–53. IBTA—OSTROM, La
Paz, Bolivia.
Tenge, A. J., De Graaff, J., and Hella, J. P. (2004). Social and economic factors affecting the
adoption of soil and water conservation in west Usambara highlands, Tanzania. Land
Degrad. Dev. 15, 99–114.
Terrazas, F., Suarez, G., Gardner, G., Thiele, G., Devaux, A., and Walker, T. (1998). Diagnosing
Potato Productivity in Farmers Fields in Bolivia. International Potato Center (CIP), Lima.
Thamer, S., Schädler, M., Bonte, D., and Ballhorn, D. J. (2011). Dual benefit from a
belowground symbiosis: Nitrogen fixing rhizobia promote growth and defense against
a specialist herbivore in a cyanogenic plant. Plant Soil 341, 209–219.
Thibeault, J., Seth, A., and Garcia, M. (2010). Changing climate in the Bolivian Altiplano:
CMIP3 projections for temperature and precipitation extremes. J. Geophys. Res. 115, 1–18.
Thies, J., and Grossman, J. (2006). The soil habitat and soil ecology. In “Biological
Approaches to Sustainable Soil Systems” (N. Uphoff, Ed.), pp. 59–78. CRC Press,
Boca Raton.
Thies, J., and Rillig, M. (2009). Characteristics of Biochar: Biological Properties. In “Bio-
char for Environmental Management”, (J. Lehmann and S. Joseph, Eds), pp. 85–105.
Earthscan, London.
Tisdall, J. M., and Oades, J. M. (1982). Organic matter and water-stable aggregates in soils.
J. Soil Sci. 62, 141–163.
Tittonell, P., Vanlauwe, B., Leffelaar, P. A., Shepherd, K. D., and Giller, K. E. (2005).
Exploring diversity in soil fertility management of smallholder farms in western Kenya: II.
Within-farm variability in resource allocation, nutrient flows and soil fertility status. Agric.
Ecosyst. Environ. 110, 166–184.
Tonneijck, F., and Jongmans, G. (2008). The influence of bioturbation on the vertical
distribution of soil organic matter in volcanic ash soils: A case study in northern Ecuador.
Eur. J. Soil Sci. 59, 1063–1075.
Andean Soil Fertility 183
Tonneijck, F., Jansen, B., Niero, K. G. J., Verstraten, J. M., Sevink, J., and De Lange, L.
(2010). Towards understanding of carbon stocks and stabilization in volcanic ash soils in
natural Andean ecosystems of northern Ecuador. Eur. J. Soil Sci. 61, 392–405.
Tourneux, C., Devaux, A., Camacho, M. R., Mamani, P., and Ledent, J. F. (2003a). Effect
of water shortage on six potato genotypes in the highlands of Bolivia (II): Water relations,
physiological parameters. Agronomie 23, 181–190.
Tourneux, C., Devaux, A., Camacho, M. R., Mamani, P., and Ledent, J. F. (2003b). Effects
of water shortage on six potato genotypes in the highlands of Bolivia (I): Morphological
parameters, growth and yield. Agronomie 23, 169–179.
Urgiles, N., Loján, P., Aguirre, N., Blaschke, H., Günter, S., Stimm, B., and Kottke, I.
(2009). Application of mycorrhizal roots improves growth of tropical tree seedlings in the
nursery: A step towards reforestation with native species in the Andes of Ecuador. New
Forest. 38, 229–239.
Valdivia, C., and Quiroz, R. (2003). Coping and adapting to increased climate variability in
the Andes. In “Annual Meeting, Presentation at the American Agricultural Economics
Association”.
Valdivia, C., Seth, A., Gilles, J., Garcia, M., Jiménez, E., Cusicanqui, J., Navia, F., and
Yucra, E. (2010). Adapting to climate change in Andean ecosystems: Landscapes, capitals,
and perceptions shaping rural livelihood strategies and linking knowledge systems. Ann.
Assoc. Am. Geogr. 100, 818–834.
Valente, J. F., and Oliver, R. (1993). Fertisuelos: Evaluacion de la fertilidad de los suelos del
antiplano, valle central y los llanos de Bolivia. FAO, Rome.
van de Kop, P. (1996). Regional Scale Nutrient Balances for Agro-Ecosystems in Ecuador.
Department of Agronomy. Wageningen Agricultural University, Wageningen. p. 43.
van Groenigen, K. J., Six, J., Hungate, B. A., de Graaff, M., van Breemen, N., and van
kessel, C. (2006). Elemental interactions limit soil carbon storage. Proc. Natl. Acad. Sci.
USA 103, 6571–6574.
Vanek, S. (2010). Legume-Phosphorus Synergies in Mountain Agroecosystems: Field
Nutrient Balances, Soil Fertility Gradients, and Effects on Legume Attributes and Nutri-
ent Cycling in the Bolivian Andes. Cornell University, Ithaca, NY.
Vanlauwe, B., Wendt, J., and Diels, J. (2001). Combined application of organic matter and
fertilizer. In “Sustaining Soil Fertility in West-Africa” (G. Tian, Ed.), pp. 247–279. Soil
Science Society of America, Madison, WI.
Vanlauwe, B., Tittonell, P., and Mukalama, J. (2006). Within-farm soil fertility gradients
affect response of maize to fertiliser application in western Kenya. Nutr. Cycl. Agroecosyst.
76, 171–182.
Vargas, C. A. (2009). Sistema de innovación agropecuaria y forestal. Revista de Agricultura
(Cochabamba, Bolivia) 45, 2–9.
Velasquez, E., Lavelle, P., and Andrade, M. (2007). GISQ, a multifunctional indicator of soil
quality. Soil Biol. Biochem. 39, 3066–3080.
Verma, M., Brar, S. K., Tyagi, R. D., Surampalli, R. Y., and Valero, J. R. (2007). Antagonistic
fungi, Trichoderma spp.: Panoply of biological control. Biochem. Eng. J. 37, 1–20.
Vinale, F., Krishnapillai, S., Ghisalberti, E. L., Marra, R., Woo, S. L., and Lorito, M. (2008).
Trichoderma-plant-pathogen interactions. Soil Biol. Biochem. 40, 1–10.
Vitousek, P. M., Naylor, R., Crews, T., David, M. B., Drinkwater, L. E., Holland, E.,
Johnes, P. J., Katzenberger, J., Martinelli, L. A., Matson, P. A., Nziguheba, G.,
Ojima, D., et al. (2009). Nutrient imbalances in agricultural development. Science 324,
1519–1520.
Wall, P. (1999). Experiences with crop residue cover and direct seeding in the Bolivian
highlands. Mt. Res. Dev. 19, 313–317.
Wang, K. H., Sipes, B. S., and Schmitt, D. P. (2002). Crotalaria as a covercrop for nematode
management: Review. Nematrópica 32, 35–57.
184 Steven J. Fonte et al.
Warnock, D. D., Lehmann, J., Kuyper, T. W., and Rillig, M. C. (2007). Mycorrhizal
responses to biochar in soil—Concepts and mechanisms. Plant Soil 300, 9–20.
Wheeler, T. R., Keatinger, H., Ellis, R. H., and Summerfield, R. J. (1999). Selecting
legume cover crops for hillside environments in Bolivia. Mt. Res. Dev. 19, 318–324.
Wiegers, E. S., Hijmans, R. J., Herve, D., and Fresco, L. O. (1999). Land use intensification
and disintensification in the Upper Canete valley, Peru. Hum. Ecol. 27, 319–357.
Winklerprins, A. (1999). Local soil knowledge: A tool for sustainable land management. Soc.
Nat. Resour. 12, 151–161.
Winters, P., Espinosa, P., and Crissman, P. (1998). Manejo de los Recursos en los Andes
Ecuatorianos: Revisión de Literatura y Evaluación del Proyecto Manejo del Uso Sos-
tenible de Tierras Andinas (PROMUSTA) de CARE. International Potato Center
(CIP), Quito, Ecuador.
Woomer, P., and Swift, M. J. (1994). The Biological Management of Tropical Soil Fertility.
John Wiley, Chichester, UK.
Wu, L., Vomocil, J. A., and Childs, S. W. (1990). Pore-size, particle-size, aggregate size, and
water-retention. Soil Sci. Soc. Am. J. 54, 952–956.
Young, A. (1997). Agroforestry for Soil Management. CAB International, Oxon, UK.
Zebarth, B. J., Tarn, T. R., de Jong, H., and Murphy, A. (2008). Nitrogen use efficiency
characteristics of andigena and diploid potato selections. Am. J. Potato Res. 85, 210–218.
Zehetner, F., and Miller, W. P. (2006). Soil variations along a climatic gradient in an Andean
agro-ecosystem. Geoderma 137, 126–134.
Zhu, J. M., Kaeppler, S. M., and Lynch, J. P. (2005). Topsoil foraging and phosphorus
acquisition efficiency in maize (Zea mays). Funct. Plant Biol. 32, 749–762.
Zimmerer, K. (1993). Soil erosion and labor shortages in the Andes with special reference to
Bolivia, 1953-91: Implications for “conservation-with-development. World Dev. 21,
1659–1675.
Zimmerer, K. S. (2011). The landscape technology of spate irrigation amid development
changes: Assembling the links to resources, livelihoods, and agrobiodiversity-food in the
Bolivian Andes. Glob. Environ. Chang. 21, 917–934.
Zurita, G. L. (1997). Composición taxonómica y abundancia poblacional de Lombrices en
sistemas de Monocultivo y Rotación de cultivos en Suka Kollus. Universidad Mayor de
San Andrés, La Paz, Bolivia.
C H A P T E R F I V E
Contents
1. Introduction 186
2. Defining PUE: Terms, Units, and Assumptions 189
2.1. Criteria with agronomic implications 189
2.2. Criteria with physiological implications 192
2.3. Defining the utilization of P as “efficiency” 195
3. Quantifying PUE of Crop Genotypes Using Criteria with
Physiological Implications 196
3.1. Screening for vegetative PUE 197
3.2. Screening for grain PUE 200
4. P-Stress Levels in Screening Studies and the Utility of PUE in Low,
Medium, and High P Input Systems 201
4.1. P-deficient crops suffer a range of stress levels 201
4.2. What are the likely outcomes of improved PUE in
P-deficient plants? 202
5. Mechanisms and Physiology of PUE 205
5.1. Remobilization and scavenging of P 205
5.2. Alternative glycolytic pathways and mitochondrial electron
transport pathways 206
5.3. Exploiting P-deficiency stress response mechanisms 206
6. Conclusions and Future Prospective 208
6.1. PAE versus PUE—Does one offer better chances of success? 208
6.2. Screening methods, targets, and possible results 208
6.3. Marker-assisted selection—A paradigm shift in breeding
suited for PUE 209
6.4. Remaining questions 210
References 211
* Southern Cross Plant Science, Southern Cross University, Lismore, NSW, Australia
{
Japan International Research Center for Agricultural Sciences (JIRCAS), Crop Production and Environment
Division, Ohwashi, Tsukuba, Ibaraki, Japan
Abstract
Grain crops are a key driver of the current global phosphorus (P) cycle through
their continued demand for P fertilizer, and the subsequent removal of P from
fields in the harvested grain. Breeding crops that can yield well with fewer P
inputs (i.e., P-efficient crops) may reduce the impact of grain crops of the P cycle,
but to date breeding P-efficient cultivars has focused on enhancing P acquisition
efficiency (PAE). While the literature abounds in reported genotypic differences in
internal P utilization efficiency (PUE) across a range of crops, there has been little
progress in breeding crop cultivars with high PUE. This review critically analyzes
why drawing conclusions from the body of research on PUE over the past few
decades remains difficult and how progress in breeding crop cultivars high in PUE
has been impeded. Four aspects of research on PUE are highlighted as being
critical in limiting our understanding and exploitation of PUE in grain crops:
(i) poor definition of PUE and inconsistent use of terminology, (ii) inappropriate
methods used in genotypic screening for PUE that fail to account for the con-
founding effects of PAE on PUE, (iii) inadequate discussion on the level of P stress
suffered by plants and its influence on potential mechanisms conferring high PUE
and their utility in cropping systems, and (iv) a focus on P-stress response
mechanisms rather than mechanisms conferring genotypic P-tolerance when
investigating PUE. These factors are discussed in detail and new approaches
and future areas of research on PUE are proposed.
1. Introduction
Phosphorus (P) is the second most limiting macronutrient to plant
growth and lack of plant-available P constrains plant growth in over 5.7
billion ha of land worldwide (Batjes, 1997). Large amounts of soil P are
“locked up” in recalcitrant organic P fractions or nonlabile inorganic P
pools in complexes with iron/aluminum in acid soils or with calcium in
alkaline soils. While application of phosphorus fertilizer can overcome these
constraints, the lack of locally available P fertilizer sources and the high cost
of importing and transporting P fertilizers typically prevent resource-poor
farmers in developing countries from doing so (Wissuwa and Ae, 2001).
Even in Western countries where food and fiber production relies heavily
on the application of nonrenewable P fertilizers, much of this fertilizer is
gradually rendered unavailable to plants over time as it reacts with soil
constituents (Richardson et al., 2011).
In addition to P immobilization in soils, a second frequently overlooked
factor is that the demand for P fertilizer is further driven by the removal of
almost 10million tons of P across the globe each year in harvested produce
(Lott et al., 2000), with grain crops (cereals, oilseeds, and pulses) being
responsible for the vast majority of this P removal at harvest (Lott et al.,
2002). Estimates suggest that as much as 85% of P applied as fertilizer can be
Internal Phosphorus Utilization Efficiency in Grain Crops 187
removed from fields in harvest product each year (Lott et al., 2009), though
regional imbalances in P budgets are apparent such that excess P inputs are
applied in many areas while 30% of the globe suffer from deficits in P
budgets (MacDonald et al., 2011). This inefficiency of P use in agriculture
has been highlighted in many recent reviews because of environmental
concerns regarding P from field runoff and human waste entering water-
ways and because finite global rock phosphate reserves are being depleted
(e.g., Cordell et al., 2009; Richardson et al., 2011; Shen et al., 2011).
Improving the P efficiency of farming systems (higher crop yields per
unit of P fertilizer applied) can be achieved using agronomic strategies, for
example, increasing P fertilizer availability to crops using liquid fertilizers
(Holloway et al., 2001) or strategic fertilizer placement (Ma et al., 2009).
In addition, breeding P-efficient crop cultivars has been advocated for
improving the P efficiency of cropping systems because it is relatively low
cost and, unlike optimizing fertilizer formulations and application strategies, it
can provide benefits both in high-input systems and in low-input systems
where fertilizer application may be rare or nonexistent due to its high cost
(Rose et al., 2010). Scope also exists to breed more P-efficient cultivars
because modern cultivars are generally not highly efficient at acquiring and
utilizing P, having been bred under optimal conditions that selected against
P-efficiency traits often present in landrace genotypes (Wissuwa et al., 2009).
Traits that confer P efficiency have typically been divided into those that
improve P acquisition efficiency (PAE) from soils or those that enhance
(internal) P utilization efficiency (PUE) (Wang et al., 2010). However, to
the best of our knowledge, all advances in breeding P-efficient grain crop
cultivars have involved the exploitation of PAE traits (Wissuwa et al., 2009),
and the genetics and mechanisms conferring PAE have been widely reviewed
elsewhere (Hinsinger, 2001; Ramaekers et al., 2010; Shen et al., 2011; Vance
et al., 2003). In contrast, PUE is poorly understood, despite a plethora of
studies reporting variation in PUE among grain crop genotypes (Table 1), the
mapping of several QTLs for PUE in a range of grain crops (Chen et al., 2009;
Hammond et al., 2009; Su et al., 2006, 2009; Wissuwa et al., 1998; Yang et al.,
2011), and discussion on PUE in many reviews on P efficiency (Ahmad et al.,
2001; Batten, 1992; Richardson et al., 2011; Shen et al., 2011; Shenoy and
Kalagudi, 2005; Vance et al., 2003; Wang et al., 2010).
A number of factors have contributed to our poor understanding of PUE
and lack of advances in breeding crops with enhanced PUE. First and
foremost, PUE is not clearly defined. Typically, PUE is expressed as biomass
per unit P, but biomass may refer to entire plant biomass or just grain yield,
whereas the unit P may refer to fertilizer P applied, P taken up by a plant, or
P present in specific tissues. The lack of a single definition being adopted
across studies has made it difficult to draw conclusions from the literature.
Second, PUE has long been the “‘poor cousin” in studies on P efficiency,
typically investigated as an additional component in studies primarily
188 Terry J. Rose and Matthias Wissuwa
Crop Authors
Barley Górny (1999) and Römer and Schenk (1998)
Canola Akhtar et al. (2008, 2009), Aziz et al. (2006), Duan et al.
(2009), Hammond et al. (2009), and Yang et al. (2011)
Common bean Araújo et al. (1997) and Fageria and da Costa (2000)
Faba bean Stelling et al. (1996) and T. Rose (unpublished data)
Maize Chen et al. (2009), Corrales et al. (2007), Fageria and Baligar
(1997a), and Parentoni and Júnior (2008)
Pigeon pea Adu-Gyamfi et al. (1989) and Subbarao et al. (1997)
Rice Aziz et al. (2005), Fageria and Baligar (1997b), Fageria et al.
(1988), Gill et al. (2002), Hafeez et al. (2010), Hedley et al.
(1994), Saleque et al. (1998), Sahrawat et al. (1997), and
Wissuwa and Ae (2001)
Triticale Oracka and Lapinski (2006)
Wheat Batten (1986), Batten and Khan (1987), Cao et al. (2009),
Fageria and Baligar (1999), Gill et al. (2004), Górny and
Garczy nski (2008), Gunes et al. (2006), Jones et al. (1989,
1992), Korkmaz et al. (2009), Manske et al. (2001, 2002),
Osborne and Rengel (2002a,b), Ozturk et al. (2005),
Sepehr et al. (2009), Su et al. (2006, 2009), Wang et al.
(2005), and Yaseen and Malhi (2009a)
Soybean Furlani et al. (2002), Kakar et al. (2002), Li et al. (2005), and
Zhang et al. (2009)
modern varieties (MVs) of rice typically had higher PERs than tradi-
tional or landrace cultivars in an upland (aerobic) screening trial: how-
ever, the PER was significantly correlated to harvest index (HI) such
that the low PER of many traditional varieties simply reflected their
much lower HI and grain yield potential (T. Rose and M. Wissuwa,
unpublished data). Similarly, studies with wheat also concluded that
PER was more highly correlated to HI than with grain yield (Batten
and Khan, 1987). Perhaps the most telling study was that of Manske et al.
(2002): these authors studied a number of P-efficiency traits in near
isogenic wheat lines differing in dwarfing characteristics and showed
that the yield-increasing effect of two dwarfing genes led to higher
PER. Thus, the apparent improvements in PER in MVs often reflect
their higher HI rather than specific selection of physiological traits that
improve internal P utilization in plants.
Calculating the tissue P concentration required for a given percentage of
maximal grain yield, referred to as the “critical” tissue P concentration if this
is 90% of maximal yield (White and Hammond, 2008), avoids the issue of
confounding P efficiency with HI because HI becomes inconsequential
when the yield of a genotype is expressed as a percentage of its own
maximum yield. However, this criterion may only be useful when all
genotypes screened have a similar yield potential: comparing the critical
tissue P concentration of a landrace with low yield potential (e.g., 4tha1)
to that of a high-yielding MV (e.g., 8tha1) is likely to produce misleading
results that may favor low-yielding genotypes.
A further issue with using grain yield as a parameter in defining internal
PUE, either as PER or as “critical” tissue P concentration, is that yield
formation not only depends on the yield potential of a given genotype
but also equally on location-specific effects such as day length, tempera-
ture, or disease pressure. Thus, while critical tissue P concentrations and
PER could be used in breeding nurseries to evaluate locally adapted high-
yielding breeding lines, neither criterion would be suitable to evaluate P
efficiency across a broad range of genotypes. Further, it remains question-
able whether sufficient genetic variation exists for these P-efficiency traits
among a set of rather similar breeding lines that have typically been
developed in well-fertilized nurseries (Wissuwa et al., 2009). Ideally,
screening studies investigating internal P utilization traits should maximize
the portion of the gene pool assessed within a crop species in search of
novel genes/alleles, many of which may only be present in older varieties
that were developed prior to the widespread use of P fertilizers. To enable
valid comparisons of older varieties and MVs, and to investigate physio-
logical mechanisms, it would be preferable to define PUE as the biomass
produced per unit P accumulated in tissue (gDMmg1 P) and to dissect
PUE into components such as grain PUE, shoot PUE, and root PUE and
look to improve them individually.
192 Terry J. Rose and Matthias Wissuwa
Table 4 Correlation coefficients between shoot P content, biomass, and shoot PUE
from four experiments conducted over a range of P supply from deficient to fully
fertilized
2.5
y = 2.31x −0.49
R 2 = 0.96
PUE shoot (g DM mg P-1) 2.0
1.5
1.0
0.5
0.0
0 10 20 30 40
Shoot P content (mg plant −1)
Figure 1 Shoot PUE as affected by shoot P content over a wide range of P supply
levels. Data presented are based on experimental data of Wissuwa et al. (2005). Briefly,
the experiment compared biomass accumulation of two closely related genotypes
differing for the presence of the Pup1 locus in a nutrient solution experiment under
four levels of P supply (3.2, 6.4, 9.6, and 100mMP). Genotypic differences were not
significant, so data from both genotypes were combined. A power function provided
the best fit to the experimental data.
4 10
4
3
2
PUE (mg P g–1 DM)
0
0 1 2 3 4 5 6
–1
P content (mg plant )
R 2 = 0.64
0
0 1 2 3 4 5 6
P content (mg plant–1)
Shoot
biomass
Assimilate distribution
PUE
tx : +20%
tx+1: +27.2%
tx+2: +29.8%
70%
Root:shoot
P uptake
P distribution
30% tx : -
tx+1: +6%
tx+2: +8.2%
PAE PUE
Soil Root
Architecture
exploration biomass
tx : +20%
tx+1: +20% +(6*1.2)%
tx+2: +20% +(8.2*1.2)%
would suggest because increased PUE would also enhance P uptake (Rose
et al., 2011; Wissuwa, 2003). Small genotypic differences in PUE as detected
in our modified screening experiment (Rose et al., 2011) could therefore be
well worth exploiting in plant breeding, at least in P-deficient scenarios where
growth is limited directly by insufficient P availability rather than by assimilate
supply (Wissuwa et al., 2005).
In addition to the obvious reason for focusing on PUE, the efficient use
of a limited and costly resource, improving PUE rather than PAE alone
may be more sustainable since relying solely on increasing P uptake would
pose the danger of overmining soil-P reserves, particularly in low or zero
input agricultural systems (Henry et al., 2010). The theoretical considera-
tions put forward in Fig. 3, as well as additional experimental data on low-
cost root systems and their effect on soil exploration (Zhu and Lynch,
2004), indicate that a strict separation of PUE and P uptake may be
impossible since enhanced PUE, especially in roots, would also enhance
P uptake and mining of soil-P.
Thus, improving grain PUE is theoretically the only option for reducing
the amount of P taken up by crops in P-deficient scenarios, with the added
benefit of minimizing the removal of P from fields, which would be beneficial
in low-input and high-input cropping systems. However, any potential
reductions in P uptake by crops with high grain PUE will depend on whether
a reduction in competition for P between the grain and vegetative tissues
triggers a decline or cessation in P uptake from soil. Research to date has
produced no firm conclusions: for example, early work suggested that wheat
crops acquire sufficient P for maximum grain yields by anthesis (Batten et al.,
1986; Boatwright and Viets, 1966), yet studies have shown that wheat plants
may cease (net) soil uptake at anthesis, accumulate small additional amounts of
P beyond anthesis, or take up a substantial proportion of total plant P after
anthesis (Manske et al., 2002; Mohamed and Marshall, 1979; Römer and
Schilling, 1986; Rose et al., 2007) and this pattern is genotype specific ( Jones
et al., 1992). Indeterminate crops such as canola, or crops that are partially
perennial such as rice, continue to acquire P from the soil during grain filling
to satisfy the competing demands of actively growing vegetative tissue and the
developing grain (Rose et al., 2009, 2010).
Thus, it is unclear whether determinate or indeterminate crop species (or
genotypes within crop species) would reduce their net P uptake if postanthesis
P demand was reduced. Whether regulatory mechanisms exist that reduce P
uptake from soils when P demand has been satisfied during postanthesis
growth is therefore a key question that remains to be addressed in future
research. Interestingly, among the suite of mutants with reduced phytic acid
levels in grains (lpa mutants) developed over the past decade, one barley lpa
mutant also has lower total grain P (Raboy, 2009). However, to the best of
our knowledge, it is not known whether this mutant has lower postanthesis P
uptake from soil compared to wild-type barley plants.
Internal Phosphorus Utilization Efficiency in Grain Crops 205
overexpression of genes related to the synthesis of organic acids has not yet
proved to be a successful strategy for increasing the PAE of plants (summar-
ized in Richardson et al., 2011).
Similarly, the known genetic and physiological responses of plant to
P-deficiency stress related to PUE have yet to have made the leap from
the laboratory to the field. As yet, we are unaware of any studies that
have conclusively shown that the overexpression of any P-stress response
genes confers higher PUE in plants. Wang et al. (2009) overexpressed an
Arabidopsis purple acid phosphatase gene (AtPAP15) in soybean and
reported higher yields in the transgenic plants when grown on a low-P
acid soil, and concluded that higher internal PUE was the likely cause.
However, no data on P uptake by plants or tissue P concentrations were
presented, and in earlier experiments, the transgenic plants showed
greater phytase secretion from roots and enhanced P acquisition from
organic P sources; hence, improved P uptake may have been the causal
factor in yield differences.
The lack of progress in improving P efficiency through manipulation
of genes that are responsive to P stress suggests that in many instances
P-stress response mechanisms are not necessarily those that confer toler-
ance to P stress. A microarray study comparing a P-efficient rice near
isogenic line containing the Pup1 locus with its P-inefficient parent
Nipponbare came to similar conclusions with regard to many PUE stress
response genes (Pariasca-Tanaka et al., 2009). While many of the afore-
mentioned PUE-related genes were upregulated in root and shoot tissue,
virtually all were induced to a similar or higher degree in the P-inefficient
Nipponbare than the P-efficient NIL6-4 (Pariasca-Tanaka et al., 2009).
Thus, even genotypes that show poor tolerance to P deficiency in the field
have evolved many of the well-known adaptations to P stress. Questions
remain as to whether the manipulation of P-stress response genes using
biotechnology can facilitate improvements in the PUE of field-grown
crops, but given the rate at which global P resources are being depleted,
further research is warranted.
In terms of breeding crops higher in PAE, the most promising option to
date has been to exploit natural variation among genotypes (Lynch, 2007;
Wissuwa et al., 2009). It is therefore surprising that although the physiolog-
ical and molecular responses to P stress have been widely studied, little is
known regarding mechanisms that confer higher PUE in genotypes of the
same species. While there still remains much to be elucidated about the
genetics and physiology of P-stress responses, further research also needs to
target the identification of valid QTLs from appropriate screening studies
to determine the genes and mechanisms that confer enhanced genotypic
PUE. Genes identified in such studies may prove useful in breeding pro-
grams but may also provide targets for genetic manipulation in addition to
the well-known genes that are affected by P-deficiency stress.
208 Terry J. Rose and Matthias Wissuwa
REFERENCES
Adu-Gyamfi, J. J., Fujita, K., and Ogata, S. (1989). Phosphorus absorption and utilization
efficiency of pigeon pea (Cajanus cajan (L) Millsp.) in relation to dry matter production
and dinitrogen fixation. Plant Soil 119, 315–324.
Ahmad, Z., Gill, M. A., and Qureshi, R. H. (2001). Genotypic variations of phosphorus
utilization efficiency of crops. J. Plant Nutrition 24, 1149–1171.
Akhtar, M. S., Oki, Y., and Adachi, T. (2008). Genetic variability in phosphorus acquisition
and utilization efficiency from sparingly soluble P-sources by Brassica cultivars under
P-stress environment. J. Agron. Crop Sci. 194, 380–392.
Akhtar, M. S., Oki, Y., and Adachi, T. (2009). Mobilization and acquisition of sparingly
soluble P-sources by Brassica cultivars under P-starved environment I. Differential
growth response, P-efficiency characteristics and P-remobilization. J. Integr. Plant Biol.
51, 1008–1023.
Araújo, A. P., Teixeira, M. G., and De Almeida, D. L. (1997). Phosphorus efficiency of wild
and cultivated genotypes of common bean (Phaseolus vulgaris L.) under biological nitro-
gen fixation. Soil Biol. Biochem. 29, 951–957.
Aziz, T., Gill, M. A., Rahmatullah, M. A., and Mansoor, T. (2005). Differences in
phosphorus absorption, transport and utilization by twenty rice (Oryza saliva L.) cultivars.
Pak. J. Agric. Sci. 42, 8–15.
Aziz, T., Rahmatullah, M. A., Tahir, M. A., Ahmad, I., and Cheema, M. A. (2006).
Phosphorus utilization by six Brassica cultivars (Brassica juncea L.) from tri-calcium
phosphate; a relatively insoluble compound. Pak. J. Bot. 38, 1529–1538.
Batjes, N. H. (1997). A world data set of derived soil properties by FAO-UNESCO soil unit
for global modeling. Soil Use Manag. 13, 9–16.
Batten, G. D. (1986). The uptake and utilization of phosphorus and nitrogen by diploid,
tetraploid and hexaploid wheats (Triticum spp.). Ann. Bot. 58, 49–59.
Batten, G. D. (1992). A review of phosphorus efficiency in wheat. Plant Soil 146, 163–168.
Batten, G. D., and Khan, M. A. (1987). Uptake and utilisation of phosphorus and nitrogen
by bread wheats grown under natural rainfall. Aust. J. Exp. Agric. 27, 405–410.
212 Terry J. Rose and Matthias Wissuwa
Batten, G. D., Wardlaw, I. F., and Aston, M. J. (1986). Growth and the distribution of
phosphorus in wheat developed under various phosphorus and temperature regimes.
Aust. J. Agric. Res. 37, 459–469.
Bentsink, L., Yuan, K., Koornneef, M., and Vreugdenhil, D. (2003). The genetics of phytate
and phosphate accumulation in seeds and leaves of Arabidopsis thaliana, using natural
variation. Theor. Appl. Genet. 106, 1234–1243.
Boatwright, G. O., and Viets, J. F. G. (1966). Phosphorus absorption during various growth
stages of spring wheat and intermediate wheatgrass. Agron. J. 58, 185–188.
Calderón-Vázquez, C., Ibarra-Laclette, E., Caballero-Perez, J., and Herrera-Estrella, L.
(2008). Transcript profiling of Zea mays roots reveals gene responses to phosphate
deficiency at the plant- and species-specific levels. J. Exp. Bot. 59, 2479–2497.
Calderón-Vázquez, C., Sawers, R. J. H., and Herrera-Estrella, L. (2011). Phosphate depri-
vation in maize: Genetics and genomics. Plant Physiol. 156, 1067–1077.
Cao, H. X., Zhang, Z. B., Sun, C. X., Shao, H. B., Song, W. Y., and Xu, P. (2009).
Chromosomal location of traits associated with wheat seedling water and phosphorus use
efficiency under different water and phosphorus stresses. Int. J. Mol. Sci. 10, 4116–4136.
Chen, J., Xu, L., Cai, Y., and Xu, J. (2009). Identification of QTLs for phosphorus
utilization efficiency in maize (Zea mays L.) across P levels. Euphytica 167, 245–252.
Cordell, D., Drangert, J., and White, S. (2009). The story of phosphorus: Global food
security and food for thought. Glob. Environ. Chang. 19, 292–305.
Corrales, I., Amenós, M., Poschenreider, C., and Barceló, J. (2007). Phosphorus efficiency
and root exudates in two contrasting tropical maize varieties. J. Plant Nutr. 30, 887–900.
Denton, M. D., Veneklaas, E. J., Freimoser, F. M., and Lambers, H. (2007). Banksia species
(Proteaceae) from severely phosphorus-impoverished soils exhibit extreme efficiency in
the use and re-mobilisation of phosphorus. Plant Cell Environ. 30, 1557–1565.
Ding, G., Yang, M., Hu, Y., Liao, Y., Shi, L., Xu, F., and Meng, J. (2010). Quantitative trait
loci affecting seed mineral concentrations in Brassica napus grown with contrasting
phosphorus supplies. Ann. Bot. 105, 1221–1234.
Dobermann, A., and Fairhurst, T. H. (2000). Nutrient Disorders and Nutrient Management.
Potash and Phosphate Institute of Canada and International Rice Research Institute,
Singapore.
Duan, H. Y., Shi, L., Ye, X. S., Wang, Y. H., and Xu, F. S. (2009). Identification of
phosphorous efficient germplasm in oilseed rape. J. Plant Nutr. 32, 1148–1163.
Fageria, N. K., and Baligar, V. C. (1997a). Phosphorus-use efficiency by corn genotypes.
J. Plant Nutr. 20, 1267–1277.
Fageria, N. K., and Baligar, V. C. (1997b). Upland rice genotypes evaluation for phosphorus
use efficiency. J. Plant Nutr. 20, 499–509.
Fageria, N. K., and Baligar, V. C. (1999). Phosphorus-use efficiency in wheat genotypes.
J. Plant Nutr. 22, 331–340.
Fageria, N. K., and da Costa, J. G. C. (2000). Evaluation of common bean genotypes for
phosphorus use efficiency. J. Plant Nutr. 23, 1145–1152.
Fageria, N. K., Wright, R. J., and Baligar, V. C. (1988). Rice cultivar evaluation for
phosphorus use efficiency. Plant Soil 111, 105–109.
Furlani, A. M. C., Furlani, P. R., Tanaka, R. T., Mascarenhas, H. A. A., and
Delgado, M. D. P. (2002). Variability of soybean germplasm in relation to phosphorus
uptake and use efficiency. Sci. Agric. 59, 529–536.
Garcia, R. L., and Hanway, J. J. (1976). Foliar fertilization of soybean during the seed-filling
period. Agron. J. 68, 653–657.
George, T. S., Richardson, A. E., Smith, J. B., Hadobas, P. A., and Simpson, R. J. (2005).
Limitations to the potential of transgenic Trifolium subterraneum L. plants that exude
phytase, when grown in soils with a range of organic phosphorus content. Plant Soil
278, 263–274.
Internal Phosphorus Utilization Efficiency in Grain Crops 213
Ghandilyan, A., Ilk, N., Hanhart, C., Mbengue, M., Barboza, L., Schat, H., Koornneef, M.,
El-Lithy, M., Vreugdenhil, D., Reymond, M., and Aarts, M. G. M. (2009). A strong
effect of growth medium and organ type on the identification of QTLs for phytate and
mineral concentrations in three Arabidopsis thaliana RIL populations. J. Exp. Bot. 60,
1409–1425.
Gill, M. A., Mansoor, S., Aziz, T., Rahmatullah, M. A., and Akhtar, M. S. (2002).
Differential growth response and phosphorus utilization efficiency of rice genotypes.
Pak. J. Agric. Sci. 39, 83–87.
Gill, H. S., Singh, A., Sethi, S. K., and Behl, R. K. (2004). Phosphorus uptake and use efficiency
in different varieties of bread wheat (Triticum aestivum L.). Arch. Agron. Soil Sci. 56, 563–572.
Górny, A. G. (1999). Inheritance of nitrogen and phosphorus utilization efficiency in spring
barley at the vegetative growth stages under high and low nutrition. Plant Breed. 118,
511–516.
Górny, A. G., and Garczy nski, S. (2008). Nitrogen and phosphorus efficiency in wild and
cultivated species of wheat. J. Plant Nutr. 31, 263–279.
Gourlay, C. J. P., Allan, D. L., and Russelle, M. P. (1994). Plant nutrient efficiency:
A comparison of definitions and suggested improvement. Plant Soil 158, 29–37.
Gunes, A., Inal, A., Alpaslan, M., and Cakmak, I. (2006). Genotypic variation in phosphorus
efficiency between wheat cultivars grown under greenhouse and field conditions. Soil Sci.
Plant Nutr. 52, 470–478.
Hafeez, F., Aziz, T., Maqsood, M. A., Ahmed, M., and Farooq, M. (2010). Differences in
rice cultivars for growth and phosphorus acquisition from rock phosphate and mono-
ammonium phosphate sources. Int. J. Agric. Biol. 12, 907–910.
Hammond, J. P., Bennett, M. J., Bowen, H. C., Broadley, M. R., Eastwood, D. C.,
May, S. T., Rahn, C., Swarup, R., Woolaway, K. E., and White, P. J. (2003). Changes
in gene expression in Arabidopsis shoots during phosphate starvation and the potential for
developing smart plants. Plant Physiol. 132, 578–596.
Hammond, J. P., Broadley, M. R., and White, P. J. (2004). Genetic responses to phosphorus
deficiency. Ann. Bot. 94, 323–332.
Hammond, J. P., Broadley, M. R., White, P. J., King, G. J., Bowen, H. C., Hayden, R.,
Meacham, M. C., Mead, A., Overs, T., Spracklen, W. P., and Greenwood, D. J. (2009).
Shoot yield drives phosphorus use efficiency in Brassica oleracea and correlates with root
architecture traits. J. Exp. Bot. 60, 1953–1968.
Hedley, M. J., Kirk, G. J. D., and Santos, M. B. (1994). Phosphorus efficiency and the forms
of soil phosphorus utilized by upland rice cultivars. Plant Soil 158, 53–62.
Henry, A., Chaves, N. F., Kleinman, P. J. A., and Lynch, J. P. (2010). Will nutrient-efficient
genotypes mine the soil? Effects of genetic differences in root architecture in common
bean (Phaseolus vulgaris L.) on soil phosphorus depletion in a low-input agro-ecosystem in
Central America. Field Crop Res. 115, 67–78.
Hinsinger, P. (2001). Bio-availability of soil inorganic P in the rhizosphere as affected by root
induced chemical changes: A review. Plant Soil 237, 173–195.
Holloway, B., Bertrand, I., Frischke, A. J., Brace, D., McLaughlin, M., and Shepperd, W.
(2001). Improving fertiliser efficiency on calcareous and alkaline soils with fluid sources
of P, N and Zn. Plant Soil 236, 209–219.
Ismail, A. M., Heuer, S., Thomson, J. T., and Wissuwa, M. (2007). Genetic and genomic
approaches to develop rice germplasm for problem soils. Plant Mol. Biol. 65, 547–570.
Jones, G. P. D., Blair, G. J., and Jessop, R. S. (1989). Phosphorus efficiency in wheat:
A useful selection criterion? Field Crop Res. 21, 257–264.
Jones, G. P. D., Jessop, R. S., and Blair, G. J. (1992). Alternative methods for the selection of
phosphorus efficiency in wheat. Field Crop Res. 30, 29–40.
Kakar, K. M., Tariq, M., Taj, F. H., and Nawab, K. (2002). Phosphorus use efficiency
of soybean as affected by phosphorus application and inoculation. Pak. J. Agron. 1, 49–50.
214 Terry J. Rose and Matthias Wissuwa
Korkmaz, K., Ibrikci, H., Karnez, E., Buyuk, G., Ryan, J., Ulger, A. C., and Oguz, H.
(2009). Phosphorus use efficiency of wheat genotypes grown in calcareous soils. J. Plant
Nutr. 32, 2094–2106.
Lambers, H., Shane, M. W., Cramer, M. D., Pearse, S. J., and Veneklaas, E. J. (2006). Root
structure and functioning for efficient acquisition of phosphorus: Matching morphologi-
cal and physiological traits. Ann. Bot. 98, 693–713.
Li, Y. D., Wang, Y. J., Tong, Y. P., Gao, J. G., and Zhang, J. S. (2005). QTL mapping of
phosphorus deficiency tolerance in soybean (Glycine max L. Merr.). Euphytica 142, 137–142.
Li, M., Welti, R., and Wang, X. (2006). Quantitative profiling of Arabidopsis polar
glycerolipids in response to phosphorus starvation: Roles of phospholipases D z1 and D
z2 in phosphatidylcholine hydrolysis and digalactosyldiacylglycerol accumulation in
phosphorus-starved plants. Plant Physiol. 142, 750–761.
Lott, J. N. A., Ockenden, I., Raboy, V., and Batten, G. D. (2000). Phytic acid and
phosphorus in crop seeds and fruits: A global estimate. Seed Sci. Res. 10, 11–33.
Lott, J. N. A., Ockenden, I., Raboy, V., and Batten, G. D. (2002). A global estimate of
phytic acid and phosphorus in crop grains, seeds, and fruits. In “Food Phytates”
(N. R. Reddy and S. K. Sathe, Eds.), pp. 7–24. CRC Press, Boca Raton.
Lott, J. N. A., Bojarski, M., Kolasa, J., Batten, G. D., and Campbell, L. C. (2009). A review
of the phosphorus content of dry cereal and legume crops of the world. Int. J. Agric.
Resour. Gov. Ecol. 8, 351–370.
Lynch, J. P. (2007). Roots of the second green revolution. Aust. J. Bot. 55, 493–512.
Lynch, J. P., and Ho, M. D. (2005). Rhizoeconomics: Carbon costs of phosphorus acquisi-
tion. Plant Soil 269, 45–56.
Ma, Q., Rengel, Z., and Rose, T. J. (2009). The effectiveness of deep placement of fertilisers
is determined by crop species and edaphic conditions in Mediterranean-type environ-
ments: A review. Aust. J. Soil Res. 47, 19–32.
MacDonald, G. K., Bennett, E. M., Philip, A., Potter, P. A., Ramankutty, N., and
Agronomic phosphorus imbalances across the world’s croplands (2011). Proc. Natl.
Acad. Sci. USA 108, 3086–3091. 10.1073/pnas.1010808108.
Manske, G. G. B., Ortiz-Monasterio, J. I., van Ginkel, M., Gonzalez, R. M., Fischer, R. A.,
Rajaram, S., and Vlek, P. L. G. (2001). Importance of P uptake efficiency versus P utilization
for wheat yield in acid and calcareous soils in Mexico. Eur. J. Agron. 14, 261–274.
Manske, G. G. B., Ortiz-Monasterio, J. I., van Ginkel, M., Rajaram, S., and Vlek, P. L. G.
(2002). Phosphorus use efficiency in tall, semi-dwarf and dwarf near-isogenic lines of
spring wheat. Euphytica 125, 113–119.
Mohamed, G. E. S., and Marshall, C. (1979). The pattern of distribution of phosphorus and
dry matter with time in spring wheat. Ann. Bot. 44, 721–730.
Morcuende, R., Bari, R., Gibon, Y., Zheng, W., Pant, B. D., Blasing, O., Usadel, B.,
Czechowski, T., Mudvardi, M. K., Stitt, M., and Scheible, W. R. (2007). Genome-wide
reprogramming of metabolism and regulatory networks of Arabidopsis in response to
phosphorus. Plant Cell Environ. 30, 85–112.
Oracka, T., and Lapinski, B. (2006). Nitrogen and phosphorus uptake and utilization
efficiency in D(R) substitution lines of hexaploid triticale. Plant Breed. 125, 221–224.
Osborne, L. D., and Rengel, Z. (2002a). Screening cereals for genotypic variation in the
efficiency of phosphorus uptake and utilization. Aust. J. Agric. Res. 53, 295–303.
Osborne, L. D., and Rengel, Z. (2002b). Genotypic differences in wheat for uptake and
utilisation of P from iron phosphate. Aust. J. Agric. Res. 53, 837–844.
Ozturk, L., Eker, S., Torun, B., and Cakmak, I. (2005). Variation in phosphorus efficiency
among 73 bread and durum wheat genotypes grown in a phosphorus-deficient calcareous
soil. Plant Soil 269, 69–80.
Parentoni, S. N., and Júnior, C. L. S. (2008). Phosphorus acquisition and internal utilization
efficiency in tropical maize genotypes. Pesqui. Agropecu. Bras. 43, 893–901.
Internal Phosphorus Utilization Efficiency in Grain Crops 215
Pariasca-Tanaka, J., Satoh, K., Rose, T., Mauleon, R., and Wissuwa, M. (2009). Stress
response versus stress tolerance: A transcriptome analysis of two rice lines contrasting
in tolerance to phosphorus deficiency. Rice 2, 167–185.
Pratt, J., Boisson, A.-M., Gout, E., Bligny, R., Douce, R., and Aubert, S. (2009). Phosphate
(Pi) starvation effect on the cytosolic Pi concentration and Pi exchanges across the
tonoplast in plant cells: An in vivo 31P-nuclear magnetic resonance study using methyl-
phosphonate as a Pi analog. Plant Physiol. 151, 1646–1657.
Raboy, V. (2009). Approaches and challenges to engineering seed phytate and total phos-
phorus. Plant Sci. 177, 281–296.
Rae, A. L., Jarmey, J. M., Mudge, S. R., and Smith, F. W. (2004). Over-expression
of a high-affinity phosphate transporter in transgenic barley plants does not enhance
phosphate uptake rates. Funct. Plant Biol. 31, 141–148.
Raghothama, K. G., and Karthikeyan, A. S. (2005). Phosphate acquisition. Plant Soil 274,
37–49.
Ramaekers, L., Remans, R., Rao, I., Blair, M., and Vanderleyden, J. (2010). Strategies for
improving phosphorus acquisition efficiency of crop plants. Field Crop Res. 117, 169–176.
Reuter, D. J., and Robinson, J. B. (1997). Plant Analysis: An Interpretation Manual.
CSIRO, Collingwood, Victoria.
Richardson, A. E., Lynch, J. P., Ryan, P. R., Delhaize, E., Smith, F. A., Smith, S. E.,
Harvey, P. R., Ryan, M. H., Veneklaas, E. J., Lambers, H., Oberson, A.,
Culvenor, R. A., et al. (2011). Plant and microbial strategies to improve the phosphorus
efficiency of agriculture. Plant Soil 349, 121–156. 10.1007/s11104-011-0950-4.
Römer, W., and Schenk, H. (1998). Influence of genotype on phosphate uptake and
utilization efficiencies in spring barley. Eur. J. Agron. 8, 215–224.
Römer, W., and Schilling, G. (1986). Phosphorus requirements of the wheat plant in various
stages of its life cycle. Plant Soil 91, 221–229.
Rose, T. J., Rengel, Z., Ma, Q., and Bowden, J. W. (2007). Differential accumulation
patterns of phosphorus and potassium by canola cultivars compared to wheat. J. Plant
Nutr. Soil Sci. 170, 404–411.
Rose, T. J., Rengel, Z., Ma, Q., and Bowden, J. W. (2008). Post-flowering supply of P, but
not K, is required for maximum canola seed yields. Eur. J. Agron. 28, 371–379.
Rose, T. J., Rengel, Z., Ma, Q., and Bowden, J. W. (2009). Phosphorus accumulation
by field-grown canola crops and the potential for deep phosphorus placement in a
Mediterranean-type climate. Crop Pasture Sci. 60, 987–994.
Rose, T. J., Pariasca-Tanaka, J., Rose, M. T., Fukuta, Y., and Wissuwa, M. (2010).
Genotypic variation in grain phosphorus concentration; and opportunities to improve
P-use efficiency in rice. Field Crop Res. 119, 154–160.
Rose, T. J., Rose, M. T., Pariasca-Tanaka, J., Heuer, S., and Wissuwa, M. (2011). The
frustration with utilization: Why have improvements in internal phosphorus utilization
efficiency in crops remained so elusive? Front. Plant Nutr. 2, 1–5.
Sahrawat, K. L., Jones, M. P., and Diatta, S. (1997). Direct and residual fertilizer phosphorus
effects on yield and phosphorus efficiency of upland rice in an Ultisol. Nutr. Cycl.
Agroecosyst. 48, 209–215.
Saleque, M. A., Abedin, M. J., Panaullah, G. M., and Bhuiyan, N. I. (1998). Yield and
phosphorus efficiency of some lowland rice varieties at different levels of soil-available
phosphorus. Commun. Soil Sci. Plant Anal. 29, 2905–2916.
Sepehr, E., Malakouti, M. J., Kholdebarin, B., Samadi, A., and Karimian, N. (2009).
Genotypic variation in P efficiency of selected Iranian cereals in greenhouse experiment.
Int. J. Plant Prod. 3, 17–28.
Septiningsih, E. M., Pamplona, A. M., Sanchez, D. L., Neeraja, C. N., Vergara, G. V.,
Heuer, S., Ismail, A. M., and Mackill, D. J. (2009). Development of submergence-
tolerant rice cultivars: The Sub1 locus and beyond. Ann. Bot. 103, 151–160.
216 Terry J. Rose and Matthias Wissuwa
Sharpley, A. N., McDowell, R. W., and Kleinman, P. J. A. (2001). Phosphorus loss from
land to water: Integrating agricultural and environmental management. Plant Soil 237,
287–307.
Shen, J., Yuan, L., Zhang, J., Li, H., Bai, Z., Chen, X., Zhang, W., and Zhang, F. (2011).
Phosphorus dynamics: From soil to plant. Plant Physiol. 156, 997–1005.
Shenoy, V. V., and Kalagudi, G. M. (2005). Enhancing plant phosphorus use efficiency for
sustainable cropping. Biotechnol. Adv. 23, 501–513.
Siddiqi, M. Y., and Glass, A. D. (1981). Utilization index: A modified approach to the
estimation and comparison of nutrient utilization efficiency in plants. J. Plant Nutr. 4,
289–302.
Stangoulis, J. C. R., Huynh, B., Welch, R. M., Choi, E., and Graham, R. D. (2007).
Quantitative trait loci for phytate in rice grain and their relationship with grain micronu-
trient content. Euphytica 154, 289–294.
Stelling, D., Wang, S. H., and Römer, W. (1996). Efficiency in the use of phosphorus,
nitrogen and potassium in topless faba bean (Vicia faba L.)—Variability and inheritance.
Plant Breed. 115, 361–366.
Su, J. Y., Xiao, Y., Li, M., Liu, Q., Li, B., Tong, Y., Jia, J., and Li, Z. (2006). Mapping
QTLs for phosphorus-deficiency tolerance at wheat seedling stage. Plant Soil 281, 25–36.
Su, J. Y., Zheng, Q., Li, H. W., Li, B., Jing, R. L., Tong, Y. P., and Li, Z. S. (2009). Detection
of QTLs for phosphorus use efficiency in relation to agronomic performance of wheat
grown under phosphorus sufficient and limited conditions. Plant Sci. 176, 824–836.
Subbarao, G. V., Ae, N., and Otani, T. (1997). Genetic variation in acquisition, and
utilization of phosphorus from iron-bound phosphorus in pigeonpea. Soil Sci. Plant
Nutr. 43, 511–519.
Theodorou, M. E., and Plaxton, W. C. (1993). Metabolic adaptations of plant respiration to
nutritional phosphate deprivation. Plant Physiol. 101, 339–344.
Theodorou, M. E., and Plaxton, W. C. (1996). Purification and characterization of
pyrophosphate-dependent phosphofructokinase from phosphate-starved Brassica nigra
suspension cells. Plant Physiol. 112, 343–351.
Vance, C. P., Uhde-Stone, C., and Allan, D. L. (2003). Phosphorus acquisition and use: Critical
adaptations by plants securing a non-renewable resource. New Phytol. 157, 423–457.
Wang, Q. R., Li, J. Y., Li, Z. S., and Christie, P. (2005). Screening Chinese wheat
germplasm for phosphorus efficiency in calcareous soils. J. Plant Nutr. 28, 489–505.
Wang, X., Wang, Y., Tian, J., Lim, B. L., Yan, X., and Liao, H. (2009). Overexpressing
AtPAP15 enhances phosphorus efficiency in soybean. Plant Physiol. 151, 233–240.
Wang, X., Shen, J., and Liao, H. (2010). Acquisition or utilization, which is more critical for
enhancing phosphorus efficiency in modern crops? Plant Sci. 179, 302–306.
White, P. J., and Hammond, J. P. (2008). Phosphorus nutrition of terrestrial plants. In “The
Ecophysiology of Plant–Phosphorus Interactions” (P. J. White and J. P. Hammond,
Eds.), pp. 51–81. Springer, The Netherlands.
Wissuwa, M. (2003). How do plants achieve tolerance to phosphorus deficiency? Small
causes with big effects. Plant Physiol. 133, 1947–1958.
Wissuwa, M., and Ae, N. (1999). Molecular markers associated with phosphorus uptake and
internal phosphorus-use efficiency in rice. In “Plant Nutrition—Molecular Biology and
Genetics” (G. Gissel-Nielsen and A. Jensen, Eds.), pp. 433–439. Kluwer Academic
Publishers, The Netherlands.
Wissuwa, M., and Ae, N. (2001). Genotypic variation for tolerance to phosphorus defi-
ciency in rice and the potential for its exploitation in rice improvement. Plant Breed. 120,
43–48.
Wissuwa, M., Yano, M., and Ae, N. (1998). Mapping of QTLs for phosphorus-deficiency
tolerance in rice (Oryza sativa L.). Theor. Appl. Genet. 97, 777–783.
Internal Phosphorus Utilization Efficiency in Grain Crops 217
Wissuwa, M., Gamat, G., and Ismail, A. (2005). Is root growth under phosphorus deficiency
affected by source or sink limitations? J. Exp. Bot. 56, 1943–1950.
Wissuwa, M., Mazzola, M., and Picard, C. (2009). Novel approaches in plant breeding for
rhizosphere-related traits. Plant Soil 321, 409–430.
Yang, M., Ding, G., Shi, L., Xu, F., and Meng, J. (2011). Detection of QTL for phosphorus
efficiency at vegetative stage in Brassica napus. Plant Soil 339, 97–111.
Yaseen, M., and Malhi, S. S. (2009a). Variation in yield, phosphorus uptake, and physiolog-
ical efficiency of wheat genotypes at adequate and stress phosphorus levels in soil.
Commun. Soil Sci. Plant Anal. 40, 3104–3120.
Yaseen, M., and Malhi, S. S. (2009b). Differential growth response of wheat genotypes
to ammonium phosphate and rock phosphate phosphorus sources. J. Plant Nutr. 32,
410–432.
Zhang, D., Cheng, H., Geng, L. Y., Kan, G. Z., Cui, S. Y., Meng, Q. C., Gai, J. Y., and
Yu, D. Y. (2009). Detection of quantitative trait loci for phosphorus deficiency tolerance
at soybean seedling stage. Euphytica 167, 313–322.
Zhao, J., Jamar, D. C. L., Lou, P., Wang, Y., Wu, J., Wang, X., Bonnema, G.,
Koornneef, M., and Vreugdenhil, D. (2008). Quantitative trait loci analysis of phytate
and phosphate concentrations in seeds and leaves of Brassica rapa. Plant Cell Environ. 31,
887–900.
Zhu, J., and Lynch, J. P. (2004). The contribution of lateral rooting to phosphorus acquisi-
tion efficiency in maize (Zea mays) seedlings. Funct. Plant Biol. 31, 949–958.
C H A P T E R S I X
Computer Simulation
in Plant Breeding
Xin Li,* Chengsong Zhu,* Jiankang Wang,† and Jianming Yu*
Contents
1. Introduction 220
1.1. An urgent need in plant breeding 220
1.2. Computer simulation 221
1.3. Joining computer simulation with plant breeding 222
2. Applying Computer Simulation to Plant Breeding 224
2.1. Comparing breeding methods 225
2.2. Gene mapping 237
2.3. Gene network and genotype-by-environment interaction 246
2.4. Crop physiology and crop modeling 247
3. Computation and Software Issues 254
4. Summary and Perspectives 255
Acknowledgments 257
References 257
Abstract
As a bridge between theory and experimentation, computer simulation has
become a powerful tool in scientific research, providing not only preliminary
validation of theories but also guidelines for empirical experiments. Plant
breeding focuses on developing superior genotypes with available genetic
and nongenetic resources, and improved plant-breeding methods maximize
genetic gain and cost-effectiveness. Computer simulation can lay out the
breeding process in silico and identify optimal candidates for various scenarios;
empirical validation can then follow. Insights gained from empirical studies, in
turn, can be incorporated into computer simulations. In this review, we discuss
different applications of computer simulation in plant breeding. First, we briefly
summarize the history of plant breeding and computer simulation and how
computer simulation can facilitate the breeding process. Next, we partition the
utility of computer simulation into different research areas of plant breeding,
including breeding method comparison, genetic mapping, gene network and
219
220 Xin Li et al.
1. Introduction
1.1. An urgent need in plant breeding
1.1.1. Three eras in plant-breeding history
What is plant breeding? A definition in Bernardo (2002) captured its
essence: “Plant breeding is the science, art, and business of improving plants
for human benefit.” Plant breeding is a discipline initiated by our ancestors
tens of thousands years ago. Without any knowledge of biology or genetics,
our ancestors provided domesticated crops for us today based only on
phenotypic selection (PS), or “art.” But scientific elements are also involved
in this process because the superior characteristics can be inherited from
generation to generation. Crossing “good” by “good,” chances are we get
better progeny next season. Comparing the crops we have today with their
wild relatives, we can see what great progress our ancestors achieved, even
though it took a time as long as civilization itself.
In the early twentieth century, with the rediscovery of Mendel’s Law and
the milestone paper of R. A. Fisher, plant breeding entered the scientific era.
Geneticists began to investigate the underlying principles of plant improve-
ment, especially for quantitative traits, which are the major features of most
economically important characteristics such as yield, biotic and abiotic stress
resistance, and flowering time. Classical textbooks of Mather (1949), Allard
(1960), and Falconer (1960) provide systematic guidelines for plant breeding.
With scientific theories in hand, plant breeders made significant progress in
answering questions pertinent to the plant-breeding process: Which parents
should I use to get the most variance and simultaneously a higher mean value?
Which is the best, single, double, or triple cross? How many generations does
it take to get the desired cultivar or population? What about the selection
intensity? Early testing or late testing? How do I control genotype-by-
environment interaction (GEI)? The complexity of the mechanisms of quan-
titative traits, unsurprisingly, means that these questions have no absolute
answers, which is why plant breeding is often called “a numbers game.” The
more plants, the more likely we can find a superior genotype. But exact
answers do not exist; they depend on the type of crops, populations, and most
importantly, the genetics of the traits. Probabilities are intrinsic to all of the
elements controlling the final result.
The third era came along with the discovery of molecular markers.
Although phenotypic or biochemical markers can also be used indirectly
to select the desired traits, most of which are difficult or expensive to
Computer Simulation in Plant Breeding 221
measure; they are not diverse enough and are too tedious and expensive to
implement into plant-breeding practice. Molecular markers (AFLF, RFLP,
RAPD, SSR, SNP, etc.) are abundant in organisms and easy to apply to
selection and gene mapping (Mackay, 2001). Thanks to those molecular
makers, today we have a better understanding of the genetic mechanism
controlling many quantitative traits, both in humans and plants. At the
beginning of the second decade of the twenty-first century, we are
equipped with powerful tools such as the high-throughput sequencing
platform and bioinformatics to help us tap into research areas impossible
only decades ago.
1.2.3. An example
To illustrate the process and components involved in computer simulation,
we provide a simple computer simulation to sample progeny phenotypic
values (Fig. 1A). Two parental inbred lines are crossed to form segregating
populations (for simplicity, only F1 and F2 are shown). In a completely
additive genetic model, the phenotypic mean of the F1 generation should
be the mean of the two parental lines. In this model, what we can control
are the overall mean and genetic effects. Other nongenetic effects are
assumed to follow a normal distribution with a mean of zero and predefined
variance. Computer simulation can be used to randomly sample nongenetic
effects and add them into the genetic model. Other subsequent generations,
such as BC1, DH, or RIL, also can be generated when recombination is
simulated to control for genetic effects.
Thanks to this random sampling and repeat feature of computer simula-
tion, the models used to simulate a real system are good approximations
of what would happen in reality. That is also why plant breeding, which
is time-consuming and resource-intensive, can benefit from computer
simulation.
A B
P1 ⫻ P2
P1 F1 P2 F1
F2
Density
F3
F2 F4
F5
F6
F7
F8
Phenotype value
Computer simulation Plant breeding
C 90
Genotype value
80
(% of max)
70
H=0.1 ES
H=0.1 SSD
H=0.5 ES
H=0.5 SSD
60
50
0 100 200 300 400
Number of F2 plants
Comparative simulation of ES and SSD
Figure 1 Joining computer simulation with plant breeding. (A) Computer simulation
can be used to generate random numbers from the underlying distribution. One
random sample is an approximation of what would happen in reality. (B) The aim of
plant breeding is to create superior genotypes for new cultivars. The more plants are
available for selection, the more likely we are to obtain superior genotypes. A quantita-
tive assessment can be very helpful for directing plant breeding. (C) Applying computer
simulation to plant breeding can integrate the number of plants, heritability, number of
genes, selection method and intensity, and number of generations to predict possible
outcomes of a breeding program and help us find the optimal combination of these
factors to accelerate the breeding process. The graph in (C) is based on data from
Vanoeveren and Stam (1992).
Breeding method
• Compare breeding strategy
• Assess factors influencing
marker-assisted selection
Gene mapping
Crop modeling
• Assess factors influencing Computer • Use gene information as
mapping power simulation in model parameters
• Determine significance
plant breeding • Predict crop performance in
threshold and confidence
target environments
interval of QTL position
marker effects was less than using individual effects. With an effective
population size of about 60, selecting large effect markers could not gener-
ate significant genetic drift, so crossing the selected lines to regenerate LD is
not effective in producing further genetic response. Zhang and Smith
suggested that the issue of decreased LD with selection could be solved by
identifying markers that are themselves the target QTL.
Population size was important in determining the efficiency of MAS
(Gimelfarb and Lande, 1994) for two possible reasons. First, with a large
population, the estimation of QTL effects is more accurate; second, finding
the desirable genotype is more likely with a large population. For a popula-
tion of at least 100 individuals of each sex, MAS is most efficient if the
phenotype does not contribute to the selection index. When several QTLs
are linked in coupling phase, the regression method may detect the QTLs as
a block, but when QTLs are linked in repulsion, the combined effects of
these QTLs decrease, making detection of the underlying QTLs difficult.
The optimal number of markers depends on the degree of LD between the
marker and QTL. LD is determined by the average number of recombinations
per generation in that region of the genome, the number of generations since
the original mutation, and the population size (Mackay, 2001). If LD is large,
more markers will not necessarily lead to a higher selection response. This
result occurs for two possible reasons. First, when LD is large, multiple makers
are in LD with each other, creating colinearity, meaning that their effects
overlap; second, covariance in these markers makes estimating QTL effects
inaccurate, thus decreasing the efficiency of MAS. Increasing the number of
markers in a finite population could even result in spurious linkages between
markers and unlinked QTLs (Gimelfarb and Lande, 1995).
For most MAS experiments, the effect of QTL is fixed at the beginning
of a breeding program; however, because of recombination between
marker and QTL, epistasis, or gene-by-environment interaction, the con-
tribution of QTL to a specific trait is not necessarily the same over the entire
breeding process. Some researchers (Gimelfarb and Lande, 1994; Peccoud
et al., 2004; Podlich et al., 2004) suggest reevaluating the QTL effect every
generation would be more efficient. This also highlights the dynamic
property of breeding methods involving MAS.
Podlich et al. (2004) proposed a method called mapping as you go
(MAYG) to tackle the issue of context dependency. Because plant breeding
is a dynamic process, the QTL values used in MAS would change from
generation to generation (or cycle to cycle) because of gene-by-gene and
gene-by-environment interaction. Computer simulation based on the E(N:
K) model (see Section 3) showed that compared with the mapping start only
(MSO) approach, which keeps the initial QTL values unchanged in the
breeding process, the MAYG method led to higher genetic gains in almost
all simulated scenarios. When an additive model was assumed (both E and K
are zero), these two methods do not differ significantly, but when epistasis
228 Xin Li et al.
or GEI effects occurred, the MAYG approach has the advantage because, in
different genetic backgrounds or different environments, the genetic effect
of a QTL can be distinct or even reversed. The MAYG approach can take
this into account and update the gene effect in every cycle or generation.
Thus, the gene effects remain relevant to the current breeding population.
A B
4.2 4.2 128 markers 48 DHs
128 markers 96 DHs
256 markers 48 DHs
256 markers 96 DHs
3.8 3.8
Response
Response
20 QTL MARS
3.4 20 QTL GS 3.4
100 QTL MARS
100 QTL GS
3.0 3.0
32 64 128 256 512 768 288 576 1152
Number of markers used Number of plants used
in two selection cycles in genomic selection
cross was made between adapted and exotic parents. Then multiple cycles of
recurrent selection were used to select superior lines. Surprisingly, simula-
tion showed that starting from an F2, not a BC1 or BC2 population, could
lead to higher selection response because this increases the frequency of
favorable alleles inherited from the adapted parent while maintaining a
moderate frequency of favorable alleles inherited from the exotic parent.
On the other hand, selection starting from a BC1 or BC2 population, which
has a lower frequency of exotic genes, could lose opportunities to incorpo-
rate favorable alleles from the exotic parents. As shown in another study
(Bernardo and Yu, 2007), responses increased as heritability increased, but
unlike the previous study, which showed that 144 individuals in cycle 0 was
sufficient for adapted by adapted cross, more selection response was
observed with larger populations for adapted by exotic cross.
Selection responses to GS for self-pollinated crops were 81–87% of the
corresponding responses to GS for cross-pollinated crops on an equal-time basis
(Bernardo, 2010). To achieve approximately the same response as GS in cross-
pollinated crops, the number of individuals that should be genotyped in self-
pollinated crops should be roughly twice as large as in cross-pollinated crops.
2.2.1.1. Population size and marker density Increasing sample size and
testing environments could make estimating QTL effects more accurate.
With large populations, more recombination events can be sampled, thus
increasing mapping power. With small sample sizes, the QTL effects are
inflated and mapping results are less likely to be repeated in other environ-
ments or experiments (Beavis, 1998; Schon et al., 2004). A population of at
least 2000 individuals is necessary to map a QTL with a small additive effect
(variance explained is 1%; Ooijen, 1992). Ooijen’s (1992) simulation results
showed that a minimum population of 200 backcross or F2 individuals is
needed to detect a QTL that explains at least 5% of the total phenotypic
variance. Moreover, the larger the genetic effect of the QTL, the more
likely it can be located precisely.
In a classic computer simulation in QTL mapping study, Beavis (1998)
investigated the potential power, precision, and accuracy of QTL analysis
(Fig. 4). He used three factors in his simulation: number of QTLs (10 or 40),
phenotypic variability explained by these QTLs (30%, 63%, 95%), and
number of F2 progeny (100, 500, 1000). These 18 simulation combinations
were repeated 200 times each, so the total number of simulations was 3600.
Simulation results showed that identifying all QTLs was possible if the
number of QTLs was small and the population size was as large as 1000
(Fig. 4A). For more QTLs (e.g., 40), even with 1000 progeny and a
heritability of 0.95, detecting all the QTLs was impossible. With large
numbers of progeny and higher heritability, the standard error of the
estimated genetics effect of any QTL decreased. The average estimated
magnitudes of genetic effects associated with a correctly identified QTL
were greatly overestimated if only 100 progeny were evaluated, slightly
overestimated with 500 progeny, and fairly close to the actual magnitude
when 1000 progeny were evaluated (Fig. 4B).
Computer Simulation in Plant Breeding 239
100 2
10 QTL h = 0.3 2
2 10 QTL h = 0.3
10 QTL h = 0.95
60 2
40
1
20
0 0
Figure 4 Effects of heritability and sample size on the power, precision, and accuracy
of QTLs identified in F2 progeny with either 10 or 40 QTLs. (A) Power to detect
simulated QTLs given a¼0.25. Power is the proportion of QTLs detected among all
the simulated QTLs. (B) Differences between estimated and simulated additive genetic
effects and standard errors. The genetic effect of each QTL is the percentage of the
phenotypic variability that each contributes. Only additive genetic effects are simulated,
and all QTLs have equal genetic effects. This figure is based on data from Beavis (1998).
even in the absence of known coancestry, some parents may carry identical-in-
state alleles. Third, QTL-by-environment interaction may affect QTL substi-
tution effects in experiments conducted in different environments. Finally,
QTL-by-genetic background interaction may affect QTL substitution effects
in experiments conducted with different populations. These sources of differ-
ence not only explain problems with replication but also could explain the
discrepancies between mapping results and MAS results.
From these possible causes of unrepeatable results, we can logically
assume that more testing environments and multiple mapping populations
could increase mapping accuracy. Simulation and experimental results both
show that more samples are more important than more testing environ-
ments (Schon et al., 2004).
Meta-analysis combines multiple QTL studies to improve the estimation
of QTL effects and positions. Xu (1998) pointed out that using more than a
single family for QTL mapping would allow detection of some QTLs that
may not be segregating in a particular family. In addition, mapping popula-
tions from a pair of parents with opposite extreme values for the trait of
interest are commonly used for population development. This is a powerful
method of mapping a single trait, but it may not help in detecting QTLs
responsible for other traits. For multiple-family QTL mapping, a random
model approach that treats each effect of gene substitution as a random
variable is preferred.
from a normal distribution. This assumes no QTLs in the genome (the null
hypothesis of maximum likelihood test). If this process is repeated a large
number of times (e.g., 10,000), all these LOD scores can be used to generate
a distribution of test statistics under the null hypothesis. Subsequently, this
distribution can be used to determine the critical value.
The threshold deduced in this approach is case-sensitive, that is, a unique
character of the genetic model under study (Churchill and Doerge, 1994;
Van Ooijen, 1999). Churchill and Doerge (1994) suggested the permuta-
tion test to approximate the distribution of LOD statistics under the null
hypothesis. The genetic information is kept untapped, the trait values are
randomly assigned to the genotype without replacement, and the LOD
scores are calculated. This process is repeated 1000 times (or more, depend-
ing on the Type I error rate that investigators use; the lower the Type I error
rate, the more repetitions) to obtain the distribution of LOD scores.
Because the genetic maps and trait values are not altered by the shuffling
procedure, this distribution automatically takes into account the particular
characteristics of the experiment at hand.
To get the LOD score that can be applied to various mapping studies, Van
Ooijen (1999) used computer simulation to determine the cumulative distri-
bution function of the maximum LOD score under the null hypothesis that
no segregating QTL is present. The simulation was based on a diploid species
with one chromosome on which no QTL was segregating. The chromosome
map length was 50–250cM with a fully informative marker every 1cM. Four
different kinds of population (BC1, F2, RIL, FS [full sibling] family of a cross
between non-inbred genotypes of an outbreeding species) were used with a
population size of 100. The individuals were assigned a random phenotypic
value sampled from a normal distribution regardless of the genotype, that is,
no QTL. This method means calculating the LOD score, in effect, as if you
were looking up the test statistics distribution tables at the back of a statistics
textbook, and it needs less computation time.
The false discovery rate (FDR) is a concept usually confused with P-
value. P-value is used in hypothesis tests to control Type I error rates (i.e.,
rejecting the null hypothesis when it is true). FDR is a measure of the
quality of a mapping experiment. FDR is the number of false-positive
QTLs/number of total QTLs detected. In extreme cases, finding one
significant QTL that turned out to be a false positive gives an FDR of
100%, even though we may have used 0.01 as the P-value. Bernardo (2004)
used computer simulation to test the relationship between these two distinct
criteria for QTL detection. The results showed that a P-value of 0.05 led to
a low FDR when many QTLs (30 or 100) controlled the trait, but this
lower FDR was accompanied by a diminished power to detect QTLs.
Larger mapping populations led to both a lower FDR and an increased
power. To control false positives, Bernardo (2004) recommended a P-value
of 0.0001 to determine a QTL as true positive.
Computer Simulation in Plant Breeding 243
under multiple, diverse environments. Third, the hybrids and inbreds tested
typically represent a broader genetic background. Fourth, the data used for
in silico mapping are available at no extra cost. Yu et al. (2005) investigated
the statistical power of in silico mapping in a hybrid breeding program via
computer simulation. They found that a large sample size, high marker
density, high heritability, and small numbers of QTLs led to the highest
power for in silico mapping via a mixed-model approach, but a compromise
between the power of detecting QTL and FDR would be necessary.
of all the unknowns. Because the joint posterior distribution does not have a
standard form, MCMC samplers (e.g., Jibbs sampler) can generate samples
from the joint posterior distribution. Bayesian mapping can perform linkage
analysis with any number of marker loci, multiple trait loci, and multiple
genomic segments (Xu, 2010).
conditions (Keating et al., 2003; McCown et al., 1996). Asseng et al. (2003)
used a derivative of APSIM, APSIM-Nwheat, to investigate the relationship
between specific leaf area (SLA) and yield under the environmental condi-
tions of Australia. Simulation results showed that increasing SLA could
increase yield only on the sandy, low water-holding soils in a Mediterranean
environment when the N supply was unlimited. On loamy soils, the
increased SLA trait actually reduced grain yields.
Using sorghum as an example, Chapman et al. (2003) simulated the
selection for yield in three drought environment types in sorghum produc-
tion regions in Australia. In their simulation, 15 genes were assumed to have
an influence on yield. These 15 genes controlled transpiration efficiency
coefficient (TE), flowering time (FT), osmotic adjustment (OA), and stay
green (SG). APSIM generated the yield value for all 4235 genotypes in the
target environments, and QU-GENE monitored the selection response of
the 15 genes in different drought conditions in an S1 family recurrent
selection-breeding program (Fig. 5). In general, selection for higher values
Selection in MET
Soil
Gene number
Water
Gene effect
QU-GENE CO2
Epistasis Trait value APSIM
engine Nitrogen
Environment
Radiation
GEI
Temperature
for these four traits resulted in higher yield, but not in all environments.
This study demonstrated the advantages of combining a crop physiology
model with a breeding simulation program: to investigate selection response
in different environments. Without the biophysical crop simulation model,
and assuming that all the 15 genes have equal additive effects on yield in
different environments, each of these genes would be fixed at a similar rate.
The introduction of crop models has modified the relative yield value of the
different genes in the environment types and thus has influenced the rate
and timing of fixation of the favorable alleles for these four traits. Simulation
results showed that increasing the level of trait expression (by increasing the
number of favorable alleles for a trait) led to higher grain yield, except for trait
PH in the Severe-Terminal drought environment type. Higher values for the
PH and TE traits had more effect on yield as the environments changed from
Severe-Terminal stress to Mid-season stress to Mild-Terminal stress. On the
other hand, higher values of TE and OA had more effect on yield in the
Severe-Terminal drought environment type. This is a kind of epistasis effect,
which has important implications for MAS programs for the sequence of gene
fixation that can result in the highest selection response.
Moreover, crop simulation models can be used to characterize environ-
ments based on crop performance data by connecting GIS systems and crop
models (Chapman et al., 2000; Loffler et al., 2005). With 108 years of
weather and soil data from six locations in major sorghum growth regions
in Australia, Chapman et al. (2000) used a simulation model to characterize
the environments into three drought patterns: Mid-season, Mild-Terminal,
and Severe-Terminal. The frequency with which these three types of
environment occur is an approximation of TPE. In another later study,
based on geographic information for 50 years (1952–2002) for each U.S.
Corn Belt Township, six major environment classes (EC) were identified
(Loffler et al., 2005). This classification of maize growth environments,
rather than climate and soil data, can be useful in interpreting GEI and
predicting cultivar performance in TPE.
(IPCC) report projects that the global surface temperature is likely to rise a
further 1.1–6.4 C (2.0–11.5F) during the twenty-first century (Christensen
et al., 2007). How would the crops respond to this climate change, and what
kind of cultivars could adapt to the changing environment? These questions
are of great concern to plant breeders and policymakers because food
security is strategically important to society.
Climate change research in plant breeding is not intended to provide
information about how to improve the efficiency of plant breeding, but to
determine our direction. For example, the advanced global general circula-
tion models (GCMs) can assess the effects of different emissions on climate
change. These data can be used to determine the possible future importance
of specific traits in a geographic region (Masutomi et al., 2009). GIS
technologies could benefit plant breeding by improving our understanding
of crop production areas and environments at both the global and local
scales, providing insights into potential shifts in production environments
with changing climate and predicting the possible spread of epidemics
(Hodson and White, 2010). Modeling, using this kind of information,
would allow us to make early decisions and prepare for the likely influences
of climate change on crop growth.
Regression-based models and dynamic process-based models are used to
assess the influence of climate change on crop production. Masutomi et al.
(2009) used 49 GCM models to predict the impact of climate change on rice
production in Asia. A large number of GCMs were used to account for the
inherent process/parameter uncertainty in GCMs. Under all future climate
scenarios, assuming no technological development, average production fell
for 2020 and 2080. CO2 fertilization was offset by the detrimental effect of
climate change. Under the scenario with the lowest atmospheric CO2
concentration, however, the simulation showed a small decrease in produc-
tion and a much lower probability of production decrease. Technological
development, when included in assessing the effects of climate change,
influenced production by only 10%.
Zou and Zeng (2009) investigated the combined effects of climatic con-
ditions, CO2 concentration, and technology development on wheat produc-
tion in Europe from 2000 to 2080. Technology development refers to all
measures related to crop management (e.g., improved machinery, pesticides,
herbicides, etc., and agronomic knowledge of farmers) and breeding (devel-
opment of higher yielding varieties through improved stress resistance and/or
yield potential) that result in yield increase. The estimates suggested increases
in crop productivity ranging from 25% to 163% depending on time slice and
scenario compared to the baseline year of 2000. The predicted changes in
productivity were mostly due to the effects of technology development.
Surprisingly, the CO2 increase did not have pronounced effects on wheat
production, and the differences between regions were also not significant,
even though climate change might make the northern part of Europe more
254 Xin Li et al.
amenable to wheat growth than the south. Typically, the further the predic-
tions are in the future, the less accurate the estimate. Both these studies (Lobell
et al., 2008; Masutomi et al., 2009) used multiple GCMs (20 and 49, respec-
tively) to control the uncertainty in the estimations.
Lobell et al. (2008) showed that South Asia and Southern Africa would
suffer from climate change by 2030 based on different models of climate
change and crop response. Some priorities were set: Southern Africa: maize,
wheat, and sugarcane; Western Africa: yams and groundnut; and Sahel: wheat.
Commonly used variables for different runs include the number of QTLs and
markers, QTL position on the genetic map, genome size, the effect of each
QTL, population size, heritability, degree of epistasis, and environment types.
After meeting the computational challenge, researchers must analyze and
interpret computer output. As with statistical software, computer simulation
gives us only the numbers, but the numbers require interpretation to be
meaningful and practical. Computer simulation must be coupled with
human intelligence to be useful.
ACKNOWLEDGMENTS
This work was supported by the Agriculture and Food Research Initiative Competitive
Grant (2011-03587) from the USDA National Institute of Food and Agriculture, the Plant
Feedstock Genomics Program (DE-SC0002259) of the U.S. Department of Energy, and the
Plant Genome Program (DBI-0820610) of the National Science Foundation, the Targeted
Excellence Program of Kansas State University, and the Kansas State University Center for
Sorghum Improvement.
REFERENCES
Ahuja, I., de Vos, R. C. H., Bones, A. M., and Hall, R. D. (2010). Plant molecular stress
responses face climate change. Trends Plant Sci. 15(12), 664–674.
Allard, R. W. (1960). Principles of Plant Breeding. John Wiley and Sons Inc., New York.
Asseng, S., Turner, N. C., Botwright, T., and Condon, A. G. (2003). Evaluating the impact
of a trait for increased specific leaf area on wheat yields using a crop simulation model.
Agron. J. 95(1), 10–19.
Barczi, J. F., Rey, H., Caraglio, Y., De Reffye, P., Barthelemy, D., Dong, Q. X., et al.
(2008). AmapSim: A structural whole-plant simulator based on botanical knowledge and
designed to host external functional models. Ann. Bot. 101(8), 1125–1138.
Bauer, A. M., Hoti, F., Korff, M.v., Pillen, K., Leon, J., and Sillanpaa, M. J. (2009).
Advanced backcross-QTL analysis in spring barley (H. vulgare ssp. spontaneum) com-
paring a REML versus a bayesian model in multi-environmental field trials. Theor. Appl.
Genet. 119(1), 105–123.
Beavis, W. D. (1998). QTL analyses: Power, precision, and accuracy. In “Molecular Dissec-
tion of Complex Traits” (A. H. Paterson, Ed.), pp. 145–162. CRC press, New York.
Beavis, W. (1994). The power and deceit of QTL experiments: Lessons from comparative
QTL studies. In “Proceedings of the 49th Annual Corn and Sorghum Industry Research
Conference,” pp. 250–266.
Bernacchi, D., Beck-Bunn, T., Emmatty, D., Eshed, Y., Inai, S., Lopez, J., et al. (1998a).
Advanced backcross QTL analysis of tomato. II. Evaluation of near-isogenic lines
carrying single-donor introgressions for desirable wild QTL-alleles derived from lyco-
persicon hirsutum and L. pimpinellifolium. Theor. Appl. Genet. 97(1), 170–180.
Bernacchi, D., Beck-Bunn, T., Eshed, Y., Lopez, J., Petiard, V., Uhlig, J., et al. (1998b).
Advanced backcross QTL analysis in tomato. I. Identification of QTLs for traits of
agronomic importance from lycopersicon hirsutum. Theor. Appl. Genet. 97(3), 381–397.
Bernardo, R., and Yu, J. (2007). Prospects for genomewide selection for quantitative traits
in maize. Crop Sci. 47(3), 1082–1090.
Bernardo, R. (1999a). Marker-assisted best linear unbiased prediction of single-cross perfor-
mance. Crop Sci. 39(5), 1277–1282.
Bernardo, R. (1999b). Selection response with marker-based assortative mating. Crop Sci. 39
(1), 69–73.
Bernardo, R. (2001). What if we knew all the genes for a quantitative trait in hybrid crops?
Crop Sci. 41(1), 1–4.
Bernardo, R. (2002). Breeding for Quantitative Traits in Plants. Stemma Press, Woodbury,
MN.
Bernardo, R. (2003). Parental selection, number of breeding populations, and size of each
population in inbred development. Theor. Appl. Genet. 107(7), 1252–1256.
Bernardo, R. (2004). What proportion of declared QTL in plants are false? Theor. Appl.
Genet. 109(2), 419–424.
258 Xin Li et al.
Bernardo, R. (2008). Molecular markers and selection for complex traits in plants: Learning
from the last 20 years. Crop Sci. 48(5), 1649–1664.
Bernardo, R. (2009). Genomewide selection for rapid introgression of exotic germplasm in
maize. Crop Sci. 49(2), 419–425.
Bernardo, R. (2010). Genomewide selection with minimal crossing in self-pollinated crops.
Crop Sci. 50(2), 624–627.
Bernardo, R., and Charcosset, A. (2006). Usefulness of gene information in marker-assisted
recurrent selection: A simulation appraisal. Crop Sci. 46(2), 614–621.
Bernardo, R., Moreau, L., and Charcosset, A. (2006). Number and fitness of selected
individuals in marker-assisted and phenotypic recurrent selection. Crop Sci. 46(5), 1972–
1980.
Bogdan, M., Ghosh, J. K., and Doerge, R. W. (2004). Modifying the schwarz bayesian
information criterion to locate multiple interacting quantitative trait loci. Genetics 167(2),
989–999.
Botstein, D., White, R. L., Skolnick, M., and Davis, R. W. (1980). Construction of a
genetic-linkage map in man using restriction fragment length polymorphisms. Am. J.
Hum. Genet. 32(3), 314–331.
Chapman, S., Cooper, M., Podlich, D., and Hammer, G. (2003). Evaluating plant
breeding strategies by simulating gene action and dryland environment effects. Agron.
J. 95(1), 99–113.
Chapman, S. C. (2008). Use of crop models to understand genotype by environment
interactions for drought in real-world and simulated plant breeding trials. Euphytica 161
(1), 195–208.
Chapman, S. C., Cooper, M., Hammer, G. L., and Butler, D. G. (2000). Genotype by
environment interactions affecting grain sorghum. II. Frequencies of different seasonal
patterns of drought stress are related to location effects on hybrid yields. Aust. J. Agr. Res.
51(2), 209–221.
Chapman, S. C., Wang, J., Rebetzke, G. J., and Bonnett, D. G. (2007). Accounting for
variability in the detection and use of markers for simple and complex traits. In “Scale and
Complexity in Plant Systems Research: Gene-Plant-Crop Relations” ( J. H. J. Spiertx,
P. Struik, and H. H. van Laar, Eds.), pp. 37–44. Springer, Dordrecht, Netherlands.
Charmet, G., Robert, N., Perretant, M. R., Gay, G., Sourdille, P., Groos, C., et al. (1999).
Marker-assisted recurrent selection for cumulating additive and interactive QTLs in
recombinant inbred lines. Theor. Appl. Genet. 99(7–8), 1143–1148.
Chen, X., Zhao, F., and Xu, S. (2010). Mapping environment-specific quantitative trait loci.
Genetics 186(3), 1053–1066.
Chenu, K., Chapman, S. C., Tardieu, F., McLean, G., Welcker, C., and Hammer, G. L.
(2009). Simulating the yield impacts of organ-level quantitative trait loci associated with
drought response in maize: A “gene-to-phenotype” modeling approach. Genetics 183(4),
1507–1523.
Chesler, E. J., Rodriguez-Zas, S. L., and Mogil, J. S. (2001). In silico mapping of mouse
quantitative trait loci. Science 294(5551), 2423.
Christensen, J., Hewitson, B., Busuioc, A., Chen, A., Gao, X., Held, R., et al. (2007).
Regional climate projections. In “Climate change, 2007: The physical science basis.
Contribution of working group I to the fourth assessment report of the intergovernmen-
tal panel on climate change” (S. Solomon, D. Qni, M. Manning, Z. Chen, M. Marquis,
K. B. Averyt, M. Tignor, and H. L. Miller, Eds.), Cambridge University Press, Cam-
bridge, United Kingdom and New York, NY, USA.
Churchill, G. A., and Doerge, R. W. (1994). Empirical threshold values for quantitative trait
mapping. Genetics 138(3), 963–971.
Cooper, M., Chapman, S. C., Podlich, D. W., and Hammer, G. L. (2002). The GP
problem: Quantifying gene-to-phenotype relationships. In Silico Biol. 2(2), 151–164.
Computer Simulation in Plant Breeding 259
Cooper, M., Podlich, D. W., and Luo, L. (2007). Modeling QTL effects and MAS in plant
breeding. In “Genomics-Assisted Crop Improvement” (R. K. Varshney and R. Tuberosa,
Eds.), pp. 57–95. Springer, Dordrecht, The Netherlands.
Crepieux, S., Lebreton, C., Servin, B., and Charmet, G. (2004). Quantitative trait loci
(QTL) detection in multicross inbred designs: Recovering QTL identical-by-descent
status information from marker data. Genetics 168(3), 1737–1749.
Darvasi, A., and Soller, M. (1997). A simple method to calculate resolving power and
confidence interval of QTL map location. Behav. Genet. 27(2), 125–132.
Darvasi, A., Weinreb, A., Minke, V., Weller, J. I., and Soller, M. (1993). Detecting marker-
qtl linkage and estimating qtl gene effect and map location using a saturated genetic-map.
Genetics 134(3), 943–951.
de Dorlodot, S., Forster, B., Pagčs, L., Price, A., Tuberosa, R., and Draye, X. (2007). Root
system architecture: Opportunities and constraints for genetic improvement of crops.
Trends Plant Sci. 12(10), 474–481.
Doerge, R. W. (2002). Mapping and analysis of quantitative trait loci in experimental
populations. Nat. Rev. Genet. 3(1), 43–52.
Dupuy, L., Gregory, P. J., and Bengough, A. G. (2010). Root growth models: Towards a
new generation of continuous approaches. J. Exp. Bot. 61(8), 2131–2143.
Edwards, M. D., and Page, N. J. (1994). Evaluation of marker-assisted selection through
computer-simulation. Theor. Appl. Genet. 88(3–4), 376–382.
Evers, J. B., Vos, J., Yin, X., Romero, P., van der Putten, P. E. L., and Struik, P. C. (2010).
Simulation of wheat growth and development based on organ-level photosynthesis and
assimilate allocation. J. Exp. Bot. 61(8), 2203–2216.
Falconer, D. S. (1960). Introduction to Quantitative Genetics. Oliver and Boyd, London.
Frisch, M., Bohn, M., and Melchinger, A. (2000). Computer note. PLABSIM: Software for
simulation of marker-assisted backcrossing. J. Hered. 91(1), 86.
Frisch, M., Bohn, M., and Melchinger, A. E. (1999). Comparison of selection strategies for
marker-assisted backcrossing of a gene. Crop Sci. 39(5), 1295–1301.
Frisch, M., and Melchinger, A. E. (2001). Marker-assisted backcrossing for simultaneous
introgression of two genes. Crop Sci. 41(6), 1716–1725.
Frisch, M., and Melchinger, A. E. (2005). Selection theory for marker-assisted backcrossing.
Genetics 170(2), 909–917.
Gimelfarb, A., and Lande, R. (1994). Simulation of marker-assisted selection in hybrid
populations. Genet. Res. 63(1), 39–47.
Gimelfarb, A., and Lande, R. (1995). Marker-assisted selection and marker-qtl associations
in hybrid populations. Theor. Appl. Genet. 91(3), 522–528.
Grafahrend-Belau, E., Schreiber, F., Koschutzki, D., and Junker, B. H. (2009). Flux balance
analysis of barley seeds: A computational approach to study systemic properties of central
metabolism. Plant Physiol. 149(1), 585–598.
Grupe, A., Germer, S., Usuka, J., Aud, D., Belknap, J. K., Klein, R. F., et al. (2001). In silico
mapping of complex disease-related traits in mice. Science 292(5523), 1915–1918.
Hammer, G., Kropff, M., Sinclair, T., and Porter, J. (2002). Future contributions of crop
modelling—From heuristics and supporting decision making to understanding genetic
regulation and aiding crop improvement. Eur. J. Agron. 18(1–2), 15–31.
Han, L., and Xu, S. (2010). Genome-wide evaluation for quantitative trait loci under the
variance component model. Genetica 138(9–10), 1099–1109.
Hodson, D., and White, J. (2010). GIS and crop simulation modelling applications in climate
change research. In “Climate Change and Crop Production” (M. P. Reynolds, Ed.),
pp. 245–262. CABI, Oxfordshire, UK.
Holland, J. B. (2004). Implementation of molecular markers for quantitative traits in breeding
programs—challenges and opportunities. In “New Directions for a Diverse Planet: Pro-
ceedings for the 4th International Crop Science Congress, Brisbane, Australia, 26”.
260 Xin Li et al.
Hoogenboom, G., White, J. W., and Messina, C. D. (2004). From genome to crop:
Integration through simulation modeling. Field Crop Res 90(1), 145–163.
Hospital, F., and Charcosset, A. (1997). Marker-assisted introgression of quantitative trait
loci. Genetics 147(3), 1469–1485.
Hospital, F., Chevalet, C., and Mulsant, P. (1992). Using markers in gene introgression
breeding programs. Genetics 132(4), 1199–1210.
Hospital, F., Moreau, L., Lacoudre, F., Charcosset, A., and Gallais, A. (1997). More on the
efficiency of marker-assisted selection. Theor. Appl. Genet. 95(8), 1181–1189.
Hospital, F. (2001). Size of donor chromosome segments around introgressed loci and
reduction of linkage drag in marker-assisted backcross programs. Genetics 158(3),
1363–1379.
Jannink, J. L., and Jansen, R. (2001). Mapping epistatic quantitative trait loci with one-
dimensional genome searches. Genetics 157(1), 445–454.
Jannink, J. L., Lorenz, A. J., and Iwata, H. (2010). Genomic selection in plant breeding:
From theory to practice. Brief. Funct. Genomics 9(2), 166–177.
Jansen, R. C. (1993). Interval mapping of multiple quantitative trait loci. Genetics 135(1),
205–211.
Jansen, R. C., Jannink, J. L., and Beavis, W. D. (2003). Mapping quantitative trait loci in
plant breeding populations: Use of parental haplotype sharing. Crop Sci. 43, 829–834.
Jiang, C. J., and Zeng, Z. B. (1997). Mapping quantitative trait loci with dominant and
missing markers in various crosses from two inbred lines. Genetica 101(1), 47–58.
Johnson, R. (2004). Marker-assisted selection. Plant Breed. Rev. 24(1), 293–309.
Kang, M. S. (2001). Quantitative genetics, genomics and plant breeding. In “Quantitative
Genetics Genomics and Plant Breeding” (M. S. Kang, Ed.). xvi þ 400 pp. CABI,
Wallingford, UK.
Kao, C. H., Zeng, Z. B., and Teasdale, R. D. (1999). Multiple interval mapping for
quantitative trait loci. Genetics 152(3), 1203–1216.
Keating, B. A., Carberry, P., Hammer, G., Probert, M. E., Robertson, M., Holzworth, D.,
et al. (2003). An overview of APSIM, a model designed for farming systems simulation.
Eur. J. Agron. 18(3–4), 267–288.
Kuchel, H., Fox, R., Reinheimer, J., Mosionek, L., Willey, N., Bariana, H., et al. (2007).
The successful application of a marker-assisted wheat breeding strategy. Mol. Breed. 20(4),
295–308.
Kuchel, H., Ye, G., Fox, R., and Jefferies, S. (2005). Genetic and economic analysis of a
targeted marker-assisted wheat breeding strategy. Mol. Breed. 16(1), 67–78.
Lande, R., and Thompson, R. (1990). Efficiency of marker-assisted selection in the
improvement of quantitative traits. Genetics 124(3), 743–756.
Lander, E. S., and Botstein, D. (1989). Mapping mendelian factors underlying quantitative
traits using rflp linkage maps. Genetics 121(1), 185–199.
Letort, V., Mahe, P., Cournede, P. H., De Reffye, P., and Courtois, B. (2008). Quantitative
genetics and functional-structural plant growth models: Simulation of quantitative trait
loci detection for model parameters and application to potential yield optimization. Ann.
Bot. 101(8), 1243–1254.
Leuning, R., Kelliher, F., Pury, D., and Schulze, E. D. (1995). Leaf nitrogen, photosynthe-
sis, conductance and transpiration: Scaling from leaves to canopies. Plant Cell Environ. 18
(10), 1183–1200.
Li, H. H., Ye, G. Y., and Wang, J. K. (2007). A modified algorithm for the improvement of
composite interval mapping. Genetics 175(1), 361–374.
Lobell, D. B., Burke, M. B., Tebaldi, C., Mastrandrea, M. D., Falcon, W. P., and
Naylor, R. L. (2008). Prioritizing climate change adaptation needs for food security in
2030. Science 319(5863), 607–610.
Computer Simulation in Plant Breeding 261
Loffler, C. M., Wei, J., Fast, T., Gogerty, J., Langton, S., Bergman, M., et al. (2005).
Classification of maize environments using crop simulation and geographic information
systems. Crop Sci. 45(5), 1708–1716.
Lynch, M., and Walsh, B. (1997). Genetics and Analysis of Quantitative Traits. Sinauer
Associates, Inc., Sunderland, MA.
Mackay, T. F. C. (2001). The genetic architecture of quantitative traits. Annu. Rev. Genet.
35, 303–339.
Manichaikul, A., Moon, J. Y., Sen, S., Yandell, B. S., and Broman, K. W. (2009). A model
selection approach for the identification of quantitative trait loci in experimental crosses,
allowing epistasis. Genetics 181(3), 1077–1086.
Masutomi, Y., Takahashi, K., Harasawa, H., and Matsuoka, Y. (2009). Impact assessment of
climate change on rice production in Asia in comprehensive consideration of process/
parameter uncertainty in general circulation models. Agric. Ecosyst. Environ. 131(3–4),
281–291.
Mather, K. (1949). Biometrical Genetics. Methuen, London.
Maurer, H. P., Melchinger, A. E., and Frisch, M. (2008). Population genetic simulation and
data analysis with Plabsoft. Euphytica 161(1–2), 133–139.
Mayor, P. J., and Bernardo, R. (2009). Genomewide selection and marker-assisted recurrent
selection in doubled haploid versus F-2 populations. Crop Sci. 49(5), 1719–1725.
McCown, R., Hammer, G., Hargreaves, J., Holzworth, D., and Freebairn, D. (1996).
APSIM: A novel software system for model development, model testing and simulation
in agricultural systems research. Agric. Syst. 50(3), 255–271.
Messina, C., Jones, J., Boote, K., and Vallejos, C. (2006). A gene-based model to
simulate soybean development and yield responses to environment. Crop Sci. 46(1),
456–466.
Meuwissen, T. H. E., Karlsen, A., Lien, S., Olsaker, I., and Goddard, M. E. (2002). Fine
mapping of a quantitative trait locus for twinning rate using combined linkage and
linkage disequilibrium mapping. Genetics 161(1), 373–379.
Meuwissen, T. H. E., Hayes, B. J., and Goddard, M. E. (2001). Prediction of total genetic
value using genome-wide dense marker maps. Genetics 157(4), 1819–1829.
Myles, S., Peiffer, J., Brown, P. J., Ersoz, E. S., Zhang, Z. W., Costich, D. E., et al. (2009).
Association mapping: Critical considerations shift from genotyping to experimental
design. Plant Cell 21(8), 2194–2202.
Ooijen, J. W. (1992). Accuracy of mapping quantitative trait loci in autogamous species.
Theor. Appl. Genet. 84(7), 803–811.
Parisseaux, B., and Bernardo, R. (2004). In Silico Mapping of Quantitative Trait Loci in
Maize. Springer, Berlin/Heidelberg.
Peccoud, J., Vander Velden, K., Podlich, D., Winkler, C., Arthur, L., and Cooper, M.
(2004). The selective values of alleles in a molecular network model are context depen-
dent. Genetics 166(4), 1715–1725.
Peleman, J. D., and van der Voort, J. R. (2003). Breeding by design. Trends Plant Sci. 8(7),
330–334.
Pillen, K., Zacharias, A., and Leon, J. (2003). Advanced backcross QTL analysis in barley
(Hordeum vulgare L.). Theor. Appl. Genet. 107(2), 340–352.
Podlich, D., Cooper, M., Basford, K., and Geiger, H. (1999). Computer simulation of a
selection strategy to accommodate genotype environment interactions in a wheat recur-
rent selection programme. Plant Breed. 118(1), 17–28.
Podlich, D. W., and Cooper, M. (1998). QU-GENE: A simulation platform for quantitative
analysis of genetic models. Bioinformatics 14(7), 632–653.
Podlich, D. W., Winkler, C. R., and Cooper, M. (2004). Mapping as you go: An effective
approach for marker-assisted selection of complex traits. Crop Sci. 44(5), 1560–1571.
262 Xin Li et al.
Price, A. L., Patterson, N. J., Plenge, R. M., Weinblatt, M. E., Shadick, N. A., and
Reich, D. (2006). Principal components analysis corrects for stratification in genome-
wide association studies. Nat. Genet. 38(8), 904–909.
Pritchard, J. K., Stephens, M., Rosenberg, N. A., and Donnelly, P. (2000). Association
mapping in structured populations. Am. J. Hum. Genet. 67(1), 170–181.
Prusinkiewicz, P., Hammel, M., Hanan, J., and Mech, R. (1996). L-systems: From the
theory to visual models of plants. In “Proceedings of the Second CSIRO Symposium on
Computational Challenges in Life Sciences”.
Prusinkiewicz, P. (2004). Art and science for life: Designing and growing virtual plants
with L-systems. In “Nursery Crops Development, Evaluation, Production and Use”
(C. Davidson and T. Fernandez, Eds.), Vol. 630, pp. 15–28. ISHS, Leuven, Belgium.
Reymond, M., Muller, B., Leonardi, A., Charcosset, A., and Tardieu, F. (2003). Combining
quantitative trait loci analysis and an ecophysiological model to analyze the genetic
variability of the responses of maize leaf growth to temperature and water deficit. Plant
Physiol. 131(2), 664–675.
Ribaut, J., Jiang, C., and Hoisington, D. (2002). Simulation experiments on efficiencies of
gene introgression by backcrossing. Crop Sci. 42(2), 557–565.
Risch, N., and Merikangas, K. (1996). The future of genetic studies of complex human
diseases. Science 273(5281), 1516–1517.
Sahana, G., Guldbrandtsen, B., Janss, L., and Lund, M. S. (2010). Comparison of association
mapping methods in a complex pedigreed population. Genet. Epidemiol. 34(5), 455–462.
Satagopan, J. M., Yandell, Y. S., Newton, M. A., and Osborn, T. C. (1996). A Bayesian
approach to detect quantitative trait loci using Markov chain Monte Carlo. Genetics 144
(2), 805–816.
Schauer, N., Semel, Y., Roessner, U., Gur, A., Balbo, I., Carrari, F., et al. (2006).
Comprehensive metabolic profiling and phenotyping of interspecific introgression lines
for tomato improvement. Nat. Biotechnol. 24(4), 447–454.
Schon, C. C., Utz, H. F., Groh, S., Truberg, B., Openshaw, S., and Melchinger, A. E.
(2004). Quantitative trait locus mapping based on resampling in a vast maize testcross
experiment and its relevance to quantitative genetics for complex traits. Genetics 167(1),
485–498.
Slafer, G. A. (2003). Genetic basis of yield as viewed from a crop physiologist’s perspective.
Ann. Appl. Biol. 142(2), 117–128.
Soller, M., Brody, T., and Genizi, A. (1976). Power of experimental designs for detection of
linkage between marker loci and quantitative loci in crosses between inbred lines. Theor.
Appl. Genet. 47(1), 35–39.
Stich, B. (2009). Comparison of mating designs for establishing nested association mapping
populations in maize and Arabidopsis thaliana. Genetics 183(4), 1525–1534.
Takahashi, K., Ishikawa, N., Sadamoto, Y., Sasamoto, H., Ohta, S., Shiozawa, A., et al.
(2003). E-cell 2: Multi-platform E-cell simulation system. Bioinformatics 19(13), 1727–
1729.
Tanksley, S., and Nelson, J. (1996). Advanced backcross QTL analysis: A method for the
simultaneous discovery and transfer of valuable QTLs from unadapted germplasm into
elite breeding lines. Theor. Appl. Genet. 92(2), 191–203.
Tomita, M. (2001). Whole-cell simulation: A grand challenge of the 21st century. Trends
Biotechnol. 19(6), 205–210.
Tomita, M., Hashimoto, K., Takahashi, K., Shimizu, T. S., Matsuzaki, Y., Miyoshi, F., et al.
(1999). E-CELL: Software environment for whole-cell simulation. Bioinformatics 15(1),
72–84.
van Berloo, R., and Stam, P. (1998). Marker-assisted selection in autogamous RIL popula-
tions: A simulation study. Theor. Appl. Genet. 96(1), 147–154.
Computer Simulation in Plant Breeding 263
van Berloo, R., and Stam, P. (2001). Simultaneous marker-assisted selection for multiple
traits in autogamous crops. Theor. Appl. Genet. 102(6–7), 1107–1112.
Van Ooijen, J. W. (1999). LOD significance thresholds for QTL analysis in experimental
populations of diploid species. Heredity 83, 613–624.
Vanoeveren, A. J., and Stam, P. (1992). Comparative simulation studies on the effects of
selection for quantitative traits in autogamous crops—Early selection versus single seed
descent. Heredity 69, 342–351.
Vieland, V. J. (1998). Bayesian linkage analysis, or: How I learned to stop worrying and love
the posterior probability of linkage. Am. J. Hum. Genet. 63(4), 947–954.
Visscher, P. M., Haley, C. S., and Thompson, R. (1996a). Marker-assisted introgression in
backcross breeding programs. Genetics 144(4), 1923–1932.
Visscher, P. M., Thompson, R., and Haley, C. S. (1996b). Confidence intervals in QTL
mapping by bootstrapping. Genetics 143(2), 1013–1020.
Wang, J. K., Chapman, S. C., Bonnett, D. G., and Rebetzke, G. J. (2009a). Simultaneous
selection of major and minor genes: Use of QTL to increase selection efficiency of
coleoptile length of wheat (Triticum aestivum L.). Theor. Appl. Genet. 119(1), 65–74.
Wang, J. K., and Pfeiffer, W. H. (2007). Simulation modeling in plant breeding: Principles
and applications. Agric. Sci. China 6(8), 908–921.
Wang, J. K., Singh, R. P., Braun, H. J., and Pfeiffer, W. H. (2009b). Investigating the
efficiency of the single backcrossing breeding strategy through computer simulation.
Theor. Appl. Genet. 118(4), 683–694.
Wang, J. K., Eagles, H. A., Trethowan, R., and van Ginkel, M. (2005). Using computer
simulation of the selection process and known gene information to assist in parental
selection in wheat quality breeding. Aust. J. Agric. Res. 56(5), 465–473.
Wang, J. K., Chapman, S. C., Bonnett, D. G., Rebetzke, G. J., and Crouch, J. (2007).
Application of population genetic theory and simulation models to efficiently pyramid
multiple genes via marker-assisted selection. Crop Sci. 47(2), 582–590.
White, J. W., and Hoogenboom, G. (2003). Gene-based approaches to crop simulation: Past
experiences and future opportunities. Agron. J. 95(1), 52–64.
Winsberg, E. (2010). Science in the Age of Computer Simulation. University of Chicago
Press, Chicago.
Wong, C., and Bernardo, R. (2008). Genomewide selection in oil palm: Increasing selection
gain per unit time and cost with small populations. Theor. Appl. Genet. 116(6), 815–824.
Wu, R. L., and Zeng, Z. B. (2001). Joint linkage and linkage disequilibrium mapping in
natural populations. Genetics 157(2), 899–909.
Xu, S. Z. (2010). An expectation-maximization algorithm for the lasso estimation of
quantitative trait locus effects. Heredity 105(5), 483–494.
Xu, S. Z. (1998). Mapping quantitative trait loci using multiple families of line crosses.
Genetics 148(1), 517–524.
Xu, S. Z. (2003). Estimating polygenic effects using markers of the entire genome. Genetics
163(2), 789–801.
Xu, Y. (2010). Bayesian mapping. In “Molecular Plant Breeding” (Y. Xu, Ed.), pp. 219–
223. CAB International, Wallingford, UK.
Yang, J., Zhu, J., and Williams, R. W. (2007). Mapping the genetic architecture of complex
traits in experimental populations. Bioinformatics 23(12), 1527–1536.
Yin, X., Kropff, M. J., and Stam, P. (1999). The role of ecophysiological models in QTL
analysis: The example of specific leaf area in barley. Heredity 82(4), 415–421.
Yin, X., Stam, P., Kropff, M. J., and Schapendonk, A. (2003). Crop modeling, QTL
mapping, and their complementary role in plant breeding. Agron. J. 95(1), 90–98.
Yin, X., Struik, P. C., and Kropff, M. J. (2004). Role of crop physiology in predicting gene-
to-phenotype relationships. Trends Plant Sci. 9(9), 426–432.
264 Xin Li et al.
Yin, X., Struik, P. C., Van Eeuwijk, F. A., Stam, P., and Tang, J. (2005). QTL analysis and
QTL-based prediction of flowering phenology in recombinant inbred lines of barley.
J. Exp. Bot. 56(413), 967.
Yin, X., and Struik, P. C. (2010). Modelling the crop: From system dynamics to systems
biology. J. Exp. Bot. 61(8), 2171–2183.
Yu, J. (2009). Realizing the potential of ultrahigh throughput genomic technologies in plant
breeding. Plant Genome 2, 2.
Yu, J., Holland, J. B., McMullen, M. D., and Buckler, E. S. (2008). Genetic design and
statistical power of nested association mapping in maize. Genetics 178(1), 539.
Yu, J., Pressoir, G., Briggs, W. H., Bi, I. V., Yamasaki, M., Doebley, J. F., et al. (2006).
A unified mixed-model method for association mapping that accounts for multiple levels
of relatedness. Nat. Genet. 38(2), 203–208.
Yu, J., Zhang, Z., Zhu, C., Tabanao, D. A., Pressoir, G., Tuinstra, M. R., et al. (2009).
Simulation appraisal of the adequacy of number of background markers for relationship
estimation in association mapping. Plant Genome 2(1), 63–77.
Yu, J., Arbelbide, M., and Bernardo, R. (2005). Power of in silico QTL mapping from
phenotypic, pedigree, and marker data in a hybrid breeding program. Theor. Appl. Genet.
110(6), 1061–1067.
Zeng, Z. B. (1994). Precision mapping of quantitative trait loci. Genetics 136(4), 1457.
Zeng, Z. B., Kao, C. H., and Basten, C. J. (1999). Estimating the genetic architecture of
quantitative traits. Genet. Res. 74(3), 279–289.
Zhang, L. Y., Wang, S. Q., Li, H. H., Deng, Q. M., Zheng, A. P., Li, S. C., et al. (2010).
Effects of missing marker and segregation distortion on QTL mapping in F-2 populations.
Theor. Appl. Genet. 121(6), 1071–1082.
Zhang, W., and Smith, C. (1992). Computer-simulation of marker-assisted selection utiliz-
ing linkage disequilibrium. Theor. Appl. Genet. 83(6–7), 813–820.
Zhang, W., and Smith, C. (1993). Simulation of marker-assisted selection utilizing linkage
disequilibrium—The effects of several additional factors. Theor. Appl. Genet. 86(4),
492–496.
Zhu, C., Gore, M., Buckler, E. S., and Yu, J. (2008). Status and prospects of association
mapping in plants. Plant Genome 1(1), 5–20.
Zhu, C., and Yu, J. (2009). Nonmetric multidimensional scaling corrects for population
structure in association mapping with different sample types. Genetics 182(3), 875–888.
Zou, W., and Zeng, Z. B. (2009). Multiple interval mapping for gene expression QTL
analysis. Genetica 137(2), 125–134.
Index
Note: Page numbers followed by “f ” indicate figures, and “t” indicate tables.
265
266 Index