ArticlePDF Available

Multiple benefits of legumes for agriculture sustainability: an overview

Authors:

Abstract

Food security, lowering the risk of climate change and meeting the increasing demand for energy will increasingly be critical challenges in the years to come. Producing sustainably is therefore becoming central in agriculture and food systems. Legume crops could play an important role in this context by delivering multiple services in line with sustainability principles. In addition to serving as fundamental, worldwide source of high-quality food and feed, legumes contribute to reduce the emission of greenhouse gases, as they release 5–7 times less GHG per unit area compared with other crops; allow the sequestration of carbon in soils with values estimated from 7.21 g kg−1 DM, 23.6 versus 21.8 g C kg−1 year; and induce a saving of fossil energy inputs in the system thanks to N fertilizer reduction, corresponding to 277 kg ha−1 of CO2 per year. Legumes could also be competitive crops and, due to their environmental and socioeconomic benefits, could be introduced in modern cropping systems to increase crop diversity and reduce use of external inputs. They also perform well in conservation systems, intercropping systems, which are very important in developing countries as well as in low-input and low-yield farming systems. Legumes fix the atmospheric nitrogen, release in the soil high-quality organic matter and facilitate soil nutrients’ circulation and water retention. Based on these multiple functions, legume crops have high potential for conservation agriculture, being functional either as growing crop or as crop residue.
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
DOI 10.1186/s40538-016-0085-1
REVIEW
Multiple benets oflegumes
foragriculture sustainability: an overview
Fabio Stagnari1*, Albino Maggio2, Angelica Galieni1 and Michele Pisante1
Abstract
Food security, lowering the risk of climate change and meeting the increasing demand for energy will increasingly be
critical challenges in the years to come. Producing sustainably is therefore becoming central in agriculture and food
systems. Legume crops could play an important role in this context by delivering multiple services in line with sus-
tainability principles. In addition to serving as fundamental, worldwide source of high-quality food and feed, legumes
contribute to reduce the emission of greenhouse gases, as they release 5–7 times less GHG per unit area compared
with other crops; allow the sequestration of carbon in soils with values estimated from 7.21 g kg1 DM, 23.6 versus
21.8 g C kg1 year; and induce a saving of fossil energy inputs in the system thanks to N fertilizer reduction, corre-
sponding to 277 kg ha1 of CO2 per year. Legumes could also be competitive crops and, due to their environmental
and socioeconomic benefits, could be introduced in modern cropping systems to increase crop diversity and reduce
use of external inputs. They also perform well in conservation systems, intercropping systems, which are very impor-
tant in developing countries as well as in low-input and low-yield farming systems. Legumes fix the atmospheric
nitrogen, release in the soil high-quality organic matter and facilitate soil nutrients’ circulation and water retention.
Based on these multiple functions, legume crops have high potential for conservation agriculture, being functional
either as growing crop or as crop residue.
Keywords: Soil fertility, Conservation agriculture, Sustainable agricultural systems, Food security, Climate change,
Greenhouse gas, Energy
© The Author(s) 2017. This article is distributed under the terms of the Creative Commons Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium,
provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license,
and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/
publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.
Introduction
Global population will hit 9.6 billion people by 2050 [108]
and will face global challenges among which achiev-
ing food security, lowering the risk of climate change
by reducing the net release of greenhouse gases into
the atmosphere and meeting the increasing demand
for energy are the most critical ones. In particular, the
impact of climate change and associated biotic and abi-
otic stresses to which crop systems will be increasingly
exposed pose serious implications for global food pro-
duction [119].
To meet these challenges, a policy framework needs to
be developed in which the sustainability of production/
consumption patterns becomes central. In this context,
food legumes and legume-inclusive production systems
can play important roles by delivering multiple services
in line with sustainability principles. Indeed, legumes
play central roles [112]: (1) at food-system level, both for
human and animal consumption, as a source of plant pro-
teins and with an increasingly importance in improving
humans health [106]; (2) at production-system level, due
to the capacity to fix atmospheric nitrogen making them
potentially highly suitable for inclusion in low-input
cropping systems, and due to their role in mitigating
greenhouse gases emissions [53]; and (3) at cropping-
system levels, as diversification crops in agroecosystems
based on few major species, breaking the cycles of pests
and diseases and contributing to balance the deficit in
plant protein production in many areas of the world,
including Europe [43, 48, 72, 78, 116].
Leguminosae family comprises 800 genera and 20,000
species [54] and represents the third largest family of
flowering plants. Some legumes are considered weeds of
Open Access
*Correspondence: fstagnari@unite.it
1 Faculty of Biosciences and Technologies for Agriculture Food
and Environment, University of Teramo, Coste Sant’Agostino, Teramo, TE,
Italy
Full list of author information is available at the end of the article
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 2 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
cereal crops, while others are major grain crops; these
latter species are known as grain legumes, or pulses,1 and
represent the focus of this review. For some of these spe-
cies, the trends for word acreage and yield are available,
as reported in Table1.
Despite the growing trends observed during the
50-year period between 1974 and 2014 for some
warm-season legumes (e.g. soybean, cowpea, dry bean,
groundnut, and pigeon pea), the acreages of several
temperate legumes (e.g. pea, faba bean, lupin, french
bean and vetch) have declined worldwide with differ-
ences between world Regions (Table3). In any case,
food legumes occupy a minimal part of arable land,
mostly dominated by cereal crops [99]; soybean rep-
resents the most important and cultivated legume,
acreage of which reached 117.72 million ha in 2014
(steadily increased over years, see also Table3), which
is about that of the other grain legumes, but still far
below the major cereals (e.g. rice, wheat, maize). Such
trend is mainly associated to the expansion of more
specialized and intensive production systems [82].
1 Soybean and groundnuts are not defined by FAO as ‘pulse crops’.
Market forces stimulating specialization of cropping
systems as non-marketable benefits of diversification,
like cultivation/introduction of legumes in the farm-
ing system, do not deliver immediate and/or apparent
profits [82]. This is, however, not equally perceived
throughout the globe, and there is indeed a remarkable
diversity in grain legumes’ production trends across
the world (Table3).
e European decline in grain legume’s produc-
tion is not mirrored by other regions of the world such
as Canada or Australia, where legume’s cultivation
has been increasing over the last few decades. In these
areas, monoculture of cereals, which relies on frequent
summer-fallowing and use of mechanical tillage, has
been replaced by extended and diversified crop rotations
together with the use of conservation tillage [122]. Fur-
thermore, supply chains and markets are inadequately
developed for most legume crops (see also [66], for
France) with the exception of soybean, for which the
global market is well developed [85]. Nevertheless, soy-
bean areas in Europe are constrained by climatic factors
although there is considerable potential to develop new
varieties suitable to flourish under cool growing condi-
tions [123].
Table 1 Trends forword acreage (million ha) andyield (t ha1) forlegume crops included inFAOSTAT classication start-
ing from1974 to2014 [23]; the major three cereal crops are also reported, forcomparison
In Table2, for each legume crop, item name and code as well as FAO denitions are reported
a Data are referred to year 2013 (2014 data not available)
Harvested area (Million ha) Yield (t ha1)
1974 1984 1994 2004 2014 1974 1984 1994 2004 2014
Legume crops
Bambara bean 0.05 0.05 0.09 0.12 0.37 0.67 0.66 0.64 0.65 0.77
Dry bean 23.9 26.3 26.7 27.3 30.14 0.53 0.6 0.65 0.67 0.83
Faba bean 3.98 3.32 2.48 2.65 2.37 1.07 1.29 1.45 1.62 1.82
Chickpea 10.6 9.85 9.96 10.5 14.8 0.56 0.67 0.71 0.8 0.96
Cowpea 4.7 3.66 7.35 9.18 12.52 0.35 0.31 0.38 0.45 0.45
Groundnut 19.9 18.2 22 23.7 25.68 0.94 1.1 1.3 1.54 1.65
Lentil 2.03 2.56 3.43 3.85 4.52 0.61 0.68 0.81 0.93 1.08
Lupin 0.76 1.06 1.56 1.05 0.76 0.84 1.05 0.78 1.18 1.3
Pea 8.13 8.91 7.65 6.34 6.87 1.22 1.3 1.88 1.85 1.65
Pigeon pea 3.04 3.61 4.24 4.72 6.67 0.54 0.78 0.74 0.7 0.73
Soybean 37.4 52.9 62.5 91.6 117.72 1.41 1.71 2.18 2.24 2.62
French bean 0.22 0.17 0.22 0.23 0.20a5.76 6.93 7.44 9.04 9.32a
Vetch 1.52 1.29 0.93 0.89 0.52 1.24 1.21 1.12 1.43 1.71
Pulses, nes 5.67 5.65 5.33 4.27 6.1 0.54 0.58 0.65 0.82 0.84
Vegetables, leguminous nes 0.14 0.18 0.18 0.25 0.24a5.41 5.09 5.18 6.54 6.86a
Major cereal crops
Wheat 222.12 230.77 215.12 216.57 221.62 1.62 2.22 2.45 2.92 3.29
Maize 119.86 127.76 137.99 147.45 183.32 2.56 3.53 4.12 4.94 5.66
Rice (paddy) 136.89 144.24 147.29 150.58 163.25 2.43 3.23 3.66 4.03 4.54
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 3 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
e low diffusion of legumes’ cultivation is also due to
reduced and unstable yields and susceptibility to biotic
and abiotic stress conditions; the average yields for
unit area have increased (soybean +86%, lentil +77%,
groundnut +75%, chickpea +70%) less than cereal crops
(+104%, on average) (Table1). Moreover, legume cultiva-
tion depends not only on the effect of farmers’ choices,
although they play a central role for such decision, but
also on policymakers who have the responsibility to pro-
vide effective strategies to support the integration of leg-
umes into cropping systems. is aspect is particularly
relevant if the overall objective for future agricultural sys-
tems is to promote sustainability, improve resource use
efficiency and preserve the environment [82].
Grain legumes impacts onatmosphere andsoil
quality
Among the many important benefits that legumes deliver
to society, their role in contributing to climate change
mitigation has been rarely addressed. Legumes can (1)
lower the emission of greenhouse gases (GHG) such as
carbon dioxide (CO2) and nitrous oxide (N2O) compared
Table 2 Denition oflegume crops focused inTable1 andcorresponding item name inFAOSTAT
Legume crop Scientic name Corresponding FAO item name andcode FAO denition
Bambara bean Voandzeia subterranea Bambara bean [203] Bambara groundnut, earth pea. These beans
are grown underground in a similar way to
groundnuts
Dry bean Beans, dry [176] Phaseolus spp.: kidney, haricot bean (Ph. vulgaris);
lima, butter bean (Ph. lunatus); adzuki bean (Ph.
angularis); mungo bean, golden, green gram (Ph.
aureus); black gram, urd (Ph. mungo); scarlet run-
ner bean (Ph. coccineus); rice bean (Ph. calcara-
tus); moth bean (Ph. aconitifolius); tepary bean
(Ph. acutifolius). Several countries also include
some types of beans commonly classified as
Vigna (angularis, mungo, radiata, aconitifolia)
Faba bean Vicia faba Broad beans, horse beans, dry [181] Vicia faba: horse-bean (var. equina); broad bean
(var. major); field bean (var. minor)
Chickpea Cicer arietinum Chick peas [191] Chickpea, Bengal gram, garbanzos (Cicer arieti-
num).
Cowpea Vigna ungiculanta Cow peas, dry [195] Cowpea, blackeye pea/bean (Vigna sinensis; Doli-
chos sinensis)
Groundnut Arachis hypogaea Groundnuts, with shell [242] Arachis hypogaea. For trade data, groundnuts in
shell are converted at 70% and reported on a
shelled basis
Lentil Lens esculenta Lentils [201] Lens esculenta; Ervum lens
Lupin Lupins [210] Lupinus spp. Used primarily for feed, though in
some parts of Africa and in Latin America some
varieties are cultivated for human food
Pea Peas, dry [187] Garden pea (Pisum sativum); field pea (P. arvense)
Pigeon pea Cajanus cajan Pigeon peas [197] Pigeon pea, cajan pea, Congo bean (Cajanus cajan)
Soybean Glycine max Soybeans [236] Glycine soja
French bean String beans [423] Phaseolus vulgaris; Vigna spp. Not for shelling
Vetches Vicia sativa Vetches [205] Spring/common vetch (Vicia sativa). Used mainly
for animal feed
Pulses, nes Pulses, nes [211] Including inter alia: lablab or hyacinth bean
(Dolichos spp.); jack or sword bean (Canavalia
spp.); winged bean (Psophocarpus tetragonolo-
bus); guar bean (Cyamopsis tetragonoloba); velvet
bean (Stizolobium spp.); yam bean (Pachyrrhizus
erosus); Vigna spp. other than those included in
176 and 195; other pulses that are not identified
separately because of their minor relevance at
the international level. Because of their limited
local importance, some countries report pulses
under this heading that are classified individually
by FAO
Vegetables, leguminous nes Vegetables, leguminous nes [420] Vicia faba. For shelling
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 4 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
with agricultural systems based on mineral N fertiliza-
tion, (2) have an important role in the sequestration of
carbon in soils, and (3) reduce the overall fossil energy
inputs in the system.
Greenhouse gas emissions
e introduction of legumes into agricultural rota-
tions help in reducing the use of fertilizers and energy
in arable systems and consequently lowering the GHG
emissions [52]. N fertilizer savings across Europe [51],
in rotations including leguminous crops, range around
277kgha1 of CO2 per year (1kgN=3.15kg CO2, [42].
It has been reported that half of the CO2 generated dur-
ing NH3 production would be reused if the NH3 was
converted to urea. is is, however, only a time shift of
CO2 release in the atmosphere since, once the urea is
applied to the soil, the hydrolyzation activity by urease
will release CO2 originally captured during urea produc-
tion [39]. Considering an efficiency of 2.6–3.7 kg CO2
generated per kilogram of N synthesized, the annual
global fertilizer leads to a release of 300Tg of CO2 into
the atmosphere each year [42]. Some studies indicate
that at global scale, the amount of CO2 respired from
the root systems of N2-fixing legumes could be higher
than the CO2 generated during N-fertilizer production
[42]. However, it is important to emphasize that the
CO2 respired from nodulated roots of legumes comes
from the atmosphere through the photosynthesis activ-
ity. Conversely, all the CO2 released during the process
of N-fertilizer synthesis derives from fossil energy, thus
determining a net contribution to atmospheric amount
of CO2 [42].
N2O represents about 5–6% of the total atmospheric
GHG, but it is much more active2 than CO2 [21]. Agricul-
ture represents the main source of anthropogenic N2O
emissions (about 60%; [84], due to both animal and crop
production [38]). A majority of these emissions result
from the application of nitrogen fertilizers [84]: every
100kg of N fertilizer about 1.0kg of N is emitted as N2O
[42], although different amounts depend on several fac-
tors including N application rate, soil organic C content,
soil pH, and texture [78, 88]. Denitrification processes are
the most important source of N2O in most cropping and
pasture systems [76, 88, 102].
In the recent years, several studies have focalized on
the role of legumes in the reduction of GHG emissions.
Jeuffroy et al. [44] demonstrated that legume crops
2 N2O absorbs approximately 292 times as much infra-red radiation per
kilogram as CO2.
Table 3 Trend forRegion Δ acreage (%) duringthe 50-year period starting from1974 to2014 forlegume crops included
inFAOSTAT classication [23]; the major three cereal crops are also reported, forcomparison
In Table2, for each legume crop, item name and code as well as FAO denitions are reported
a Data are referred to year 2013 (2014 data not available)
Δ harvested area 1974–2014 (%)
Africa Northern America South America Asia Europe Oceania
Legume crops
Bambara bean +612 – –
Dry bean +207 +16 20 +25 84 +1778
Faba bean +7 Disappeared 53 59 54 +75,085
Chickpea +30 Appeared +1+37 35 –
Cowpea +168 Appeared Appeared +402 +153 –
Groundnut +69 10 22 +6+16 39
Lentil 20 +3376 75 +72 45 Appeared
Lupin 82 – +577 89 64 +315
Pea +49 +1119 +721 63 +578
Pigeon pea +226 – 83 +108 –
Soybean +642 +71 +882 +116 +291 10
French bean Appeareda39a+129a+66a18a+122a
Vetch +109 – 73 80 +4757
Pulses, nes +20 – 69 15 +73 +7648
Vegetables, leguminous nes +180aAppeareda+118a+23a31a52a
Major cereal crops
Wheat +11 20 +16 +39 33 +51
Maize +98 +29 +45 +76 +21 +31
Rice (paddy) +185 +15 16 +16 28 +2
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 5 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
emit around 5–7 times less GHG per unit area com-
pared with other crops. Measuring N2O fluxes, they
showed that peas emitted 69kg N2Oha1, far less than
winter wheat (368 kg N2O ha1) and rape (534 kg
N2O ha1). Clune et al. [19] reviewed different life
cycle-assessment (LCA) studies on GHG emissions
carried out from 2000 to 2015 around the world
(despite the used literature was predominately Euro-
pean centric) highlighting that pulses have a very low
Global Warming Potential (GWP) values (0.50–0.51kg
CO2 eqkg1 produce or bone-free meat3). In a compar-
ison between vetch and barley under Mediterranean
environments and alkaline soil, N2O emissions were
higher for barley than vetch; furthermore, the N2O
fluxes derived from the synthetic fertilizers added to
the crops were 2.5 times higher in barley compared
with vetch [29]. In two field experiments conducted in
a black Vertosol in sub-tropical Australia, Schwenke
etal. [95] demonstrated that the cumulative N2O emis-
sions from N-fertilized canola greatly exceeded those
from chickpea, faba bean and field pea (385 vs. 166, 166
and 135g N2O-Nha1, respectively). e same authors
highlighted that grain legumes significantly reduced
their emission factors suggesting that legume-fixed N is
a less-emissive form of N input to the soil than ferti-
lizer N.
Nevertheless, it is important to highlight that the
influence of legumes in reducing GHG depends also on
the management of agro-ecosystems in which they are
included. When faba bean was grown as mono crop-
ping, it led to threefold higher cumulative N2O emissions
than that of unfertilized wheat (441 vs. 152g N2Oha1,
respectively); conversely, when faba bean was mixed with
wheat (intercropping system), cumulative N2O emissions
fluxes were 31% lower than that of N-fertilized wheat
[96]. Anyway, the benefits derived from the introduction
of legumes in crop rotations become significant when
commercially relevant rates of N fertilizer are applied
[42].
e mitigation in terms of GHG emissions is also
obtained by adopting sustainable agricultural systems,
such as conservation tillage and conservation agriculture
systems, which are suitable for the cultivation of both
grain and green-manure legumes (see “Grain legumes
and conservation agriculture” section).
In conclusion, it is noteworthy to underline that field
tests and experimental analyses on GHG emissions, and
in particular on N2O, provided quite different results [89]
due to the influences of differences of several variables,
3 In the study of Clune et al. [19], each GWP value recorded from the
literature data was converted into a common functional unit and system
boundary in kg CO2 eqkg1 bone-free meat (BFM), using the conversion
ratios identified in the literature.
including climatic, soil and management conditions [45,
78, 88].
In general, N2O losses from soils covered with leg-
umes are certainly lower than those from both N2O
fertilized grasslands and non-legume crops, as also indi-
cated by Jensen etal. [42] who report a mean of 3.22kg
N2O-N ha1, calculated from 67 site years of data. In
addition, there is no direct association between N2O
emissions and biological nitrogen fixation [42], since
organic N from legume residues is decomposed, miner-
alized and rapidly immobilized by microorganisms [78].
Emissions of N2O could occur either during nitrifica-
tion or due to denitrification, being affected by timing
of mineralized N supply [20]: the asynchrony between N
supply and utilization from the following crops enhances
N loss, especially in winter/early spring in cold wet soils
[64].
Soil properties
Cultivation and cropping may cause significant SOC
losses through decomposition of humus [18]. Shifting
from pasture to cropping systems may result in loss of
soil C stocks between 25 and 43% [101].
Legume-based systems improve several aspects of
soil fertility, such as SOC and humus content, N and P
availability [42]. With respect to SOC, grain legumes
can increase it in several ways, by supplying biomass,
organic C, and N [27, 53], as well as releasing hydro-
gen gas as by-product of BNF, which promotes bacte-
rial legume nodules’ development in the rhizosphere
[49].
In sandy soils, the beneficial effect of grain legumes
after three years of study was registered in terms of
higher content of SOC compared with soils with oats
(7.21 g kg1 DM, on average). Specifically, cultiva-
tion of pea exerted the most positive action to organic
carbon content (7.58 g kg1, after harvest, on aver-
age), whereas narrow-leaved lupin had the least effect
(7.23 g kg1, on average) [30]. In southern America
(Argentina), the intercropping of soybean with maize
at different rates favoured a SOC accumulation of
23.6g C kg1 versus 21.8g C kg1 of the sole maize;
the greatest potential for enhancing SOC stocks
occurred in the 2:3 (maize:soybean) intercrop configu-
ration [11]. Furthermore, just only amending the soil
with soybean residues allows to obtain an increase of
38.5% in SOC [11].
anks to BNF, legumes also affect significantly soil
N availability; by using legumes as winter crops in rice–
bean and rice–vetch combination, rice residue N con-
tent is enhanced by 9.7–20.5%, with values ranging from
1.87 to 1.93gNkg1 soil [120]. It needs to be underlines
that a majority of studies on the role of legumes for soil
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 6 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
N fertility have investigated the shoot N content. In this
regard, Carranca et al. [15] found that 7–11% of total
legume N was associated with root and nodules and an
allocation of 11–14kg N fixed t1 belowground dry mat-
ter, representing half the amount of total aboveground
plant.
In intercropping cowpea–maize, Latati et al. [50]
found an increase in P availability at rhizosphere level
associated with significant acidification (0.73U) than
in sole cropping. Wang etal. [115], assessing proper-
ties related to N and P cycling in the rhizosphere of
wheat and grain legumes (faba bean and white lupin)
grown in monoculture or in wheat/legume mixtures,
found that the less-labile organic P pools (i.e. NaOH-
extractable P pools and acid-extractable P pools) sig-
nificantly accumulated in the rhizosphere of legumes.
However, the P uptake and the changes in rhizosphere
soil P pools seem to depend also on legume species.
Compared with the unplanted soil, the depletion of
labile P pools (resin P and NaHCO3-P inorganic) was
the greatest in the rhizosphere of faba bean (54 and
39%) with respect to chickpea, white lupin, yellow lupin
and narrow-leafed lupin [31]. Of the less-labile P pools,
NaOH-P inorganic was depleted in the rhizosphere of
faba bean, while NaOH-P organic and residual P were
most strongly depleted in the rhizosphere of white
lupin [31].
Also in North Rift, Kenya Region, in well-drained,
extremely deep, friable clay, acid humic top soil, the
effects of cultivation and incorporation of lupine and gar-
den pea were significant in terms of soil-available P with
respect to fallow, with lupine showing higher P availabil-
ity than pea (from 20.3 to 31.0% higher).
Although there is a general agreement on the influence
of grain legumes on rhizosphere properties in terms of
N supply, SOC and P availability, the magnitude of the
impact varied across legume species, soil properties and
climatic conditions. Among these, soil type represents
the major factor determining plant growth, rhizosphere
nutrient dynamics and microbial community structure.
e pattern of depletion and accumulation of some
macro- and micronutrients differed also between crop-
ping systems (i.e. monoculture, mixed culture, narrow
crop rotations) as well as among soil management strate-
gies (i.e. tillage, no-tillage).
Role ofgrain legumes incropping systems
Legumes could be competitive crops, in terms of envi-
ronmental and socioeconomic benefits, with potential
to be introduced in modern cropping systems, which
are characterized by a decreasing crop diversity [24, 80]
and an excessive use of external inputs (i.e. fertilizers and
agrochemicals).
Grain legumes intocrop‑sequences
In the recent years, many studies have focused on the
sustainable re-introduction of grain legumes into crop
rotations,4 based on their positive effects on yield and
quality characteristics on subsequent crops [46, 82, 103].
However, assessment of the rotational advantages/disad-
vantages should be based on a pairwise comparison
between legume and non-legume pre-crops [82]. Some
experimental designs involving multi-year and multispe-
cies rotations do not provide information on yield bene-
fits to the subsequent species in the rotation sequence.
erefore, it is difficult to formulate adequate conclu-
sions [2].
e agronomic pre-crop benefits of grain legumes can
be divided into a ‘nitrogen effect’ component and ‘break
crop effect’ component. e ‘nitrogen effect’ component
is a result of the N provision from BNF [77], which is
highest insituations of low N fertilization to subsequent
crop cycles [82]. e second one (break crop effect)
includes non-legume-specific benefits, such as improve-
ments of soil organic matter and structure [34], phospho-
rus mobilization [98], soil water retention and availability
[2], and reduced pressure from diseases and weeds [87].
In this case, benefits are highest in cereal-dominated
rotations [82].
Several authors have reviewed the yield benefits of leg-
umes for subsequent cereal crops.
In Australia, Angus etal. [2] reported higher yield of
wheat after legumes (field peas, lupins, faba beans, chick-
peas and lentils) than those of wheat after wheat. In par-
ticular for a wheat–wheat yield of 4.0t ha1, the mean
grain legume-wheat yield was 5.2tha1 (+30% on aver-
age). Other studies from Australia quantified yield ben-
efits compared to pure cereal crop sequences at 40–50%
for low N levels and 10–17% for high N levels [3].
In Europe yields benefits of grain legumes have been
shown to strongly depend on climatic factors which
affect N dynamics in soils [52]. In temperate environ-
ments, cereals yield is on average 17 and 21% higher in
grain-legume based systems than wheat monocropping,
under standard and moderate fertilization levels, respec-
tively [40]. Conversely, yield benefits are lower in Medi-
terranean climates where water availability is the limiting
factor to cereal yields [46, 61, 62].
e yield advantage to subsequent cereal crops pro-
vided by legumes depends also on the species and
amounts of fixed N [114, 121]. Field pea and faba bean
4 According to Angus etal. [2], crop-sequences experiments can be classi-
fied into rotation experiments and break crop experiments. Rotation strictly
defined, refers to a recurring sequence of crops, forages and fallows, or
more loosely defined, to a cropping sequence that contains fallows, or crops
and forages in addition to the locally dominant species. A break crop gener-
ally refers to a single alternative crop followed by the dominant species.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 7 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
accumulate about 130 and 153kgNha1 in their above-
ground biomass, respectively [77] and significant quan-
tities (30–60% of the accumulated total N) may also be
stored in belowground biomass [77]. Differences in BNF
patterns are also found between the same species. For
example, Mokgehle etal. [69] compared 25 groundnut
varieties for plant BNF at three differing agro-ecologies
in South Africa, highlighting N-fixed range between 76
and 188 kg ha1, depending also on soil and environ-
mental conditions as well as on N-uptake. Other factors
influencing BNF include salinity and sodicity (alkalinity)
of soils, as observed in chickpea [83], common bean [22]
and faba bean [109].
It is, however, rather difficult to quantify the legume
dependent increase in N uptake in subsequent crops,
versus other sources of N [46, 77]. In temperate envi-
ronments of Australia, measurements of the additional
N-nitrate available to wheat crops following legumes
instead of cereals, averaged around 37kgNha1 [17]. In
Denmark, nitrogen uptake in crops that follow legume
crops has been reported to increase by 23–59% after field
pea and narrow-leafed lupin on different soil types [40],
but only 14–15% for durum wheat following vetch in a
semi-arid Mediterranean environment [28]. Increased N
uptake of crops after grain legumes reached up to 61%
or 36kgha1 for a vetch-barley rotation in Cyprus [74].
Further, some legume residues have beneficial effects on
some quality aspects of the subsequent crops in southern
Italy [104].
Among other beneficial effects brought about by leg-
umes, the production of hydrogen gas (H2) as a by-prod-
uct of BNF greatly affects the composition of the soil
microbial population, further favouring the development
of plant growth-promoting bacteria [2].
Some grain legumes, including chickpea, pigeon
pea and white lupin can mobilize fixed forms of soil P
through the secretion of organic acids such as citrate
and malate and other P mobilizing compounds from
their roots [36]. Among grain legumes, white lupin most
strongly solubilize P, a function that can be facilitated by
its proteoid roots that may englobe small portions of soil
[2]. Glasshouse experiments using a highly P-fixing soil
showed better wheat growth following white lupin than
soybean [37], suggesting that the cereal was able to access
P made available by the previous white lupin break crop.
‘Break crop’ effects also include increased soil water con-
tent, since the break-crop stubble can affect retention
of soil water and infiltration and retention of rain water
[47]. A species-specific response has also been docu-
mented. Soil profiles after pea field can be wetter than
after a wheat crop [2]. In Saskatchewan, Canada, Miller
et al. [68] reported that post-harvest soil water status
up to 122cm-depth was 31 and 49mm greater for all
legumes (field pea, lentil and chick pea) with respect of
wheat under loam and clay soils, respectively. is was
primarily due to increased plant water use efficiency.
Lentil in rotation with cereals has been shown to increase
total grain production by increasing residual soil water in
dry areas of Saskatchewan [25].
In general, grain legumes are not susceptible to the
same pests and diseases as the main cereal crops (non-
host), resulting suitable as break crops in wheat-based
rotations [121]. Grain legumes as break crops can also
contribute to weed control [97] by contrasting their spe-
cialization and helping stabilizing the agricultural crop
weed community composition [7].
Despite the described beneficial effects, there are still
concerns on the introduction of grain legumes into crop-
ping sequences. Cropping systems that include legume
crops in farm rotations must be supported by best crop-
management practices (e.g. N fertilization rates and tim-
ing, soil management, weeding, irrigation), which often
do not match standard techniques normally applied by
farmers. For example, some possible risks in terms of
nitrate leaching associated to grain legumes cultivation
can be counteracted by including cover crops in the sys-
tem [33, 81]. Additional reasons may explain why grain
legumes are not very common in high-input cropping
systems. ese include (1) their low and unstable yields
[16, 86]; (2) inadequate policy support [14]; (3) lack of
proper quantification (and recognition) of long-term
benefits of legumes within cropping systems [82]. How-
ever, other efforts could be addressed, for example, to
breeding programs for improved crop cultivars, to better
sustain livelihood and increase the economic return to
farmers. Indeed, during last years significant progresses
in breeding for quality traits for food [110] and feed uses
[79], as well as for resistances to biotic [91] and abiotic
stresses [4] are being achieved, but several others, many
of which are controlled quantitatively by multiple genes,
have been more difficult to achieve.
Grain legumes inintercropping
Intercropping systems consist in simultaneous growth of
two or more crop species on the same area and at the
same time [13]. Intercropping is widely used in develop-
ing countries or in low-input and low-yield farming sys-
tems [73]. Despite several recognized beneficial aspects
of intercropping such as better pest control [60], com-
petitive yields with reduced inputs [70, 107], pollution
mitigation [63], more stable aggregate food or forage
yields per unit area [100], there are a number of con-
strains that make intercropping not common in modern
agriculture, such as example the request of a single and
standardized product and the suitability for mechaniza-
tion or use of other inputs as a prerogative in intensive
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 8 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
farming system [13]. It is therefore necessary to opti-
mize intercropping systems to enhance resource-use
efficiency and crop yield simultaneously [55], while also
promoting multiple ecosystem services (see also [13]).
Most recent research has focalized on the potential of
intercropping in sustainable productions and in par-
ticular on grain legumes that can fix N2 through bio-
logical mechanisms (BNF). Indeed, legumes are pivotal
in many intercropping systems, and of the top 10 most
frequently used intercrop species listed by Hauggaard-
Nielsen and Jensen [32], seven are legumes One of the
basic spatial arrangements used in intercropping is
strip intercropping, in which two or more crops grow
together in strips wide enough to permit separate crop
production using inputs but close enough for the crops
to interact. e current challenge is how to determine an
optimal intercropping width to maximise the resources
use efficiency and, consequently, the crop productivity.
In a maize-bean strip intercropping, Mahallati etal. [65]
suggested that strip width of 2 and 3 rows was superior
compared with monoculture and other strip intercrop-
ping combinations in terms of radiation absorption,
radiation use efficiency and biological yields of both
species, also allowing to an improve of total land pro-
ductivity and land equivalent ratio (1.39 and 1.37). Gao
etal. [26] showed a total yield increase of 65 and 71% in
a system of 1 and 2 rows of maize (planted at a higher
density in intercropping) alternated with 3 rows of soy-
bean compared with both crops grown as monoculture.
However, Liu etal. [59] showed a reduction in the pho-
tosynthetically active radiation and R:FR ratio at the
top of soybean canopy intercropped with maize - under
two intercropping patterns: 1 row of maize with 1 row
of soybean; 2 rows of maize with 2 rows of soybean -
leading to increased internode lengths, plant height and
specific leaf area (SLA), but reduced branching of soy-
bean plants. In order to gain sufficient light in the most
shaded border rows of the neighbouring, shorter crops,
efforts could be addressed to (i) the selection of highly
productive maize cultivars with reduced canopy height
and LAI; (ii) the increase of the strip width under a
higher fraction of direct PAR; (iii) the selection of crops
and cultivars suitable under the shade levels that likely
occur in strip-intercropping systems with maize [71].
e increase in N availability in intercrops hosting leg-
umes occurs because the competition for soil N from
legumes is weaker than from other plants. Moreover,
non-legumes obtain additional N from that released by
legumes into the soil [56, 117] or via mycorrhizal fungi
[113]. Legumes can contribute up to 15% of the N in an
intercropped cereal [57], thus increasing biomass pro-
duction and carry-over effects [75], reducing synthetic
mineral N-fertilizer use and mitigating N2O fluxes [9,
96]. However, the adoption of grain legume intercrop-
ping systems should benefit from the identification of
suitable legumes that are less susceptible to N fertilizer-
induced inhibition of BNF—that is, legumes that sustain
higher %BNF in the presence of increasing soil mineral
N. To this purpose, Rose etal. [90] indicated that faba
bean is more suitable as intercrop than chickpea when
supplementary N fertilizer additions are required, with
about 40%BNF and 29%BNF maintained in faba bean
and chickpea, respectively, supplying both crops with
150kgNha1.
BNF represents the most common plant growth stim-
ulating factor that can also improve crop competition
with respect to weeds in both organic and sustainable
farming systems [10]. Grain legumes are weak suppres-
sors of weeds, but mixing species in the same crop-
ping system could represent a valid way to improve the
ability of the crop itself to suppress weeds [41, 94]. In
a wheat-chickpea intercropping system (20cm spacing
without weeding treatment) it was observed a 69.7%
reduction in weed biomass and 70% in weed population
as compared to un-weeded monocrop wheat at 20cm
spacing [6]. Similar results on weed smothering have
been obtained by Midya et al. [67] in rice-blackgram
(20 cm) intercropping system although the deferred
seeding of blackgram in rice field (30 cm) with one
weeding may be recommended for both better yield and
weed suppression.
Direct mutual benefits in cereal-legumes intercropping
involve below-ground processes in which cereals while
benefiting of legumes-fixed N, increase Fe and Zn bio-
availability to the companion legumes [118].
Physiology, agronomy and ecology can simultaneously
contribute to the improvement of intercropping systems,
allowing to enhance crop productivity and resource-use
efficiency, so making intercropping a viable approach for
sustainable intensification, particularly in regions with
impoverished soils and economies where measured ben-
efits have been greatest [93]. But to realize these goals,
major efforts in research programs still remain. For exam-
ple: (1) breeding for intercrops; (2) better understanding
of the interactions between plants and other organisms in
crop systems, focusing on the roles of above- and below-
ground interactions of plants with other organisms; (3)
improving agricultural engineering and management, i.e.
developing new machinery that can till, weed and harvest
at small spatial scales and in complex configurations to
encourage the uptake of intercropping without greater
demands for labour [58]; (4) adoption of a wider ‘systems
thinking’ through the enactment of schemes, including
payment for ecosystem services [105].
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 9 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
Grain legumes andconservation agriculture
Legumes have some characteristics particularly suitable
for sustainable cropping systems and conservation agri-
culture, and making them functional either as growing
crop or as crop residue. Conservation agriculture is based
on minimal soil disturbance and permanent soil cover
combined with rotations [35]. As previously described,
major advantages of legumes include the amount of
nitrogen fixed into the soil and the high quality of the
organic matter released to the soil in term of C/N ratio.
Some legume species have also deep root systems, which
facilitate nutrients solubilization by root exudates and
their uptake/recycling as well as water infiltration in
deeper soil layers.
Many countries already rely on conservation agricul-
ture. Brazil has implemented conservation agriculture
systems using soybean as legume crop. Grain legumes
like lentil, chickpea, pea and faba bean play a major role
in conservation agriculture in North America, Australia,
and Turkey. In Australia, some advantages of minimum
tillage for grain legumes have been quantified for water-
limited environments. Some studies indicate that the
majority of grain-legumes producers use direct seeding
after a legume pre-crop [1]. is change from conven-
tional tillage (CT) to reduced or no tillage (NT) systems
(with at least 30% of the soil surface covered) would lead
to significant positive impacts on SOC [18]. In contrast,
other results indicate that such positive effects are lim-
ited to the first 20cm depth, while little or no difference
between CT and NT in total SOC can be seen lower
down the soil profile [5, 111]. Such findings suggest that
C stock changes in the soil are mainly dependent to the
net N-balance in the system. With high N harvest index
legumes, SOC stocks are not preserved due to the high
amount of N taken off from the field into the grain [42].
Conversely, the effect of legumes on soil carbon seques-
tration is more detectable for forage, green-manures
and cover-crops which return to the soil large amounts
of organic C and N [52]. Boddey etal. [12] indicate that
vetch under no tillage may increase SOC stocks under
NT (0–100 cm) at a rate between 0.48 and 1.53Mg C
ha1 per year [42].
e implementation of practices of conservation till-
age could significantly reduce the GWP, especially when
a grain legumes is added to the rotation. In Mediter-
ranean agro-ecosystems, Guardia et al. [29] compared
three tillage treatments (i.e. no tillage: NT, minimum
tillage: MT, conventional tillage: CT) and two crops (i.e.
vetch, barley) and recorded the emission of N2O, CH4
and CO2 during one year. Authors found a significant
‘tillage×crop’ interaction on cumulative N2O emissions
with vetch releasing higher N2O amount than barley only
in CT and MT, whereas similar fluxes were observed
under NT. is was attributable to the soil water-filled
pore space, dissolved organic carbon content and denitri-
fication losses, in spite of the presumable predominance
of nitrification. In any case, the most sustainable crop
and tillage treatments in terms of GWP were represented
by the non-fertilized vetch and NT, due to higher carbon
sequestration, lower fuel consumption and the absence
of mineral N fertilizers [29]. In subtropical Ultisol, under
legume cover crops, NT soil exhibited increased N2O
emissions with respect to CT soil (531 vs. 217kg CO2
eqha1 year1); however, emissions of this gas from NT
soil were fully offset by CO2 retention in soil organic mat-
ter (2063 to 3940kg CO2 ha1 year1) [8]. Moreover,
NT soil under legume cover crops behaved as a net sink
for GHG (GWP ranged from 971 to 2818kg CO2 eq
ha1 year1) [8].
e expansion of ecological-based approaches like con-
servation agriculture opens opportunities to food leg-
umes to be profitably included in sustainable cropping
systems. ere are still major challenges for conserva-
tion agriculture that need to be overcome, including the
development of effective methods for weed control (see
also [92]) that can avoid the use of herbicides or tillage.
Overall conservation agriculture is an environmentally
sustainable production system that may boost the incor-
poration of grain legumes within large and small-scale
farming.
Conclusion
e roles and importance of grain legumes in a context
of sustainability in agriculture could be enhanced by the
emerging research opportunities for the major topics dis-
cussed above.
A major task in the future will be the selection of leg-
ume species and cultivars which could be effectively
introduced across different cropping systems. An impor-
tant point concerns balancing yield, which gives eco-
nomic return, with the environmental and agronomic
benefits.
Some priority areas seem emerge. Nitrogen fixation
activity of grain legumes should be evaluated in relation
with soil, climatic, plant characteristics and management
conditions to find the suitable approach to achieve the
best improvements. With this respect, the ability of the
host plant to store fixed nitrogen appears to be a major
component of increasing nitrogen fixation input. A par-
ticular focus should be paid also to the study of abiotic
stress limitations and in particular water deficit, salinity
and thermal shocks require extensive investigation.
Legumes that can recover unavailable forms of soil
phosphorus could be major assets in future cropping
systems. Consequently, those legumes which are able to
accumulate phosphorus from forms normally unavailable
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 10 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
need to be further studied, since phosphorus represents
an expensive and limiting resource in several cropping
systems.
Because of the growing request for plant products, i.e.
protein and oils, and to the increased economic and envi-
ronmental pressures on agro-eco systems, it emerges that
grain legumes would play a major role in future cropping
systems.
Abbreviations
GHG: greenhouse gases; CO2: carbon dioxide; N2O: nitrous oxide; LCA: life
cycle assessment; GWP: global warming potential; BNF: biological nitrogen
fixation; NT: no tillage; CT: conventional tillage.
Authors’ contributions
FS participated in the topic literature view and selection, and in drafting
of the manuscript. AM drafted the manuscript and revised it critically. AG
participated in the topic literature view and selection, and in the drafting of
the manuscript. MP drafted the manuscript and revised it critically. All authors
read and approved the final manuscript.
Author details
1 Faculty of Biosciences and Technologies for Agriculture Food and Environ-
ment, University of Teramo, Coste Sant’Agostino, Teramo, TE, Italy. 2 Depart-
ment of Agricultural Science, University of Naples Federico II, Via Università
100, Portici, Italy.
Authors information
FS: Associate Professor of Agronomy and Crop Sciences at University of
Teramo (ITALY). AM: Associate Professor of Agronomy and Crop Sciences at
University Federico II°, Naples (ITALY). AG: Researcher in Agronomy and Crop
Sciences at University of Teramo (ITALY). MP: Full Professor of Agronomy and
Crop Sciences at University of Teramo (ITALY).
Availability of data and materials
Data will not be shared, because they care adapted from other works of
several different authors.
Competing interests
The authors declare that they have no competing interests.
Consent for publication
Each author give his personal consent for publication.
Funding
No funds were received for this work.
Received: 27 September 2016 Accepted: 15 December 2016
References
1. Alpmann D, Braun J, Schäfer BC. Analyse einer Befragung unter erfol-
greichen Körnerleguminosen anbauern im konventionellen Landbau.
Erste Ergebnisse aus dem Forschungsprojekt LeguAN. In: Wintertagung
DLG, Im Fokus: Heimische Körnerleguminosen vom Anbau bis zur
Nutzung. Berlin; 2013.
2. Angus JF, Kirkegaard JA, Hunt JR, Ryan MH, Ohlander L, Peoples MB.
Break crops and rotations for wheat. Crop Pasture Sci. 2015;66:523–52.
3. Angus JF, van Herwaarden AF, Howe GN. Productivity and break crop
effects of winter-growing oilseeds. Animal Prod Sci. 1991;31:669–77.
4. Araújo SS, Beebe S, Crespi M, Delbreil B, González EM, Gruber V, et al.
Abiotic stress responses in legumes: strategies used to cope with
environmental challenges. Crit Rev Plant Sci. 2014;34:237–80.
5. Baker JM, Ochsner TE, Venterea RT, Griffis TJ. Tillage and carbon seques-
tration-what do we really know? Agric Ecosyst Environ. 2007;118:1–5.
6. Banik P, Midya A, Sarkar BK, Ghose SS. Wheat and chickpea intercrop-
ping systems in an additive series experiment: advantages and weed
smothering. Eur J Agron. 2006;24:325–32.
7. Barbery P. Weed management in organic agriculture: are we addressing
the right issues. Weed Res. 2002;42:177–93.
8. Bayer C, Gomes J, Zanatta JA, Vieira FCB, Dieckow J. Mitigating green-
house gas emissions from a subtropical Ultisol by using long-term
no-tillage in combination with legume cover crops. Soil Tillage Res.
2016;161:86–94.
9. Beaudette C, Bradley RL, Whalen JK, McVetty PBE, Vessey K, Smith DL.
Tree-based intercropping does not compromise canola (Brassica napus
L.) seed oil yield and reduces soil nitrous oxide emissions. Agric Ecosyst
Environ. 2010;139:33–9.
10. Berry PM, Sylvester-Bradley R, Philipps L, Hatch SP, Cuttle FW, Gosling P.
Is the productivity of organic farms restricted by the supply of available
nitrogen? Soil Use Manag. 2002;18:248–55.
11. Bichel A, Oelbermann M, Voroney P, Echarte L. Sequestration of native
soil organic carbon and residue carbon in complex agroecosystems.
Carbon Manag. 2016;7:1–10.
12. Boddey RM, Jantalia CP, Zanatta JA, Conceição PC, Bayer C, Miel-
niczuk J, et al. Carbon accumulation at depth in Ferralsols under
zero-till subtropical agriculture in southern Brazil. Global Change Biol.
2010;16:784–95.
13. Brooker RW, Bennett AE, Cong W-F, Daniell TJ, George TS, Hallett PD,
et al. Improving intercropping: a synthesis of research in agronomy,
plant physiology and ecology. New Phytol. 2015;206:107–17.
14. Bues A, Preißel S, Reckling M, Zander P, Kuhlmann T, Topp K, et al. The
Environmental Role of Protein Crops in the New Common Agricultural
Policy. European Parliament, Directorate General for Internal Policies,
Policy Department B: Structural and Cohesion Policies, Agricultural and
Rural DevelopmentIP/B/AGRI/IC/2012-067; 2013. Access www.europarl.
europa.eu/studies.
15. Carranca C, Torres MO, Madeira M. Underestimated role of legume roots
for soil N fertility. Agron Sustain Dev. 2015;35:1095–102.
16. Cernay C, Ben-Ari T, Pelzer E, Meynard J-M, Makowski D. Estimating vari-
ability in grain legume yields across Europe and the Americas. Sci Rep.
2015;5:11171.
17. Chalk PM. Dynamics of biologically fixed N in legume-cereal rotations: a
review. Aust J Agric Res. 1998;49:303–16.
18. Christopher SF, Lal R. Nitrogen management affects carbon sequestra-
tion in North American cropland soils. Crit Rev Plant Sci. 2007;26:45–64.
19. Clune S, Crossin E, Verghese K. Systematic review of greenhouse
gas emissions for different fresh food categories. J Clean Prod.
2017;140:766–83.
20. Crews TE, Peoples MB. Can the synchrony of nitrogen supply and crop
demand be improved in legume and fertilizer-based agroecosystems?
A Review. Nutr Cycl Agroecosyst. 2005;72:101–20.
21. Crutzen PJ, Mosier AR, Smith KA, Winiwarter W. N2O release from agro-
biofuel production negates global warming reduction by replacing
fossil fuels. Atmos Chem Phys Discuss. 2007;7:11191–205.
22. Faghire M, Mohamed F, Taoufiq K, Faghire R, Bargaz A, Mandri B, et al.
Genotypic variation of nodules’enzymatic activities in symbiotic nitro-
gen fixation among common bean (Phaseolus vulgaris L.) genotypes
grown under salinity constraint. Symbiosis. 2013;60:115–22.
23. FAOSTAT. www.faostat.fao.org (visited 18 November 2016).
24. FAO. The State of the World’s Land and Water Resources for Food and
Agriculture (SOLAW)—Managing Systems at Risk. Food and Agriculture
Organization of the United Nations, Rome and Earthscan, London;
2011.
25. Gan Y, Hamel C, Kutcher HR, Poppy L. Lentil enhances agroecosystem
productivity with increased residual soil water and nitrogen. Renew Agr
Food Syst 2016:1–12. doi:10.1017/S1742170516000223.
26. Gao Y, Duan A, Qiu X, Liu Z, Sun J, Zhang J, Wang H. Distribution of roots
and root length density in a maize/soybean strip intercropping system.
Agric Water Manag. 2010;98:199–212.
27. Garrigues E, Corson MS, Walter C, Angers DA, van der Werf H. Soil-
quality indicators in LCA: method presentation with a case study. In:
Corson MS, van der Werf HMG, editors. Proceedings of the 8th interna-
tional conference on life cycle assessment in the agri-food sector, 1–4
October 2012, INRA, Saint Malo; 2012. p 163–68.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 11 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
28. Giambalvo D, Stringi L, Durante G, Amato G, Frenda AS. Nitrogen
efficiency component analysis in wheat under rainfed. Mediterra-
nean conditions: effects of crop rotation and nitrogen fertilization. In:
Cantero-Martínez C, Gabiña D, editors. Mediterranean rainfed agricul-
ture: strategies for sustainability. Mediterranean Agronomic Institute of
Zaragoza, Zaragoza; 2004. p. 169–73.
29. Guardia G, Tellez-Rio A, García-Marco S, Martin-Lammerding D, Tenorio
JL, Ibáñez MÁ, Vallejo A. Effect of tillage and crop (cereal versus legume)
on greenhouse gas emissions and Global Warming Potential in a non-
irrigated Mediterranean field. Agric Ecosyst Environ. 2016;221:187–97.
30. Hajduk E, Właśniewski S, Szpunar-Krok E. Influence of legume crops on
content of organic carbon in sandy soil. Soil Sci Ann. 2015;66:52–6.
31. Hassan HM, Hasbullah H, Marschner P. Growth and rhizosphere P
pools of legume–wheat rotations at low P supply. Biol Fertil Soils.
2013;49:41–9.
32. Hauggaard-Nielsen H, Jensen ES. Facilitative root interactions in inter-
crops. Plant Soil. 2005;274:237–50.
33. Hauggaard-Nielsen H, Mundus S, Jensen ES. Nitrogen dynamics fol-
lowing grain legumes and catch crops and the effects on succeeding
cereal crops. Nutr Cycl Agroecosyst. 2009;84:281–91.
34. Hernanz JL, Sanchez-Giron V, Navarrete L. Soil carbon sequestration
and stratification in a cereal/leguminous crop rotation with three tillage
systems in semiarid conditions. Agric Ecosyst Environ. 2009;133:114–22.
35. Hobbs PR, Sayre K, Gupta R. The role of conservation agriculture in
sustainable agriculture. Philos Trans R Soc Lond, Ser B. 2008;363:543–55.
36. Hocking PJ. Organic acids exuded from roots in phosphorus uptake and
aluminum tolerance of plants in acid soils. Adv Agron. 2001;74:63–97.
37. Hocking PJ, Randall PJ. Better growth and phosphorus nutrition of
sorghum and wheat following organic acid secreting crops. In: Horst
WJ, et al., editors. Proceedings of the 14th international plant nutrition
colloquium Germany. Dordrecht: Kluwer Academic Publishers; 2001. p.
548–9.
38. IPCC. Climate change 2007: Synthesis report. summary for policymak-
ers. Intergovernmental panel on climate change (IPCC); 2007.
39. Jenkinson DS. The impact of humans on the nitrogen cycle, with focus
on temperate agriculture. Plant Soil. 2001;228:3–15.
40. Jensen CR, Joernsgaard B, Andersen MN, Christiansen JL, Mogensen
VO, Friis P, Petersen CT. The effect of lupins as compared with peas and
oats on the yield of the subsequent winter barley crop. Eur J Agron.
2004;20:405–18.
41. Jensen ES, Ambus P, Bellostas N, Boisen S, Brisson N, Corre-Hellou
G, et al. Intercropping of cereals and grain legumes for increased
production, weed control, im-proved product quality and prevention
of N–losses in European organic farming systems. In: International
conferences: joint organic congress - Theme 4: crop systems and soils, 9
May 2006.
42. Jensen ES, Peoples MB, Boddey RM, Gresshoff PM, Hauggaard-Nielsen
H, Alves BJ, Morrison MJ. Legumes for mitigation of climate change
and the provision of feedstock for biofuels and biorefineries. A review.
Agron Sustain Dev. 2012;32:329–64.
43. Jensen ES, Peoples MB, Hauggaard-Nielsen H. Faba bean in cropping
systems. Field Crops Res. 2010;115:203–16.
44. Jeuffroy MH, Baranger E, Carrouée B, Chezelles ED, Gosme M, Hénault
C. Nitrous oxide emissions from crop rotations including wheat, oilseed
rape and dry peas. Biogeosciences. 2013;10:1787–97.
45. Jones SK, Rees RM, Skiba UM, Ball BC. Influence of organic and mineral
N fertiliser on N2O fluxes from a temperate grassland. Agric Ecosyst
Environ. 2007;121:74–83.
46. Kirkegaard JA, Christen O, Krupinsky J, Layzell DB. Break crop benefits in
temperate wheat production. Field Crop Res. 2008;107:185–95.
47. Kirkegaard JA, Ryan MH. Magnitude and mechanisms of persistent crop
sequence effects on wheat. Field Crops Res. 2014;164:154–65.
48. Köpke U, Nemecek T. Ecological services of faba bean. Field Crop Res.
2010;115:217–33.
49. La Favre JS, Focht DD. Conser vation in soil of H2 liberated from N2 fixa-
tion by H up-nodules. Appl Environ Microb. 1983;46:304–11.
50. Latati M, Bargaz A, Belarbi B, Lazali M, Benlahrech S, Tellah S. The inter-
cropping common bean with maize improves the rhizobial efficiency,
resource use and grain yield under low phosphorus availability. Eur J
Agron. 2016;72:80–90.
51. Legume Futures Report 4.2. Reckling M, Schläfke N, Hecker J-M,
Bachinger J, Zander P, Bergkvist G, et al. Generation and evaluation
of legume-supported crop rotations in five case study regions across
Europe; 2014. Available from www.legumefutures.de.
52. Legume Futures Report 1.6. Reckling M, Preissel S, Zander P, Topp CFE,
Watson CA, Murphy-Bokern D, Stoddard FL. Effects of legume cropping
on farming and food systems; 2014. Available from www.legumefu-
tures.de.
53. Lemke RL, Zhong Z, Campbell CA, Zentner RP. Can pulse crops play a
role in mitigating greenhouse gases from North American agriculture?
Agron J. 2007;99:1719–25.
54. Lewis G, Schrire B, Mackinder B, Lock M. Legumes of the World. Kew:
Royal Botanic Gardens; 2005.
55. Li L, Tilman D, Lambers H, Zhang F-S. Biodiversity and over yielding:
insights from belowground facilitation of intercropping in agriculture.
New Phytol. 2014;203:63–9.
56. Li L, Zhang L-Z, Zhang F-Z. Crop mixtures and the mechanisms of
overyielding. In: Levin SA, editor. Encyclopedia of biodiversity, vol. 2.
2nd ed. Waltham: Academic Press; 2013. p. 382–95.
57. Li YF, Ran W, Zhang RP, Sun SB, Xu GH. Facilitated legume nodulation,
phosphate uptake and nitrogen transfer by arbuscular inoculation
in an upland rice and mung bean intercropping system. Plant Soil.
2009;315:285–96.
58. Lithourgidis AS, Dordas CA, Damalas CA, Vlachostergios D. Annual
intercrops: an alternative pathway for sustainable agriculture. Aust J
Crop Sci. 2011;5:396.
59. Liu X, Rahman T, Song C, Su B, Yang F, Yong T, et al. Changes in light
environment, morphology, growth and yield of soybean in maize-
soybean intercropping systems. Field Crop Res. 2017;200:38–46.
60. Lopes T, Hatt S, Xu Q, Chen J, Liu Y, Francis F. Wheat (Triticum aestivum
L.)-based intercropping systems for biological pest control: a review.
Pest Manag Sci. 2016;72:2193–202.
61. López-Bellido L, Munoz-Romero V, Benítez-Vega J, Fernández-García
P, Redondo R, López-Bellido RJ. Wheat response to nitrogen splitting
applied to a Vertisols in different tillage systems and crop-ping rota-
tions under typical Mediterranean climatic conditions. Eur J Agron.
2012;43:24–32.
62. López-Bellido L, Munoz-Romero V, López-Bellido RJ. Nitrate accumula-
tion in the soil profile: long-term effects of tillage, rotation and N rate in
a Mediterranean Vertisol. Soil Tillage Res. 2013;130:18–23.
63. Luo S, Yu L, Liu Y, Zhang Y, Yang W, Li Z, Wang J. Effects of reduced nitro-
gen input on productivity and N2O emissions in a sugarcane/soybean
intercropping system. Eur J Agron. 2016;81:78–85.
64. Magid J, Henriksen O, Thorup-Kristensen K, Mueller T. Dispropor-
tionately high N-mineralisation rates from green manures at low
temperatures-implications for modelling and management in cool
temperate agro-ecosystems. Plant Soil. 2001;228:73–82.
65. Mahallati MN, Koocheki A, Mondani F, Feizi H, Amirmoradi S. Deter-
mination of optimal strip width in strip intercropping of maize (Zea
mays L.) and bean (Phaseolus vulgaris L.) in Northeast Iran. J Clean Prod.
2015;106:343–50.
66. Meynard JM, Messéan A, Charlier A, Charrier F, Farès M, Le Bail M, et al.
Crop diversification: obstacles and levers. Study of farms and supply
chains. Synopsis of the study report. INRA, Paris. 2013.
67. Midya A, Bhattacharjee K, Ghose SS, Banik P. Deferred seeding of black-
gram (Phaseolus mungo L.) in rice (Oryza sativa L.) field on yield advan-
tages and smothering of weeds. J Agron Crop Sci. 2005;191:195–201.
68. Miller PR, Gan Y, McConkey BG, McDonald CL. Pulse crops for the North-
ern Great Plains: I. Grain productivity and residual effects on soil water
and nitrogen. Agron J. 2003;95:972–9.
69. Mokgehle SN, Dakora FD, Mathews C. Variation in N2 fixation and N
contribution by 25 groundnut (Arachis hypogaea L.) varieties grown
in different agro-ecologies, measured using 15 N natural abundance.
Agric Ecosyst Environ. 2014;195:161–72.
70. Monti M, Pellicanò A, Santonoceto C, Preiti G, Pristeri A. Yield com-
ponents and nitrogen use in cereal-pea intercrops in Mediterranean
environment. Field Crop Res. 2016;196:379–88.
71. Munz S, Graeff-Hönninger S, Lizaso JI, Chen Q, Claupein W. Modeling
light availability for a subordinate crop within a strip–intercropping
system. Field Crop Res. 2014;155:77–89.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 12 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
72. Nemecek T, von Richthofen J-S, Dubois G, Casta P, Charles R, Pahl H.
Environmental impacts of introducing grain legumes into European
crop rotations. Eur J Agron. 2008;28:380–93.
73. Ngwira AR, Aune JB, Mkwinda S. On-farm evaluation of yield and
economic benefit of short term maize legume intercropping
systems under conservation agriculture in Malawi. Field Crop Res.
2012;132:149–57.
74. Papastylianou I. Effect of rotation system and N fertiliser on barley and
vetch grown in various crop combinations and cycle lengths. J Agric
Sci. 2004;142:41–8.
75. Pappa VA, Rees RM, Walker RL, Baddeley JA, Watson CA. Legumes
intercropped with spring barley contribute to increased biomass
production and carry-over effects. J Agric Sci. 2012;150:584–94.
76. Peoples MB, Boyer EW, Goulding KWT, Heffer P, Ochwoh VA, Vanlauwe B,
et al. Pathways of nitrogen loss and their impacts on human health and
the environment. In: Mosier AR, Syers KJ, Freney JR, editors. Agriculture
and the nitrogen cycle, the Scientific Committee on Problems of the
Environment (SCOPE). Covelo: Island Press; 2004. p. 53–69.
77. Peoples MB, Brockwell J, Herridge DF, Rochester IJ, Alves BJR, Urquiaga
S, et al. The contributions of nitrogen-fixing crop legumes to the pro-
ductivity of agricultural systems. Symbiosis. 2009;48:1–17.
78. Peoples MB, Hauggaard-Nielsen H, Jensen ES. The potential environ-
mental benefits and risks derived from legumes in rotations. In: Emerich
DW, Krishnan HB, editors. Nitrogen fixation in crop production. Madi-
son: American Society of Agronomy, Crop Science Society of America,
Soil Science Society of America; 2009. p. 349–85.
79. Phelan P, Moloney AP, McGeough EJ, Humphreys J, Bertilsson J,
O’Riordan E, O’Kiely P. Forage legumes for grazing and conserving in
ruminant production systems. Crit Rev Plant Sci. 2015;34:281–326.
80. Plaza-Bonilla D, Nolot J-M, Raffaillac D, Justes E. Innovative cropping
systems to reduce N inputs and maintain wheat yields by inserting
grain legumes and cover crops in southwestern France. Eur J Agron.
2016. doi:10.1016/j.eja.2016.05.010.
81. Plaza-Bonilla D, Nolot JM, Raffaillac D, Justes E. Cover crops mitigate
nitrate leaching in cropping systems including grain legumes: field
evidence and model simulations. Agric Ecosyst Environ. 2015;212:1–12.
82. Preissel S, Reckling M, Schläfke N, Zander P. Magnitude and farm-
economic value of grain legume pre-crop benefits in Europe: a review.
Field Crop Res. 2015;175:64–79.
83. Rao DLN, Giller KE, Yeo AR, Flowers TJ. The effects of salinity and sodicity
upon nodulation and nitrogen fixation in chickpea (Cicer arietinum).
Ann Bot Lond. 2002;89:563–70.
84. Reay DS, Davidson EA, Smith KA, Smith P, Melillo JM, et al. Global agri-
culture and nitrous oxide emissions. Nat Clim Change. 2012;2:410–6.
85. Reckling M, Hecker J-M, Bergkvist G, Watson CA, Zander P, Schläfke N,
et al. A cropping system assessment framework—evaluating effects of
introducing legumes into crop rotations. Eur J Agron. 2016;76:186–97.
86. Reckling M, Döring T, Stein-Bachinger K, Bloch R, Bachinger J. Yield
stability of grain legumes in an organically managed monitoring
experiment. Aspects Appl Biol. 2015;128:57–62.
87. Robson MC, Fowler SM, Lampkin NH, Leifert C, Leitch M, Robinson
D, et al. The agronomic and economic potential of breakcrops for
ley/arable rotations in temperate organic agriculture. Adv Agron.
2002;77:369–427.
88. Rochester IJ. Estimating nitrous oxide emissions from flood irrigated
alkaline grey clays. Aust J Soil Res. 2003;41:197–206.
89. Rochette P, Janzen HH. Towards a revised coefficient for estimating N2O
emissions from legumes. Nutr Cycl Agroecosyst. 2005;73:171–9.
90. Rose TJ, Julia CC, Shepherd M, Rose MT, Van Zwieten L. Faba bean is
less susceptible to fertiliser N impacts on biological N2 fixation than
chickpea in monoculture and intercropping systems. Biol Fert Soils.
2016;52:271–6.
91. Rubiales D, Fondevilla S, Chen W, Gentzbittel L, Higgins TJV, Castillejo
MA, et al. Achievements and challenges in legume breeding for pest
and disease resistance. Crit Rev Plant Sci. 2014;34:195–236.
92. Rühlemann L, Schmidtke K. Evaluation of monocropped and inter-
cropped grain legumes for cover cropping in no-tillage and reduced
tillage organic agriculture. Eur J Agron. 2015;65:83–94.
93. Rusinamhodzi L, Corbeels M, Nyamangara J, Giller KE. Maize–grain leg-
ume intercropping is an attractive option for ecological intensification
that reduces climatic risk for smallholder farmers in central Mozam-
bique. Field CropRes. 2012;136:12–22.
94. Šarūnaitė L, Deveikytė I, Kadžiulienė Ž. Intercropping spring wheat with
grain legume for increased production in an organic crop rotation.
Žemdirbystė = Agric. 2010;97:51–8.
95. Schwenke GD, Herridge DF, Scheer C, Rowlings DW, Haigh BM, McMul-
len KG. Soil N2O emissions under N2-fixing legumes and N-fertilised
canola: a reappraisal of emissions factor calculations. Agric Ecosyst
Environ. 2015;202:232–42.
96. Senbayram M, Wenthe C, Lingner A, Isselstein J, Steinmann H, Kaya
C, Köbke S. Legume-based mixed intercropping systems may lower
agricultural born N2O emissions. Energy Sustain Soc. 2016;6:2.
97. Seymour M, Kirkegaard JA, Peoples MB, White PF, French RJ. Break-crop
benefits to wheat in Western Australia-insights from over three decades
of research. Crop Pasture Sci. 2012;63:1–16.
98. Shen J, Yuan L, Zhang J, Li H, Bai Z, Chen X, et al. Phosphorus dynamics:
from soil to plant. Plant Physiol. 2011;156:997–1005.
99. Siddique KH, Johansen C, Turner NC, Jeuffroy MH, Hashem A, Sakar
D, Gan Y, Alghamdi SS. Innovations in agronomy for food legumes: a
review. Agron Sustain Dev. 2012;32:45–64.
100. Smith J, Pearce BD, Wolfe M, Martin S. Reconciling productivity with
protection of the environment: Is temperate agroforestry the answer?
Renew Agric Food Syst. 2013;28:80–92.
101. Soussana JF, Loiseau P, Vuichard N, Ceschia E, Balesdent J, Chevallier T,
Arrouays D. Carbon cycling and sequestration opportunities in temper-
ate grasslands. Soil Use Manag. 2004;20:219–30.
102. Soussana JF, Tallec T, Blanfor t V. Mitigating the greenhouse gas balance
of ruminant production systems through carbon sequestration in
grasslands. Animal. 2010;4:334–50.
103. St. Luce M, Grant CA, Zebarth BJ, Ziadi N, O’Donovan JT, Blackshaw
RE, et al. Legumes can reduce economic optimum nitrogen rates
and increase yields in a wheat-canola cropping sequence in western
Canada. Field Crop Res. 2015;179:12–25.
104. Stagnari F, Pisante M. Managing faba bean residues to enhance the fruit
quality of the melon (Cucumis melo L.) Crop. Sci. Hort. 2010;126:317–23.
105. Swinton SM, Lupi F, Robertson GP, Hamilton SK. Ecosystem services and
agriculture: cultivating agricultural ecosystems for diverse benefits. Ecol
Econ. 2007;64:245–52.
106. Tharanathan RN, Mahadevamma S. Grain legumes—a boon to human
nutrition. Trends Food Sci Tech. 2003;14:507–18.
107. Tosti G, Guiducci M. Durum wheat–faba bean temporary intercrop-
ping: effects on nitrogen supply and wheat quality. Eur J Agron.
2010;33:157–65.
108. United Nations: World population prospects: The 2012 revision, key
findings and advance tables. Working paper no. ESA/P/WP.227; 2013
(United Nations, Department of Economic and Social Affairs, Population
Division, New York).
109. Van Zwieten L, Rose T, Herridge D, Kimber S, Rust J, Cowie A, Morris S.
Enhanced biological N2 fixation and yield of faba bean (Vicia faba L.) in
an acid soil following biochar addition: dissection of causal mecha-
nisms. Plant Soil. 2015;395:7–20.
110. Vaz Patto MC, Amarowicz R, Aryee AN, Boye JI, Chung HJ, Martín-Cabre-
jas MA, Domoney C. Achievements and challenges in improving the
nutritional quality of food legumes. Crit Rev Plant Sci. 2015;34:105–43.
111. VandenBygaart A J, Gregorich EG, Angers DA. Influence of agricultural
management on soil organic carbon: a compendium and assessment
of Canadian studies. Can J Soil Sci. 2003;83:363–80.
112. Voisin AS, Guéguen J, Huyghe C, Jeuffroy MH, Magrini MB, Meynard JM,
et al. Legumes for feed, food, biomaterials and bioenergy in Europe: a
review. Agron Sustain Dev. 2014;34:361–80.
113. Wahbi S, Maghraoui T, Hafidi M, Sanguin H, Oufdou K, Prin Y, et al.
Enhanced transfer of biologically fixed N from faba bean to inter-
cropped wheat through mycorrhizal symbiosis. Appl Soil Ecol.
2016;107:91–8.
114. Walley FL, Clayton GW, Miller PR, Carr PM, Lafond GP. Nitrogen economy
of pulse crop production in the Northern Great Plains. Agron J.
2007;99:1710–8.
115. Wang Y, Marschner P, Zhang F. Phosphorus pools and other soil
properties in the rhizosphere of wheat and legumes growing in three
soils in monoculture or as a mixture of wheat and legume. Plant Soil.
2012;354:283–98.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 13 of 13
Stagnari et al. Chem. Biol. Technol. Agric. (2017) 4:2
116. Westhoek H, Rood T, van den Berg M, Janse J, Nijdam D, Reudink M,
Stehfest E. The protein puzzle. The consumption and production of
meat, dairy and fish in the European Union. Netherlands Environmental
Assessment Agency (PBL); 2011.
117. White PJ, George TS, Gregory PJ, Bengough AG, Hallett PD, McKen-
zie BM. Matching roots to their environment. Ann Bot Lond.
2013;112:207–22.
118. Xue Y, Xia H, Christie P, Zhang Z, Li L, Tang C. Crop acquisition of phos-
phorus, iron and zinc from soil in cereal/legume intercropping systems:
a critical review. Ann Bot Lond 2016;117:363–77. doi:10.1093/aob/
mcv182..
119. Yadav SS, Hunter D, Redden B, Nang M, Yadava DK, Habibi AB. Impact of
climate change on agriculture production, food, and nutritional secu-
rity. In: Redden R, Yadav SS, Maxted N, Dulloo MS, Guarino L, Smith P,
editors. Crop wild relatives and climate change. New Jersey, USA: Wiley;
2015. p. 1–23.
120. Yu Y, Xue L, Yang L. Winter legumes in rice crop rotations reduces
nitrogen loss, and improves rice yield and soil nitrogen supply. Agron
Sustain Dev. 2014;34:633–40.
121. Zander P, Amjath-Babu TS, Preissel S, Reckling M, Bues A, Schläfke N,
et al. Grain legume decline and potential recovery in European agricul-
ture: a review. Agron Sustain Dev. 2016;36:26.
122. Zentner RP, Wall DD, Nagy CN, Smith EG, Young DL, Miller PR, et al.
Economics of crop diversification and soil tillage opportunities in the
Canadian prairies. Agron J. 2002;94:216–30.
123. Zimmer S, Messmer M, Haase T, Piepho HP, Mindermann A, Schulz H,
et al. Effects of soybean variety and Bradyrhizobium strains on yield,
protein content and biological nitrogen fixation under cool growing
conditions in Germany. Eur J Agron. 2016;72:38–46.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... This legumebased nitrogen use efficiency offers a sustainable and cost-effective alternative. Pulses also foster other benefits like maintaining soil health, improving biodiversity, enhancing soil organic carbon, improving the water retention capacity of the soil, and reducing greenhouse gas emissions (Foyer et al., 2016;Peoples et al., 2019;Stagnari et al., 2017). Globally, legume crops are reported to fix about 21 metric tons of nitrogen, out of which 5 to 7 metric tons are returned to the soil by pulse crops, which saves 8 to 12 billion USD (Christine et al., 2019). ...
... Globally, legume crops are reported to fix about 21 metric tons of nitrogen, out of which 5 to 7 metric tons are returned to the soil by pulse crops, which saves 8 to 12 billion USD (Christine et al., 2019). Reports also confirmed that Field peas have a positive effect in improving soil organic carbon through their ability of nitrogen fixation (Stagnari et al., 2017). They also have a positive effect in improving cereal productivity when used as crop rotation. ...
Article
Full-text available
This study intended to measure the technical efficiency and sources of inefficiency for field pea producers in Ethiopia. The study used primary data collected from 207 smallholder farmers. A multi-stage sampling technique was followed to select sample households. A stochastic frontier was the model used, and a single-step estimation approach was followed. The result indicated that the mean technical efficiency is 49.23 %, and the average productivity of the crop is 0.96 tons/hectare. This indicates that the technical efficiency could be increased by 51.77 % through proper utilization of production variables, i.e. agrochemicals, oxen power, plot size, and seed rate. The inefficiency model result also confirmed that the technical inefficiency of field pea production was negatively affected by access to off-farm income, access to credit services, extension contacts, adoption of new technologies, access to training, and participation in contract farming of the crop. Therefore, this study recommends agricultural policies favoring vocational training to farmers on the new agricultural technology packages, availing access to off-farm incomes, facilitating credit services to farmers, and establishing legal frameworks that encourage contract farming.
... Pulses are often grown in rotation with cereal crops and provide a range of agronomic benefits that increase the sustainability of cropping systems. Such agronomic benefits include increased crop diversity, the breaking of cereal disease and weed cycles, improving soil health, fixing of atmospheric nitrogen in the soil and reducing the need for fertilisers (Boye, Zare, and Pletch 2010;Stagnari et al. 2017). ...
Article
Full-text available
Plant‐based proteins continue to gain popularity as a sustainable alternative to animal proteins due to their nutritional, functional and health benefits. Dry and wet fractionation methods are increasingly being used for the production of value‐added pulse protein ingredients. The objective of this study was to compare the nutritional properties, protein quality, physicochemical properties and secondary structure of Australian faba bean and yellow pea protein concentrates and protein isolates obtained by dry and wet fractionation. Amino acid scores highlighted that faba bean and yellow pea protein isolates and concentrates were deficient in the sulphur‐containing amino acids, methionine and cysteine and tryptophan. Faba bean (63.7%) and yellow pea protein (63.0%) isolates had higher in vitro protein digestibility‐corrected amino acid scores compared to the protein concentrates, being 50.7% and 54.4%, respectively. Fourier‐transform infrared spectroscopy revealed different secondary structures between protein concentrates and isolates, especially for the amide I region. Faba bean and yellow pea protein concentrates had higher protein solubility (46.2% and 50.1%, respectively) and higher foaming capacity (65% and 59%, respectively) compared to the protein isolates. Water‐holding capacity was higher for faba bean and yellow pea protein isolates, being 4.3% and 4.0 g/g, respectively. These findings demonstrate that faba bean and yellow pea protein concentrates and isolates have unique functional properties that can be exploited for use in a diverse range of new and existing food applications.
... Additional research has shown that lima bean can be used effectively as cover crops and green manure to improve soil quality, such as with nutrients recycling, increasing soil natural matter, decreasing soil pH, and improving soil porosity, soil structure, microscopic biodiversity, and sustainability [61,62]. The nitrogen-fixing capabilities of LB-rhizobia symbioses, as well as the plant's ability to suppress weeds and improve soil structure, make it a valuable component in sustainable agricultural systems [63]. A total of 69.6% of the fixed total nitrogen is attributable to changes in soil quality, according to an analysis of the Pearson correlation between total nitrogen and soil quality, which revealed a substantial and positive linear association. ...
Article
Full-text available
Citation: Abdulraheem, M.I.; Moshood, A.Y.; Li, L.; Taiwo, L.B.; Oyedele, A.O.; Ezaka, E.; Chen, H.; Farooque, A.A.; Raghavan, V.; Hu, J. Abstract: Background: This study explores the role of leguminous crops like lima bean in enhancing soil quality and ecosystem stability. Despite existing studies on agronomic aspects, there is a significant research gap on its impact on soil organic matter level, microbial activity, soil health, and nutrient availability. Therefore, this study examines the capacity of lima bean to reactivate soil quality, focusing on its impact on soil organic matter level, microbial activity, soil health, and nutrient availability. Methods: The experimental area was set up in 2023 using three replicates and a randomized block design. Two treatments were used: lima bean-planted plots and control plots with various weeds and without lima bean. Post-harvest soil samples were collected from various agroecological zones and sterilely packed, and physical, chemical, and biological indices were examined. Results: lima bean significantly affected nutrients, enzymes, soil microbial respiration, and other markers. Amylase activity (0.41**) was positively correlated with urease activity (0.73**), while dehydrogenase activity positively correlated with both. Dehydrogenase activity was negatively correlated with total nitrogen (0.66**) and sulfur (0.60**). Lima bean significantly affected soil quality, with all locations showing higher ratings (55-77%) than wild land, except for location D (Ilora). A total of 70% of total nitrogen variation may be attributed to soil quality (r 2 = 0.696). Lima bean enhanced soil quality, potentially enhancing productivity and reducing dependence on inorganic nitrogen inputs. Conclusions: The symbiotic relationship between lima bean and nitrogen-fixing bacteria improves nutrient cycling, enhancing agricultural productivity and environmental conservation. Future research should explore the economic viability of integrating lima bean into crop rotations or agroforestry systems for sustainable agricultural practices, providing valuable information for farmers.
... Similarly, the dominant abiotic stresses include heat, drought, salinity, acidity, frost, and others (Maalouf et al., 2021). Besides serving as a fundamental source of high-quality food and feed, legumes also highly contribute to reducing greenhouse gas emissions, the sequestration of carbon into the soils, fixing the atmospheric nitrogen, and releasing high-quality soil organic matter which helps in facilitating the circulation of soil nutrients and water retention, and this will contribute much in sustainable food production (Stagnari et al., 2017). Based on national and international data, Ethiopia is on good progress in fababean production. ...
Article
This study uses the profit frontier model to assess the profit efficiency and determinants of inefficiency using primary data collected from 334 fababean producers selected from the Amhara and Oromia regions of Ethiopia. The result confirmed that fababean production in Ethiopia is profitable as the profit efficiency score is 86.10%, indicating 14% profit inefficiency. This implies that the enhancement of all the inefficiencies will result in the achievement of the potential profit of 9,360.48 birr/ha. Sex, access to off-farm income, training on pulse production, participation in improved fababean seed production, and the total land owned by the household head were the determinant factors for profit inefficiency. This study suggests that the profit efficiency of fababean farmers will remarkably improve if vocational training is given to farmers, production is planned based on available resources to the farmers, appropriate prices are available for their produce, and households get access to off-farm incomes.
... Additional research has shown that lima bean can be used effectively as cover crops and green manure to improve soil quality, such as with nutrients recycling, increasing soil natural matter, decreasing soil pH, and improving soil porosity, soil structure, microscopic biodiversity, and sustainability [61,62]. The nitrogen-fixing capabilities of LB-rhizobia symbioses, as well as the plant's ability to suppress weeds and improve soil structure, make it a valuable component in sustainable agricultural systems [63]. A total of 69.6% of the fixed total nitrogen is attributable to changes in soil quality, according to an analysis of the Pearson correlation between total nitrogen and soil quality, which revealed a substantial and positive linear association. ...
Article
Full-text available
This study explores the role of leguminous crops like lima bean in enhancing soil quality and ecosystem stability. Despite existing studies on agronomic aspects, there is a significant research gap on its impact on soil organic matter level, microbial activity, soil health, and nutrient availability. Therefore, this study examines the capacity of lima bean to reactivate soil quality, focusing on its impact on soil organic matter level, microbial activity, soil health, and nutrient availability. The experimental area was set up in 2023 using three replicates and a randomized block design. Two treatments were used: lima bean-planted plots and control plots with various weeds and without lima bean. Post-harvest soil samples were collected from various agroecological zones and sterilely packed, and physical, chemical, and biological indices were examined. Lima bean significantly affected nutrients, enzymes, soil microbial respiration, and other markers. Amylase activity (0.41**) was positively correlated with urease activity (0.73**), while dehydrogenase activity positively correlated with both. Dehydrogenase activity was negatively correlated with total nitrogen (0.66**) and sulfur (0.60**). Lima bean significantly affected soil quality, with all locations showing higher ratings (55-77%) than wild land, except for location D (Ilora). A total of 70% of total nitrogen variation may be attributed to soil quality (r 2 = 0.696). Lima bean enhanced soil quality, potentially improving productivity and reducing dependence on inorganic nitrogen inputs. The symbiotic relationship between lima beans and nitrogen-fixing bacteria improves nutrient cycling, enhancing agricultural productivity and environmental conservation. Future research should explore the economic viability of integrating lima beans into crop rotations or agroforestry systems for sustainable agricultural practices, providing valuable information for farmers.
... Legumes, when included in cropping systems, improve soil fertility and crop yield through nitrogen fixation (Abd El-hady et al., 2022), biological nitrogen fixation (BNF)symbiotic association with microorganisms like rhizobia (Kebede, 2021). Benefits of legumes include increased nutrient availability and uptake for subsequent crops (Sinclair and Vadez, 2012;Hauggaard-Nielsen et al., 2008), improvement of soil properties (Jena et al., 2022), breaking pests' cycles (Stagnari et al., 2017), and enhancement of soil microbial activity (Yang et al., 2020). After chickpeas (Cicer arietinum L.) and peas (Pisum sativum L.), lentil is the 3 rd most important grain legume in the globe (Sehgal et al., 2021), and has acquired the first position among the pulse crops considering area (3,60,699 acres) and production (1,85,500 MT) in 2020-21 in Bangladesh (BBS, 2022) but its production lower than the neighboring countries . ...
Article
Full-text available
It is critical that Bangladesh faces water scarcity during the dry season, affecting lentil (Lens culinaris Medik.) yield and some yield components during seedling and flowering stages. Thus, a two-factorial pot experiments (The experiment comprises Factor A: three fertilization levels i.e. F1 = Control [inorganic], F2 = poultry litter-based compost [20 ton/ha], F3 = poultry litter-based compost [30 ton/ha]; Factor B: two irrigation levels such as W1 = 100% field capacity [FC] and W2 = 70% FC) were designed at Hajee Mohammad Danesh Science and Technology University, Dinajpur, from November 2018 to April 2019. And it was investigated how the poultry litter-based composts affected the morpho-physiology, yield and yield components of the lentil (BARI Masur-4) variety under different irrigation stress levels. Obtained results revealed that the tallest plant (30.7 cm at 75 DAS) and maximum branch number per plant (14.1 at 65 DAS), leaf chlorophyll a (0.30 mg/g), highest RLWC (70.28%), lowest proline content (1.57 µ moles g-1 FW), maximum number of pods per plant (39.4 at 75 DAS) and total grain yield (3.62 kg/m2) were recorded from compost F3 (poultry litter-based compost 30 tons/ha) with W1 (100% FC). Results also showed that the yield contributing attributes and yield of lentils were drastically reduced by water stress conditions with different rates of fertilization. In drought conditions (W2 = 70% FC), F3 (30 ton/ha poultry litter-based compost) fertilization produced the highest plant height (30.20 cm at 75 DAS), number of branches (11.5 at 65 DAS), stem dry weight (0.35 g), lowest proline (3.88 µ moles g-1 FW), highest pod number per plant (33.1), weight of 100-seed (2.36 g), total grain weight (2.77 kg/m2), harvest index (58.84%) compared to other fertilizations (F1 and F2). In summary, F3 (30 tons), a compost made from poultry litter, provides better soil conditions under drought conditions compared to F1 and F2 in the year of 2018-19 at the 0 and 20 tons/ha, respectively under the field conditions.
Article
Full-text available
Co-pyrolysis of chicken manure with tree bark was investigated to mitigate salinity and potentially toxic element (PTE) concentrations of chicken manure-derived biochar. The effect of tree bark addition (0, 25, 50, 75 and 100 wt%) on the biochar composition, surface functional groups, PTEs and polycyclic aromatic hydrocarbons (PAH) concentration in the biochar was evaluated. Biochar-induced toxicity was assessed using an in-house plant growth assay with Arabidopsis thaliana. This study shows that PTE concentrations can be controlled through co-pyrolysis. More than 50 wt% of tree bark must be added to chicken manure to reduce the concentrations below the European Biochar Certificate-AGRO (EBC-AGRO) threshold. However, the amount of PAH does not show a trend with tree bark addition. Furthermore, co-pyrolysis biochar promotes plant growth at different application concentrations, whereas pure application of 100 wt% tree bark or chicken manure biochar results in decreased growth compared to the reference. In addition, increased plant stress was observed for 100 wt% chicken manure biochar. These data indicate that co-pyrolysis of chicken manure and tree bark produces EBC-AGRO-compliant biochar with the potential to stimulate plant growth. Further studies need to assess the effect of these biochars in long-term growth experiments.
Article
Interactions between arbuscular mycorrhizal fungi (AMF), plants, and the soil microbial community have the potential to increase the availability and uptake of phosphorus (P) and nitrogen (N) in agricultural systems. Nutrient exchange between plant roots, AMF, and the adjacent soil microbes occurs at the interface between roots colonized by mycorrhizal fungi and soil, referred to as the mycorrhizosphere. Research on the P exchange focuses on plant–AMF or AMF–microbe interactions, lacking a holistic view of P exchange between the plants, AMF, and other microbes. Recently, N exchange at both interfaces revealed the synergistic role of AMF and bacterial community in N uptake by the host plant. Here, we highlight work carried out on each interface and build upon it by emphasizing research involving all members of the tripartite network. Both nutrient systems are challenging to study due to the complex chemical and biological nature of the mycorrhizosphere. We discuss some of the effective methods to identify important nutrient processes and the tripartite members involved in these processes. The extrapolation of in vitro studies into the field is often fraught with contradiction and noise. Therefore, we also suggest some approaches that can potentially bridge the gap between laboratory-generated data and their extrapolation to the field, improving the applicability and contextual relevance of data within the field of mycorrhizosphere interactions. Overall, we argue that the research community needs to adopt a holistic tripartite approach and that we have the means to increase the applicability and accuracy of in vitro data in the field.
Article
Full-text available
This study provides an overview of the development and environmental effects of protein crop production in Europe. Nine policy options for supporting protein crops are presented: six inside the CAP, and three outside. We recommend an integrated policy approach combining the inclusion of protein crops into greening measures, investment in research and constraints on the use of synthetic nitrogen fertiliser. We conclude that increasing the production of protein crops would be an important contribution to the sustainable development of European agricultural and food systems.
Article
Full-text available
Knowing short-term gains and losses of soil organic carbon (SOC) is crucial for understanding the role of different land management practices in climate change mitigation. This study evaluated the flow of carbon (C) in soil from two differently configured intercrops [1:2 (one row of maize and two rows soybean); 2:3 (two rows of maize and three rows of soybean)] compared to a maize and soybean sole crop as a result of residue addition. Addition of soybean or maize residues significantly increased (p < 0.05) SOC, light fraction (LF-C), and soil microbial biomass (SMB). Soil organic C from native sources was significantly greater (p < 0.05) than C from new (residue) sources. The LF had a significantly greater (p < 0.05) C content from new sources. Treatments amended with soybean residue had a significantly greater (p < 0.05) contribution from new C sources for SOC and LF than treatments amended with maize residue. The SMB-C was significantly greater (p < 0.05) in the 2:3 intercrop. Cumulative soil CO2 emission was significantly lower (p < 0.05) in intercrops than in sole crops. CO2 emissions derived from new C sources was significantly greater (p < 0.05) than that derived from native sources in maize amended treatments; and not significantly different (p < 0.05) for treatments amended with soybean residues.
Chapter
The strategy of most concern in this chapter is the importance of agricultural biodiversity, utilization, and maintenance of plant genetic resources for crop improvement and diversification of agricultural and food systems. This chapter reviews the current global food security context and the need to feed a growing global population with limited access to natural resources in the context of significant climate change. Considering the nature of this chapter, a brief summary of area, production, and productivity of major field crops is provided. Creating more options for climate change adaptation and improving the adaptive capacity in the agricultural sector will be crucial for improving food security and preventing an increase in global inequality in living standards in the future. The chapter talks about the availability of crop wild relatives (CWR) of various field crops for utilization in regular crop breeding programs for the development of new varieties.
Article
The maize-soybean intercropping system has become increasingly popular in many areas of the world, particularly in China, due to its high productivity and the harvest of two different grains. While efforts have been made to maintain the yield of the taller maize crop, there is limited understanding of how the morphology, growth and yield of the lower soybean crop changes in response to the shading by maize. We therefore conducted a three-year field experiment from 2013 to 2015 to investigate the changes in light environment, growth of individual organs, biomass, and grain yield of soybean under two intercropping patterns (1M1S, one row of maize with one row of soybean; 2M2S, two rows of maize with two rows of soybean) as compared to monoculture. Our results showed that at soybean flowering stage, the R:FR ratio at the top of soybean canopy was reduced 17–21% more than the photosynthetically active radiation (PAR) under intercropping compared to monoculture, with 15–19% more reduction under 1M1S than 2M2S. This led to increased internode lengths, plant height and specific leaf area (SLA), but reduced branching of soybean plants under intercropping. These morphological changes enabled the crop to intercept relatively more light and the shading also increased the light use efficiency (LUE) of soybean. However, these positive responses were not able to compensate the effect of reduced leaf area (due to smaller leaf size and less branching) and total light interception, leading to reduced biomass and grain. The reduction in grain yield was mainly caused by the reduced number of grains (particularly on the middle nodes) produced by the intercropped soybean plants, while the grain size remained unchanged. The data and results of this study may be used to develop and parameterize crop models for simulating development and growth of soybean crop in response to changes in the light environment under intercropping.
Article
Lentil ( Lens culinaris Medikus) may have a potential to enhance the productivity of agroecosystems in dry areas where water and nutrients are limited. This study quantified soil water, residual soil nitrogen (N), and crop yields in lentil-based systems in comparison with continuous cereal and conventional summerfallow systems. A 3-yr cropping sequence study was conducted for three cycles in Saskatchewan (50.28°N, 107.79°W) from 2007 to 2011. On average, soil retained 187, 196 and 337 mm of water in the 0–1.2 m depth at crop harvest in 2008, 2009 and 2010, respectively. Summerfallow contained the same amounts of water as the cropped treatments at the harvest in 2009 and 2010. However, in 2008, summerfallow contained more soil water than the cropped treatments. The effect of lentil cultivar on soil water conservation varied with years; the cultivars Glamis, Laird and Sedley conserved highest amounts of soil water by the planting time of 2009 and 2010, but no differences were found among cultivars in 2011. Soil available N (NO 3 − + NH 4 + ) at spring planting time was 50.4 kg ha −1 in the preceding lentil treatments, which was 44% higher compared with preceding barley or flax, but was 25% lower compared with preceding summerfallow. Lentil cultivars had a similar amount of soil residual N. Grain production in the 3-yr rotation averaged 6.3 t ha −1 per rotation for the wheat–lentil–durum system and 6.8 t ha −1 for the wheat–cereal–durum monoculture, averaging 36% greater compared with wheat–summerfallow–durum system. The lentil system increased total grain production through the access of residual soil water and biologically fixed N, whereas continuous cereal system relies on inorganic fertilizer input for yield. Summerfallow system relies on ‘mining’ the soil for nutrients. We conclude that the adoption of lentil systems will enhance grain production through the use of residual soil water and available N.