Symbiotic dinitrogen fixation by trees: an

0 downloads 0 Views 913KB Size Report
Nov 16, 2012 - super family) with certain a- and b-Proteobacteria. (rhizobia), are ..... diversity of rhizobia (Bala and Giller 2001; Odee et al. 1995 .... Casida (1982). Dreyfus et ...... The contribution of PN was funded by the Academy of Finland.
Symbiotic dinitrogen fixation by trees: an underestimated resource in agroforestry systems? Pekka Nygren, María P. Fernández, Jean-Michel Harmand & Humberto A. Leblanc Nutrient Cycling in Agroecosystems (formerly Fertilizer Research) ISSN 1385-1314 Volume 94 Combined 2-3 Nutr Cycl Agroecosyst (2012) 94:123-160 DOI 10.1007/s10705-012-9542-9

1 23

Your article is protected by copyright and all rights are held exclusively by Springer Science +Business Media Dordrecht. This e-offprint is for personal use only and shall not be selfarchived in electronic repositories. If you wish to self-archive your work, please use the accepted author’s version for posting to your own website or your institution’s repository. You may further deposit the accepted author’s version on a funder’s repository at a funder’s request, provided it is not made publicly available until 12 months after publication.

1 23

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 DOI 10.1007/s10705-012-9542-9

REVIEW ARTICLE

Symbiotic dinitrogen fixation by trees: an underestimated resource in agroforestry systems? Pekka Nygren • Marı´a P. Ferna´ndez • Jean-Michel Harmand • Humberto A. Leblanc

Received: 24 March 2012 / Accepted: 24 October 2012 / Published online: 16 November 2012 Ó Springer Science+Business Media Dordrecht 2012

Abstract We compiled quantitative estimates on symbiotic N2 fixation by trees in agroforestry systems (AFS) in order to evaluate the critical environmental and management factors that affect the benefit from N2 fixation to system N economy. The so-called ‘‘N2-fixing tree’’ is a tripartite symbiotic system composed of the plant, N2-fixing bacteria, and mycorrhizae-forming fungi. Almost 100 recognised rhizobial species associated with legumes do not form an evolutionary homologous clade and are functionally diverse. The global bacterial diversity is still unknown. Actinorrhizal symbioses in AFS remain almost unstudied. Dinitrogen fixation in AFS should be quantified using N isotopic methods or

P. Nygren Department of Forest Sciences, P.O. Box 27, 00014 University of Helsinki, Finland P. Nygren (&) Finnish Society of Forest Science, P.O. Box 18, 01301 Vantaa, Finland e-mail: [email protected] M. P. Ferna´ndez Ecologie Microbienne, UMR5557, USC 1193, Universite´ Lyon1, 43 boulevard du 11 novembre 1918, 69622 Villeurbanne Cedex, France J.-M. Harmand CIRAD, UMR Eco&Sols, 2 Place Viala, 34060 Montpellier Cedex 01, France H. A. Leblanc EARTH University, 4442-1000 San Jose´, Costa Rica

long-term system N balances. The general average ± standard deviation of tree dependency on N2 fixation (%Ndfa) in 38 cases using N isotopic analyses was 59 ± 16.6 %. Under humid and sub-humid conditions, the percentage was higher in young (69 ± 10.7 %) and periodically pruned trees (63 ± 11.8 %) than in freegrowing trees (54 ± 11.7 %). High variability was observed in drylands (range 10–84 %) indicating need for careful species and provenance selection in these areas. Annual N2 fixation was the highest in improved fallow and protein bank systems, 300–650 kg [N] ha-1. General average for 16 very variable AFS was 246 kg [N] ha-1, which is enough for fulfilling crop N needs for sustained or increasing yield in low-input agriculture and reducing N-fertiliser use in large-scale agribusiness. Leaf litter and green mulch applications release N slowly to the soil and mostly benefit the crop through long-term soil improvement. Root and nodule turnover and N rhizodeposition from N2-fixing trees are sources of easily available N for the crop yet they have been largely ignored in agroforestry research. There is also increasing evidence on direct N transfer from N2fixing trees to crops, e.g. via common mycelial networks of mycorrhizal fungi or absorption of tree root exudates by the crop. Research on the below-ground tree-cropmicrobia interactions is needed for fully understanding and managing N2 fixation in AFS. Keywords 15N  Actinorrhizal trees  Legume trees  Management practices  Nitrogen balance  Rhizobial symbiosis

123

Author's personal copy 124

Introduction Nitrogen is the first plant growth-limiting factor after water in most ecosystems. Agroecosystems may be even more N limited than natural ecosystems because of heavy N export in crop harvest (Nair et al. 1999). The atmospheric N2 is the biggest pool of N in the world but only some prokaryotic microbes are able to reduce it, thus playing a key role in both terrestrial and marine ecosystems. Crops that form symbiosis with N2-fixing microbes, most notably legumes (Fabaceae super family) with certain a- and b-Proteobacteria (rhizobia), are an alternative to cope with N deficiencies in agroecosystems. Legume crops and forages respond only a part of human needs of plant food and fibre and, thus, vast agricultural areas depend on industrially-fixed N fertilizers. These are often unavailable for small-scale farmers in the developing world and may cause environmental problems such as contamination of water sources. Agroforestry systems (AFS) with ‘‘N2fixing trees’’1 provide alternatives to alleviate these problems if managed properly. These systems are diverse including but not restricted to cultivation of cereals and other crops between rows of periodicallypruned trees in alley cropping (Akinnifesi et al. 2010; Kang et al. 1981; Rowe et al. 1999), shade trees with perennial crops (Beer et al. 1998; Soto-Pinto et al. 2010), improvement of fallow phase with N2-fixing trees (Chikowo et al. 2004; Harmand et al. 2004; Sta˚hl et al. 2002, 2005), living supports for climbing crops (Salas et al. 2001), and simultaneous cultivation of fodder trees and grass (Blair et al. 1990; Dulormne et al. 2003). Many tree species used in AFS provide multiple products including fuelwood, fodder, or several nontimber forest products. Dinitrogen-fixing trees are often preferred in comparison to other multi-purpose species because of the assumed benefit to the whole system N balance. Thus, symbiotic N2 fixation in AFS was enthusiastically studied in the 1970s and 1980s but it almost disappeared from research agenda in the

1

In fact, no tree fixes atmospheric N2 because all organisms capable of N2 fixation are Bacteria or Archae. However, in order to avoid repeating the long correct expression ‘‘trees forming N2-fixing symbiosis with bacteria’’ we use the common though inaccurate term ‘‘N2-fixing trees’’ for referring to these trees as a group.

123

Nutr Cycl Agroecosyst (2012) 94:123–160

1990s because of studies suggesting little benefit of N2 fixation to system level N balance in AFS (Fassbender 1987; Garrity and Mercado 1994; van Kessel and Roskoski 1981) or sustainability (Kass 1995). Recent research based on N isotopic relations in whole plant (Leblanc et al. 2007; Peoples et al. 1996; Sta˚hl et al. 2002, 2005) or compilation of whole AFS N balance (Dulormne et al. 2003) indicate that symbiotic N2 fixation may have been underestimated as a N source for AFS. The N2-fixing symbiosis is regulated by both intrinsic and environmental factors. The intrinsic physiological and morphological factors form the basis of the functional plant groups as they result in differences in resource requirements, seasonality of growth, and life history (Tilman et al. 1997), i.e. responses to the environment and interactions with other organisms. In spite of the recent advances in biotechnology, our ability to manage the intrinsic factors of plants and N2-fixing bacteria are still quite limited. In practical agroforestry, these intrinsic factors may be taken into account only by selecting suitable tree species (Aronson et al. 2002) and combinations of trees and bacterial strains (AcostaDuran and Martı´nez-Romero 2002; Bala and Giller 2001; Bala et al. 2003). The macroenvironment is out of human control, in spite of activities such as development of vegetationbased C sequestration to mitigate the global climate change (Soto-Pinto et al. 2010). On the other hand, agroforestry offers a wide variety of tools for managing the microenvironment for better sustainability and productivity, starting with simple techniques like optimisation of tree spacing in alley cropping for best N benefit and minimal crop shading (Akinnifesi et al. 2008). It is also important to understand the effects of the agroforestry management on the N2-fixing symbiosis; e.g., the green pruning of trees practiced for increasing nutrient recycling and reducing crop shading may also disturb nodulation (Nygren and Ramı´rez 1995). Although several reviews on N2 fixation in AFS have been published (Bryan 2000; Giller 2001; Kass et al. 1997; Khanna 1998; Mafongoya et al. 2004; Sanginga et al. 1995), they are mostly descriptive compilations of relevant data. Thus, critical analysis on issues such as the methods used in N2 fixation research within the agroforestry context, functional importance of N2-fixing symbiosis for the legumes and actinorrhizal plants, and effects of the different AFS

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

125

management practices on symbiotic N2 fixation seems timely. Recent research (Andre´ et al. 2005; Cardoso and Kuyper 2006; Duponnois and Plenchette 2003; Ingleby et al. 2001; Lesueur and Sarr 2008) also indicates the importance of the tripartite symbiosis between plants, N2-fixing bacteria, and mycorrhizaeforming fungi on N2 fixation. Our aim is to analyse symbiotic N2 fixation by agroforestry trees as a part of their functions and indicate the critical environmental and management aspects needed to benefit from N2 fixation at AFS level. We also evaluate the research methodologies and compile quantitative data on N2 fixation in AFS based on the most reliable methods only. The specific objectives of the review are: revise current knowledge on N2-fixing microbia relevant for understanding the functioning of AFS; evaluate the methods for estimating N2 fixation suitable for use in AFS; evaluate the published data on symbiotic N2 fixation in AFS in the light of current knowledge on ecophysiology of legumes and actinorrhizal plants; and evaluate the importance of symbiotic N2 fixation for AFS functions and productivity.

Dinitrogen-fixing organisms

Biological N2 fixation and N2-fixing organisms

Dinitrogen-fixers have been reported among most of the taxonomic divisions of Prokaryota and the methanogenic Archae, thus presenting large genetic and physiological diversity. The N2-fixing bacteria are currently divided into symbiotic and free-living N2fixers according to their capacity to form mutualistic association with eukaryotic higher organisms, either plants or animals, or their saprophytic life as components of environmental microflora. Most symbiotic N2fixing bacteria also have a saprophytic stage. However, some symbiotic strains that remain non-isolated could represent obligate symbionts or intermediate stages in evolution towards a greater symbiotic dependence. Dinitrogen-fixers associated with plants comprise 3 main groups: (1) Cyanobacteria that establish ectosymbiosis (non-intracellular location) with a large diversity of fungi and various plant groups, including mostly bryophytes and cycads but very few higher plants; (2) a- and b-Proteobacteria that form symbiosis in specialised structures, nodules, within roots or stems of legumes and Parasponia spp.; and (3) Frankiaceae (Actinobacteria) that nodulate plants belonging to 25 genera within eight Angiosperm families. Plants forming symbiosis with Frankiaceae are collectively called actinorrhizal plants. They are mostly trees or shrubs. (Vessey et al. 2004).

Dinitrogen fixation process

Cyanobacteria and other free-living diazotrophs

All N2-fixing Prokaryota reduce atmospheric N2 using the nitrogenase enzyme:

Taxonomy of Cyanobacteria is problematic (Castenholz 2001) and recent molecular studies indicate considerable biodiversity. Among plants symbiotically associated with Cyanobacteria, only the aquatic ferns of the genus Azolla have economic importance in farming systems (Herridge et al. 2008) and none is important in AFS. Endophytic free-living and associative diazotrophic bacteria within the rhizosphere of Gramineae may make substantial contributions to N balance of an agroecosystem. The best-studied case is probably sugar cane (Saccharum spp.; Baldani et al. 1997), which is never used in AFS as far as we know. Some other plants known to harbour N2-fixing endophytes such as forage grasses (Herridge et al. 2008), coffee (Coffea arabica L.; Fuentes-Ramı´rez et al. 2001; Jime´nez-Salgado et al. 1997), and banana (Musa spp.; Martı´nez et al. 2003) are often used in AFS. As far as we know, the role of the free-living N2-fixing bacteria in crop N supply has not been estimated in any AFS.

N2 þ 8Hþ þ 8e þ 16ATP

nitrogenase ! 2NH3 þ H2

þ 16ADP þ 16Pi ð1Þ Thus, reducing a N2 molecule to two NH3 molecules requires 8 protons (H?) and electrons (e-) and it releases a hydrogen molecule (H2). The energy for N2 fixation is released by breaking 16 adenosinetriphosphate (ATP) molecules to adenosinediphosphate (ADP) and inorganic phosphate (Pi). Details of the N2 fixation process can be found in most textbooks of plant physiology (e.g. Taiz and Zeiger 2006) or soil microbiology (e.g. Paul and Clark 1996) and they will not be repeated here, except for issues needed for better understanding the discussion on symbiotic N2 fixation in AFS.

123

Author's personal copy 126

Symbiotic N2-fixing organisms The nodules within which the symbiotic N2 fixation occurs are a plant organ. Nodule growth is stimulated by the symbiotic N2-fixing bacteria, and it contains a bacteroid space where the actual N2 fixation occurs (Minchin 1997). The symbiotic bacteria survive as free-living in the soil but many fix N2 only in symbiosis with a host plant. The endophytic bacteria, called bacteroids, lose part of their cellular organelles and, thus, become completely dependent on the host plant. Synthesis and catalytic activity of the nitrogenase enzyme are inhibited by oxygen. Consequently, the bacteroid space is anaerobic, protected by an O2 diffusion barrier in the inner cortex of the legume nodules (Minchin 1997) or by specialised structures, such as vesicles of Frankia spp. The bacteroids are, however, aerobic and the transport compound leghemoglobin supplies O2 to the bacteroids without damaging the nitrogenase. Ammonia produced by symbiotic N2 fixation is assimilated to amino acids or ureides in the root nodules and transferred in this form to the plant’s vascular system (Vessey et al. 2004).

Dinitrogen-fixing bacteria in agroforestry systems Legume symbioses The bacterial species associated with legumes described so far are diverse and do not form an evolutionary homologous clade. They are classified into 13 genera belonging to two distinct phylogenetic branches where rhizobia (i.e. bacteria forming N2fixing nodules with legumes) are intermingled with many non-symbiotic bacteria. Earlier, rhizobia were assumed to belong to five genera of a-Proteobacteria recognized as Azorhizobium, Bradyrhizobium, Rhizobium, Mesorhizobium, and Ensifer (formerly Sinorhizobium) (Young and Haukka 1996; Zakhia and de Lajudie 2001). Besides these genera, several other a-Proteobacteria (Allorhizobium, Devosia, Methylobacterium, Ochrobactrum, Phyllobacterium, and Shinella spp.) and b-Proteobacteria (Burkholderia, Cupriavidus (ex. Waustersia ex. Ralstonia), and Herbaspirillum spp.) have been isolated from legume nodules (reviewed by Balachandar et al. 2007). Several of these species (Blastobacter, Burkholderia, Cupriavidus, Devosia,

123

Nutr Cycl Agroecosyst (2012) 94:123–160

Methylobacterium, Ochrobactrum, Phyllobacterium, and Shinella spp.) were recently shown to be rhizobia by the presence of nod and nif genes, which encode nodulation and N2 fixation. Some bacteria found in nodules, such as Agrobacterium spp., were shown to be devoid of nodulation and N2 fixation genes. These strains often form mixed populations with nodulating rhizobial strains in the nodules (reviewed by Balachandar et al. 2007). An up-to-date list of rhizobial taxa with recommended nomenclature is maintained by the ‘‘ICSP Subcommittee on the taxonomy of Rhizobium and Agrobacterium’’ (http://edzna.ccg. unam.mx/rhizobial-taxonomy/node/4) or the New Zealand rhizobia website (http://www.rhizobia.co.nz/ taxonomy/rhizobia.html). In July 2012, the New Zealand rhizobia website recognised 98 rhizobial species. Nevertheless, the global diversity is still unknown because less than 30 % of the 728 genera and the ca. 19,325 species of the Fabaceae have been tested for nodulation (Sprent 2009). Although considerable effort has been made during the last 20 years to describe rhizobial diversity from unstudied wild legumes, the symbionts have been identified to species level only in a few studies. Many of the recently described ‘‘new rhizobia’’ have been isolated from tropical legume trees. They seem to be a source of large microbial diversity, particularly among the fast-growing rhizobia. In an extensive numerical analysis of 115 phenotypic characteristics of ca. 130 rhizobial strains, 12 clusters composed of tree rhizobia were found among the 19 clusters obtained (Zhang et al. 1991). An important diversity is found in the diversification centres of the legumes, most of which are tropical (Lie et al. 1987); e.g. in the sub-Saharan African centre of diversity of legume trees (The Sudan, Ethiopia, and Kenya), a wide phylogenetic bacterial diversity was found in a relatively low number of host species (Nick et al. 1999; Odee et al. 1997, 2002; Wolde-Meskel et al. 2005; Zhang et al. 1991). These studies suggest that studying symbionts of unexplored wild legumes from new biogeographical areas will reveal additional diversity. Before the modern phylogenetic techniques became widely available, rhizobia were commonly classified to fast- and slow-growing strains; the latter were later classified to the Bradyrhizobium genus. It seems that often a given tree species recognises preferentially either slow- or fast-growing symbionts (Turk and Keyser 1992; Wolde-Meskel et al. 2005);

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

e.g., strains isolated from Acacia senegal (L.) Willd. and Prosopis chilensis (Molina) Stuntz in the Sudan were all fast-growing whereas strains from Acacia mangium Willd. in Thailand were slow-growing (Zhang et al. 1991). The phenomenon has not been widely studied but Wolde-Meskel et al. (2005) noted that in a few cases a tree species was nodulated with both fast- and slow-growing rhizobia. However, the classification to slow- or fast-growing strains may not be genetically relevant; e.g., fast-growing strains from Inga edulis Mart. were genetically associated with slow-growing bradyrhizobia (Leblanc et al. 2005). Because there is no specific selective medium for isolating rhizobia directly from soil, most of the biodiversity studies have been conducted sampling nodules from field-grown plants or from trap plant experiments. The natural rhizobial populations obtained from field-collected nodules seem to be more diverse than the trapped ones (Liu et al. 2005; Wang et al. 1999, 2002a). Indeed, the latter approach only allows studying compatible nodulating strains under the specific conditions used and estimates depend mostly of what and how many species are used for trapping, thus resulting in an underestimation of the global biodiversity. In addition to taxonomic diversity, rhizobia that nodulate legume trees are functionally very diverse, i.e. with respect to their ecological, physiological, and biochemical properties (Zhang et al. 1991). Particularly, their cross-nodulation patterns vary remarkably (Table 1). Fast-growers seem to be more specific in comparison to slow-growers that are generally found promiscuous. Each rhizobial species has a defined host range, varying from very narrow to very broad (Duhoux and Dommergues 1985; Graham and Hubbell 1975; Trinick 1982); e.g. Calliandra calothyrsus Meisn. (sub-family Mimosoideae, tribe Ingeae), Leucaena leucocephala (Lam.) de Wit (Mimosoideae: Mimoseae), and Gliricidia sepium (Jacq.) Kunth ex Walp. (Papilionoideae: Robineae) were able to nodulate in most of the soils tested indicating that their symbiotic partners are widely distributed (Bala and Giller 2001). These phylogenetically diverse hosts also shared their symbionts, because same rhizobial species were isolated from their nodules (Bala and Giller 2001; Moreira et al. 1998; Oyaizu et al. 1993). Sesbania sesban (L.) Merr. (Papilionoideae: Sesbanieae) is a common species in AFS that failed to nodulate in most of the soils tested suggesting a higher symbiotic specificity (Bala and Giller 2001).

127

The specificity apparently varies depending on the level, at which it is analysed (nodulation, effectiveness, or both), and the applied methodology. Rhizobia may be promiscuous for nodulation and have high specificity for effectiveness as was demonstrated in Robinia pseudoacacia L., Acacia mearnsii De Wild. (Turk and Keyser 1992), and A. mangium (Galiana et al. 1990; Prin et al. 2003). Globally, there is no obvious correlation between phylogenies of rhizobia and the hosts, from which they were isolated. A given tropical legume tree may be nodulated by a wide diversity of rhizobia (Bala and Giller 2001; Odee et al. 1995, 1997) and several rhizobia isolated from tropical trees are also able to nodulate herbaceous legumes (Herrera et al. 1985; Zhang et al. 1991). Actinorrhizal symbiosis All N2-fixing actinorrhizal symbionts belong to the unique genus Frankia, which forms nodules in the roots of ca. 280 nodulating non-legume species that belong to 25 plant genera. This low number of species and genera compared with legumes does not imply low genetic diversity at the plant level because eight Angiosperm families are concerned (Dawson 2008). The main species used in tropical AFS belong to the genera Casuarina and Allocasuarina (Casuarinaceae), and Alnus (Betulaceae). Alnus spp. and Hippophae rhamnoides L. (Elaeagnaceae) are also used in some temperate AFS. The first successful isolation of a Frankia sp. occurred only 30 years ago. Today at least 11 species have been reliably isolated and identified, and several others require verification. However, the majority of Frankia diversity remains to be described because almost half of the known Frankia strains have not been successfully grown ex-planta. Modern methods of molecular biology allow the characterisation of strains directly from the nodules, thus avoiding the limiting culture step. Due to the difficulties to cultivate Frankia spp. and lack of reproducibility of some classical taxonomical methods, most of the Frankia genomic species described are not yet named. The only recognised named species is F. alni (Table 1). Frankia spp. are probably distributed in all continents except Antarctica under very diverse soil and environmental conditions, including areas devoid of actinorrhizal plants. The capacity of Frankia spp. to fix N2 saprophytically and sporulate may explain their survival out of the normal distribution of the host

123

Author's personal copy 128 Table 1 Symbiotic N2fixing bacteria found with legume and actinorrhizal trees commonly used in agroforestry systems

Nutr Cycl Agroecosyst (2012) 94:123–160

Host plant genus

Symbionts

References

Sinorhizobium fredii

Wolde-Meskel et al. (2005)

S. terangae

Boivin and Giraud (1999)

S. saheli

Lortet et al. (1996)

S. kostiense

Nick et al. (1999)

S. americanum

Toledo et al. (2003)

S. arboris

Nick et al. (1999)

Ensifer mexicanus

Llore al. (2007)

Allorhizobium undicola

De Lajudie et al. (1998)

Mesorhizobium plurifarium

De Lajudie et al. (1998)

M. huakuii

Sprent (2009)

Bradyrhizobium sp.

Dupuy et al. (1994)

Albizia spp.

M. albiziae

Wang FQ et al. (2007)

Calliandra spp.

Rhizobium tropici

Martı´nez-Romero et al. (1991)

R. gallicum

Zurdo-Pin˜eiro et al. (2004)

R. mongolense

Wolde-Meskel et al. (2005)

Rhizobium sp.

Zhang et al. (1991)

Legume tree genera Acacia spp.

Erythrina spp. Faidherbia spp.

Gliricidia spp.

Inga spp. Leucena spp.

Prosopis spp.

Robinia pseudoacacia Sesbania spp.

123

B. liaoningense

Wolde-Meskel et al. (2005)

Allorhizobium undicola

De Lajudie et al. (1998)

Ochrobactrum sp.

Ngom et al. (2004)

B. elkanii R. etli

Wolde-Meskel et al. (2005) Herna´ndez-Lucas et al. (1995) Herna´ndez-Lucas et al. (1995)

R. tropici Sinorhizobium sp.

Acosta-Dura´n and Martı´nez-Romero (2002)

B. japonicum

Leblanc et al. (2005)

B. liaoningense

Leblanc et al. (2005)

R. tropici

Martı´nez-Romero et al. (1991)

R. etli biovar phaseoli R. gallicum

Segovia et al. (1993) Herna´ndez-Lucas et al. (1995)

R. giardinii

Amarger et al. (1997)

S. morelense

Wang ET et al. (2002b)

M. plurifarium

De Lajudie et al. (1998)

M. albiziae

Wang FQ et al. (2007)

M. plurifarium M. chacoense

De Lajudie et al. (1998) Vela´squez et al. (2001)

S. kostiense

Sprent (2009)

S. arboris

Sprent (2009)

R. multihospitium

Han et al. (2008)

R. huautlense

Wang ET et al. (1998),

S. saheli

De Lajudie et al. (1994)

S. terangae

De Lajudie et al. (1994)

Ensifer adhaerens

Casida (1982)

Azorhizobium caulinodans

Dreyfus et al. (1988)

Azorhizobium johannense

Moreira et al. (2000)

S. meliloti

Wolde-Meskel et al. (2005)

S. fredii

Wolde-Meskel et al. (2005)

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 Table 1 continued

Host plant genus Tephrosia candida

129

Symbionts

References

R. hainanense

Chen et al. (1997)

3 genomic species

An et al. (1985) Ferna´ndez et al. (1989)

Actinorrhizal tree genera Alnus spp.

(3 to be confirmed)

Akimov and Dobritsa (1992) Shi and Ruan (1992) Casuarina spp.

1 genomic species

Ferna´ndez et al. (1989)

Elaeagnus spp and Hippophae spp

5 genomic species

Ferna´ndez et al. (1989)

(5 to be confirmed)

Akimov and Dobritsa (1992) Lumini et al. (1996) An et al. (1985)

plants (Burleigh and Dawson 1994; Maunuksela et al. 2000, 2006; Paschke and Dawson 1992). Consequently, many actinorrhizal plants are known for their high capacity to acclimate to environments different from their natural range and they are, thus, widely used as exotics, e.g. in plantation forestry. The nodulation out of the natural range is generally good in Alnus spp., Myricaceae, and Elaeagnaceae (Dawson 2008). Family Casuarinaceae that originates from Australia and New Zealand and has been introduced in Africa, the Americas, and Asia, is a good example on how distribution history can bias the knowledge about strain diversity and plant-strain specificity. The first studies indicated a strong genetic homogeneity of Casuarina-nodulating strains at both intra- and interspecific level all around the world (Ferna´ndez et al. 1989; Nazaret et al. 1989; Rouvier et al. 1992). However, these studies concerned only the introduction areas and often the most widely distributed species Casuarina equisetifolia L. Further studies within the native range indicated genetic diversity of the strains and a high specificity between the plant species and the symbiont genotype (Rouvier et al. 1992). Globally, the actinorrhizal symbioses vary from very promiscuous to very specific associations even within a family, like in the case of Myricaceae (Huguet et al. 2005).

Estimation of N2 fixation Need to quantify N2 fixation When preparing this review, it became evident that symbiotic N2 fixation is often cited in agroforestry literature without actual estimates on N2 fixation rate

or contribution of N2 fixation to plant or system N economy. In order to really evaluate the N benefit from N2-fixing trees to an AFS, N2 fixation must be estimated. Not all legume trees fix N2, e.g. Senna spp. that are common in AFS (Ladha et al. 1993; Sta˚hl et al. 2005). Non-nodulating species are most common among the Caesalpinoideae but exist also in Mimosoideae and Papilionoideae (Sprent 2009). Further, tree management may have important effects on nodulation and N2 fixation by trees reported as N2fixers (Nygren and Ramı´rez 1995). Several excellent reviews on the methods for quantifying N2 fixation have been published and they will be referred to in the subchapters dealing with relevant techniques. Unkovich et al. (2008), which is freely available in electronic form, is a good general starting point but the authors largely deal with crop legumes. Boddey et al. (2000) and Domenach (1995) provide reviews targeting N2-fixing trees. Before dealing with methods suitable for AFS, it is necessary to note limitations of some old methods. Acetylene reduction assay (ARA) is based on the multiaffinity of nitrogenase to break triple bounds in gas molecules (see Giller 2001 or Hunt and Layzell 1993 for a review). Nodules are incubated in acetylene (C2H2) enriched atmosphere and the production of ethylene (C2H4) is measured. Theoretically, C2H4 production rate is 1/4 of N2 fixation rate (Giller 2001). However, the ratio seems to be highly variable in natural systems and ARA requires calibration with other methods in order to be quantitative (Hunt and Layzell 1993). Further, the incubation conditions may cause serious disturbance to nodule functions leading to underestimation of the actual N2 fixation rate (Giller 2001; Minchin et al. 1983; 1986). Thus, earlier

123

Author's personal copy 130

reports on low N2 fixation rate by individual agroforestry trees (Lindblad and Russo 1986; Roskoski and Van Kessel 1985) and extrapolations to AFS level of ca. 40 kg ha-1 year-1 (Roskoski 1982) based on ARA should be dealt with caution. Unfortunately, they have been quite influential among agroforestry researchers in Latin America (Beer et al. 1998). While the ARA must be dismissed as a quantitative field method, it has undeniable value as a rapid method for checking if nitrogenase is active (Roggy et al. 1999b; Unkovich et al. 2008). Another method unsuitable in agroforestry research is the xylem solute method (Peoples et al. 1996). It is used for ureide-transporting crop legumes such as soybean (Glycine max L.): xylem sap is collected and relative amounts of ureides, amino acids, and nitrate (NO3-) out of total xylem N are calculated. The ureide proportion is used as an indicator of N2 fixation, which may be calibrated against another quantitative method (Unkovich et al. 2008). However, it has little value in agroforestry research because very few trees transport ureides; only two tree species of the legume tribe Desmodiae were found to be ureide transporters in an extensive review by Giller (2001). Thus, the oftenrepeated statement that temperate legumes transport fixed N in the form of amino acids and tropical legumes as ureides (Vessey et al. 2004) may be a false perception based on a few crop legumes.

Controlled and semi-controlled conditions We call laboratory and greenhouse experiments where the growth conditions are mostly under the researchers’ control as controlled condition experiments. Semi-controlled conditions refer to field experiments where some environmental factors are restricted, typically root growing space by physical barriers (Kurppa et al. 2010; Leblanc et al. 2007; Sta˚hl et al. 2005). Under these conditions, using the stable heavy 15 N isotope as a tracer may be the most adequate method. The method is based on the stable proportion of 15N out of atmospheric N, with mean 0.3663 % and standard deviation 0.0002 % (Mariotti 1983). If 15N enriched fertilizer is applied to the substrate, a N2-fixing plant becomes less enriched with 15N than a non-N2-fixing plant because the N taken up from the substrate is diluted with atmospheric N2 in the former. The symbiotic dependence of the N2-fixing plant

123

Nutr Cycl Agroecosyst (2012) 94:123–160

(Chalk and Ladha 1999) is estimated as the percentage of N derived from atmosphere out of total plant N (%Ndfa): %Ndfa ¼ 1  %15 Nex;fix =%15 Nex;ref



ð2Þ

where %15Nex,fix and %15Nex,ref are 15N atom excess percentage with respect to atmosphere in the N2-fixing and non-N2-fixing reference plant, respectively. The %15Nex is calculated simply by subtracting the atmospheric 15N percentage from the measured 15N percentage in the sample, determined by an isotope ratio mass spectrometer. The pros and cons of the 15N dilution method have been extensively revised by Unkovich et al. (2008). They also present different variants of the basic method for particular experimental conditions. The main advantage of the method is that it provides both yield-independent and time-integrated estimates of %Ndfa. Main difficulties are related to temporal and spatial non-uniformity of the distribution of the 15 N label in the substrate. When small plants in small pots are studied, the 15N label may be efficiently mixed in the substrate (Unkovich et al. 2008). Studies on N2 fixation by seedlings or saplings are of little use for agroforestry research because juvenile trees may behave differently from mature trees in many aspects, as we will show in the chapter ‘‘Symbiotic N2 fixation in whole plant physiology’’. The limits of complete mixing of 15N with substrate are soon met: when 200 l barrels buried in soil were used for controlling root growth and 15N distribution in the field, the top 0–15 cm of the barrel soil was significantly more enriched than the deeper layers (Kurppa et al. 2010). Under semi-controlled field conditions, the only possibility to deal with the problems caused by the uneven distribution of the 15N tracer is to use an adequate reference plant. Unkovich et al. (2008) list characteristics, which the reference plant should have in common with the N2-fixing plant: (1) same life form (i.e. only trees should be used as references for trees); (2) adaptation to similar environmental conditions; (3) rooting zone and relative N uptake pattern within the root system; (4) mycorrhizal status; and (5) growth habit and phenology. Further, the N2-fixing and nonN2-fixing reference plant should prefer the same inorganic N form for uptake (Schimann et al. 2008). In natural ecosystems, plants tend to differentiate with

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

respect to these factors in order to share soil resources within the same site, which results in niche differentiation (McKane et al. 2002). Ecologically, the N2fixing and the reference plant should share the same niche except for N2 fixation. Following to that, the crop species associated with a N2-fixing tree in an AFS is seldom a good reference because species selection in agroforestry targets to complementary resource use (Ong et al. 2004) and, thus, niche differentiation. Field studies When N2 fixation is estimated under field conditions, the problems associated with the uniformity of 15N labelling and match between the characteristics of the N2-fixing and reference tree become manifold in comparison to controlled-environment studies. It is almost impossible to distribute the 15N tracer uniformly in a soil profile. Injecting the 15N tracer to different soil depths has been used for studying rooting depths and N uptake in AFS (Lehmann et al. 2001; Rowe et al. 1999; 2001). The injection is always punctual and extensive sampling is required for assuring a reliable recovery of the 15N signal in the plants (Rowe and Cadisch 2002). Although the deep injection labelling has some promise for agroforestry research, it has not yet been used for estimations of the contribution of N2 fixation to N economy of the N2fixing trees or the whole AFS. It seems that selection of a ‘‘correct’’ reference tree species that uses the same form of N from same depth and horizontal area as the N2-fixing tree is the only method to overcome the problem of uneven distribution of the 15N tracer in soil (Chalk and Ladha 1999). However, often too little preinformation is available on these tree characteristics. Thus, using several reference species is recommended (Unkovich et al. 2008). Given the problems of the 15N dilution method under field conditions, the 15N natural abundance method (Shearer and Kohl 1986) is often recommended for quantifying N2 fixation in AFS. Handley and Raven (1992), Ho¨gberg (1997), and Martinelli et al. (1999) reviewed the behaviour of 15N in plant metabolism, soil–plant systems, and natural forests, respectively. Boddey et al. (2000), Domenach (1995), and Gehring and Vlek (2004) reviewed the application of the 15N natural abundance method for estimating N2 fixation by trees. Here we revise the basic idea and some agroforestry applications.

131

In the basic application of the 15N natural abundance method, the deviation of the sample 15N content from that of atmosphere (d15N) is calculated (Shearer and Kohl 1986): 15

d15 N ¼

N=14 Nsa  15 N=15 Nat  1; 000 & 15 N=14 N at

ð3Þ

where 15N/14N is the ratio of 15N to 14N and subscripts sa and at refer to the sample and atmosphere, respectively. In the 15N natural abundance method, %Ndfa is estimated by comparing the d15N values in a N2-fixing plant growing in the field (d15Nf), a non-N2fixing reference plant growing in the same soil (d15Nr), and the N2-fixing plant growing in N-free environment (d15N0), i.e. depending only on N2 fixation for N supply (Shearer and Kohl 1986): %Ndfa ¼

d15 Nr  d15 Nf  100% d15 Nr  d15 N0

ð4Þ

While the 15N natural abundance in the atmosphere is very stable and uniform around the world (Mariotti 1983), it varies in the soils and plants because plant metabolism and several microbiological processes in the soil cause isotopic fractionation (Handley and Raven 1992). Most soil microbiological processes discriminate against 15N (Ho¨gberg 1997), thus, enriching soil with this isotope. This causes a slight difference in the d15N of plants depending on soil N and plants partially fixing atmospheric N2. The difference can be detected with modern isotope ratio mass spectrometers but all work phases must be conducted with great care because the differences are only a few %-units. The general trend of the soil becoming naturally 15 N enriched does not hold in all soils. If the non-N2fixing trees have a low soil-derived d15N that cannot be distinguished from the d15N of N2-fixing trees, application of the 15N natural abundance method is impossible (Gehring and Vlek 2004; Roggy et al. 1999a). In some cases, the soil may be so depleted of 15 N and, consequently, the reference trees have so much lower d15N than the N2-fixing trees that the 15N natural abundance method is applicable because active N2-fixers tend to have d15N close to 0, in this case significantly higher than the non-N2-fixing reference (Augusto et al. 2005; Domenach 1995; Domenach et al. 1989). Further, it was shown among non-N2-fixing caesalpinoids that Eperua falcata Aubl., which prefers

123

Author's personal copy 132

NO3- as the inorganic N source, had significantly lower d15N than Dicorynia guianensis Amshoff, which uses ammonium (NH4?) for N supply in the same soil (Schimann et al. 2008). This suggests that the N2-fixing tree and the reference tree should use the same inorganic N form. Most legumes form symbiosis, in addition to rhizobia, with arbuscular mycorrhizaeforming fungi (AMF) and some with ectomycorrhizaeforming fungi (ECM; Sprent and James 2007). According to a global analysis, AMF deplete the host plant with 15N and ECM even more (Craine et al. 2009) yet the difference between AMF and ECM plants decreases when they grow in the same site (Hobbie and Ho¨gberg 2012). It seems that the effect of mycorrhizae on plant d15N depends on an interaction between fungal strain, environment, and host plant (Boddey et al. 2000). The mycorrhizal effects are further complicated by the fungal uptake of organic N (Boddey et al. 2000; Na¨sholm et al. 2009). However, before a better understanding on these effects arises, we recommend using reference trees that have the same mycorrhizal type and preference for soil N as the studied N2-fixing tree. As in the case of 15N dilution method, the reference plant should have similar rooting pattern and phenology as the N2-fixing plant. The question whether to use the associated non-N2fixing plant or an independent reference growing in similar soil but without contact with the N2-fixing plant is more important in the case of 15N natural abundance than the 15N dilution method. In the latter case, the selection is an issue of ecological and physiological similarity, which is seldom met between the trees and crops in AFS. In the field, a N2-fixing tree may slowly decrease the soil d15N by recycling litter and root exudates depleted of 15N. Consequently, the d15N of the associated plants also decreases over time (van Kessel et al. 1994; Sierra and Nygren 2006). Leucaena leucocephala altered so much the soil isotopic signature in six years that by the end of the study period, associated weeds had almost the same isotopic signature as the tree itself (van Kessel et al. 1994). The d15N gradually increased in Dichanthium aristatum (Poir.) C.E. Hubb. grass brought from a field site with Gliricidia sepium trees to a greenhouse, probably because of removal of N transferred from the tree by successive grass cuttings (Sierra and Nygren 2006). Thus, the reference trees should grow without direct contact with the N2-fixing trees but in a similar soil.

123

Nutr Cycl Agroecosyst (2012) 94:123–160

Use of several reference tree species is recommendable also in 15N natural abundance method because often too little is known about the characteristics described above for selecting an ecologically and physiologically similar reference to a N2-fixing tree (Boddey et al. 2000; Unkovich et al. 2008). Several candidate references can also be screened in a pre-trial (Nygren and Leblanc 2009). This will allow the removal of apparent outliers; e.g. in a comparison of five non-N2-fixing tree species, Nygren and Leblanc (2009) found that one species had much lower d15N than the four other species, which had little variation among themselves. Non-N2-fixing legume trees may be the best references for legume trees if available (Roggy et al. 1999b). Further, if the research involves comparisons over different sites, some of the reference species should be used in all sites if possible (Roggy et al. 1999b). In some cases, it may also be possible to use a nonnodulating phenophase of the studied species as the reference (Salas et al. 2001). This method requires that the isotopic signature of the soil N is stable over time, sufficient variation in nodulation exists between the phenophases, and the temporal variation in N2 fixation of the nodules is small. The approach was successfully applied for estimating rainy season N2 fixation in Erythrina lanceolata Standl. using the dry season nonnodulating phenophase as the reference (Salas et al. 2001). Sampling procedures Whole tree harvesting is recommended for seedlings and saplings because the isotopic signal may vary in different parts of a tree (Kurppa et al. 2010; Leblanc et al. 2007). In the case of big trees, this is often not possible. Typically, both N2-fixing and non-N2-fixing trees tend to have the highest d15N in leaves and the lowest in stem and coarse roots (Boddey et al. 2000). Leaves are often the preferred sink for recently fixed N (Domenach 1995) and they are, thus, the best choice for sampling if the objective is to follow seasonal or other variation in N2 fixation in time scale of weeks or few months. Large structural parts of a tree, stem and coarse roots, may be envisioned to integrate the isotopic signature of N sources over time. They may be the choice if the objective is to understand the longterm role of N2 fixation in trees’ N economy (Nygren and Leblanc 2009). Fine roots have seldom been

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

sampled. In a study using 15N enrichment, higher proportion of 15N than total N was retained in the fine roots of Gliricidia sepium suggesting that soil N may be preferentially used for fine roots and fixed N is transported to the shoot. However, the root:shoot sharing of 15N and total N was the same in the legume tree Inga edulis and non-N2-fixing crop Theobroma cacao L. (Kurppa et al. 2010). Thus, dynamics of fine roots require further study for understanding the whole plant isotopic relationships. Sanginga et al. (1995) claimed that the 15N natural abundances in different organs tend to be proportionally equal in N2-fixing and non-N2-fixing trees and, consequently, estimates of the %Ndfa are about the same if the same organ of both trees is used. Thus, they proposed that leaf sampling would result in reliable %Ndfa estimates. This hypothesis held between G. sepium and five different reference tree species in a cacao plantation but not with I. edulis (Nygren and Leblanc 2009). The estimates of %Ndfa based on 15N enrichment (Eq. 2) were about the same using leaves only and whole sapling harvesting for three agroforestry tree species, including I. edulis (Leblanc et al. 2007). Leaves were recommended for 15N sampling in trees but not in herbaceous legumes by Unkovich et al. (2008). Isotopic fractionation within a tree resulting in differences in N isotopic relationships between tree organs may affect the usability of the 15N natural abundance method in some cases. Some nodulated Inga spp., in which ARA indicated nitrogenase activity, had high leaf d15N values typical for nonN2-fixing trees in a rain forest (Roggy et al. 1999b) and a tropical fresh water swamp forest (Koponen et al. 2003) in French Guiana. In both cases, most Inga spp. had leaf d15N values typical for N2-fixing trees. In a Costa Rican cacao plantation, I. edulis shade trees had a high d15N in leaves but low stem and coarse root d15N values were typical for N2-fixers (Nygren and Leblanc 2009). These data may indicate strong isotopic fractionation in the N transport and metabolism of some Inga spp. Similar observations have been made in Acacia mangium (Bouillet et al. 2008) and A. senegal (Isaac et al. 2011a). Although the strong within-tree isotopic fractionation may be a rare phenomenon, its possible occurrence should be taken into account when analysing 15N natural abundance data and nodulation must be revised in all trees sampled. The 15N dilution method does not seem to be

133

affected by isotopic fractionation, indicating active N2 fixation in I. edulis (Kurppa et al. 2010; Leblanc et al. 2007).

Symbiotic N2 fixation in whole tree physiology Conceptual model on nitrogen acquisition in the tripartite plant-bacteria-mycorrhizae symbiotic system It may be assumed that legumes (Bethlenfalvay 1992; Bethlenfalvay and Newton 1991) and actinorrhizal plants (Gardner and Barrueco 1999) are mycorrhizal, except Lupinus spp. (Sprent and James 2007). Thus, ecophysiology of N2 fixation must be evaluated within the context of the tripartite symbiosis formed by plants, N2-fixing bacteria, and mycorrhizae in all natural and man-made ecosystems (Barea et al. 1992; Bethlenfalvay and Newton 1991; Cardoso and Kuyper 2006; Kuyper et al. 2004). Most legumes seem to form symbiosis with AMF (Bethlenfalvay 1992; Cardoso and Kuyper 2006; Kuyper et al. 2004; Sprent and James 2007). A minority of legume species, notably Acacia spp. and caesalpinoids, may form symbiosis with ECM or both AMF and ECM (Haselwandter and Bowen 1996; Sprent and James 2007). More than 40 actinorrhizal species have been recorded to support ECM, AMF, or both symbioses (Gardner and Barrueco 1999). A simplified scheme of potential N acquisition strategies by the tripartite symbiosis is presented in Fig. 1. The soil N forms available for the mycorrhizal fungi vary: AMF may take up soil NH4?, NO3-, and amino acids, while some ECM cannot use NO3- but they are able to use, in addition to amino acids, some other organic N forms in the soil such as simple proteins and oligopeptides (Lambers et al. 2008; Smith and Read 2008). Mycorrhizae also enhance plant N nutrition by largely increasing the soil volume available to uptake of N and other nutrients because of the extensive extraradical mycelium. Fungal hyphae also access smaller soil pores than plant roots (Smith and Read 2008). The N2 fixation process by symbiotic bacteria within the root nodules has been described above. Soil inorganic N taken up by the mycorrhizal fungus is processed in the fungal hyphae (Smith and Read 2008). Nitrate is reduced to NH4?, which is further assimilated to amino acids together with NH4?

123

Author's personal copy 134

Nutr Cycl Agroecosyst (2012) 94:123–160

Fig. 1 A simplified scheme of nitrogen acquisition strategies of a plant, which forms symbiosis with N2-fixing bacteria and mycorrhiza-forming fungi. The potential N sources (clouds) are soil ammonium and nitrate, which the plant may take up directly via its own fine roots or via mycorrhizal symbiosis, soil amino acids (AA), which may be taken up by both arbuscularmycorrhizal and ectomycorrhizal fungi, other soil organic N (No), which may be taken up by the ectomycorrhizal fungi, and atmospheric N2, which is fixed by the symbiotic bacteria in the root nodules. Flows of inorganic N are indicated by thin arrows,

flows of organic N by thick arrows, flows of organic carbon (Co) by double arrows, feedback loops by dashed arrows, metabolic processes by valves, and NH4? and NO3- transporters with ellipsoids. The metabolic processes are breakdown of AA to NH4?, NO3- reduction to NH4?, assimilation of NH4? to AA, and biological N2 fixation. There is evidence on N transfer from the plant to mycorrhizal fungus as either NH4? or AA, thus double-ended arrows. Details of leaf N metabolism and photochemistry that also affect plant N status are beyond the scope of this review

taken up from the soil. The plant may reduce NO3- in the roots or transport it to the leaves where NO3- is reduced using NADH and NADPH produced in photosynthesis (Oaks 1992; Zerihun et al. 1998). Free NH4? is toxic to plants and must be assimilated in the site of reduction. Carbon transported in phloem from the leaves is needed for NH4? assimilation in the roots or mycorrhizal mycelium. There is also evidence on N transfer from the plant to mycorrhizal fungus, especially from N-rich roots of N2-fixing plants (Arnebrant et al. 1993; He et al. 2003; Simard et al. 2002). Parsons et al. (1993) suggested that nodulation and N2 fixation may be regulated by the concentration of

reduced N in the phloem flow. Nitrogen is transported to leaves as NO3- or amino acids in the xylem, and amino acids may be transported from leaves back to roots in phloem (Fig. 1). Comparison of nodulation and canopy N flows in the agroforestry tree Erythrina poeppigiana (Walp.) O.F. Cook suggested that nodulation was regulated by N needs of the canopy with reduced nodulation when net N flow turned downward from canopy to roots (Nygren 1995). Accumulation of amino acids in the roots controls the NH4? and NO3transporters in all plants (Amtmann and Blatt 2009) and the blockage of the transporters by high root amino acid concentration may result in N exudation

123

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

from roots to soil observed in many plants (Fustec et al. 2010; Wichern et al. 2008). Thus, root amino acid pool may be envisioned to play a key role in regulating the N acquisition of the whole tripartite symbiotic system (Fig. 1). It may be hypothesised that the root pool size is maintained by the balance between the influxes from the tripartite symbiosis and efflux to foliage. Thus, N acquisition may be ultimately regulated by N needs in the plant canopy. Not all evidence to support this hypothesis is available but we feel that it is a fruitful starting point for further research. Energetics of N2 fixation Symbiotic N2 fixation is quite an energy intensive strategy for acquiring N (Vance and Heichel 1991; Vitousek et al. 2002). Following the biochemicallybased estimation method of Thornley and Johnson (1990), 4.29 g of glucose is needed to reduce 1 g of N to NH4? from NO3- in the roots and 5.00 g of glucose for fixing 1 g of N from the atmosphere. Thus, the C cost of N2 fixation is only about 17 % higher than that of NO3- reduction in roots. However, glucose consumption per gram of protein produced is considerably lower when amino acids (on average 0.89 g glucose) or NH4? (1.77 g) are available as the N source (Zerihun et al. 1998). At whole plant level, acquisition of soil N also requires C for growth and maintenance of the root system. Legumes probably evolved in the humid tropical forests (McKey 1994) where NH4? is often a much more abundant inorganic N form than NO3(Boddey et al. 2000). Under these conditions, N2-fixing symbiosis may have been energetically more appropriate N acquisition strategy than growing roots for competing for the scarce soil NO3- resources. Carbon consumption to N2 fixation may be compensated by increased photosynthetic rate; photosynthetic rate of several nodulated herbaceous crop legumes was increased on average by 28 % in comparison to non-N2fixing plants (Kaschuk et al. 2009). In most cases, increase in photosynthetic production was higher than the C allocation to the rhizobial symbiont (4–16 % of C fixed in photosynthesis). We are not aware of any respective studies on N2-fixing trees. Phosphorus in the N2-fixing trees Phosphorus has been considered a limiting nutrient for symbiotic N2 fixation process because of the high ATP

135

requirements (Eq. 1). For evaluating the importance of the P on N2 fixation, we searched the CAB Abstracts database for articles published 2000–2009 on the effects of P fertilisation on nodulation and N2 fixation in trees. All studies dealing with tropical tree seedlings under greenhouse or nursery conditions (Binkley et al. 2003; Diouf et al. 2008; Pons et al. 2007; Uddin et al. 2008) indicated positive effects of P fertilisation on N2 fixation. All field studies were conducted on actinorrhizal trees in temperate forests. These studies indicated no direct P effect on nitrogenase activity (Uliassi and Ruess 2002) or N2 fixation rate (Augusto et al. 2005; Cavard et al. 2007; Gokkaya et al. 2006) but enhanced nodulation was observed in one case (Uliassi and Ruess 2002). In a recent study, however, Isaac et al. (2011a) found that P supply rates did not markedly affect N2 fixation rates of Acacia senegal seedlings under nonlimiting N supply, but higher P supply stimulated growth, which resulted in greater mineral N uptake from soil solution. On the other hand, the N2 fixation rate of A. senegal increased with increasing soil P availability in natural stands of the Rift Valley in Kenya (Isaac et al. 2011b). Application of rock phosphate did not significantly affect N2 fixation by Inga edulis in a P-poor soil in the humid Atlantic lowlands of Costa Rica (Leblanc 2004). Severe P deficiency markedly impaired symbiotic N2 fixation in the tropical Casuarinaceae species (Sanginga et al. 1989) and other actinorrhizal plants (Gardner and Barrueco 1999). Several actinorrhizal genera such as Comptonia and Myrica (Myricaceae) develop cluster (proteoid-like) roots, which could enhance P uptake in P poor soils (Berliner and Torrey 1989). Both plant roots and hyphae of mycorrhizal fungi secrete extracellular phosphatases to the rhizosphere (Grierson et al. 2004; Louche et al. 2010; Treseder and Vitousek 2001). Phosphatases hydrolyse the esterphosphate bonds in soil organic P, thus releasing phosphate to soil solution where it may be taken by plant roots or hyphae of mycorrhizal fungi. Extracellular phosphatases in the rhizosphere assure the breakdown of organic P compounds in the proximity of roots. Phosphatases contain 8–32 % of N (Treseder and Vitousek 2001), and it has been proposed that N2fixing plants have an advantage in producing these N-rich compounds (Wang et al. 2001). Houlton et al. (2008) presented a hypothesis that improved P

123

Author's personal copy 136

nutrition because of a higher production rate of extracellular phosphatases may explain the abundance of legume trees in P-poor humid tropical forests. Higher phosphatase activity has been observed in the soil below both actinorrhizal (Giardina et al. 1995; Zou et al. 1995) and legume trees (Allison et al. 2006; Zou et al. 1995) than below non-N2-fixing trees. As far as we know, no studies on the phosphatase production in AFS exist. However, it may be an important yet at the moment unstudied positive interaction between N2-fixing trees and crops in AFS. The fact that P fertilisation enhanced growth of tree seedlings in pot experiments (Diouf et al. 2008; Pons et al. 2007; Uddin et al. 2008) but results on mature trees in the field were more variable (Gokkaya et al. 2006; Uliassi and Ruess 2002) suggest that caution should be applied for extrapolating the pot results to the field. In one case, the same tree species, Falcataria moluccana (Miq.) Barnaby & J.W. Grimes, was studied both in a pot culture and in a forest. Phosphorus fertilisation enhanced N2 fixation by seedlings in the pot study (Binkley et al. 2003), while higher phosphatase activity was observed in soils under F. moluccana than non-N2-fixing trees in Hawaii (Allison et al. 2006). It is possible that mature trees and associated mycorrhizae produce more phosphatases than seedlings and, thus, better cope with soil P deficit. Interaction between N2-fixing bacteria and mycorrhizae Mycorrhizae may also alleviate the effects of soil P deficit on N2 fixation through enhanced P supply to the tripartite symbiotic system. Best growth and highest shoot nutrient concentrations were observed in P fertilised Acacia senegal seedlings inoculated with rhizobia and dual inoculation with rhizobia and AMF was the second best treatment (Colonna et al. 1991). In a sterile, low P substrate, AMF colonisation in roots of Calliandra calothyrsus seedlings decreased with increasing P addition (Ingleby et al. 2001). Biomass production was not affected by AMF colonisation or P level but higher nodulation was observed in the AMFcolonised plants. Rhizobia-ECM dual inoculated Acacia holosericea A. Cunn. ex. G. Don seedlings had the highest biomass production (Andre´ et al. 2005). Both AMF and ECM colonisation enhanced nodulation in the same tree species (Duponnois and Plenchette 2003). Inoculation with AMF enhanced

123

Nutr Cycl Agroecosyst (2012) 94:123–160

nodulation of Leucaena leucocephala seedlings and dual inoculation was recommended (Arau´jo et al. 2001). Dual rhizobia-AMF inoculation also enhanced initial growth of C. calothyrsus under field conditions but statistically significant differences between various inoculation treatments disappeared after 24 months (Lesueur and Sarr 2008). It has also been shown that the percentage of roots infected by mycorrhizae is markedly higher in nodulated than non-nodulated actinorrhizal plants (Chatarpaul et al. 1989). Based on their own observations and earlier work (Habte 1995; Habte and Turk 1991; Manjunath and Habte 1992), Ingleby et al. (2001) concluded that among the common agroforestry trees, N2-fixing C. calothyrsus, Gliricidia sepium, and L. leucocephala, and the non-N2-fixing Senna siamea (Lam.) H.S. Irwin & Barneby are highly responsive to mycorrhizae while the non-N2-fixing Senna reticulata (Willd.) H.S. Irwin & Barneby and N2-fixing Sesbania pachycarpa DC. are less responsive. Most studies revised indicate beneficial effects of the tripartite symbiosis on plant development and N2 fixation. The beneficial effect of mycorrhizae seems to be most important under P limitation supporting Bethlenfalvay’s (1992) suggestion that the main function of the mycorrhizal symbiont is P supply to the tripartite system. The observations that AMF colonisation was more beneficial than P fertilisation (Arau´jo et al. 2001) and that dual inoculation enhanced biomass production independently of P level (Weber et al. 2005) suggest, however, that mycorrhizal colonisation may enhance the functioning of the tripartite symbiotic system also in ways other than just increased P supply. It should also be noted that mature trees might function differently as host than seedlings or saplings.

Estimates of N2 fixation in agroforestry systems Field estimates on the dependence of agroforestry trees on symbiotic N2 fixation When compiling quantitative data on N2 fixation in AFS, we excluded seedling studies, because we have shown in the previous chapter that seedlings may behave quite differently from mature trees. On the other hand, many legume trees used in AFS are fastgrowing and some management practices, notably green pruning, maintain them in a physiological state

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

resembling juvenile trees, e.g. by impeding flowering (Nygren et al. 2000). Thus, we accepted data from experiments with at least 1-year-old saplings. We also included data on tree plantations if the tree species is known to be used in agroforestry. Data were most abundant on the %Ndfa including 38 data points (Table 2). We classified the data according to management practices and climate. Climate was classified into three broad categories based on our interpretation of the data given in original articles: dry (water deficit prevails most of the year), subhumid (seasonal climate with water deficit for six months or less), and humid (short water deficit periods or none at all). Two irrigated experiments (Gauthier et al. 1985; Kadiata et al. 1997) were pooled with humid climate in the following analyses. We pooled data on 15N dilution (Eq. 2) and natural abundance (Eq. 4) methods. The general mean of the %Ndfa of the compiled data (Table 2) was 59 % (Standard deviation, SD 16.6 %) and the range was from 10 to 85 %. We used the mean of minimum and maximum in the calculation of the general mean when a range was given in Table 2. Five studies were conducted in dry environments (Aronson et al. 2002; Gueye et al. 1997; Isaac et al. 2011b; Raddad et al. 2005; Unkovich et al. 2000) with the mean %Ndfa of 49 % (SD 25.3 %, N = 8). The data compiled also included ten cases of young (1–2 years) trees under humid or sub-humid conditions with mean %Ndfa of 69 % (SD 10.7 %). The remaining 20 studies were conducted with at least 2-year-old trees under humid or subhumid conditions with mean %Ndfa of 57 % (SD 12.1 %). However, only three of these studies were conducted in systems that had been subjected to the same management for several years, i.e. a 15-yearold shaded cacao plantation (Nygren and Leblanc 2009), a 12-year-old improved fallow of Acacia mangium (Mercado et al. 2011), and an 8-year-old cut-and-carry forage production system under partial pruning regime (Nygren et al. 2000). The studies on at least 2-year-old systems under humid or subhumid conditions were further divided into pruned and unpruned systems, with the mean %Ndfa of 63 % (SD 12.0 %, N = 7) and 54 % (SD 11.7 %, N = 13), respectively. The popular agroforestry tree Gliricidia sepium appeared eight times in Table 2 being the only species, for which we calculated the species mean for %Ndfa, 67 % (SD 13.0 %). All these means are higher than the general mean for legume trees in natural ecosystems according to the

137

review by Andrews et al. (2011), 42 % (SD 5.4 %). Because competition with non-N2-fixing plants seems to increase the legume dependence on N2 fixation (Andrews et al. 2011) competition with crops may enhance N2 fixation by trees in AFS. Soil N depletion due to N export in crop harvest may also partially explain the higher legume dependence on N2 fixation in AFS than natural ecosystems. Because of the high within-group variation and low number of studies under dry conditions, we compared statistically only the three distinct cases under humid or subhumid conditions: young trees (1–2 years) and older systems ([2 years) further divided into pruned and unpruned practices. Juvenile trees had a significantly higher %Ndfa than mature trees in unpruned systems while mature trees in pruned systems did not differ significantly from either juvenile or unpruned mature trees (analysis of variance followed by Duncan’s Multiple Range Test at 5 %). The ANOVA had a low r2 value, 0.258 with significant P value (0.0176). Thus, no strong conclusions may be drawn but the differences reveal some possible trends that may be fruitful for further research together with causes for the ranges observed in several cases included in Table 2. Especially a meta-analysis on these effects will be interesting when more data become available. The ranges reported in Table 2 were caused by five factors. Differences because of reference species (Sta˚hl et al. 2005) or tree organ (Chikowo et al. 2004) are methodological issues. A range of N2 fixation estimates based on various reference species would be more reliable than single value because the range also gives an estimate of the reliability of the estimates. However, the range was reported only by Sta˚hl et al. (2005). Interestingly, the same reference species gave very similar N2 fixation estimates in another study and the seasonality had a much stronger effect on the N2 fixation estimates (Sta˚hl et al. 2002). An estimate based on the average d15N of several nonN2-fixing references was reported in three other articles (Aronson et al. 2002; Lesueur and Sarr 2008; Nygren and Leblanc 2009). The estimates based on one or several herbaceous references (Chikowo et al. 2004; Mercado et al. 2011; Unkovich et al. 2000) should be dealt with caution; we agree with Unkovich et al. (2008) that the reference should be of the same life form as the N2-fixer. The wide range of estimates for different tree organs (Chikowo et al. 2004) may be caused by the use of the Hyparrhenia rufa (Nees)

123

Associated species

Zea mays

None

Eucalyptus grandis

None

None

None

Grass

None

None

None

Khaya senegalensis, Ziziphus mauritaniana and Anacardium occidentale

Eucalyptus x robusta

None

None

Sequential Lupinus angustifolius and Avena sativa

None

None

N2-fixing species

Acacia angustisima

123

Acacia caven

Acacia mangium

Acacia mangium

Acacia mangium

Acacia senegal

Acacia senegal

Albizia lebbeck

Calliandra calothyrsus

Calliandra calothyrsus

Calliandra calothyrsus

Casuarina equisetifolia

Casuarina equisetifolia

Chamaecytisus proliferus

Chamaecytisus proliferus

Codariocalyx gyroides

Erythrina fusca

Vochysia guatemalensis

Senna spectabilis

Mean of Ptilotus polystachus and annual weeds

Schinus polyganus and Fraxinus excelsior

None

Humid

Subhumid

Dry

Dry

Irrigated

Sudhumid-humid

Not reported (subhumid assumed)

Subhumid

Subhumid

Irrigated

Dry

Dry

Humid

Humid

Subhumid

Dry

Subhumid

Climate

14-month-old saplings

Protein bank

Experimental alley cropping, 4 years

Experimental plantation; 6 years

1-year experiment

Mixed tree plantation

Mixed experimental plantation; 2 years

Protein bank

Improved fallow

16-month greenhouse experiment

Tree-grass savanna

Experimental arabic gum production

Improved fallow, 12 years

Experimental plantation

Mixed plantation

Experimental plantation

Improved fallow

System

Free growth

7 prunings in 2 years

Pruned

Free growth

Free growth

Free growth

Free growth

7 prunings in 2 years

Free growth

Variable pruning frequency

Free growth

Free growth

Free growth

Free growth

Free growth

Free growth

Unpruned

Management information

N enrichment; whole tree

15

N natural abundance; whole tree

15

N natural abundance; coppice biomass

15

N natural abundance; leaf biomass

15

N dilution; shoot biomass

15

N enrichment; whole tree biomass

15

N natural abundance; leaf biomass

15

N natural abundance; whole tree

15

N enrichment; whole tree biomass

15

N enrichment; shoot biomass

15

N natural abundance; leaf biomass

15

N natural abundance; leaf biomass

15

N natural abundance; whole tree

15

N natural abundance; sampling not reported

15

N enrichment; whole tree biomass

15

N natural abundance; leaf biomass

15

64

48–86f

83

84

55

59

30–60b

24–84f

5–54f

74–83e

33–39d

24–61c

57

42–62b

59

50

48–79a

15

N natural abundance; whole tree

%Ndfa

Estimation method

Leblanc et al. (2007)

Peoples et al. (1996)

Unkovich et al. (2000)

Aronson et al. (2002)

Gauthier et al. (1985)

Parrotta et al. (1996)

Lesueur and Sarr (2008)

Peoples et al. (1996)

Sta˚hl et al. (2002)

Kadiata et al. (1997)

Isaac et al. (2011b)

Raddad et al. (2005)

Mercado et al. (2011)

Galiana et al. (1998)

Bouillet et al. (2008)

Aronson et al. (2002)

Chikowo et al. (2004)

References

138

Eucalyptus x robusta

Mean of associated tree species

Senna spectabilis

Eucalyptus deglupta and Grevillea robusta

Senna siamea

Balanites aegyptiaca

Balanites aegyptiaca

Understorey weeds

Eucalyptus urophylla

Eucalyptus grandis

Schinus polyganus and Fraxinus excelsior

Hyparrhenia rufa

Reference species

Table 2 Percentage of nitrogen derived from atmosphere out of total nitrogen (%Ndfa) in several tree species used in agroforestry

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

None

None

Dichanthium aristatum

Dichanthium aristatum

None

Sequential Zea mays and Oryza sativa

Arachis pintoi (?)

Theobroma cacao

Theobroma cacao, Bactris gasipaes, non-leg. trees

None

None

Theobroma cacao

Theobroma cacao, Cordia alliodora

Eucalyptus x robusta

Erythrina poeppigiana

Faidherbia albida

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Gliricidia sepium

Inga edulis

Inga edulis

Inga edulis

Leucaena leucocephala

None

Vanilla planifolia

Erythrina lanceolata

Leucaena leucocephala

Associated species

N2-fixing species

Table 2 continued

Senna siamea

Eucalyptus x robusta

Cordia alliodora

Theobroma cacao

Vochysia guatemalensis

Senna siamea

Mean of 4 non-legume tree spp.

Theobroma cacao

Peltophorum dasyrrachis

Senna spectabilis

Senna spectabilis

Gmelina arborea

Gmelina arborea

Parkia biglobosa

Vochysia guatemalensis

Non-nodulated phenophase

Reference species

Irrigated

Subhumid

Humid

Humid

Humid

Irrigated

Humid

Humid

Subhumid

Humid

Subhumid

Subhumid

Subhumid

Dry

Humid

Humid

Climate

16-month greenhouse experiment

Mixed tree plantation

Shaded cacao plantation; 15 years

14-month-old saplings

14-month-old saplings

16-month greenhouse experiment

Shaded cacao plantation; 12 years

14-month-old saplings

Alley cropping

Alley cropping

Protein bank

Cut-and-carry fodder production; 8 years

Cut-and-carry fodder production

15-month-old saplings

14-month-old saplings

Living support

System

Variable pruning frequency

Free growth

Free growth

Free growth

Free growth

Variable pruning frequency

Free growth

Free growth

Complete pruning, frequency not reported

4 prunings per year at 50 cm

7 prunings in 2 years

50 % pruning every 2 months

100 % pruning every 6 months

Free growth

Free growth

Free growth

Management information

N enrichment; shoot biomass

15

N enrichment; whole tree biomass

15

N natural abundance; stem biomass

15

N enrichment; whole tree

15

N enrichment; whole tree

15

N enrichment; shoot biomass

15

N natural abundance; leaf biomass

15

N enrichment; whole tree

15

N enrichment; prunings

15

N natural abundance; shoot biomass

15

N natural abundance; whole tree

15

N natural abundance; shoot biomass

15

N natural abundance; shoot biomass

15

N enrichment; whole tree

15

N enrichment; whole tree

15

N natural abundance; leaf biomass

15

Estimation method

g

80–84e

38

50–63f

74

57

69–75d

57–67f

85

50

35–59f

58–89f

60–87f

54–92f

54

59

53

%Ndfa

Kadiata et al. (1997)

Parrotta et al. (1996)

Nygren and Leblanc (2009)

Kurppa et al. (2010)

Leblanc et al. (2007)

Kadiata et al. (1997)

Nygren and Leblanc (2009)

Kurppa et al. (2010)

Rowe et al. (1999)

Ladha et al. (1993)

Peoples et al. (1996)

Nygren et al. (2000)

Nygren et al. (2000)

Gueye et al. (1997)

Leblanc et al. (2007)

Salas et al. (2001)

References

Author's personal copy

Nutr Cycl Agroecosyst (2012) 94:123–160 139

123

123

None

Zea mays (sequential)

None

None

Prosopis chilensis

Sesbania sesban

Sesbania sesban

Sesbania sesban

h

g

f

e

d

c

b

Depending on reference species

Rainy season (non-nodulated in dry season)

Depending on sampling time

Depending on pruning frequency

Depending on tree age

Depending on tree provenance

Depending on rhizobial strain

Depending on organ

None

Prosopis alba

a

Associated species

N2-fixing species

Table 2 continued

Eucalyptus deglupta and Grevillea robusta

Eucalyptus deglupta and Grevillea robusta

Hyparrhenia rufa

Schinus polyganus and Fraxinus excelsior

Schinus polyganus and Fraxinus excelsior

Reference species

Subhumid

Subhumid

Subhumid

Dry

Dry

Climate

Improved fallow

Improved fallow

Improved fallow

Experimental plantation; 6 years

Experimental plantation; 6 years

System

Free growth

Free growth

Free growth

Free growth

Free growth

Management information

N enrichment; whole tree biomass

15

N enrichment; whole tree biomass

15

N natural abundance; whole tree

15

N natural abundance; leaf biomass

15

N natural abundance; leaf biomass

15

Estimation method

61–71

70–81f h

42–73a

30

10

%Ndfa

Sta˚hl et al. (2005)

Sta˚hl et al. (2002)

Chikowo et al. (2004)

Aronson et al. (2002)

Aronson et al. (2002)

References

Author's personal copy

140 Nutr Cycl Agroecosyst (2012) 94:123–160

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

Stapf grass as the reference; thus, organs of the reference are not comparable with tree organs. Sampling time (Ladha et al. 1993; Nygren and Leblanc 2009; Nygren et al. 2000; Peoples et al. 1996; Sta˚hl et al. 2002) refers to the seasonality of N2 fixation, which is mostly out of the control of agronomic management. The influence of tree provenance (Raddad et al. 2005) and rhizobial strain (Galiana et al. 1998; Lesueur and Sarr 2008) imply intrinsic factors that regulate N2 fixation. These influences may be managed by proper selection of tree sources, strains of N2-fixing bacteria, and their combinations. Finally, pruning frequency (Kadiata et al. 1997; Nygren et al. 2000) is essentially a management issue under the control of the farmers. Dryland agroforestry It is obvious from the data collected in Table 2 that dry environment poses challenges for the use of N2-fixing trees in AFS. Although the dry environment seems to reduce the average dependence on N2 fixation, the high within-group variation in the data indicates that some woody legumes may form N2-fixing symbiosis well adapted to dry habitats. In fact, the mean for dry environments was biased upwards because of high %Ndfa observed in Chamaecytisus proliferus (L. f.) Link in dry Central Chile (84 %; Aronson et al. 2002) and Australia (83 %; Unkovich et al. 2000). These numbers were comparable only with juvenile G. sepium under constantly humid conditions in Costa Rica (85 %; Kurppa et al. 2010). The adverse effect of dry environment on N2 fixation was also observed in a study of 11 Mimosoideae species in natural ecosystems of Namibia with %Ndfa typically varying 10–30 % (Schulze et al. 1991). They also observed considerable variation between species with %Ndfa of 49 % in two species, the maximum of 71 %, and the lowest indicating no N2 fixation. High variation (%Ndfa 24–61 %) was also observed between Acacia senegal provenances in the Sudan (Raddad et al. 2005). Faidherbia albida (Delile) A. Chev., a tree species now strongly recommended for AFS in semi-arid Africa (Garrity et al. 2010), was cited as a N2-fixer in 39 references found in CAB Abstracts data base but only twice N2 fixation was estimated in the field. Active N2 fixation with %Ndfa 54 % at 15-month-age was observed in Senegal (Gueye et al. 1997), while no

141

N2 fixation was observed in mature trees over a rainfall gradient 30–400 mm year-1 in Namibia (Schulze et al. 1991). The former result is based on a provenance that was a superior N2-fixer in a pot study and the trees were apparently very small, with total N accumulation of only 1.41 g [N] per seedling (Gueye et al. 1997). Functional nodules seem to be present in F. albida only during 2-3 months each year when the soil is superficially humid at the end of the rainy season and beginning of the dry season when tree foliage grows (Roupsard 1997). High variation has been observed in N2 fixation of F. albida seedlings in controlled environments (Gueye et al. 1997; Ndoye et al. 1995; Sellstedt et al. 1993). Thus, field estimates on N2 fixation in different environments and AFS are needed for this important tree species. Under field conditions, vigorous provenances from Burundi fixed more N2 than provenances from Niger and Burkina Faso (Roupsard 1997). Among actinorrhizal plants, Casuarinaceae species, mostly used in AFS in arid and semi-arid regions, are known for their drought and temperature resistance and Frankia spp. isolates obtained from these species have high optimal growth temperatures (Dawson 2008). These results indicate that it is possible to find tree species and provenances that are active N2-fixers in dry environments. Thus, we recommend screening trials of N2-fixing tree species and provenances for selecting suitable germplasm for dryland agroforestry. No potentially ‘‘N2-fixing tree’’ should be recommended for dryland agroforestry before its N2 fixation capacity is verified. Dinitrogen-fixing trees in acidic soils Acidic soils of the humid tropics are often cited as a major constraint to N2 fixation (Kass 1995) because most rhizobia and many Frankia strains do not tolerate high acidity and consequent solubilisation of aluminium. However, stains tolerant to acidity exist, especially among Bradyrhizobium spp. (Graham 1992) and N2-fixing trees are abundant in the humid tropical forests with acidic soils (McKey 1994; Roggy et al. 1999a, b). Although some Frankia strains survive in various acidic soils, a negative correlation was found between soil acidity and nodulation in Alnus spp. (Smolander and Sundman 1987) and Elaeagnus angustifolia L. (Jamann et al. 1992; Zitzer

123

Author's personal copy 142

Nutr Cycl Agroecosyst (2012) 94:123–160

Table 3 Nitrogen fixation per hectare and year in some agroforestry systems with legume or actinorrhizal trees based on measurement of whole plant N isotopic relations or whole system N balance N2-fixing species

Associated species

Reference species

Climate

System

Management information

Estimation method

N2 fixation kg ha-1 year-1

References

Acacia angustisima

Zea mays

Hyparrhenia rufa

Subhumid

Improved fallow

Free growth

15

N natural abundance; whole tree

91a

Chikowo et al. (2004)

Acacia auriculiformis

None

Eucalyptus sp.

Humid

Improved fallow, 7 years

Free growth

Whole system N balance

140

BernhardReversat (1996)

Acacia mangium

None

Understorey weeds

Humid

Improved fallow, 12 years

Free growth

15

N natural abundance

129b

Mercado et al. (2011)

Acacia mangium

None

Eucalyptus sp.

Humid

Improved fallow, 7 years

Free growth

Whole system N balance

115

BernhardReversat (1996)

Acacia polyacantha

None

Eucalyptus camaldulensis

Subhumid

Improved fallow, 7 years

Free growth

Whole system N balance

90

Harmand (1998)

Calliandra calothyrsus

None

Eucalyptus deglupta and Grevillea robusta

Subhumid

Improved fallow

Free growth

15

122–171c

Sta˚hl et al. (2002)

Calliandra calothyrsus

None

Cassia spectabilis

Subhumid

Protein bank

7 prunings in 2 years

15

670d

Peoples et al. (1996)

Casuarina equisetifolia

Eucalyptus x robusta

Eucalyptus x robusta

Subhumid

Mixed tree plantation

Free growth

15

73

Parrotta et al. (1996)

Chamaecytisus proliferus

Sequential Lupinus angustifolius and Avena sativa

Mean of Ptilotus polystachus and annual weeds

Dry

Experimental alley cropping, 4 years

Pruned

15

83a

Unkovich et al. (2000)

Chamaecytisus proliferus

None

Mean of Ptilotus polystachus and annual weeds

Dry

Experimental plantation, 4 years

Pruned

15

N natural abundance

390a

Unkovich et al. (2000)

Gliricidia sepium

Dichanthium aristatum (grass)

None

Subhumid

Cut-and-carry fodder production

50 % pruning every 2-4 months

Whole system N balance

555e

Dulormne et al. (2003)

Gliricidia sepium

None

Cassia spectabilis

Subhumid

Protein bank

7 prunings in 2 years

15

675d

Peoples et al. (1996)

Leucaena leucocephala

Eucalyptus x robusta

Eucalyptus x robusta

Subhumid

Mixed tree plantation

Free growth

15

74

Parrotta et al. (1996)

Sesbania sesban

Zea mays (sequential)

Hyparrhenia rufa

Subhumid

Improved fallow

Free growth

15

56

Chikowo et al. (2004)

Sesbania sesban

None

Eucalyptus deglupta and Grevillea robusta

Subhumid

Improved fallow

Free growth

15

282–363c

Sta˚hl et al. (2002)

Sesbania sesban

None

Eucalyptus deglupta and Grevillea robusta

Subhumid

Improved fallow

Free growth

15

310–356f

Sta˚hl et al. (2005)

a

N enrichment N natural abundance N enrichment N natural abundance

N natural abundance N enrichment N natural abundance; whole tree N enrichment N enrichment

Including fixed N in litter

b

Fixed N in roots estimated not measured

c

Depending on sampling time

d

The authors did not show the whole tree values but provided data on N partitioning between shoot and root that we used for recalculating the estimate

e

All fixed N in the agroecosystem, including net transfer to the crop and soil accumulation

f

Depending on reference species

123

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

and Dawson 1992). We did not observe any trend related to soil acidity when compiling data for Table 2. In a screening of N2-fixing potential of legume trees, Roggy et al. (1999b) classified 110 species as supposed N2-fixers and 33 species as supposed nonN2-fixers in a rain forest in French Guiana. In a selection trial of 25 legume tree species for forestry and agroforestry in Costa Rica, 18 species were inspected for nodulation and 9 for nitrogenase activity; all of them appeared to be N2-fixers, including Acacia, Albizia, Dahlbergia, Erythrina, Inga, and Pithecellobium spp. (Tilki and Fisher 1998). As is unfortunately common, no attempts to estimate N2 fixation were made in these studies but they indicate that a large collection of potential N2-fixers exists for acidic soils. Promising N2-fixers for AFS under these conditions include Acacia mangium, Codariocalyx gyroides (Roxb. ex Link) X.Y. Zhu, Erythrina fusca Lour., E. poeppigiana, Inga edulis, and G. sepium that has been introduced from seasonal to the humid tropics (Tables 2 and 3). Another introduction from seasonal climate, C. calothyrsus, has variable performance (Table 2; Peoples et al. 1996).

Host-bacteria interaction Rhizobial strain may have a significant effect on N2 fixation by the symbiotic system (Lesueur and Sarr 2008). Many of the studies on the host 9 rhizobia interaction have been conducted with seedlings under controlled conditions (Andre´ et al. 2005; Bala and Giller 2001; Makatiani and Odee 2007; Weber et al. 2005). Although they clearly indicate the importance of the interaction, practical conclusions are hard to make beyond nursery production and early establishment in the field. Lesueur and Sarr (2008) observed that apparent growth differences between triple symbiotic systems of C. calothyrsus host and rhizobial and AMF strains disappeared after 5 months in the field. However, plants inoculated with AMF had higher foliar N, P and K content until 12 months after transplanting. According to the genetic analysis, one of the rhizobial strains persisted in the nodules of C. calothyrsus also in the field. Rhizobia-inoculated Acacia mangium grew better than non-inoculated trees up to 39 months after transplanting with detection of the efficient strains in nodules 42 months after

143

transplanting (Galiana et al. 1998). The longest persisting effect of which we are aware, is the detection of rhizobial strains used for inoculating Leucaena leucocephala 10 years after introduction in a soil with only a few native rhizobia capable of nodulating it (Sanginga et al. 1994). Table 1 indicates high diversity of rhizobia and Frankia spp. nodulating different tree genera. Further, both controlled-environment studies (Bala and Giller 2006) and genetic analyses of the rhizobia infecting legumes in the field (Diouf et al. 2007) indicate a high intra-specific genetic diversity among the rhizobia. Thus, the importance of rhizobial inoculation, at least with selected commercial strains, is questionable in small-holder agriculture and scarce resources may be better used for management improvements (Giller and Cadisch 1995). In large-scale nursery production, however, good legume-rhizobia combinations may improve the seedling establishment in the field (cf. Galiana et al. 1998; Lesueur and Sarr 2008). The observation that Rhizobium tropici type rhizobia are the most efficient symbionts in acidic soils, Mesorhizobium strains in intermediate, and Sinorhizobium strains in alkaline soils (Bala and Giller 2006) indicates that the host 9 rhizobia interactions may differ in different soils. Thus, only local trials may be valid for a particular environment. Use of selected inoculants may be necessary when N2-fixing trees are introduced to new areas. When exotic legume or actinorrhizal trees have been planted in new locations where they did not occur previously, scarce nodulation or absence of nodulation has been observed (Diem and Dommergues 1990; Sanginga et al. 1994; Woomer et al. 1988). It has been postulated that in the absence of a compatible host, rhizobial or Frankia strains are confronted with competition from other soil bacteria and cannot maintain their population or can lose host specificity encoding genes (Barnet 1991). Green pruning Green pruning of N2-fixing trees is a common practice in many AFS for avoiding excessive shading of the crop (Kang et al. 1981; Akinnifesi et al. 2008), enhancing nutrient cycling (Beer et al. 1998; Kass et al. 1997), or harvesting tree fodder (Peoples et al. 1996). Symbiotic N2 fixation is often mentioned as an important benefit in these systems because it is

123

Author's personal copy 144

assumed to contribute to increase in soil organic matter and N reserves (Beer et al. 1998; Haggar et al. 1993) or to produce protein-rich fodder (Peoples et al. 1996). Data in Table 2 indicate that pruned trees are often active N2-fixers with higher %Ndfa in pruned than unpruned systems. Because green pruning may result in partial rejuvenation of the trees (Nygren et al. 2000), the higher mean %Ndfa in pruned systems may be related to the higher dependence on N2 fixation in young trees also observable in Table 2. The placement of the pruned trees as an intermediate group between the juvenile and unpruned mature trees in the statistical analysis of Table 2 provides partial support for this ‘‘rejuvenation hypothesis’’. Not all trees fix actively N2 under a periodic pruning regime. Complete nodule turnover was observed in Erythrina poeppigiana in 2 weeks after complete pruning. Renodulation initiated at 10 weeks after pruning and nodulation was abundant at 14 weeks after pruning (Chesney and Nygren 2002; Nygren and Ramı´rez 1995). Leaving only 5 % of foliage in prunings twice-a-year was enough to retain low nodulation in 2-year-old trees and caused stabilisation of nodule growth rather than turnover in 8-year-old trees (Chesney and Nygren 2002). The nodule turnover after complete pruning of E. poeppigiana seemed to be a response to C starvation because of cessation of photosynthesis. Nodulation after foliage regrowth appeared to be regulated by canopy N needs with high nodulation when N flow was unidirectional from roots to foliage and reduction in nodule biomass when N flow to roots initiated (Nygren 1995), in line with the theoretical scheme on the regulation of N2 fixation presented in Fig. 1. Even stronger negative response to green pruning was observed in Erythrina lanceolata, in which nodulation did not recover from complete pruning twice-a-year and partial pruning every three months caused a significant reduction in nodulation and %Ndfa in comparison to unpruned control (Salas et al. 2001). In contrast, pruning regime did not affect the %Ndfa in Gliricidia sepium but total N2 fixation in mass terms was reduced by bimonthly partial prunings in comparison to twice-a-year complete pruning (Nygren et al. 2000). Calliandra calothyrsus, Codariocalyx gyroides (Peoples et al. 1996), Albizia lebbeck (L.) Benth., and Leucaena leucocephala (Kadiata et al. 1997) also actively fixed N2 when pruned heavily. These data indicate that responses to

123

Nutr Cycl Agroecosyst (2012) 94:123–160

green pruning are species-specific and N2 fixation should be estimated in all trials for selecting trees for AFS managed by pruning. Whole tree N2 fixation Much less data were available on the N2 fixation by agroforestry trees in terms of kg [Ndfa] ha-1 year-1, including fixed N in the root system (Table 3) than on the %Ndfa. Dulormne et al. (2003) reported the estimate based on all fixed N tracked in a cut-and-carry system of G. sepium and fodder grass Dichanthium aristatum, i.e. in the trees, grass, and soil over 12 years. The whole system N balance was also used to estimate N2 fixation by 7-year-old Acacia mangium, Acacia auriculiformis A. Cunn. ex. Benth. (BernhardReversat, 1996), and Acacia polyacantha Willd. (Harmand 1998). Other data refer to fixed N in the trees and litter during a time range from 18 months (Sta˚hl et al. 2005) to 12 years (Mercado et al. 2011). All data were converted to correspond to the annual N2 fixation. Table 3 indicates a significant contribution of N2 fixation to the N economy of the trees or whole system. Trees pruned in AFS seem to fix more N than unpruned trees also in mass terms as the top three systems (Dulormne et al. 2003; Peoples et al. 1996; Unkovich et al. 2000) were managed by tree pruning. Mafongoya et al. (2004) cite a ‘‘typical’’ range of N2 fixation by trees in AFS to be 70–200 kg [Ndfa] ha-1 year-1. Six out of the 16 cases compiled in Table 3 exceeded this range, probably because we accepted only estimates based on whole tree harvesting—including root system – or whole system N balance. Schroth et al. (1995) estimated that the N reserves in roots of 0–5 mm diameter of six legume tree species were 60–133 kg [N] ha-1 in 5-year-old fallows in a sub-humid area in Coˆte d’Ivoire. A similar amount, 71.5 kg [N] ha-1, was observed in fine roots of 0–2 mm diameter of Inga edulis shade trees in an organically-grown cacao plantation under humid conditions in Costa Rica (Go´mez Luciano 2008). Higher amount, 190 [N] ha-1, was observed in fine roots of 0–2 mm in a 7-year-old Acacia polyacantha fallow in the subhumid zone of Cameroon (Harmand et al. 2004); whole root system of A. polyacantha contained 342 kg [N] ha-1. Giller (2001) refers to the range of 26–60 % of legume tree N to be belowground, under variable field and experimental conditions.

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

Based on harvestable biomass only, it was estimated that G. sepium would fix only 147 kg [N] ha-1 year-1 (Nygren et al. 2000) in the system studied by Dulormne et al. (2003). However, N2 fixation by G. sepium is the only N input to this system and N balance taking into account N export in fodder harvest, N uptake by the companion grass, and soil N accumulation revealed the considerably higher estimate (Table 3). Thus, it is obvious that below-ground N as well as fixed N released by the trees to the environment should be taken into account when estimating the total amount of N2 fixation in mass terms. These data are still scarce in agroforestry literature.

Agronomic importance of N2 fixation by agroforestry trees State-of-art in agroforestry systems Our review on the estimates of N2 fixation by several tree species used in AFS shows that many of them are active N2-fixers (Table 2) and symbiotic N2 fixation may contribute annually tens or hundreds of kg [N] ha-1 to the farming system (Table 3). A critical review of the original literature in the light of current knowledge on the methods for quantifying the N2 fixation resulted in 16 estimates on the annual N2 fixation per hectare that we considered reliable enough for Table 3. Obviously, more data are needed at fieldand farm-level. Even less data are available for evaluating the agronomic importance of N2 fixation in AFS at landscape, regional, or global level. Two recent reviews on the effects of ‘‘fertiliser trees’’ on crop productivity (Akinnifesi et al. 2010; Garrity et al. 2010) suggest that inclusion of N2-fixing trees into traditional cropping systems may significantly improve crop yields yet evaluation of the contribution of the symbiotic N2 fixation to the benefits was not evaluated. The agronomic importance of N2 fixation by the trees depends on the function of the trees in an AFS. Nitrogen in soil and plant residues, including roots, after cutting an improved fallow is important for the succeeding crop (Chikowo et al. 2004; Mercado et al. 2011; Sta˚hl et al. 2002; 2005). Nitrogen recycling in the pruning residues potentially benefits the crop in alley cropping and other green manure systems (Haggar et al. 1993; Ladha et al. 1993; Unkovich et al. 2000), while N2 fixation by the trees reduces

145

competition for soil N with the crop in unpruned shade tree systems and part of the fixed N is recycled to the crop in leaf and root litter of the trees (Escalante et al. 1984; Nygren and Leblanc 2009; Salas et al. 2001; Santana and Rosand 1985). Animal diet is enhanced by protein-rich tree foliage of the N2-fixing trees in fodder production systems, (Blair et al. 1990; Peoples et al. 1996). All N2-fixing plants form the symbiosis for their own N supply and release N to the environment only when they have it in excess (Fig. 1). Nitrogen fixed by the trees may become available for the crop via indirect or direct pathways. Here, the indirect pathway refers to the complete microbial N cycle, including mineralisation of organic N in the litter or pruning residues to NH4?, partial immobilisation by soil microbiota and mobilisation because of microbial turnover, and nitrification of NH4?. While NH4? is efficiently fixed by cation exchange on soil colloids, NO3- may be lost from the system via leaching to deep soil layers or denitrification, especially if it is available in excess with respect to plant N needs. Soil N cycle has been widely studied – also in the AFS with legume trees (Babbar and Zak 1994, 1995; Dulormne et al. 2003; Hergoualc’h et al. 2007, 2008; Kanmegne et al. 2006; Mafongoya et al. 1998)—and a discussion of the vast literature on the topic is out of our scope. Direct pathway refers here to the transfer of N from N2-fixing trees to crops either via a common mycelial network (CMN) of mycorrhizae-forming fungi colonising both trees and crops (He et al. 2003; Jalonen et al. 2009b) or absorption of tree root exudates by the crop before the complete decomposition cycle (Fustec et al. 2010; Jalonen et al. 2009a). The latter pathway requires that either crop plants absorb simple amino acids (Lambers et al. 2008; Na¨sholm et al. 2009) from the root exudates or the N2-fixing plants exude inorganic N (Fustec et al. 2010). Fate of fixed nitrogen In spite of the low C:N ratio (Mafongoya et al. 1998), legume tree mulches seem to have a relatively low efficiency as an immediate N source for a crop. Maize (Zea mays L.) gained 11 % of its N, ca. 10 kg [N] ha-1, from pruning residues of Gliricidia sepium and Erythrina poeppigiana in an alley cropping trial under humid tropical conditions (Haggar et al. 1993). In another experiment, maize used ca. 10 % of the N

123

Author's personal copy 146

available in the pruning residues of Leucaena leucocephala (Vanlauwe et al. 1996). This apparently low efficiency depends on a cascade of factors. First, although amino acids and oligopeptides decompose fast, some proteins decompose slowly or may even be recalcitrant (Paul and Clark 1996). Thus, not all residual N becomes available for an annual crop during a cropping cycle. Second, the N release is also restricted by the general litter quality: if the mulch C is not easily available for the decomposer microbes as an energy source, also N release from the residue slows down. Cellulose and hemicellulose are easily decomposable C compounds while lignin is recalcitrant (Paul and Clark 1996). Further, many legume mulches have high content of polyphenols that may be toxic to the decomposers. Nitrogen release rate is significantly higher from the legume mulches with a high N to polyphenols ratio than from residues with a low ratio (Barrios et al.1997, Ndufa et al. 2009). Following to these factors, 10–30 % of N in legume mulches is released in a month (Mafongoya et al. 1998) typically followed by an exponential decrease in the N release rate with time. Third, soil microbiota are strong competitors for soil N with the plants; free N in soil solution is absorbed by microbes on average within 24 h after its release (Jones et al. 2005). Mobility of N in soil depends on soil and N type. In a loamy soil, the 24-h diffusion distance of NH4?, amino acids, and NO3- is ca. 1.7, 1.5, and 10.2 mm, respectively (Jones et al. 2005). Thus, N mineralisation must occur in the rhizosphere to be useful for a plant. Otherwise, it will be absorbed by microbiota. Based on the 24-h diffusion distance and data on fine root densities of cacao in a plantation with Inga edulis shade (Go´mez Luciano 2008), we estimated that the cacao rhizosphere for NH4? absorption was 45.6 % out of total soil volume in the 0–2 cm soil depth but only 16.4 % in the 2–10 cm depth (Nygren and Leblanc, unpublished). If the soil has a high nitrification rate, the more mobile NO3- is in better supply for plants but the high mobility makes it also subject to leaching to deeper soil layers, where it may become unavailable for plant roots (Babbar and Zak 1995; Harmand et al. 2007a). The apparently low efficiency of new mulch as a crop N source has led some authors to argue that legume tree mulch is more important for long-term build-up of soil organic matter and N reserves than

123

Nutr Cycl Agroecosyst (2012) 94:123–160

immediate crop N nutrition (Beer et al. 1998; Haggar et al. 1993; Kass et al. 1997). We partially share this opinion. However, the role of root litter in the N cycle of AFS has been neglected, although Nye and Greenland (1960) proposed more than 50 years ago that dead tree roots may be an important nutrient source for associated crops because they decompose in close proximity of the absorbing crop roots. Pot experiments with agroforestry combinations of maize with Paraserianthes falcataria (L.) I.C. Nielsen (Chintu and Zaharah 2003) and cacao with I. edulis (Ka¨hko¨la¨ et al. 2012) suggest that root litter of the legume tree may be a more efficient N source for the crop than leaf mulch or litter. Further, ca. 300 kg ha-1 year-1 of N fixed by G. sepium was recycled below-ground to the soil and the associated fodder grass Dichanthium aristatum in a cut-and-carry fodder production system, where above-ground litter was eliminated by intensive fodder harvesting (Dulormne et al. 2003). Direct N transfer between plants Reviews of N transfer via CMN (He et al. 2003; Simard et al. 2002) and N exudation from legume roots (Fustec et al. 2010; Wichern et al. 2008) have been recently published but data on AFS are scarce. Direct N flow between plants is bidirectional and its net effect may be almost nil to both components (He et al. 2004; Johansen and Jensen 1996). Net flow seems to be more common from a N2-fixing to a non-N2-fixing plant; e.g. from soybean to maize (Bethlenfalvay et al. 1991), from Alnus subcordata C.A. Mey. and Elaeagnus angustifolia to Prunus avium L. (Roggy et al. 2004), and from soybean and peanut (Arachis hypogaea L.) to associated weeds (Moyer-Henry et al. 2006). In a pot study, direct N transfer from A. senegal to durum wheat (Triticum turgidum L.) was enhanced when crop N uptake was stimulated by high P availability and competition was low (Isaac et al. 2012). In several coffee plantations in Burundi with different legume shade tree species, 6–22 % of N in coffee leaves was of atmospheric origin (Snoeck et al. 2000). In a cut-and-carry fodder production system, 27–35 % of N in D. aristatum grass (53–68 kg ha-1) was of atmospheric origin. Nitrogen isotopic data suggested that atmospheric N was directly transferred from the associated heavily pruned G. sepium (Sierra and Nygren 2006). In a cacao plantation with I. edulis shade, 8–25 % of N in cacao leaves was of

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

atmospheric origin (Nygren and Leblanc 2009). Data on N isotopic relationships also suggest the possibility of direct N transfer from L. leucocephala to understorey weeds in a short-rotation plantation (van Kessel et al. 1994) and from Inga oerstediana Benth. ex Seem. to coffee and non-legume shade tree Liquidambar styraciflua L. in organic coffee farms in Chiapas, Mexico, (Grossman et al. 2006) although the authors themselves did not draw this conclusion. Identification of the direct N transfer pathways is difficult under AFS field conditions. Root exudates of many legume trees have a remarkably low C:N ratio: 3.3–5.7 in G. sepium (Jalonen et al. 2009b), 4.7–6.8 in Robinia pseudoacacia (Uselman et al. 1999), and 5.2 in I. edulis (Nygren and Leblanc, unpublished data). Thus, they are potentially good N sources for an associated crop but tree and crop rhizospheres should overlap for the tree root exudates to be useful for the crop. Formation of an effective CMN between trees and crops in an AFS requires that both share compatible strains of mycorrhizal fungi: anastomoses, functional connections that allow material transfer between mycelia of two fungi, have been observed only between fungi of the same population yet they may differ genetically (Croll et al. 2009). Glomus etunicatum Becker & Gerdemann and Gigaspora albida Schenck & Perez used for inoculating Calliandra calothyrsus spread their mycelia also to associated maize and common bean (Phaseolus vulgaris L.) (Ingleby et al. 2007). Millet (Pennisetum americanum (L.) Leeke) was effectively colonised by inocula from Acacia nilotica (L.) Willd. ex Delile, A. tortilis (Forssk.) Hayne, and Prosopis juliflora (Sw.) DC. (Diagne et al. 2001). Inga edulis positively responded to AMF originating from roots of cacao in a crossinoculation trial indicating the potential of formation of a CMN between these species (Iglesias et al. 2011). In a pot study, N was transferred from G. sepium to D. aristatum both via the CMN and root exudates (Jalonen et al. 2009b). Under field conditions, both species were colonised by Rhizosphagus intraradices (ex. Glomus intraradices) and grass colonisation was more abundant between the tree rows than in an adjacent pure grass plot (Jalonen et al. 2012). It seems obvious that the potential for direct N transfer from legume trees to crops in AFS exists and positive evidence on this interaction has been found in a few cases. However, the phenomenon requires

147

further study before we well understand its occurrence and importance in different AFS.

Management of N2 fixation Possibilities to manage N2 fixation and N recycling are quite limited in AFS with unpruned N2-fixing trees. According to the data in Table 2, sampling time was one of the factors affecting the %Ndfa. We interpreted this to reflect the seasonality of N2 fixation. If the climate is seasonal, the trees and crops often follow the same phenological cycle with higher metabolic activity and productivity during the same season. Because N2 fixation seems to be connected at least to some extend to the N needs of the canopy (Fig. 1; Nygren 1995), it is most active during the time of the highest general metabolic activity. Follow-up of annual biomass and N2 fixation dynamics of Erythrina lanceolata used as living support for vanilla (Vanilla planifolia Andrews) under a seasonal climate in Costa Rica indicated that main litterfall occurred during the dry season (Berninger and Salas 2003) when no N2 fixation was detected (Salas et al. 2001). Litterfall was close to nil during time of the most active foliage growth and N2 fixation. Thus, although the timing of N2 fixation and N recycling in unpruned multistrata AFS has not received much attention, we may expect the N2 fixation to occur at the time of the highest productivity of both trees and crops, thus, reducing competition between trees and crops. An interesting case is the ‘‘reverse phenology’’ of Faidherbia albida in the African drylands (Garrity et al. 2010): the tree is leafy during the dry season. The low water consumption during the cropping season and the use of deep-water reserves out of cropping season reduce competition for water with the associated crops (Roupsard et al. 1999). However, N2 fixation is effective only 2–3 months during a year when foliage growth begins (Roupsard 1997). Because N2 fixation occurs out of the cropping season F. albida adds N for crop via litterfall at the beginning of cropping season. Green pruning may also reverse phenology; Central American cattle ranchers prune G. sepium at the beginning of dry season so that the trees remain leafy and provide dry season fodder (Herna´ndez and Benavides 1994). Green pruning is generally timed according to the crop phenology. Costa Rican coffee farmers

123

Author's personal copy 148

traditionally prune shade trees leaving only a few branches (ca. 5–10 % of foliage) to promote coffee flowering at the end of the drier season and make a complete pruning for promoting the ripening of berries about half year later. The highest coffee N demand occurs 6–17 weeks after blossoming when the growing berries may consume 95 % of newly assimilated N (DaMatta et al. 2007). The popular shade tree E. poeppigiana remains unnodulated about 10 weeks following a complete pruning (Nygren and Ramı´rez 1995) but it retains reduced nodulation after a partial pruning (Chesney and Nygren 2002). Similar pattern has been observed in G. sepium but its renodulation rate is higher after a pruning (Nygren and Cruz 1998). Thus, in coffee AFS we may envision partial synchrony between coffee N needs and N2 fixation and N recycling from the shade trees. The green mulch from the pruning to promote blossoming may release N for active coffee foliage growth that occurs at the time of coffee flowering (DaMatta et al. 2007) but trees fix only limited amount of N2 for their own foliage regrowth, thus, potentially competing for soil N and reabsorbing N released from the pruning residues. Nitrogen released from nodule (Escalante et al. 1984; Nygren and Ramı´rez 1995) and fine root turnover (Chesney and Nygren 2002) forms an additional N source, which may be more efficient than above-ground mulch (Chintu and Zaharah 2003; Ka¨hko¨la¨ et al. 2012). During the high coffee N need for growing berries (DaMatta et al. 2007), the shade trees probably fix actively N2, which reduces competition between the trees and the crop. The coffee N needs are low during the second annual pruning for promoting berry ripening (DaMatta et al. 2007), and the residual N probably contributes to regrowth of the trees themselves, accumulation of soil organic N, or N losses through NO3- leaching. In alley cropping, hedgerow trees are pruned before sowing of the annual crop and, if the crop requires and trees tolerate, in a later phase of cropping season (Akinnifesi et al. 2010; Haggar et al. 1993; Kang et al. 1981; Vanlauwe et al. 1996). The pruning often refers to complete defoliation. Thus, we may expect that a complete nodule turnover follows (Nygren and Cruz 1998; Nygren and Ramı´rez 1995) and trees do not fix N2 at the beginning of the cropping cycle. Nitrogen recycled in pruning residues is a strong input to the soil when the crop is relatively small. Many authors recommend a combination of rapidly and slowly

123

Nutr Cycl Agroecosyst (2012) 94:123–160

decomposing tree mulches (‘‘high-quality’’ and ‘‘lowquality’’ mulch) for enhancing the synchrony between the crop N demand and N release (Mafongoya et al. 1998; Nair et al. 1999; Palm 1995; van Noordwijk et al. 1996). These authors, however, ignored the effect of nodule turnover on trees’ N uptake. We argue that temporarily non-N2-fixing trees enforce the root safety-net (van Noordwijk et al. 1996) by at least partially reabsorbing N leaching from the mulch application. Further, the nodule and root turnover following the pruning may form a ‘‘high-quality mulch’’ that releases N at the beginning of the cropping cycle. The trees renodulate and fix N2 – unless too heavily pruned – during the crop flowering and seed production, thus, reducing competition with the crop. Although at least partial synchrony between N2 fixation, N recycling, and crop N use is possible in AFS with both perennial and annual crops, it must be noted that precision agriculture is not possible in green manure systems. Thus, the main benefit of N2-fixing trees in AFS is probably the long-term accumulation of soil N reserves (Haggar et al. 1993) although belowground N recycling and direct N transfer between plants probably provide N for more immediate crop N needs than foliage litter or mulch applied on soil surface (Chintu and Zaharah 2003; Dulormne et al. 2003; Jalonen et al. 2009b; Ka¨hko¨la¨ et al. 2012). The long-term accumulation of organic matter and nutrients to soil is essential in improved fallow systems, in which cropping relies on the residual N after the fallow enhanced with N2-fixing trees (Chikowo et al. 2004; Mercado et al. 2011; Sta˚hl et al. 2002, 2005). Few attempts have been made for studying the fate of the high N release to soil after clearing the fallow for cropping. Nitrate leaching rates two year after clearing 7-year-old Acacia polyacantha, Eucalyptus camaldulensis Dehnh., and Senna siamea fallows were similar, ca. 10–15 kg [NO3–N] year-1 (Oliver et al. 2000).

Crop N needs and tree N2 fixation Perennial crops coffee, cacao, and tea (Camellia sinensis (L.) Kuntze) cover globally ca. 21.6 Mha of agricultural land (Table 4). Perennial crops are often grown in AFS with N2-fixing shade trees (Beer et al. 1998; Kass et al. 1997) although global estimates are missing for other perennials than cacao; 7.8 Mha of

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160

cacao is cultivated in AFS (Zomer et al. 2009), which is 89 % out of the global cacao cultivation area of 8.73 Mha (Table 4). Central American, Caribbean, and Andean coffee is mostly shaded while full-sun cultivation prevails in large Brazilian plantations. We calculated the average yield of the perennial crops based on the 2009 cultivation area and production data (FAOStat 2011) and estimated the N export in crop harvest using the estimates on N needed for producing 1 Mg of harvest according to Bertsch (2003): 30, 36, and 57 kg Mg-1 for green coffee, cacao beans, and tea leaves, respectively. The average yields of all these three crops were relatively modest and N export in the harvest was, consequently, quite low (Table 4). In addition to the N export in the harvest, part of N is used for producing permanent biomass. We compiled data on N accumulation to permanent biomass of coffee, cacao, non-legume shade trees Cordia alliodora (Ruiz & Pav.) Oken and Eucalyptus deglupta Blume, and legume shade tree Erythrina poeppigiana in Table 5. Combining these few data with statistics in Table 4, we can roughly estimate that on average ca. 40–50 kg [N] ha-1 year-1 is exported in the average harvest or immobilised in permanent biomass in unshaded cacao and coffee plantations and ca. 60–80 kg [N] ha-1 year-1 in shaded plantations. Table 3 shows that many trees commonly used in AFS may fix enough N for fulfilling the current or even increasing N needs of perennial crops. As green pruning seems to enhance N2 fixation both proportionally (Table 2) and absolutely (Table 3), it is a recommended practice for shade trees over perennial crops. Green pruning of legume trees Erythrina fusca, E. poeppigiana, and I. edulis under typical management potentially recycles 80, 70–115, and 100 kg ha-1 year-1, respectively, of N fixed from atmosphere to the soil (Leblanc et al. 2007), which is enough for compensating N accumulation to permanent crop biomass and harvest loss in1.5–3 Mg ha-1 yield. Intensively managed large-scale perennial crop plantations may be over-fertilised; e.g. the typical coffee fertilisation rate in Costa Rica ranging from 150 (Harmand et al. 2007a) to 350 kg [N] ha-1 year-1 (Hergoualc’h et al. 2007) would be enough for compensating the N export in the harvest of 4–11 Mg ha-1 of green coffee yet the average yield in Costa Rica is 0.93 Mg ha-1 (calculated from cropping area and production data in FAOStat 2011).

149

This results in NO3- leaching loss equivalent to N needed for producing 1–4 Mg of green coffee (Babbar and Zak 1995; Harmand et al. 2007a, b) and increased emissions of nitrous oxide (N2O) (Hergoualc’h et al. 2008), which is a greenhouse gas with 206 times higher atmospheric forcing potential than CO2. Including N2-fixing trees to highly fertilised coffee farms increases N2O emissions but the emissions peak after fertilisation in both shaded and unshaded systems (Hergoualc’h et al. 2008). Perhaps the best known practice for combining N2fixing trees with annual crops is alley cropping (Akinnifesi et al. 2008; Kang et al. 1981) that was intensively studied in the 1980s and 1990s but later abandoned. However, it showed some promise in certain humid and subhumid areas (Akinnifesi et al. 2008; Kass et al. 1997) and an improved version is now widely adopted by small-scale farmers in Southern Africa for maize cropping (Akinnifesi et al. 2008, 2010). The main improvement was achieved by modifying the spatial arrangement of the trees for reducing interference competition between trees and maize. Gliricidia sepium has been the most successful tree species in these systems (Akinnifesi et al. 2008), which have been adopted by over 120,000 Malawian farmers (Garrity et al. 2010). Palm (1995) estimated that 40 kg of N is needed for 1 Mg of maize yield. Gliricidia sepium may recycle 70–126 kg ha-1 year-1 of N fixed from atmosphere (estimated from alley cropping data in Ladha et al. 1993 and Rowe et al. 1999), which is sufficient for harvesting ca. 1.7–3.1 Mg ha-1 year-1 of maize. Akinnifesi et al. (2010) report even higher sustained maize yields with G. sepium without fertiliser in Malawi (around 4 Mg ha-1 year-1, depending on the year vs. ca. 1.5 Mg ha-1 year-1 in pure maize with low N fertilisation). This may imply that also the tree root safety-net, which retains nutrients in the system out of the cropping season (van Noordwijk et al. 1996), is important for these systems. Further, N2 fixation is probably underestimated because the estimates of Ladha et al. (1993) and Rowe et al. (1999) are based on tree prunings only. Dulormne et al. (2003) found that 160 kg ha-1 year-1 of N fixed by G. sepium accumulated in soil. Direct N transfer to a non-N2-fixing crop may also occur (Sierra and Nygren 2006). As far as we know, the studies by Dulormne et al. (2003) and Bernhard-Reversat (1996) are the only ones that account for the N fixed from atmosphere in the soil

123

Author's personal copy 150

Nutr Cycl Agroecosyst (2012) 94:123–160

Table 4 Global cultivation area and production of perennial crops in 2009 (FAOStat 2011), and an estimate of annual N export in the harvest of these crops Commodity and area

Yield (Mg ha-1)

N export (kg ha-1 year-1)

Total N export (Mg year-1)

Total area (ha)

Total production (Mg)

Sub-Saharan Africa

5,935,274

2,639,788

0.445

16.0

95,032

Tropical Americas

1,585,728

548,360

0.346

12.4

19,741

Cacao beans

South and East Asia

1,068,476

837,766

0.784

28.2

30,160

8,733,093

4,082,270

0.467

16.8

146,962

Sub-Saharan Africa

2,060,527

1,006,318

0.488

14.7

30,190

Tropical Americas

5,474,941

4,807,644

0.878

26.3

144,229

South and East Asia

2,250,612

2,468,351

1.097

32.9

74,051

0.848

25.4

250,279 30,267

Worlda Green coffee

a

World

9,841,317

8,342,636

Tea leaves Sub-Saharan Africa

283,161

530,992

1.875

106.9

Tropical Americas

45,539

85,477

1.877

107.0

4,872

2,680,869

3,326,333

1.241

70.7

189,601

3,014,909

3,950,047

1.310

74.7

225,153

South and East Asia Worlda a

Includes areas of low production not listed separately

and associated crop in addition to the trees. These kinds of studies would be most welcome for different AFS but the main constraint is the lack of sufficient time series for constructing the system N balance over years.

Concluding remarks Symbiotic N2 fixation in AFS should not be studied as an isolated process because even the ‘‘N2-fixing component’’ in an AFS is a tripartite symbiotic system based on a plant capable of forming symbiosis with both N2-fixing bacteria and mycorrhizae-forming fungi. The bacterial partners of the symbiosis are highly diverse both taxonomically and functionally. So far, 98 rhizobial species forming N2-fixing symbioses with legumes have been identified with most of the diversity probably still remaining unrecognised. Actinorrhizal symbioses are much less studied than legume systems and they form an unexplored resource for AFS. Plants rely more on the N2-fixing symbiosis when the plant N needs exceed soil N supply. Thus, soil N supply reduced in AFS by the crop N use and N export in harvest may enhance the N2 fixation above levels observed in natural ecosystems. Mycorrhizae enhance P nutrition but they also appear to have other

123

still poorly understood functions in the tripartite symbiosis. They also form CMN between trees’ and crops’ symbionts, which create a direct N transfer pathway. Quantification of N2 fixation under field conditions is a challenging yet necessary task as controlledenvironment studies with seedlings provide little information on functioning of mature trees in an AFS. Nitrogen isotopic analyses or detailed, long-term whole system N balances may provide the most reliable estimates on the N2 fixation by trees. The general average of %Ndfa for 19 tree species used in AFS was 59, with a higher proportion in juvenile and pruned trees and lower in unpruned trees. High variability was observed in drylands while N2 fixation was active in most of the AFS in the humid and subhumid areas. In mass terms, N2-fixation may annually add from tens to hundreds of kilograms of N per hectare to an AFS. Few data were available on fixed N in system components other than the trees. The reports published so far indicate that estimates on annual N2 fixation based on trees only may hide considerable direct transfer of fixed N from trees to crops and an important rhizodeposition of tree N to the soil. It seems that symbiotic N2 fixation is indeed an underestimated resource in AFS.

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 Table 5 Nitrogen accumulation to the permanent woody biomass in coffee (Coffea arabica L.) and cacao (Theobroma cacao L.) plantations

Species

151

Management

N accumulation (kg ha-1 year-1)

References

Fassbender (1987)

Coffea arabica

Shaded

11

Cordia alliodora

Free growth

25

System Coffea arabica

Shaded

36 11

Erythrina poeppigiana

Pruned

11

System

Fassbender (1987)

22

Coffea arabica

Unshaded

15

Harmand et al. (2007a, b)

Coffea arabica

Shaded

13

Harmand et al. (2007a, b)

Eucalyptus deglupta

Free growth

System Theobroma cacao

Shaded

Cordia alliodora

Free growth

System Shaded

Erythrina poeppigiana

Free growth

Variation in the N2 fixation by trees in AFS depends on both intrinsic factors of the trees and their symbionts and the environment. Tree species selection and in some cases inoculation with compatible bacteria are the main options for managing the intrinsic factors. Species and provenance selection is also the only available response to the challenges caused by the macro environment, especially in dryland AFS, while several options exist for managing the microenvironment such as proper spacing of trees and crops. Green pruning that is practiced for reducing crop shading and enhancing nutrient cycling, seems to rejuvenate the trees and maintain high levels of N2 fixation. However, green pruning may temporarily impede N2 fixation by simultaneously reducing both the C flow to the roots and microsymbionts and tree N requirements. This may reduce the synchrony of N2 fixation with crop N needs. The N2-fixing trees mostly improve the crop N supply by long-term build-up of soil N reserves. We may envision a role for N2-fixing trees in different kinds of AFS. In small-scale, low-input perennial cropping systems, N2-fixing trees may provide enough N for sustained or increasing yields. In large-scale, highly fertilised perennial cropping systems, they contribute to the reduction in the use of industrial fertilisers. Recent developments of legume trees mixed with annual crops show much promise for low-input farming. The highest N2 fixation activity

9

Fassbender et al. 1988

28 37

Theobroma cacao System

7 20

9

Fassbender et al. 1988

30 39

was observed in improved fallows and intensive tree fodder production systems. Thus, we may envision significant contribution of N2 fixation to the production of N-rich mulch for soil improvement in the improved fallows and to the high yields of N-rich browse for domestic animals in protein banks and other tree fodder systems. In order to manage the symbiotic N2 fixation by trees in AFS, basic research is needed on the functioning of the tripartite tree-rhizobia-mycorrhizae symbiosis; on the tree N rhizodeposition to soil via exudates and root turnover; on the direct N transfer from trees to crops; and if the trees really enhance P supply by excretion of extracellular phosphatases. Applied research is needed on tree and symbiont selection, especially for drylands, and on the effects of AFS management on N2 fixation. Acknowledgments An early version of this review was presented in the 2nd World Congress of Agroforestry (Nairobi, August 2009). We thank Dr Anne-Marie Domenach for inspiring discussions and comments on a draft of this review. The contribution of PN was funded by the Academy of Finland (grant 129166).

References Acosta-Dura´n C, Martı´nez-Romero E (2002) Diversity of rhizobia from nodules of the leguminous tree Gliricidia sepium, a natural host of Rhizobium tropici. Arch Microbiol 178:161–164

123

Author's personal copy 152 Akimov V, Dobritsa S (1992) Grouping of Frankia strains on the basis of DNA relatedness. Syst Appl Microbiol 15:372–379 Akinnifesi FK, Chirwa PW, Ajayi OC, Sileshi G, Matakala P, Kwesiga FR, Harawa H, Makumba W (2008) Contributions of agroforestry research to livelihood of smallholder farmers in Southern Africa: 1. Taking stock of the adaptation, adoption and impact of fertilizer tree options. Agric J 3:58–75 Akinnifesi FK, Ajayi OC, Sileshi G, Chirwa PW, Chianu J (2010) Fertiliser trees for sustainable food security in the maize-based production systems of East and Southern Africa. A review. Agron Sustain Dev 30:615–629 Allison SD, Nielsen C, Hughes RF (2006) Elevated enzyme activities in soils under the invasive nitrogen-fixing tree Falcataria moluccana. Soil Biol Biochem 38:1537–1544 Amarger N, Macheret V, Laguerre G (1997) Rhizobium gallicum sp. nov. and Rhizobium giardinii sp. nov., from Phaseolus vulgaris nodules. Int J Syst Bacteriol 47:996–1006 Amtmann A, Blatt MR (2009) Tansley review. Regulation of macronutrient transport. New Phytol 181:35–52 An C, Riggsby W, Mullin B (1985) Relationships of Frankia isolates based on deoxyribonucleic acid homology studies. Int J Syst Bacteriol 35:140–146 Andre´ S, Galiana A, Le Roux C, Prin Y, Neyra M, Duponnois R (2005) Ectomycorrhizal symbiosis enhanced the efficiency of inoculation with two Bradyrhizobium strains and Acacia holosericea growth. Mycorrhiza 15:357–364 Andrews M, James EK, Sprent JI, Boddey RM, Gross E, dos Reis FB (2011) Nitrogen fixation in legumes and actinorhizal plants in natural ecosystems: values obtained using 15 N natural abundance. Plant Ecol Diversity 4:131–140 Arau´jo AFS, Burity HA, Lyra M do CCP (2001) Influeˆncia de diferentes nı´veis de nitrogeˆnio e fo´sforo em leucena inoculada com Rhizobium e fungo micorrı´zico arbuscular. Ecossistema 26:35–38 Arnebrant K, Ek H, Finlay RD, So¨derstro¨m B (1993) Nitrogen translocation between Alnus glutinosa (L.) Gaertn, seedlings inoculated with Frankia sp. and Pinus contorta Doug. ex Loud seedlings connected by a common ectomycorrhizal mycelium. New Phytol 124:231–242 Aronson J, Ovalle C, Avendan˜o J, Longeri L, del Pozo A (2002) Agroforestry tree selection in central Chile: biological nitrogen fixation and early plant growth in six dryland species. Agrofor Syst 56:155–166 Augusto L, Crampon N, Saur E, Bakker MR, Pellerin S, de Lavaissiere C, Trichet P (2005) High rates of nitrogen fixation of Ulex species in the understory of maritime pine stands and the potential effect of phosphorus fertilization. Can J For Res 35:1183–1192 Babbar LI, Zak DR (1994) Nitrogen cycling in coffee agroecosystems: net nitrogen mineralization and nitrification in the presence and absence of shade trees. Agric Ecosys Environ 48:107–113 Babbar LI, Zak DR (1995) Nitrogen loss from coffee agroecosystems in Costa Rica. Leaching and denitrification in the presence and absence of shade trees. J Environ Qual 24:227–233 Bala A, Giller KE (2001) Symbiotic specificity of tropical tree rhizobia for host legumes. New Phytol 149:495–507

123

Nutr Cycl Agroecosyst (2012) 94:123–160 Bala A, Giller KE (2006) Relationships between rhizobial diversity and host legume nodulation and nitrogen fixation in tropical ecosystems. Nutr Cycling Agroecosys 76:319–330 Bala A, Murphy P, Giller KE (2003) Distribution and diversity of rhizobia nodulating agroforestry legumes in soils from three continents in the tropics. Mol Ecol 12:917–930 Balachandar D, Raja P, Kumar K, Sundaram SP (2007) Nonrhizobial nodulation in legumes. Biotech Mol Biol Rev 2:49–57 Baldani JI, Caruso L, Baldani VLD, Goi SR, Do¨bereiner J (1997) Recent advances in BNF with non-legume plants. Soil Biol Biochem 29:911–922 Barea JM, Azco´n R, Azco´n-Aguilar C (1992) Vesicular-arbuscular mycorrhizal fungi in nitrogen-fixing systems. In: Norris JR, Read DJ, Varma AK (eds) Techniques in the study of mycorrhiza. Methods in Microbiology, vol 24. Academic Press, London, pp 391–416 Barnet YM (1991) Ecology of legume root-nodule bacteria. In: Dilworth MJ, Glenn A (eds) Biology and biochemistry of nitrogen fixation. Elsevier, Amsterdam, pp 199–228 Barrios E, Kwesiga F, Buresh RJ, Sprent JI (1997) Light fraction soil organic matter and available nitrogen following trees and maize. Soil Sci Soc Am J 61:826–831 Beer J, Muschler R, Kass D, Somarriba E (1998) Shade management in coffee and cacao plantations. Agrofor Syst 38:139–164 Berliner R, Torrey J (1989) On tripartite Frankia-mycorrhizal associations in the Myricaceae. Can J Bot 67:1708–1712 Bernhard-Reversat F (1996) Nitrogen cycling in tree plantations grown on a poor sandy savanna soil in Congo. Appl Soil Ecol 4:161–172 Berninger F, Salas E (2003) Biomass dynamics of Erythrina lanceolata as influenced by shoot-pruning intensity in Costa Rica. Agrofor Syst 57:19–28 Bertsch F (2003) Absorcio´n de nutrimentos por los cultivos. ACCS, San Jose´ Bethlenfalvay GJ (1992) Vesicular-arbuscular mycorrhizal fungi in nitrogen-fixing legumes: problems and prospects. In: Norris JR, Read DJ, Varma AK (eds) Techniques in the study of mycorrhiza. Methods in microbiology vol 24. Academic Press, London, pp 375–388 Bethlenfalvay GJ, Newton WE (1991) Agro-ecological aspects of the mycorrhizal, nitrogen-fixing legume symbiosis. In: Keister DL, Cregan PB (eds) The rhizosphere and plant growth. Beltsville symposia in agricultural research 14. Kluwer, Dordrecht, pp 349–354 Bethlenfalvay GJ, Reyes-Solis MG, Camel SB, Ferrera-Cerrato R (1991) Nutrient transfer between the root zones of soybean and maize plants connected by a common mycorrhizal mycelium. Physiol Plantarum 82:423–432 Binkley D, Senock R, Cromack K Jr (2003) Phosphorus limitation on nitrogen fixation by Falcataria seedlings. For Ecol Manag 186:171–176 Blair G, Catchpoole D, Horne P (1990) Forage tree legumes: their management and contribution to the nitrogen economy of wet and humid tropical environments. Adv Agron 44:27–54 Boddey RM, Peoples MB, Palmer B, Dart PJ (2000) Use of the 15 N natural abundance technique to quantify biological nitrogen fixation by woody perennials. Nutr Cycling Agroecosys 57:235–270

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 Boivin C, Giraud E (1999) Molecular symbiotic characterization of rhizobia: towards a polyphasic approach using Nod factors and nod genes. In: Martı´nez-Romero E, Herna´ndez G (eds) Highlights of nitrogen fixation research. Kluwer/ Plenum Publishers, New York, pp 295–299 Bouillet JP, Laclau JP, Gonc¸alves JLM, Moreira MZ, Trivelin PCO, Jourdan C, Silva EV, Piccolo MC, Tsai SM, Galiana A (2008) Mixed-species plantations of Acacia mangium and Eucalyptus grandis in Brazil 2: Nitrogen accumulation in the stands and biological N2 fixation. For Ecol Manag 255:3918–3930 Bryan JA (2000) Nitrogen fixation of leguminous trees in traditional and modern agroforestry systems. In: Ashton MS, Montagnini F (eds) The silvicultural basis for agroforestry systems. CRC Press, Boca Raton, pp 161–182 Burleigh SH, Dawson JO (1994) Occurence of Myrica-nodulating Frankia in Hawaian volcanic soils. Plant Soil 164:283–289 Cardoso IM, Kuyper TW (2006) Mycorrhizas and tropical soil fertility. Agric Ecosyst Environ 116:72–84 Casida LE Jr (1982) Ensifer adhaerens, gen. nov., sp. nov. a bacterial predator of bacteria in soil. Int J Syst Bacteriol 32:339–345 Castenholz RW (2001) Cyanobacteria, oxygenic photosynthetic bacteria. In: Boone DR, Castenholz RW (eds) Bergey’s manual of systematic bacteriology, vol I, 2nd edn. Springer, Berlin, pp 473–599 Cavard X, Augusto L, Saur E, Trichet P (2007) Field effect of P fertilization on N2 fixation rate of Ulex europaeus. Ann For Sci 64:875–881 Chalk PM, Ladha JK (1999) Estimation of legume symbiotic dependence: an evaluation of techniques based on 15N dilution. Soil Biol Biochem 31:1901–1917 Chatarpaul L, Chakravarty P, Subramaniam P (1989) Studies in tetrapartite symbioses. I. Role of ecto- and endomycorrhizal fungi and Frankia on the growth performance of Alnus incana. Plant Soil 118:145–150 Chen WX, Tan ZY, Gao JL, Li Y, Wang ET (1997) Rhizobium hainanense sp. nov., isolated from tropical legumes. Int J Syst Bacteriol 47:870–873 Chesney P, Nygren P (2002) Fine root and nodule dynamics of periodically pruned hedgerow trees in an alley cropping system in Costa Rica. Agrofor Syst 56:259–269 Chikowo R, Mapfumo P, Nyamugafata P, Giller KE (2004) Woody legume fallow productivity, biological N2-fixation and residual benefits to two successive maize crops in Zimbabwe. Plant Soil 262:303–315 Chintu R, Zaharah AR (2003) Nitrogen uptake of maize (Zea mays. L) from isotope-labeled biomass of Paraserianthes falcataria grown under controlled conditions. Agrofor Syst 57:101–107 Colonna JP, Thoen D, Ducoussu M, Badji S (1991) Comparative effects of Glomus mosseae and P fertilizer on foliar mineral composition of Acacia seneyal seedlings inoculated with Rhizobium. Mycorrhiza 1:35–38 Craine JM, Elmore AJ, Aidar MPM, Bustamante M, Dawson TE, Hobbie EA, Kahmen A, Mack MC, McLauchlan KK, Michelsen A, Nardoto GB, Pardo LH, Pen˜uelas J, Reich PB, Schuur EAG, Stock WD, Templer PH, Virginia RA, Welker JM, Wright IJ (2009) Global patterns of foliar nitrogen isotopes and their relationships with climate,

153 mycorrhizal fungi, foliar nutrient concentrations, and nitrogen availability. New Phytol 183:980–992 Croll D, Giovannetti M, Koch AM, Sbrana C, Ehinger M, Lammers PJ, Sanders IR (2009) Nonself vegetative fusion and genetic exchange in the arbuscular mycorrhizal fungus Glomus intraradices. New Phytol 181:924–937 DaMatta FM, Ronchi CP, Maestri M, Barros RS (2007) Ecophysiology of coffee growth and production. Braz J Plant Physiol 19:485–510 Dawson JO (2008) Ecology of actinorhizal plants. In: Pawlowski K, Newton WE (eds) Nitrogen fixing actinorhizal symbioses. Springer, The Netherlands, pp 199–227 de Lajudie P, Willems A, Pot B, Dewettinck D, Maestrojuan G, Neyra M, Collins MD, Dreyfus B, Kersters K, Gillis M (1994) Polyphasic taxonomy of rhizobia: emendation of the genus Sinorhizobium and description of Sinorhizobium meliloti comb. nov., Sinorhizobium saheli sp. nov., and Sinorhizobium teranga sp. nov. Int J Syst Bacteriol 44:715–733 de Lajudie P, Willems A, Nick G, Moreira F, Molouba F, Hoste B, Torck U, Neyra M, Collins MD, Lindstro¨m K, Dreyfus B, Gillis M (1998) Characterization of tropical tree rhizobia and description of Mesorhizobium plurifarium sp. nov. Int J Syst Bacteriol 48:369–382 Diagne O, Ingleby K, Deans JD, Lindley DK, Diaite´ I, Neyra M (2001) Mycorrhizal inoculum potential of soils from alley cropping plots in Se´ne´gal. For Ecol Manag 146:35–43 Diem HG, Dommergues YR (1990) Current and potential use and management of Casuarinaceae in the tropics and subtropic. In: Swintzer CR, Tjepkema JD (eds) The biology of Frankia and actinorhizal plants. Academic Press, New York, pp 365–385 Diouf D, Samba-Mbaye R, Lesueur D, Ba AT, Dreyfus B, de Lajudie P, Neyra M (2007) Genetic diversity of Acacia seyal Del. rhizobial populations indigenous to senegalese soils in relation to salinity and pH of the samplings sites. Microbial Ecol 54:553–566 Diouf A, Diop TA, Ndoye I, Gueye M (2008) Response of Gliricidia sepium tree to phosphorus application and inoculations with Glomus aggregatum and rhizobial strains in a sub-Saharian sandy soil. Afr J Biotechnol 7:766–771 Domenach AM (1995) Approche de l’estimation de la fixation symbiotique des arbres par l’utilisation des abondances isotopiques naturelles de l’azote. In: Maillard P, Bonhomme R (eds) Utilisation des isotopes stables pour l’etude du fonctionnement des plantes, Paris 16–17 de´cembre 1993. Les Colloques 70, INRA Editions, Versailles, France, pp 159–172 Domenach AM, Kurdali F, Bardin R (1989) Estimation of symbiotic dinitrogen fixation in alder forest by the method based on natural 15N abundance. Plant Soil 118:51–59 Dreyfus B, Garcia JL, Gillis M (1988) Characterization of Azorhizobium caulinodans gen. nov., sp. nov., a stem-nodulating nitrogen-fixing bacterium isolated from Sesbania rostrata. Int J Syst Bacteriol 38:89–98 Duhoux E, Dommergues YR (1985) The use of nitrogen fixing trees in forest and soil restoration in the tropics. In: Ssali H, Keya SO (eds) Biological nitrogen fixation in Africa: proceedings of the first conference of the African association for biological nitrogen fixation. Matianum Press Consultants, Nairobi, pp 384–400

123

Author's personal copy 154 Dulormne M, Sierra J, Nygren P, Cruz P (2003) Nitrogen-fixation dynamics in a cut-and-carry silvopastoral system in the subhumid conditions of Guadeloupe, French Antilles. Agrofor Syst 59:121–129 Duponnois R, Plenchette C (2003) A mycorrhiza helper bacterium enhances ectomycorrhizal and endomycorrhizal symbiosis of Australian Acacia species. Mycorrhiza 13:85–91 Dupuy N, Willems A, Pot B, Dewettinck D, Vandenbruaene I, Maestrojuan G, Dreyfus B, Kersters K, Collins MD, Gillis M (1994) Phenotypic and genotypic characterization of bradyrhizobia nodulating the leguminous tree Acacia albida. Int J Syst Bacteriol 44:461–473 Escalante G, Herrera R, Aranguren J (1984) Fijacio´n de nitro´geno en a´rboles de sombra (Erythrina poeppigiana) en cacaotales del norte de Venezuela. Pesq. Agropec. Bras. 19(edic¸a˜o especial):223–230 FAOStat (2011) http://www.fao.org/corp/statistics/en/ Accessed on 29 July 2011 Fassbender HW (1987) Nutrient cycling in agroforestry systems of coffee (Coffea arabica) with shade trees in the Central Experiment of CATIE. In: Beer JW, Fassbender HW, Heuveldop J (eds) Advances in agroforestry research. CATIE, Turrialba, pp 155–172 Fassbender HW, Alpı´zar L, Heuveldop J, Fo¨lster H, Enrı´quez G (1988) Modelling agroforestry systems of cacao (Theobroma cacao) with laurel (Cordia alliodora) and poro (Erythrina poeppigiana) in Costa Rica III. Cycles of organic matter and nutrients. Agrofor Syst 6:49–62 Ferna´ndez M, Meugnier H, Grimont P, Bardin R (1989) Deoxyribonucleic acid relatedness among members of the genus Frankia. Int J Syst Bacteriol 39:424–429 Fuentes-Ramı´rez LE, Bustillos-Cristales R, Tapia-Herna´ndez A, Jime´nez Salgado T, Wang ET, Martı´nez-Romero E, Caballero-Mellado J (2001) Novel nitrogen-fixing acetic bacteria, Gluconacetobacter johannae sp. nov. and Gluconacetobacter azotocaptans sp. nov., associated with coffee plants. J Syst Evol Microbiol 51:1305–1314 Fustec J, Lesuffleur F, Mahieu S, Cliquet J-B (2010) Nitrogen rhizodeposition of legumes—a review. Agron Sustain Dev 30:57–66 Galiana A, Chaumont J, Diem HG, Dommergues Y (1990) Nitrogen-fixing potential of Acacia mangium and Acacia auriculiformis seedlings inoculated with Bradyrhizobium and Rhizobium spp. Biol Fertil Soils 9:261–267 Galiana A, Gnahoua GM, Chaumont J, Lesueur D, Prin Y, Mallet B (1998) Improvement of nitrogen fixation in Acacia mangium through inoculation with rhizobium. Agrofor Syst 40:297–307 Gardner IC, Barrueco CR (1999) Mycorrhizal and actinorhizal biotechnology: problems and prospects. In: Varma A, Hock B (eds) Mycorrhiza. Springer, Berlin, pp 469–499 Garrity DP, Mercado AR Jr (1994) Nitrogen fixation capacity in the component species of contour hedgerows: how important? Agrofor Syst 27:241–258 Garrity DP, Akinnefesi FK, Ajayi OC, Weldesemayat SG, Mowo JG, Kalinganire A, Larwanou M, Bayala J (2010) Evergreen agriculture: a robust approach to sustainable food security in Africa. Food Sec 2:197–214 Gauthier D, Diem HG, Dommergues YR (1985) Assessment of N2 fixation by Casuarina equisetifolia inoculated with

123

Nutr Cycl Agroecosyst (2012) 94:123–160 Frankia OR02001 using 15N methods. Soil Biol Biochem 17:375–379 Gehring C, Vlek PLG (2004) Limitations of the 15N natural abundance method for estimating biological nitrogen fixation in Amazonian forest legumes. Basic Appl Ecol 5:567–580 Giardina CP, Huffman S, Binkley D, Caldwell BA (1995) Alders increase soil phosphorus availability in a Douglasfir plantation. Can J For Res 25:1652–1657 Giller KE (2001) Nitrogen fixation in tropical cropping systems, 2nd edn. CABI Publishing, Wallingford Giller KE, Cadisch G (1995) Future benefits from biological nitrogen fixation: an ecological approach to agriculture. Plant Soil 174:255–277 Gokkaya K, Hurd TM, Raynal DJ (2006) Symbiont nitrogenase, alder growth, and soil nitrate response to phosphorus addition in alder (Alnus incana ssp. rugosa) wetlands of the Adirondack Mountains, New York State, USA. Environ Exp Bot 55:97–109 Go´mez Luciano CA (2008) Distribucio´n de raı´ces finas de Inga edulis y Theobroma cacao en el suelo de un sistema agroforestal orga´nico. Proyecto de Graduacio´n, Universidad EARTH, Gua´cimo Graham P (1992) Stress tolerance in Rhizobium and Bradyrhizobium, and nodulation under adverse soil conditions. Can J Microbiol 38:475–484 Graham PH, Hubbell DH (1975) Legume Rhizobium relationships in tropical agriculture. In: Doll EC, Mott GO (eds) Tropical forages in livestock production systems. ASA Spec Publ 24:9–21 Grierson PF, Smithson P, Nziguheba G, Radersma S, Comerford NB (2004) Phosphorus dynsmics and mobilization by plants. In: van Noordwijk M, Cadisch G, Ong CK (eds) Below-ground interactions in tropical agroecosystems, concepts and models with multiple plant components. CABI Publishing, Wallingford, pp 127–142 Grossman JM, Sheaffer C, Wyse D, Bucciarelli B, Vance C, Graham PH (2006) An assessment of nodulation and nitrogen fixation in inoculated Inga oerstediana, a nitrogen-fixing tree shading organically grown coffee in Chiapas, Mexico. Soil Biol Biochem 38:769–784 Gueye M, Ndoye I, Dianda M, Danso SKA, Dreyfus B (1997) Active N2 fixation in several Faidherbia albida provenances. Arid Soil Res Rehabil 11:63–70 Habte M (1995) Dependency of Cassia siamea on vesicular arbuscular mycorrhizal fungi. J Plant Nutr 18:2191–2198 Habte M, Turk D (1991) Response of two species of Cassia and Gliricidia sepium to vesicular-arbuscular mycorrhizal infection. Commun Soil Sci Plant Anal 22:1861–1872 Haggar JP, Tanner EVJ, Beer JW, Kass DCL (1993) Nitrogen dynamics in tropical agroforestry and annual cropping systems. Soil Biol Biochem 25:1363–1378 Han TX, Wang ET, Wu LJ, Chen WF, Gu JG, Gu CT, Tian CF, Chen WX (2008) Rhizobium multihospitium sp. nov., isolated from multiple legume species native of Xinjiang, China. Int J Syst Evol Microbiol 58:1693–1699 Handley LL, Raven JA (1992) The use of natural abundance of nitrogen isotopes in plant physiology and ecology. Plant Cell Environ 15:965–985 Harmand J-M (1998) Roˆle des espe`ces ligneuses a` croissance rapide dans le fonctionnement bioge´ochimique de la jache`re.

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 Effets sur la restauration de la fertilite´ des sols ferrugineux tropicaux. (Bassin de la Be´noue´ au Nord-Cameroun). The`se de doctorat, Universite´ de Paris VI, France, 213 p Harmand J-M, Njiti CF, Bernhard-Reversat F, Puig H (2004) Aboveground and belowground biomass, productivity and nutrient accumulation in tree improved fallows in the dry tropics of Cameroon. For Ecol Manag 188:249–265 Harmand J-M, Avila H, Dambrine E, Skiba U, de Miguel S, Renderos RV, Oliver R, Jime´nez F, Beer J (2007a) Nitrogen dynamics and soil nitrate retention in a Coffea arabicaEucalyptus deglupta agroforestry system in Southern Costa Rica. Biogeochemistry 85:125–139 ´ vila H, Dionisio L, Harmand J-M, Chaves V, Cannavo P, A Zeller B, Hergoualc’h K, Vaast P, Oliver R, Beer J, Dambrine E (2007b) Nitrogen dynamics (coffee productivity, nitrate leaching and N2O emissions) in Coffea arabica systems in Costa Rica according to edaphic conditions, fertilization and shade management. In: 2nd International Symposium on Multi-Strata Agroforestry Systems with Perennial Crops, CATIE, Turrialba, Costa Rica, 17–21 September 2007 Haselwandter K, Bowen GD (1996) Mycorrhizal relations in trees for agroforestry and land rehabilitation. For Ecol Manag 81:1–17 He X-H, Critchley C, Bledsoe C (2003) Nitrogen transfer within and between plants via common mycorrhizal networks (CMNs). Crit Rev Plant Sci 22:531–567 He X-H, Critchley C, Hock N, Bledsoe C (2004) Reciprocal N (15NH4? or 15NO3-) transfer between non-N2-fixing Eucalyptus maculata and N2-fixing Casuarina cunninghamiana linked by the ectomycorrhizal fungus Pisolithus sp. New Phytol 163:629–640 Hergoualc’h K, Skiba U, Harmand J-M, Oliver R (2007) Processes responsible for the nitrous oxide emission from a Costa Rican Andosol under a coffee agroforestry plantation. Biol Fert Soils 43:787–795 Hergoualc’h K, Skiba U, Harmand J-M, He´nault C (2008) Fluxes of greenhouse gases from andosols under coffee in monoculture or shaded by Inga densiflora in Costa Rica. Biogeochemistry 89:329–345 Herna´ndez M, Benavides JE (1994) Podas estrate´gicas en cercos vivos de Pin˜on cubano (Gliricidia sepium) para produccio´n de forraje en la e´poca seca. In: Benavides JE (ed) Arboles y arbustos forrajeros en Ame´rica Central, vol II., CATIETurrialba, Costa Rica, pp 559–582 Herna´ndez-Lucas I, Segovia L, Martı´nez-Romero E, Pueppke SG (1995) Phylogenetic relationships and host range of Rhizobium spp. that nodulate Phaseolus vulgaris L. Appl Environ Microbiol 61:2775–2779 Herrera AM, Bedmar EJ, Olivares J (1985) Host specificity of Rhizobium strains isolated from nitrogen-fixing trees and nitrogenase activities of strain GRH2 in symbiosis with Prosopis chilensis. Plant Sci 42:177–182 Herridge DF, Peoples MB, Boddey RM (2008) Global inputs of biological nitrogen fixation in agricultural systems. Plant Soil 311:1–18 Hobbie EA, Ho¨gberg P (2012) Tansley review. Nitrogen isotopes link mycorrhizal fungi and plants to nitrogen dynamics. New Phytol 196:367–382 Ho¨gberg P (1997) Tansley Review No. 95. 15N natural abundance in soil-plant systems. New Phytol 137:179–203

155 Houlton BZ, Wang YP, Vitousek PM, Field CB (2008) A unifying framework for dinitrogen fixation in the terrestrial biosphere. Nature 454:327–330 Huguet V, Gouy M, Normand P, Zimpfer JF, Fernandez MP (2005) Molecular phylogeny of Myricaceae: a reexamination of host-symbiont specificity. Mol Phyl Evol 34:557–568 Hunt S, Layzell DB (1993) Gas exchange of legume nodules and the regulation of nitrogenase activity. Annu Rev Plant Physiol Plant Mol Biol 44:483–511 Iglesias L, Salas E, Leblanc HA, Nygren P (2011) Response of Theobroma cacao and Inga edulis seedlings to crossinoculated populations of arbuscular mycorrhizal fungi. Agrofor Syst 83:63–73 Ingleby K, Fahmer A, Wilson J, Newton AC, Mason PA, Smith RI (2001) Interactions between mycorrhizal colonisation, nodulation and growth of Calliandra calothyrsus seedlings supplied with different concentrations of phosphorus solution. Symbiosis 30:15–28 Ingleby K, Wilson J, Munro RC, Cavers S (2007) Mycorrhizas in agroforestry: spread and sharing of arbuscular mycorrhizal fungi between trees and crops: complementary use of molecular and microscopic approaches. Plant Soil 294:125–136 Isaac ME, Harmand J-M, Drevon J–J (2011a) Growth and nitrogen acquisition strategies of Acacia senegal seedlings under exponential phosphorus additions. J Plant Physiol 168:776–781 Isaac ME, Harmand J-M, Lesueur D, Lelon J (2011b) Tree age and soil phosphorus conditions influence N2-fixation rates and soil N dynamics in natural populations of Acacia Senegal. For Ecol Manag 261:582–588 Isaac ME, Hinsinger P, Harmand J-M (2012) Nitrogen and phosphorus economy of a legume tree-cereal intercropping system under controlled conditions. Sci Total Environ. doi: 10.1016/j.scitotenv.2011.12.071 Jalonen R, Nygren P, Sierra J (2009a) Root exudates of a legume tree as a nitrogen source for a tropical fodder grass. Nutr Cycling Agroecosys 85:203–213 Jalonen R, Nygren P, Sierra J (2009b) Transfer of nitrogen from a tropical legume tree to an associated fodder grass via root exudation and common mycelial networks. Plant Cell Environ 32:1366–1376 Jalonen R, Timonen S, Sierra J, Nygren P (2012) Arbuscular mycorrhizal symbioses in a cut-and-carry forage production system of legume tree Gliricidia sepium and fodder grass Dichanthium aristatum. Agrofor Syst. doi:10.1007/ s10457-012-9553-1 Jamann S, Ferna´ndez MP, Moiroud A (1992) Genetic diversity of Elaeagnaceae-infective Frankia strains isolated from various soils. Acta Oecologica 13:395–405 Jime´nez-Salgado T, Fuentes-Ramı´rez LE, Tapia-Herna´ndez A, Mascarua-Esparza MA, Martı´nez-Romero E, CaballeroMellado J (1997) Coffea arabica L., a new host for Acetobacter diazotrophicus, and isolation of other nitrogenfixing Acetobacteria. Appl Environ Microbiol 63:3676– 3683 Johansen A, Jensen ES (1996) Transfer of N and P from intact or decomposing roots of pea to barley interconnected by an arbuscular mycorrhizal fungus. Soil Biol Biochem 28:73–81

123

Author's personal copy 156 Jones DL, Healey JR, Willett VB, Farrar JF, Hodge A (2005) Dissolved organic nitrogen uptake by plants—an important N uptake pathway? Soil Biol Biochem 37:413–423 Kadiata BD, Mulongoy K, Isirimah NO (1997) Influence of pruning frequency of Albizia lebbeck, Gliricidia sepium and Leucaena leucocephala on nodulation and potential nitrogen fixation. Biol Fert Soils 24:255–260 Ka¨hko¨la¨ A-K, Nygren P, Leblanc HA, Pennanen T, Pietika¨inen J (2012) Leaf and root litter of a legume tree as nitrogen sources for cacaos with different root colonisation by arbuscular mycorrhizae. Nutr Cycling Agroecosys 92:51–65 Kang BT, Wilson GF, Sipkens L (1981) Alley cropping maize (Zea mays L.) and leucaena (Leucaena leucocephala Lam.) in Southern Nigeria. Plant Soil 63:165–179 Kanmegne J, Smaling EMA, Brussaard L, Gansop-Kouomegne A, Boukong A (2006) Nutrient flows in smallholder production systems in the humid forest zone of southern Cameroon. Nutr Cycling Agroecosys 76:233–248 Kaschuk G, Kuyper TW, Leffelaar PA, Hungria M, Giller KE (2009) Are the rates of photosynthesis stimulated by the carbon sink strength of rhizobial and arbuscular mycorrhizal symbioses? Soil Biol Biochem 41:1233–1244 Kass DL (1995) Are nitrogen fixing trees a solution for acid soils? In: Evand DO, Szott LT (eds) Nitrogen fixing trees for acid soils. Nitrogen Fixing Tree Research Reports, Special Issue 1995. Winrock International Institute for Agricultural Development, Morrilton, pp 19–31 Kass DCL, Sylvester-Bradley R, Nygren P (1997) The role of nitrogen fixation and nutrient supply in some agroforestry systems of the Americas. Soil Biol Biochem 29:775–785 Khanna PK (1998) Nutrient cycling under mixed-species tree systems in southeast Asia. Agrofor Syst 38:99–120 Koponen P, Nygren P, Domenach AM, Le Roux C, Saur E, Roggy JC (2003) Nodulation and dinitrogen fixation of legume trees in a tropical freshwater swamp forest in French Guiana. J Trop Ecol 19:655–666 Kurppa M, Leblanc HA, Nygren P (2010) Detection of nitrogen transfer from N2-fixing shade trees to cacao saplings in 15N labelled soil: ecological and experimental considerations. Agrofor Syst 80:223–239 Kuyper TW, Cardoso IM, Onguene NA, Murniati, van Noordwijk M (2004) Managing mycorrhiza in tropical multispecies agroecosystems. In: van Noordwijk M, Cadisch G, Ong CK (eds) Below-ground interactions in tropical agroecosystems, concepts and models with multiple plant components. CABI Publishing, Wallingford, pp 243–261 Ladha JK, Peoples MB, Garrity DP, Capuno VT, Dart PJ (1993) Estimating dinitrogen fixation of hedgerow vegetation using the nitrogen-15 natural abundance method. Soil Sci Soc Am J 57:732–737 Lambers H, Raven JA, Shaver GR, Smith SE (2008) Plant nutrient acquisition strategies change with soil age. Trends Ecol Evol 23:95–103 Leblanc HA (2004) Evaluation of Inga spp. for dinitrogen fixation and nitrogen release in humid-tropical alley cropping. Ph.D. Thesis, Department of Agronomy, University of Missouri–Columbia, USA Leblanc HA, McGraw RL, Nygren P, Le Roux C (2005) Neotropical legume tree Inga edulis forms N2-fixing symbiosis with fast-growing Bradyrhizobium strains. Plant Soil 275:123–133

123

Nutr Cycl Agroecosyst (2012) 94:123–160 Leblanc HA, McGraw RL, Nygren P (2007) Dinitrogen-fixation by three Neotropical agroforestry tree species under semicontrolled field conditions. Plant Soil 291:99–209 Lehmann J, Muraoka T, Zech W (2001) Root activity patterns in an Amazonian agroforest with fruit trees determined by 32 P, 33P and 15N applications. Agrofor Syst 52:185–197 Lesueur D, Sarr A (2008) Effects of single and dual inoculation with selected microsymbionts (rhizobia and arbuscular mycorrhizal fungi) on field growth and nitrogen fixation of Calliandra calothyrsus Meissn. Agrofor Syst 73:37–45 Lie TA, Go¨ktan D, Engin M, Pijenborg J, Anlarsal E (1987) Coevolution of the legume-Rhizobium association. Plant Soil 100:171–181 Lindblad P, Russo R (1986) C2H2-reduction by Erythrina poeppigiana in a Costa Rican coffee plantation. Agrofor Syst 4:33–37 Liu J, Wang ET, Chen WX (2005) Diverse rhizobia associated with woody legumes Wisteria sinensis, Cercis racemosa and Amorpha fruticosa grown in the temperate zone of China. Syst Appl Microbiol 28:465–477 Lloret L, Ormeno-Orillo E, Rinco´n R, Martı´nez-Romero J, RogelHerna´ndez MA, Martı´nez-Romero E (2007) Ensifer mexicanus sp nov.: a new species nodulating Acacia angustisima (Mill.) Kuntze in Mexico. Syst Appl Microbiol 30:280–290 Lortet G, Mear N, Lorquin J, Dreyfus B, de Lajudie P, Rosenberg C, Boivin C (1996) Nod factor thin-layer chromatography profiling as a tool to characterize symbiotic specificity of rhizobial strains: application to Sinorhizobium saheli, S. teranga, and Rhizobium sp. strains isolated from Acacia and Sesbania. Mol Plant-Microbe Interact 9:736–747 Louche J, Ali MA, Sauvage FX, Cloutier-Hurteau B, Quiquampoix H, Plassard C (2010) Efficiency of acid phosphatases secreted from the ectomycorrhizal fungus Hebeloma cylindrosporum to hydrolyse organic phosphorus in podzols. FEMS Microbiol Ecol 73:323–335 Lumini E, Bosco M, Fernandez MP (1996) PCR-RFLP and total DNA homology revealed three related genomic species among broad-host-range Frankia strains. FEMS Microbiol Ecol 21:303–311 Mafongoya PL, Giller KE, Palm CA (1998) Decomposition and nitrogen release patterns of tree prunings and litter. Agrofor Syst 38:77–97 Mafongoya PL, Giller KE, Odee S, Gathumbi S, Ndufa SK, Sitompul SM (2004) Benefiting from N2-fixation and managing rhizobia. In: van Noordwijk M, Cadisch G, Ong CK (eds) Below-ground interactions in tropical agroecosystems, concepts and models with multiple plant components. CABI Publishing, Wallingford, pp 227–242 Makatiani ET, Odee DW (2007) Response of Sesbania sesban (L.) Merr. to rhizobial inoculation in an N-deficient soil containing low numbers of effective indigenous rhizobia. Agrofor Syst 70:211–216 Manjunath A, Habte M (1992) External and internal P requirement of plant species differing in their mycorrhizal dependency. Arid Soil Res Rehabil 6:271–284 Mariotti A (1983) Atmospheric nitrogen is a reliable standard for natural 15N abundance measurements. Nature 303:685–687 Martinelli LA, Piccolo MC, Townsend AR, Vitousek PM, Cuevas E, McDowell M, Robertson GP, Santos OC, Treseder K (1999) Nitrogen stable isotopic composition of

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 leaves and soil: tropical versus temperate forests. Biogeochemistry 46:45–65 Martı´nez L, Caballero-Mellado J, Orozco J, Martı´nez-Romero E (2003) Diazotrophic bacteria associated with banana (Musa spp.). Plant Soil 255:35–47 Martı´nez-Romero E, Segovia L, Mercante FM, Franco AA, Graham P, Pardo MA (1991) Rhizobium tropici, a novel species nodulating Phaseolus vulgaris L. beans and Leucaena sp. trees. Int J Syst Bacteriol 41:417–426 Maunuksela L, Hahn D, Haahtela K (2000) Effect of freezing of soils on nodulation capacities of total and specific Frankia populations. Symbiosis 29:107–120 Maunuksela L, Zepp K, Koivula T, Zeyer J, Haahtela K, Hahn D (2006) Analysis of Frankia populations in three soils devoid of actinorhizal plants. FEMS Microbiol Ecol 28:11–21 McKane RB, Johnson LC, Shaver GR, Nadelhoffer KJ, Rastetter EB, Fry B, Giblin AE, Kielland K, Kwiatkowski BL, Laundre JA, Murray G (2002) Resource-based niches provide a basis for plant species diversity and dominance in arctic tundra. Nature 415:68–71 McKey D (1994) Legumes and nitrogen: the evolutionary ecology of a nitrogen-demanding lifestyle. In: Sprent JI, McKey D (eds) Advances in legume systematics 5: the nitrogen factor. Royal Botanic Gardens, Kew, pp 211–228 Mercado AR Jr, van Noordwijk M, Cadisch G (2011) Positive nitrogen balance of Acacia mangium woodlots as fallows in the Philippines based 15N natural abundance data on N2 fixation. Agrofor Syst 81:221–233 Minchin FR (1997) Regulation of oxygen diffusion in legume nodules. Soil Biol Biochem 29:881–888 Minchin FR, Witty JF, Sheehy JE, Mu¨ller M (1983) A major error in the acetylene reduction assay: dreceases in nodular nitrogenase activity under assay conditions. J Exp Bot 34:641–649 Minchin FR, Sheehy JE, Witty JF (1986) Further errors in the acetylene reduction assay. Effects of plan disturbance. J Exp Bot 37:1581–1591 Moreira FMS, Haukka K, Young JPW (1998) Biodiversity of rhizobia isolated from a wide range of forest legumes in Brazil. Mol Ecol 7:889–895 Moreira FMS, Carvalho Y, Gonc¸alves M, Haukka K, Young JPW, de Faria SM, Franco AA, Cruz LM, Pedrosa FO (2000) Azorhizobium johannense sp. nov. and Sesbania virgata (Caz.) Pers.: a highly specific symbiosis. In: Pedrosa FO, Hungria M, Yates MG, Newton WE (eds) Nitrogen fixation: from molecules to crop productivity. Current Plant Sci Biotechnol Agric 38:197 Moyer-Henry KA, Burton JW, Israel DW, Rufty TW (2006) Nitrogen transfer between plants: a 15N natural abundance study with crop and weed species. Plant Soil 282:7–20 Nair PKR, Buresh RJ, Mugendi DN, Latt CR (1999) Nutrient cycling in tropical agroforestry systems: myths and science. In: Buck LE, Lassoie JP, Fernandes ECM (eds) Agroforestry in sustainable agricultural system. CRC Press, Boca Raton, pp 1–31 Na¨sholm T, Kielland K, Ganateg U (2009) Uptake of organic nitrogen by plants. New Phytol 182:31–48 Nazaret S, Simonet P, Normand P, Bardin R (1989) Genetic diversity among Frankia isolated from Casuarina nodules. Plant Soil 118:241–247

157 Ndoye I, Gueye M, Danso SKA, Dreyfus B (1995) Nitrogen fixation in Faidherbia albida, Acacia raddiana, Acacia senegal and Acacia seyal estimated using the 15N isotope dilution technique. Plant Soil 172:175–180 Ndufa JK, Gathumbi SM, Kamiri HW, Giller KE, Cadisch G (2009) Do mixed-species legume fallows provide longterm maize yield benefit compared with monoculture legume fallows? Agron J 101:1352–1362 Ngom A, Nakagawa Y, Sawada H, Tsukahara J, Wakabayashi S, Uchiumi T, Nuntagij A, Kotepong S, Suzuki A, Higashi S, Abe M (2004) A novel symbiotic nitrogen-fixing member of the Ochrobactrum clade isolated from root nodules of Acacia mangium. J Gen Appl Microbiol 50:17–27 Nick G, de Lajudie P, Eardly BD, Suomalainen S, Paulin L, Zhang X, Gillis M, Lindstro¨m K (1999) Sinorhizobium arboris sp. nov. and Sinorhizobium kostiense sp. nov., isolated from leguminous trees in Sudan and Kenya. Int J Syst Bacteriol 49:1359–1368 Nye PH, Greenland DJ (1960) The soil under shifting cultivation. Commonwealth Bureau of Soils, Harpenden Nygren P (1995) Above-ground nitrogen dynamics following the complete pruning of a nodulated woody legume in humid tropical field conditions. Plant Cell Environ 18:977–988 Nygren P, Cruz P (1998) Biomass allocation and nodulation of Gliricidia sepium under two cut-and-carry forage production regimes. Agrofor Syst 41:277–292 Nygren P, Leblanc HA (2009) Natural abundance of 15N in two cacao plantations with legume and non-legume shade trees. Agrofor Syst 76:303–315 Nygren P, Ramı´rez C (1995) Production and turnover of N2 fixing nodules in relation to foliage development in periodically pruned Erythrina poeppigiana (Leguminosae) trees. For Ecol Manag 73:59–73 Nygren P, Cruz P, Domenach AM, Vaillant V, Sierra J (2000) Influence of forage harvesting regimes on dynamics of biological dinitrogen fixation of a tropical woody legume. Tree Physiol 20:41–48 Oaks A (1992) A re-evaluation of nitrogen assimilation in roots. Bioscience 42:103–110 Odee DW, Sutherland JM, Kimiti JM, Sprent JI (1995) Natural rhizobial populations and nodulation status of woody legumes growing in diverse Kenyan conditions. Plant Soil 173:211–224 Odee DW, Sutherland JM, Makatiani ET, McInroy SG, Sprent JI (1997) Phenotypic characteristics and composition of rhizobia associated with woody legumes growing in diverse Kenyan conditions. Plant Soil 188:65–75 Odee DW, Haukka K, McInroy SG, Sprent JI, Sutherland JM, Young JPW (2002) Genetic and symbiotic characterization of rhizobia isolated from tree and herbaceous legumes grown in soils from ecologically diverse sites in Kenya. Soil Biol Biochem 34:801–811 Oliver R, Njiti CF, Harmand J-M (2000) Analyse de la durabilite´ de la fertilite´ acquise suite a` des jache`res arbore´es au NordCameroun. Etude et Gestion des Sols 7:287–309 Ong CK, Kho RM, Radersma S (2004) Ecological interactions in multispecies agroecosystems: concepts and rules. In: van Noordwijk M, Cadisch G, Ong CK (eds) Below-ground interactions in tropical agroecosystems, concepts and models with multiple plant components. CABI Publishing, Wallingford, pp 1–15

123

Author's personal copy 158 Oyaizu H, Matsumoto S, Minamisawa K, Gamou T (1993) Distribution of rhizobia in leguminous plants surveyed by phylogenetic identification. J Gen Appl Microbiol 39:339–354 Palm CA (1995) Contribution of agroforestry trees to nutrient requirements in intercropped plants. Agrofor Syst 30:105–124 Parrotta JA, Baker DD, Fried M (1996) Changes in dinitrogen fixation in maturing stands of Casuarina equisetifolia and Leucaena leucocephala. Can J For Res 16:1684–1691 Parsons R, Stanforth A, Raven JA, Sprent JI (1993) Nodule growth and activity may be regulated by a feedback mechanism involving phloem nitrogen. Plant Cell Environ 16:125–136 Paschke MW, Dawson JO (1992) The occurrence of Frankia in tropical forest soils of Costa Rica. Plant Soil 142:63–67 Paul EA, Clark FE (1996) Soil microbiology and biochemistry. Academic Press, San Diego Peoples MB, Palmer B, Lilley DM, Duc LM, Herridge DF (1996) Application of 15N and xylem ureide methods for assessing N2 fixation of three shrub legumes periodically pruned for forage. Plant Soil 182:125–137 Pons TL, Perreijn K, van Kessel C, Werger MJA (2007) Symbiotic nitrogen fixation in a tropical rainforest: 15N natural abundance measurements supported by experimental isotopic enrichment. New Phytol 173:154–167 Prin Y, Galiana A, Le Roux C, Me´le´ard B, Razafimaharo V, Ducousso M, Chaix G (2003) Molecular tracing of Bradyrhizobium strains helps to correctly interpret Acacia mangium response to inoculation in a reforestation experiment in Madagascar. Biol Fertil Soils 37:64–69 Raddad EAY, Salih AA, El Fadl MA, Kaarakka V, Luukkanen O (2005) Symbiotic nitrogen fixation in eight Acacia senegal provenances in dryland clays of the Blue Nile Sudan estimated by the 15N natural abundance method. Plant Soil 275:261–269 Roggy JC, Pre´vost MF, Garbaye J, Domenach AM (1999a) Nitrogen cycling in the tropical rain forest of French Guiana: comparison of two sites with contrasting soil types using d15N. J Trop Ecol 15:1–22 Roggy JC, Pre´vost MF, Gourbiere F, Casabianca H, Garbaye J, Domenach AM (1999b) Leaf natural 15N abundance and total N concentration as potential indicators of plant nutrition in legumes and pioneer species in a rain forest in French Guiana. Oecologia 120:171–182 Roggy JC, Moiroud A, Lensi R, Domenach AM (2004) Estimating N transfers between N2-fixing actinorhizal species and the non-N2-fixing Prunus avium under partially controlled conditions. Biol Fertil Soils 39:312–319 Roskoski JP (1982) Nitrogen fixation in a Mexican coffee plantation. Plant Soil 67:283–291 Roskoski JP, Van Kessel C (1985) Annual, seasonal and field variation in nitrogen fixing activity by Inga jinicuil, a tropical legume tree. Oikos 44:306–312 Roupsard O (1997) Ecophysiologie et diversite´ ge´ne´tique de Faidherbia albida (Del.) A. Chev. (syn. Acacia albida Del.), un arbre a` usages multiples d’Afrique semi-aride : fonctionnement hydrique et efficience d’utilisation de l’eau d’arbres adultes en parc agroforestier et de juve´niles en conditions semi-controˆle´es. Tome 1: Partie synthe`se. The`se de doctorat, Universite´ Nancy 1, France, 70 p

123

Nutr Cycl Agroecosyst (2012) 94:123–160 Roupsard O, Ferhi A, Granier A, Pallo F, Depommier D, Mallet B, Joly HI, Dreyer E (1999) Reverse phenology and dryseason water uptake by Faidherbia albida (Del.) A. Chev. in an agroforestry parkland of Sudanese west Africa. Funct Ecol 13:460–472 Rouvier C, Nazaret S, Fernandez MP, Picard B, Simonet P, Normand P (1992) rrn and nif intergenic spacers and isoenzyme patterns as tools to characterize Casuarina-infective Frankia strains. Acta Oecologica 13:367–516 Rowe EC, Cadisch G (2002) Implications of heterogeneity on procedures for estimating plant 15N recovery in hedgerow intercrop systems. Agrofor Syst 54:61–70 Rowe EC, Hairiah K, Giller K, van Noordwijk M, Cadisch G (1999) Testing the safety-net role of hedgerow trees by 15N placement at different soil depths. Agrofor Syst 43:81–93 Rowe EC, van Noordwijk M, Suprayogo D, Hairiah K, Giller KE, Cadisch G (2001) Root distributions partially explain 15 N uptake patterns in Gliricidia and Peltophorum hedgerow intercropping systems. Plan Soil 235:167–179 Salas E, Nygren P, Domenach AM, Berninger F, Ramı´rez C (2001) Estimating biological N2 fixation by a tropical legume tree using the non-nodulating phenophase as the reference in the 15N natural abundance method. Soil Biol Biochem 33:1859–1868 Sanginga N, Danso S, Bowen G (1989) Nodulation and growth response of Allocasuarina and Casuarina species to phosphorus fertilization. Plant Soil 118:125–132 Sanginga N, Danso SKA, Mulongoy K, Ojeifo AA (1994) Persistence and recovery of introduced Rhizobium ten years after inoculation on Leucaena leucocephala grown on an Alfisol in Southwestern Nigeria. Plant Soil 159:199–204 Sanginga N, Vanlauwe B, Danso SKA (1995) Management of biological N2 fixation in alley cropping systems: estimation and contribution to N balance. Plant Soil 174:119–141 Santana MBM, Rosand PC (1985) Reciclagem de nutrientes em uma plantac¸a˜o de cacau sombreada com eritrina. In: Proceedings of the IX international cocoa research conference, Togo 1984. Cocoa Producers’ Alliance, Lagos, Nigeria, pp 205–210 Schimann H, Ponton S, Ha¨ttenschwiler S, Ferry B, Lensi R, Domenach AM, Roggy J-C (2008) Differing nitrogen use strategies of two tropical rainforest late successional tree species in French Guiana: evidence from 15N natural abundance and microbial activities. Soil Biol Biochem 40:487–494 Schroth G, Kolbe D, Pity B, Zech W (1995) Searching for criteria for the selection of efficient tree species for fallow improvement, with special reference to carbon and nitrogen. Fertilizer Res 42:297–314 Schulze E-D, Gebauer G, Ziegler H, Lange OL (1991) Estimates of nitrogen fixation by trees on an aridity gradient in Namibia. Oecologia 88:451–455 Segovia L, Young JP, Martı´nez-Romero E (1993) Reclassification of American Rhizobium leguminosarum Biovar phaseoli Type I strains as Rhizobium etli sp. nov. Int J Syst Bacteriol 43:374–377 Sellstedt A, Sta˚hl L, Mattsson U, Jonsson K, Ho¨gberg P (1993) Can the 15N dilution technique be used to study N2 fixation in tropical tree symbioses as affected by water deficit? J Exp Bot 44:1749–1755

Author's personal copy Nutr Cycl Agroecosyst (2012) 94:123–160 Shearer G, Kohl DH (1986) N2 fixation in field settings: estimations based on natural 15N abundance. Aust J Plant Physiol 13:699–756 Shi Y, Ruan J (1992) DNA base composition and homology values in the classification of some Frankia strains. Acta Microbiol Sin 32:133–136 Sierra J, Nygren P (2006) Transfer of N fixed by a legume tree to the associated grass in a tropical silvopastoral system. Soil Biol Biochem 38:1893–1903 Simard SW, Jones MD, Durall DM (2002) Carbon and nutrient fluxes within and between mycorrhizal plants. In: van der Heijden MGA, Sanders I (eds) Mycorrhizal ecology. Ecological studies, vol 157. Springer, Heidelberg, pp 33–74 Smith SE, Read DJ (2008) Mycorrhizal symbiosis, 3rd edn. Academic Press, New York Smolander A, Sundman V (1987) Frankia in acid soils of forests devoid of actinorhizal plants. Physiol Plant 70:297–303 Snoeck D, Zapata F, Domenach A-M (2000) Isotopic evidence of the transfer of nitrogen fixed by legumes to coffee trees. Biotechnol Agron Soc Environ 4:95–100 Soto-Pinto L, Anzueto-Martı´nez M, Mendoza VJ, Jime´nezFerrer G, de Jong B (2010) Carbon sequestration through agroforestry in indigenous communities of Chiapas, Mexico. Agrofor Syst 78:39–51 Sprent JI (2009) Legume nodulation, a global perspective. Wiley-Blackwell, Oxford Sprent JI, James EK (2007) Legume evolution: where do nodules and mycorrhizas fit in? Plant Physiol 144:575–581 Sta˚hl L, Nyberg G, Ho¨gberg P, Buresh RJ (2002) Effects of planted tree fallows on soil nitrogen dynamics aboveground and root biomass, N2-fixation and subsequent maize crop productivity in Kenya. Plant Soil 243:103–117 Sta˚hl L, Ho¨gberg P, Sellstedt A, Buresh RJ (2005) Measuring nitrogen fixation by Sesbania sesban planted in fallows using 15N tracer technique in Kenya. Agrofor Syst 65:67–79 Taiz L, Zeiger E (2006) Plant physiology, 4th edn. Sinauer Associates Inc., Sunderland Thornley JHM, Johnson IR (1990) Plant and crop modelling: a mathematical approach to plant and crop physiology. Clarendon Press, Oxford Tilki F, Fisher RF (1998) Tropical leguminous species for acid soils: studies on plant form and growth in Costa Rica. For Ecol Manag 108:175–192 Tilman D, Knops J, Wedin D, Reich P, Ritchie M, Sieman E (1997) The influence of functional diversity and composition on ecosystem processes. Science 277:1300–1302 Toledo I, Lloret L, Martı´nez-Romero E (2003) Sinorhizobium americanus sp. nov., a new Sinorhizobium species nodulating native Acacia spp. in Mexico. Syst Appl Microbiol 26:54–64 Treseder KK, Vitousek PM (2001) Effects of soil nutrient availibility on investment in acquisition of N and P in Hawaiian rain forests. Ecology 82:946–954 Trinick MJ (1982) Host-Rhizobium associations. In: Vincent JM (ed) Nitrogen fixation in legumes. Academic Press, Sydney, pp 111–122 Turk D, Keyser HH (1992) Rhizobia that nodulate tree legumes: specificity of the host for nodulation and effectiveness. Can J Microbiol 38:451–460

159 Uddin MB, Khan MASA, Mukul SA, Hossain MK (2008) Effects of inorganic fertilizers on biological nitrogen fixation and seedling growth of some agroforestry trees in Bangladesh. J For Res 19:303–306 Uliassi DD, Ruess RW (2002) Limitations to symbiotic nitrogen fixation in primary succession on the Tanana River floodplain. Ecology 83:88–103 Unkovich MJ, Pate JS, Lefroy EC, Arthur DJ (2000) Nitrogen isotope fractionation in the fodder tree legume tagasaste (Chamaecytisus proliferus) and assessment of N2 fixation inputs in deep sandy soils of Western Australia. Aust J Plant Physiol 27:921–929 Unkovich M, Herridge D, Peoples M, Cadisch G, Boddey R, Giller K, Alves B, Chalk P (2008) Measuring plant-associated nitrogen fixation in agricultural systems. ACIAR, Canberra, Australia, 258 p. http://aciar.gov.au/publication/ MN136 (accessed 22 June 2012) Uselman SM, Qualls RG, Thomas RB (1999) A test of a potential short cut in the nitrogen cycle: the role of exudation of symbiotically fixed nitrogen from the roots of a N-fixing tree and the effects of increased atmospheric CO2 and temperature. Plant Soil 210:21–32 van Kessel C, Roskoski JP (1981) Nodulation and N2 fixation by Inga jinicuil, a woody legume in coffee plantations. II. Effect of soil nutrients on nodulation and N2 fixation. Plant Soil 59:207–215 van Kessel C, Farrell RE, Roskoski JP, Keane KM (1994) Recycling of the naturally-occurring 15N in an established stand of Leucaena leucocephala. Soil Biol Biochem 26:757–762 van Noordwijk M, Lawson G, Soumare´ A, Groot JJR, Hairiah K (1996) Root distribution of trees and crops: competition and/or complementarity. In: Ong CK, Huxley P (eds) Tree– crop interactions: a physiological approach. CAB International, Wallingford, pp 319–364 Vance CP, Heichel GH (1991) Carbon in N2 fixation: limitation and exquisite adaptation. Annu Rev Plant Physiol Mol Biol 42:373–392 Vanlauwe B, Swift MJ, Merckx R (1996) Soil litter dynamics and N use in a leucaena (Leucaena leucocephala Lam. (De Witt)) alley cropping system in Southwestern Nigeria. Soil Biol Biochem 28:739–749 Vela´squez E, Igual JM, Willems A, Ferna´ndez MP, Munoz E, Mateos PF, Abril A, Toro N, Normand P, Cervantes E, Gillis M, Martı´nez-Molina E (2001) Mesorhizobium chacoense sp. nov., a novel species that nodulates Prosopis alba in the Chaco Arido region (Argentina). Int J Syst Evol Microbiol 51:1011–1021 Vessey JK, Pawlowski K, Bergman B (2004) Root-based N2fixing symbioses: legumes, actinorhizal plants, Parasponia sp., and cycads. Plant Soil 266:205–230 Vitousek PM, Cassman K, Cleveland C, Crews T, Field CB, Grimm NB, Howarth RW, Marino R, Martinelli L, Rastetter EB, Sprent JI (2002) Towards an ecological understanding of biological nitrogen fixation. Biogeochemistry 57(58):1–45 Wang ET, van Berkum P, Beyene D, Sui XH, Dorado O, Chen WX, Martı´nez-Romero E (1998) Rhizobium huautlense sp. nov., a symbiont of Sesbania herbacea that has a close phylogenetic relationship with Rhizobium galegae. Int J Syst Bacteriol 48:687–699

123

Author's personal copy 160 Wang ET, van Berkum P, Sui XH, Beyene D, Chen WX, Martı´nez-Romero E (1999) Diversity of rhizobia associated with Amorpha fruticosa isolated from Chinese soil and description of Mesorhizobium amorphae sp. nov. Int J Syst Bacteriol 49:51–65 Wang Y-P, Houlton BZ, Field CB (2001) A model of biogeochemical cycles of carbon, nitrogen, and phosphorus including symbiotic nitrogen fixation and phosphatase production. Global Biogeochem Cycles 21:GB1018. doi: 10.1029/2006GB002797 Wang ET, Rogel MA, Sui XH, Chen WX, Martı´nez-Romero E, van Berkum P (2002a) Mesorhizobium amorphae, a rhizobial species that nodulates Amorpha fruticosa, is native to American soils. Arch Microbiol 178:301–305 Wang ET, Tan ZY, Willems A, Ferna´ndez-Lo´pez M, ReinholdHurek B, Martı´nez-Romero E (2002b) Sinorhizobium morelense sp. nov., a Leucaena leucocephala-associated bacterium that is highly resistant to multiple antibiotics. Int J Syst Evol Microbiol 52:1687–1693 Wang FQ, Wang ET, Lui J, Chen Q, Sui XH, Chen WF, Chen WX (2007) Mesorhizobium albiziae sp. nov., a novel bacterium that nodulates Albizia kalkora in a subtropical region of China. Int J Syst Evol Microbiol 57:1192–1199 Weber J, Ducousso M, Yee Tham F, Nourissier-Mountou S, Galiana A, Prin Y, Lee SK (2005) Co-inoculation of Acacia mangium with Glomus intraradices and Bradyrhizobium sp. in aeroponic culture. Biol Fert Soils 41:233–239 Wichern F, Eberhardt E, Mayer J, Joergensen RG, Mu¨ller T (2008) Nitrogen rhizodeposition in agricultural crops: methods, estimates and future prospects. Soil Biol Biochem 40:30–48 Wolde-Meskel E, Terefework Z, Frostega˚rd A, Lindstro¨m K (2005) Genetic diversity and phylogeny of rhizobia

123

Nutr Cycl Agroecosyst (2012) 94:123–160 isolated from agroforestry legume species in southern Ethiopia. Int J Syst Evol Microbiol 55:1439–1452 Woomer P, Singleton P, Bohlool BB (1988) Ecological indicators of native rhizobia in tropical soils. Appl Environ Microbiol 54:1112–1116 Young JPW, Haukka K (1996) Diversity and phylogeny of rhizobia. New Phytol 133:87–94 Zakhia F, de Lajudie P (2001) Taxonomy of Rhizobia. Minireview. Agronomie 21:569–576 Zerihun A, McKenzie BA, Morton JD (1998) Photosynthate costs associated with the utilization of different nitrogenforms: influence on the carbon balance of plants and shootroot biomass partitioning. New Phytol 138:1–11 Zhang X, Harper R, Karsisto M, Lindstro¨m K (1991) Diversity of Rhizobium bacteria isolated from the root nodules of leguminous trees. Int J Syst Bacteriol 41:104–113 Zitzer SF, Dawson JO (1992) Soil properties and actinorhizal vagetation influence nodulation of Alnus glutinosa and Elaeagnus angustifolia by Frankia. Plant Soil 140:197–204 Zomer RJ, Trabucco A, Coe R, Place F (2009) Trees on farm: analysis of global extent and geographical patterns of agroforestry. ICRAF Working Paper no. 89. World Agroforestry Centre, Nairobi, Kenya Zou X, Binkley D, Caldwell BA (1995) Effects of dinitrogenfixing trees on phosphorus biogeochemical cycling in contrasting forests. Soil Sci Soc Am J 59:1452–1458 Zurdo-Pin˜eiro JL, Vela´zquez E, Lorite MJ, Brelles-Marin˜o G, Schro¨der EC, Bedmar EJ, Mateos PF, Martı´nez-Molina E (2004) Identification of fast-growing rhizobia nodulating tropical legumes from Puerto Rico as Rhizobium gallicum and Rhizobium tropici. Syst Appl Microbiol 27:469–477