Mound building termites contribute to savanna vegetation ...

11 downloads 31827 Views 274KB Size Report
Mound building termites contribute to savanna vegetation heterogeneity. Authors; Authors ...... Cookies. We use cookies to improve your experience with our site.
Plant Ecol (2009) 202:31–40 DOI 10.1007/s11258-009-9575-6

Mound building termites contribute to savanna vegetation heterogeneity Stein R. Moe Æ Ragnhild Mobæk Æ Anne Kjersti Narmo

Received: 28 June 2007 / Accepted: 13 January 2009 / Published online: 11 February 2009 Ó Springer Science+Business Media B.V. 2009

Abstract With biomass densities comparable to large ungulates and megaherbivores, termites play a key functional role in many tropical savanna ecosystems. This study focuses on vegetated termite mounds (termitaria) constructed by the Termitidae species Macrotermes herus. We studied how resource rich termitaria affect graminoid herbs (Poaceae and Cyperaceae), forbs and woody species composition and diversity. The density of termitaria explained 89% of the variation in dense thickets in the area. Fire tolerant Acacia species dominated the open savanna while fire sensitive species like Grewia spp. and the succulent Euphorbia candelabrum were restricted to termite mounds. Termitaria plots had four times the mean number of woody species and supported three times as many forb species as the adjacent savanna. For woody species, both the Shannon–Wiener index and the Shannon evenness index were higher on

S. R. Moe (&)  R. Mobæk  A. K. Narmo Department of Ecology and Natural Resource Management, Norwegian University of Life Sciences, ˚ s, Norway P.O. Box 5003, A e-mail: [email protected] Present Address: R. Mobæk Department of Animal and Aquaculture Sciences, Norwegian University of Life Sciences, P.O. Box 5003, ˚ s, Norway A

temitaria than on the savanna. There were no differences for graminoid herbs, except for the Shannon evenness index which was higher on termitaria. Our results indicate that graminoid herb richness peaks at lower productivity levels than trees and forbs in savanna ecosystems, as also recently found in temperate areas. Keywords Biodiversity  Macrotermes herus  Species richness and productivity  Termitaria  Termitidae  Uganda

Introduction Termites constitute an important group of the soil fauna, accounting for 40–60% of the total soil macrofauna biomass in many tropical ecosystems (Wood and Sands 1978). Termite standing biomass in African savannas has been estimated to be between 70 and 110 kg ha-1 (Wood and Sands 1978; Ferrar 1982; Wood et al. 1982). Thus, termite biomass are comparable to the biomass of ungulates (10–80 kg ha-1, Bell 1982) as well as to savanna megaherbivores (commonly \110 kg ha-1, Owen-Smith 1988), indicating that termites play a major ecological role in the functioning of savannas (Dangerfield et al. 1998). While savanna soils generally have a low reserve of weatherable minerals, especially calcium, magnesium and phosphorus (Jones and Wild 1975), termite

123

32

activity directly or indirectly modifies the availability of nutrients for other organisms (Jones et al. 1994; Dangerfield et al. 1998). The importance of termites in the decomposition of plant material is well documented (Wood 1978; Abe 1980; Martius 1994; Mando and Brussaard 1999), and termites also contribute to carbon and nitrogen mineralization, nutrient availability and stimulation of microbial activity (Lobry de Bruyn and Conacher 1990; Martius 1994; Eggleton et al. 1996; Lawton et al. 1996; Lavelle et al. 1997; Okwakol 2000). The construction of large epigeal nests and extensive underground gallery systems has major effects on soil characteristics, especially soil physico-chemical structure and properties (Anderson and Wood 1984; Nutting et al. 1987; Dangerfield et al. 1998; Asawalam et al. 1999) as well as infiltration, water status and rainfall use efficiency (Watson 1969; Arshad 1982; Eldridge 1994; Konate` et al. 1999; Mando et al. 1999; Le`onard and Rajot 2001). Nest building activities, which are typical for higher termites (Termitidae), provide nutrient-enriched microhabitats within the savanna soils (Watson 1977; Arshad 1982; Pomeroy 1983; Nyamphene 1986; Donovan et al. 2001), which in turn may facilitate vegetation growth (Hesse 1955; Lal 1988; Watson 1976, 1977; Mando et al. 1999). This study focuses on vegetated termite mounds (termitaria) constructed by the Termitidae species Macrotermes herus. Macrotermes termites do not have symbiotic flagellates in their hindgut, and contrary to lower termites they cannot digest cellulose. Instead they grow gardens of fungi (Termitomyces) both outside and inside the mound and feed the fungi with plant material together with saliva and faeces (Lee and Wood 1971). The fungus decomposes cellulose and concentrates nitrogen, phosphorus and potassium into fungal tissue that the termites consume as part of their diet (Button et al. 1983). By transporting soil mineral particles vertically and horizontally during mound construction and by bringing surface organic material into the mound to feed the fungi, the Macrotermes species commonly enrich the nutrient status locally on and around mounds (Lee and Wood 1971). Since Macrotermes mounds often contain elevated levels of soil minerals (Lee and Wood 1971; Miedema and Van Vuure 1977; Maduakor et al. 1995; McCarthy et al. 1998) they increase plant nutrient availability and thereby influence plant spatial distribution and vegetation dynamics (Jones 1992;

123

Plant Ecol (2009) 202:31–40

Konate` et al. 1999). Watson (1976, 1977) found that Macrotermes falciger mounds in oligotroph miombo woodland (Brachystegia spp. and Julbernardia spp.) in Zimbabwe were richer in plant nutrients than adjacent savanna soil and that this mound soil strongly stimulated plant growth. Probably due to this plant nutrient enrichment, the particular Macrotermes mound vegetation is found to be preferred browse by megaherbivores in miombo woodland in Zimbabwe (Loveridge and Moe 2003) and by ungulates in eutrophic savanna in Uganda (Mobæk et al. 2005). Although earlier savanna studies do indicate that termitaria support a denser plant cover and more diverse plant communities (Lee and Wood 1971; Loveridge and Moe 2003), to our knowledge there is no comprehensive study on how termitaria contribute to the diversity of graminoid herbs, forbs and woody species and consequently how termites act as agents of savanna plant heterogeneity. This study focuses on how M. herus activity affects savanna plant species composition and diversity. Specifically we tested whether termitaria support vegetation that is richer in species and denser than the surrounding savanna vegetation and whether plant community composition differs between termitaria and savanna vegetation.

Materials and methods Study site The study was conducted in Lake Mburo National Park, which is located in the southwest of Uganda (between 00°300 and 00°450 S, and 45°000 and 31°050 E). The park covers an area of *260 km2, and the altitude varies between 1,220 and 1,450 m a.s.l. (Bloesch 2002). Lake Mburo National Park lies in a rain shadow between Lake Victoria and the Ruwenzori Mountains, and receives an average of 800 mm of rain annually, with two distinct rainy seasons (February–June and October–December). The average recorded temperature is 27.5°C with daily variation ranging from 21.5 to 34.0°C. The dominant soil types are ferrasols, histosols, vertisols and leptosols (Bloesch 2002). The vegetation in Lake Mburo National Park is dry open Acacia savanna (Langdale-Brown et al. 1964), with the main graminoid herbs being

Plant Ecol (2009) 202:31–40

Sporobolus pyramidalis, Chloris gayana, Loudetia kagerensis and Bracharia eminii. The dominant tree species is Acacia hockii. Other important Acacia species are Acacia gerrardii, Acacia sieberiana and Acacia polycantha (Bloesch 2002; Hoag et al. 1991). Taxonomy follows Palgrave (2002) and Katende et al. (1995) for woody plants, Tropin (1983) for forbs, and Haines and Lye (1983) and Phillips et al. (2003) for graminoid herbs (Poaceae and Cyperaceae). Tree cover varies from open areas with only few trees to woodland, and the more open parts of the park commonly support thickets associated with termitaria. The vegetated termitaria in the park are constructed by the termite genus Macrotermes, and the dominating termite species is the mound-building M. herus (Bakuneeta 1989). This study was conducted in open savanna in flat valley bottoms in the central and eastern part of the park. These valleys are characterized by numerous thicket-termitaria associations surrounded by open savanna (Hoag et al. 1991). The termitaria thickets consist of several species of trees, shrubs, climbers and forbs. S. pyramidalis and Themeda triandra make up the main grassland association in this area, while the dominating tree associations are mixed woodland composed of a variety of Acacia species and thickets where Rhus natalensis and Grewia spp. are common. The most numerous ungulate species in Lake Mburo National park are impala (Aepyceros melampus) and Burchell’s Zebra (Equus burchelli). Other common ungulates are eland (Taurotragus oryx), topi (Damascilus korrigum), waterbuck (Kobus ellipsiprymnus), bushbuck (Tragelaphus scriptus), warthog (Phacochoerus africanus) and the African buffalo (Syncerus caffer) (Rannestad et al. 2006). Sampling Twenty termite mounds (termitaria) and 20 savanna plots were selected randomly in different areas of the park. A table of random numbers was used to pick a random distance (0.1–1 km) on the park road network. After covering the random road distance, a random compass bearing (1–360°) was selected along with another random distance away from the road. The nearest mound from the end point was included in the survey. On each termitarium, the height of the mound and four radii (two longest radii and the two perpendicular shortest radii) were measured. Mound

33

height varied from 0.5 to 2.8 m with a mean of 1.74 m (±0.59 SD) (termitaria n = 20). The average radius of mounds was 3.7 m (±0.95 SD), ranging from 2.1 to 6.4 m. By averaging the four mound radii, we estimated the mound area and determined the radius of the corresponding circular random savanna plot. Twenty random savanna plots were selected by choosing a random distance between 20 and 80 m from the mound in a random direction. Careful inspections ensured that the savanna plots were never closer than 20 m to any termitaria and that the distance from the mound to the savanna plot never exceeded 80 m. A minimum of 20 m was selected to avoid any immediate effect of the mound, and a maximum of 80 m was selected because distances longer than 80 m would increase the risk of being too close to adjacent mounds. Vegetation was recorded in June and July (forbs and woody species) and in February (graminoid herbs). All woody, forb and graminoid herb species were recorded both on mounds and in the random plots. At both mounds and random plots, woody plant stem diameter was recorded by using a slide calliper. Woody plants below 3 m height and a stem diameter less than 6.0 cm at the base (10 cm) were categorized as shrubs. Plants with non-woody stems with stem diameter less than 1.0 cm were categorized as forbs. The woody component of the vegetation was determined by systematically counting all individuals on the plots. Individual counting was done for some forb species while the percentage cover was used for others. Graminoid herb cover was estimated visually on a percentage scale. All the mounds in used in this study were M. herus mounds and active at the time of plant census. However, the mounds are frequently abandoned and re-colonised and during some periods other termite species may occupy the whole mound or only part of it (unpubl. data). The alpha- and beta-diversity was calculated for woody, forbs and graminoid herbs. Alpha-diversity or within-habitat diversity was defined as the total number of species on the plots, whereas beta-diversity or between-habitat diversity was defined as the difference in community composition between termitaria and savanna plots. We used the Jaccard index (Jaccard 1928) to calculate the beta-diversity. Shannon–Wiener index (Whittaker 1977) and Shannon evenness index (Gurevitch et al. 2002) were calculated

123

34

Plant Ecol (2009) 202:31–40

for woody and graminoid herbs on termitaria and savanna plots. The Shannon–Wiener and Shannon evenness index were not calculated for forb species because these data comprised a mixture of individual counts and estimated percentages. In order to identify how termite density related to the spatial distribution of dense thickets found in the area, the total area occupied by termitaria and thickets was estimated from transect lines. Thickets were defined as closed patches of tall vegetation consisting of more than one plant species. A total of 17 transects each 500 by 50 m were located randomly in different parts of the open savanna. One person walked a straight line in a random compass bearing and using a GPS, while two persons counted the termitaria on each side. The number of termitaria with thickets (vegetated mounds), termitaria without thickets (bare mounds) and thickets without termitaria were recorded.

Results Woody species were present on all the 20 termitaria, and all of the mounds supported more than one species. In contrast, woody species occurred in 14 of the savanna plots and only 7 of these plots contained more than one species. We found that woody species alpha-diversity was more than four times higher on termitaria compared to savanna plots (paired t-test, t = 8.28, P \ 0.0001) (Table 1). The Shannon– Wiener index was significantly higher for woody species on termitaria than for woody species in the savanna plots (paired t-test, t = 8.54, P \ 0.0001). Also the Shannon evenness index was significantly

Table 1 Differences in number (mean ± SD) of plant species, densities and Shannon–Wiener index and Shannon evenness index between termitaria and savanna plots (paired t-tests, n1 and n2 = 20)

123

higher for termitaria woody vegetation compared to the savanna plots (Paired t-test, t = 4.72, P = 0.001) (Table 1). Tree densities on termitaria were 25 times the densities in the savanna plots (paired t-test, t = 5.02, P \ 0.0001), while shrub densities were 6 times higher on termitaria compared to savanna plots (paired t-test, t = 6.77, P \ 0.0001). Forb species were present on all of the 20 termitaria and in 19 of the savanna plots. All of the mounds and 13 of the savanna plots supported more than one forb species. The forb species alphadiversity on termitaria was more than twice as high as in the savanna plots (paired t-test, t = 6.80, P \ 0.0001) (Table 1). There were no significant differences in number of graminoid herbs between termitaria and savanna plots (paired t-test, t = 5.97, P = 0.53) (Table 1), nor did the Shannon–Wiener index for graminoid herbs differ between savanna and termitaria (paired t-test, t = 0.17, P = 0.11). However, the graminoid herbs Shannon evenness index was significantly higher on termitaria compared to the savanna (paired t-test, t = 3.49, P = 0.0003) (Table 1). A total of 40 woody species were found on termitaria, while 14 woody species were recorded in the savanna plots. Eleven species were unique for the termitaria while only A. sieberiana and A. hockii were exclusively found on the savanna plots (Fig. 1a). The most frequent woody species on termitaria were Grewia mollis, Grewia similis, Maytenus heterophylla and R. natalensis while the savanna plots were dominated by Dichrostachys cinerea, A. sieberiana, Solanum incanum and A. gerrardii (Fig. 1a). Succulent species like Cissus quadrangularis, Cyphostemma adenocaule and Euphorbia candelabrum were

Termitaria Number of woody species

10.5 ± 3.9

Savanna plots 2.5 ± 2.0

P-value P \ 0.0001

Shannon–Wiener index for woody species

1.53 ± 0.51

0.25 ± 0.38

P \ 0.0001

Shannon evenness index for woody species

0.80 ± 0.13

0.30 ± 0.42

P = 0.0001

Density of trees (m-2)

0.25 ± 0.20

0.01 ± 0.02

P \ 0.0001

Density of shrubs (m-2) Number of forb species

0.56 ± 0.33 7.6 ± 2.4

0.09 ± 0.12 2.7 ± 2.2

P \ 0.0001 P \ 0.0001

Number of graminoid herbs

6.80 ± 2.19

7.10 ± 2.07

P = 0.53

Shannon–Wiener index for graminoid herbs

1.58 ± 0.33

1.40 ± 0.43

P = 0.14

Shannon evenness index for graminoid herbs

0.52 ± 0.13

0.38 ± 0.13

P = 0.004

Plant Ecol (2009) 202:31–40

35

Fig. 1 Number of plots where a particular woody (a), forb (b) and graminoid herb (c) species was recorded on termitaria (n = 20) and in the savanna plots (n = 20). Only species present on more than 3 (woody and forb species) or 7 (graminoid herbs) termitaria or savanna plots are included in the figure

20

a)

15 10 5 0

. llis ilis ylla sis tata um eri osa nse sp rum ilis ulis nus ura sita sp. dea rdii rea ana ckii mo sim rophtalen aris can himp ent arie aris lab nob ed rica one ppo eae scoi erra cine beri ho a i ia e a e in sc om sc pp de lea ssa af ap o ac di g ys ie cia t s a a a n c ri s s ia i a ia h ew ew et n m m Gr Gr us h hus taec lanu belia aris dag C ia ca Te Ca hylu iscu lund Rub itari cac stac acia Ac r A ro c p ib os n R ae So m pp ma b a r o e A l t g a l o H H E C m r y ch A Ch ph iu Di Ma Ma Eu pis ir le T

Number of plots

20

b)

15 10 5 0

. . i i i ell nus sp aris ule aris um lius nti tata esis es lata per tus ans cta sp aris rch frica lina ngul noca gell ssim atifo gra den ngn liform ricu him gula adic arre ium line u b a me ra e fla ati n l ens vix bo bel au sc an e r era od sia o id ia f sm o a la m io st a nia us m ad ad us gr vo rag Co qu ma rag um stem B rler occ va fl grofi inu nec hori digo De ephr a a s a im no a sc In a roc ri m P sp m l y T u p s e B o P H Ja S Dy ss st As Oc ola A yth Ci pho S Er y C 20

c)

15 10 5 0

s i a a a n a a um tylo alis ym inica ens dula ndr nus nus ntha rata nsis yan iflor osa vata tidu edi at thrix ximdac amid mon yss umb pen tria pfia frica riza spe hire ga ong cem ia o s ni dbra com cno a a i s l a l d y a c b a u b l i a r a i t m on y ho a ec fi d s s a ex s or ia r ar re m ri p um od s p ria ria ia d nia me lus olu iari tis on hl itar stis ga yc tis hia ris nic Cyn bulu eta igita hiar rrhe The obo rob ach ros pog C Dig gro bild P ros Brac Chlo a a A o S D ac pa P ag or po Br rag ndro Er or Er E A Sp S Br Hy Sp

Termitaria

restricted to mounds. The Jaccard index between termitaria and savanna plots for woody species was 31.7%. The total number of forb species recorded on termitaria and in the savanna plots was 41 and 22, respectively. The most frequent forb species on termitaria were Pavonia burchellii, Asparagus

Savanna plots

africanus and Commelina sp. (Fig. 1b), while the most frequent forb species in the savanna plots were Ocimum gratissimum, Bidens grantii and Asparagus flagellaris. The Jaccard index between termitaria and savanna plots for forb species was 18.9%. Although some graminoid herbs dominated the termitaria while others dominated the savanna plots,

123

36

Plant Ecol (2009) 202:31–40

only two graminoid herbs (Setaria homonyma and Pycreus nitidus) occurred exclusively on the termitaria while none of the graminoid herbs were exclusively associated with the savanna (Fig. 1c). Panicum maximum was the most abundant graminoid herb on termitaria, while T. triandra and Sporobolus staphianus dominated open savanna plots. In total, 28 graminoid herbs were recorded on termitaria and 30 species were found in the savanna plots. The Jaccard index between termitaria and savanna plots for graminoid herbs was 62.2%. There were no correlations between number of graminoid herbs and forb species, graminoid herbs and trees or forbs and trees, neither within the mounds nor within the savanna plots (Person Product Moment Correlations, all P [ 0.05). The only significant species group correlation was between number of graminoid herbs on the mounds and the number of graminoid herbs on corresponding savanna plots (Person Product Moment Correlation, r = 0.51, P = 0.02). The density of termitaria explained 89% of the variation in the density of thickets (R2 = 0. 89, P \ 0.001, transect n = 17) (Fig. 2). The total density of termitaria was 10.1 ha-1 (±3.4 SD) and the total density of thickets was 8.9 ha-1 (±3.3 SD). When looking at vegetated and bare mounds separately, the density of termitaria with thickets was 8.5 ha-1 (±2.8 SD) while the density of termitaria without thickets was 1.6 ha-1 (±1.3 SD). The density of thickets without termitaria was 0.5 ha-1 (±0.6 SD).

Number of thickets ha-1

18

R2 = 0.89 P < 0.001

16 14 12 10 8 6 4

4

6

8

10

12

Number of termitaria

14

16

18

ha-1

Fig. 2 The relationship between density of termitaria (ha-1) and density of thickets (ha-1) (R2 = 0.89, P \ 0.001, n = 17)

123

Discussion Woody and forb species numbers, the Shannon– Wiener indexes and density were far higher on termitaria compared to savanna plots, while for graminoid herbs only Shannon evenness index was higher on termitaria. Mean number of woody species on termitaria was four times of the savanna areas while termitaria supported three times as many forb species. The fact that species richness for woody, forbs or graminoid herbs was not related indicates that the species richness of these plant species groups is determined by different resource constraints. Consistent with previous studies we found that woody species densities were significantly higher on termitaria compared to the savanna areas. Several studies have reported that termitaria support denser vegetation than surrounding savanna areas (Wild 1952; Arshad 1982; Pomeroy 1983; Loveridge and Moe 2004). In a study from miombo woodland in Zimbabwe, Macrotermes termitaria were found to have a higher number of tree species, twice the density of trees and almost three times the woody basal cover than the adjacent woodland (Loveridge and Moe 2004). In a vegetation restoration experiment from the Sahelian zone in West Africa, Mando et al. (1999) found that termite activity increased plant cover. According to Hesse (1955), crop plants also grow better on Macrotermes mounds than on soils of the surrounding savanna. Lal (1988) proposed that the improved growth on mounds is due to a combination of high water status and enhanced soil mineral contents. He suggested that termitaria are less leached than the surrounding soil and therefore contain relatively high amounts of soluble plant nutrients (Lal 1988). This is supported by Konate` et al. (1999), who found that the water storage capacity was higher in termite soil compared to surrounding soil in a West African savanna. It is well documented that termitaria are humid, well-drained patches containing high levels of important plant nutrients (e.g. Lee and Wood 1971; Lal 1988; Mahaney et al. 1999) and as such represent resource islands in relatively nutrient poor savanna environments (Frost 1996). Plant species richness generally follow a hump-shaped relationship, increasing with increasing productivity until it peaks and the diversity declines with a further productivity increase (e.g. Tilman 1988). The fact that the number of

Plant Ecol (2009) 202:31–40

woody and forb species increased between savanna and termitaria indicates that, for these species groups, the species richness peaks at relatively high productivity levels. The species richness probably peaks at lower productivity levels for graminoid herbs since we found no difference between the resource poor savannas and the resource rich termitaria. In central Europe, species richness of grasslands and wetlands has been found to peak in relatively nutrient poor soils while forest species richness peaks in relatively nutrient rich soils (Cornwell and Grubb 2003). Similarly, when fertilizer was experimentally added to temperate grasslands the graminoid herbs richness was reduced (Crawley et al. 2005; Gross et al. 2005). Thus, our findings indicate that these relationships between soil nutrients and species richness may also be applicable to tropical savannas. The plant species community composition differed between termitaria and savanna vegetation. Although many woody and forb species grew on both mounds and in the savanna plots, many plant species were unique for termitaria, while only a few species occurred exclusively in the savanna plots. Succulent species were restricted to mounds. Succulent growth forms are commonly associated with high soil nutrient levels under arid conditions (Knight et al. 1989), and termitaria seems to be the only suitable sites for these nutrophilic species in Lake Mburo National Park. Furthermore, since succulents are highly sensitive to fire (Pe´rez-Garcı´a and Meave 2006), termitaria may provide protection from grass fires for these species. Individuals of R. natalensis, Embelia schimperi, Trilepisium madagascarensis and Hibiscus aponeura found in the savanna were all saplings (\50 cm), but on termitaria these species also survived to mature trees, indicating high seedling predation levels on the savanna. Dense termitaria vegetation may to some extent protect woody species from browsing animals. Furthermore, woody plants growing on nutrient rich sites have sufficient carbon and nutrient reserves to replace lost tissue and thus compensate for ungulate browsing (Bryant et al. 1983). The mounds in Lake Mburo National Park are elevated spots and accordingly vegetation on termitaria may be less subjected to ground fires (Lee and Wood 1971; Lind and Morrison 1974). Both A. sieberiana and D. cinerea, dominating the savanna

37

plots, are well adapted to fire (Gillon 1983; Sabiiti and Wein 1987). When investigating seeds responses to fire in Acacia sieberiana in Queen Elizabeth National Park in Uganda, Sabiiti and Wein (1987) found that fire stimulated germination from the seed pool in the soil, and that the emergence of seedlings was higher on burned plots than on unburned plots. Acacias are also generally shade intolerant species with no seedling establishment under the canopies of mature trees (Owen-Smith 1988). For other woody species, fire may limit both tree recruitment and the progression of an individual tree from the seedling stage to adult (Higgins et al. 1999). The number of thickets in Lake Mburo National Park can almost exclusively be explained by the density of termite mounds. Expanses of thicket associations are found in the south-western part of the park, which is less disturbed by fire and grazing (Hoag et al. 1991). In a recent study, including our study area, Bloesch (2008) concluded that thickets on seasonally waterlogged plains were established on Macrotermitinae mounds because of fire protection, increased soil fertility and good drainage, while on stony hillsides thicket clump establishment was a function of the fire regime and the browsing intensity. Smaller termite mounds of more recent origin commonly lack vegetation. According to Bloesch (2008) recent mounds are effectively maintained and woody vegetation is only able to established in periods were the mounds are not active. In a study from Serengeti, Sharam et al. (2006) found that Croton thickets were associated with hilltops or riparian forests. In Serengeti, Euclea divinorum acted as a pioneer species facilitating thicket establishment. The establishment of Euclea divinorum was hampered by frequent fires, browsing and dense graminoid herbs cover (Sharam et al. 2006). In Lake Mburo National Park, Rhus spp. and Grewia spp. are suggested to be instrumental to thicket establishment (Hoeg et al. 1991; Bloesch 2002). While termite mounds benefit vegetation establishment because of fire protection, increased soil fertility and good drainage (Bloesch 2008), the vegetation also provide benefit to termites in providing detritus and dead wood. In conclusion, Macrotermes termitaria, occupying about 5% of the area in this savanna landscape, are important contributors to vegetation heterogeneity within the Lake Mburo National Park. Explaining

123

38

89% of the variation in dense thickets, termitaria represent resource islands for many plant species found exclusively on termitaria. Fire tolerant Acacia species dominate the open savanna while fire sensitive species like Grewia spp. and succulents like E. candelabrum are restricted to mounds. While the Shannon–Wiener index and the Shannon evenness index on termitaria far exceed the adjacent savanna for woody and forb species, there were no such differences for graminoid herbs, except Shannon evenness index which was higher on termitaria. Our results support findings from temperate areas in that graminoid herb richness peaks at lower productivity levels than trees and forbs. Accordingly the highly productivity termitaria only contribute to a species richness increase of trees and forbs. Acknowledgements We thank the Uganda Wildlife Authority and the Uganda National Council for Science and Technology for permission to conduct this study. J. Colman and O. G. Støen commented on earlier drafts of this manuscript. Dr. C. Bakuneeta and D. K. Aleper assisted during the study. O. Wanyana Magyani and Prof. K. A. Lye helped in identifying plant species. We are grateful to the park staff for their hospitality and support. In particular, we would like to thank K. T. Anderson for assistance in the field. This study was supported by grants from Noragric at the Norwegian University of Life Sciences and by NUFU.

References Abe T (1980) Studies on the distribution and ecological role of termites in a lowland rain forest of West Malaysia. 4. The role of termites in the process of wood decomposition in the Pasho Forest Reserve. Rev d’Ecol et de Biol du Sol 17:23–40 Anderson JM, Wood TG (1984) Mound composition and soil modification by two soil-feeding termites (Termitinae, Termitidae) in a riparian Nigerian forest. Pedobiologia 26:77–82 Arshad MA (1982) Influence of the termite Macrotermes michaelseni (Sjo¨st) on soil fertility and vegetation in a semi-arid savannah ecosystem. Agro-Ecosyst 8:47–58 Asawalam DO, Osodeke VE, Kamalu OJ, Ugwa IK (1999) Effects of termites on the physical and chemical properties of the acid sandy soils of Southern Nigeria. Commun Soil Sci Plant Anal 30:1691–1696 Bakuneeta CM (1989) The importance of Macrotermes herus (Sjostedt) to the ecosystems in Queen Elizabeth and Lake Mburo National Parks in Uganda. MSc Thesis, Makerere University, Uganda Bell RHV (1982) The effect of soil nutrient availability on community structure in African ecosystems. In: Huntley BJ, Walker BH (eds) Ecology of tropical savannas. Springer, Berlin, pp 193–216

123

Plant Ecol (2009) 202:31–40 Bloesch U (2002) The dynamics of thicket clumps in the Kagera savanna landscape, East Africa. PhD thesis, Swiss Federal Institute of Technology, Zurich, Switzerland Bloesch U (2008) Thicket clumps: a characteristic feature of the Kagera savanna landscape, East Africa. J Veg Sci 19: 31–44 Bryant JP, Stuart Chapin F, Klein DR (1983) Carbon/nutrient balance of boreal plants in relation to vertebrate herbivory. Oikos 40:357–368 Button TW, Arshad MA, Tieszen LL (1983) Stable isotope analysis of termite food habits in East African grasslands. Oecologia 59:1–6 Cornwell WK, Grubb PJ (2003) Regional and local patterns in plant species richness with respect to resource availability. Oikos 100:417–428 Crawley MJ, Johnston AE, Silvertown J, Dodd M, de Mazancourt C, Heard MS, Henman DF, Edwards GR (2005) Determinants of species richness in the park grass experiment. Am Nat 165:179–192 Dangerfield JM, McCarthy TS, Ellery WN (1998) The moundbuilding termite Macrotermes michaelseni as an ecosystem engineer. J Trop Ecol 14:507–520 Donovan SE, Eggleton P, Dubbin WE, Batchelder M, Dibog L (2001) The effect of a soil-feeding termite, Cubitermes fungifaber (Isoptera: Termitidae) on soil properties: termites may be an important source of soil microhabitat heterogeneity in tropical forests. Pedobiologia 45:1–11 Eggleton P, Bignell DE, Sands WA, Mawdsley NA, Lawton JH, Wood TG, Bignell NC (1996) The diversity, abundance and biomass of termites under differing levels of disturbance in the Mbalmayo Forest Reserve, southern Cameroon. Phil Trans Royal Soc Lond 351:51–68 Eldridge DJ (1994) Nests of ants and termites influence infiltration in a semi-arid woodland. Pedobiologia 38:481–492 Ferrar P (1982) Termites of a South African savanna IV. Subterranean populations, mass determinations and biomass estimations. Oecologia 52:147–151 Frost PGH (1996) The ecology of miombo woodlands. In: Campbell BM (ed) The miombo transition: woodland and welfare in Africa. Centre for International Forestry Research, Bogor, Indonesia, pp 11–57 Gillon D (1983) The fire problem in tropical savannas. In: Bourlie`re F (ed) Tropical savannas. Ecosystems of the world 13. Elsevier, New York, pp 617–641 Gross KL, Mittelbach GG, Reynolds HL (2005) Grassland invasiblity and diversity: responses to nutrients, seed input, and disturbance. Ecology 86:476–486 Gurevitch J, Scheiner SM, Fox GA (2002) The ecology of plants. Sinauer Associates Inc., Sunderland, Massachusetts, USA Haines RW, Lye KA (1983) The sedges and rushes of East Africa. The East Africa Natural History Society, Nairobi, Kenya Hesse PR (1955) A chemical and physical study of the soils of termite mounds in East Africa. J Ecol 43:449–461 Higgins SI, Shackleton CM, Robinson ER (1999) Changes in woody community structure and composition under contrasting landuse systems in a semi-arid savanna, South Africa. J Biogeogr 26:619–627 Hoag A, Clements A, Monday G (1991) Lake Mburo National Park Habitat Project. A study of the vegetation

Plant Ecol (2009) 202:31–40 composition and animal use of the major terrestrial habitats in the park. USAID, USA Jaccard P (1928) Die statistisch-floristische Methode als Grundlage der Pflanzensoziologie. In: Abderhalden (ed) Handbuch biologischer Arbeitsmethoden, vol 11, pp 165–202 Jones JA (1992) Soil formation and nutrient dynamics at the woodland-savanna boundary in East Africa. In: Furley PA, Proctor J, Ratter JA (eds) Nature and dynamics of forest-savanna boundaries. Chapman & Hall, London, UK, pp 185–212 Jones MJ, Wild A (1975) Soils of the west African savanna. Commonwealth Africultural Bureax (CAB) Technical Communication, vol 55. Commonwealth Agricultural Bureau, Harpenden, pp 1–246 Jones CG, Lawton JH, Shachak M (1994) Organisms as ecosystem engineers. Oikos 69:373–386 Katende AB, Birnie A, Tengna¨ B (1995) Useful trees and shrubs for Uganda. Regional Soil Conservation Unit (RSCU/SIDA), Nairobi, Kenya Knight RS, Rebelo AG, Siegfried WR (1989) Plant assemblages on Mima-like earth mounds in the Clanwille district, South Africa. S Afr J Bot 55:465–472 Konate` S, Le Roux X, Tessier D, Lepage M (1999) Influence of large termitaria on soil characteristics, soil water regime, and tree leaf shedding pattern in a West African savanna. Plant Soil 206:47–60 Lal R (1988) Effects of macrofauna on soil properties in tropical ecosystems. Agric Ecosyst Environ 24:101–116 Langdale-Brown I, Osmaston HA, Wilson JG (1964) The vegetation of Uganda and its bearing on the land use. Government Printer, Entebbe, Uganda Lavelle P, Bignell D, Lepage M, Wolters V, Roger P, Inneson P, Heal OW, Dhillion S (1997) Soil function in a changing world: the role of invertebrate ecosystem engineers. Eur J Soil Biol 33:159–193 Lawton JH, Bignell DE, Bloemers GF, Eggleton P, Hodda ME (1996) Carbon flux and diversity of nematodes and termites in Cameroon forest soils. Biodiv Cons 5:261–273 Lee KE, Wood TG (1971) Termites and soils. Academic press, London Le`onard J, Rajot JL (2001) Influence of termites on runoff and infiltration: quantification and analysis. Geoderma 104:17–40 Lind EM, Morrison MES (1974) East African vegetation. Longman, London, UK Lobry de Bruyn LA, Conacher AJ (1990) The role of termites and ants in soil modification: a review. Austr J Soil Res 28:55–93 Loveridge JP, Moe SR (2004) Termitaria as browsing hotspots for African megaherbivores. J Trop Ecol 20:337–342 Maduakor HO, Okerere AN, Onyeanuforo CC (1995) Termite mounds in relation to surrounding soils in the forest and the derived savanna zones of southeastern Nigeria. Biol Fert Soils 20:157–162 Mahaney WC, Zippin J, Millner MW, Sanmugadas K, Hancock RGV, Aufreiter S, Cambell S, Huffman MA, Wink M, Malloch D, Kalm V (1999) Chemistry, mineralogy and microbiology of termite mound soil eaten by chimpanzees of the Mahale Mountains, Western Tanzania. J Trop Ecol 15:565–588

39 Mando A, Brussaard L (1999) Contribution of termites to the breakdown of straw under Sahelian conditions. Biol Fert Soils 29:332–334 Mando A, Brussaard L, Stroosnijder L (1999) Termite- and mulch-mediated rehabilitation of vegetation on crusted soil in West Africa. Restor Ecol 7:33–41 Martius C (1994) Diversity and ecology of termites in Amazonian forests. Pedobiol 38:407–428 McCarthy TS, Ellery WN, Dangerfield JM (1998) The role of biota in the initiation and growth of islands on the floodplain of the Okavango alluvial fan, Botswana. E Surf Proc Landf 23:291–316 Miedema R, Van Vuure W (1977) The morphological, physical and chemical properties of two mounds of Macrotermes bellicosus (Smeathman) compared with surrounding soils in Sierra Leone. J Soil Sci 28:112–124 Mobæk R, Narmo AK, Moe SR (2005) Termitaria are focal feeding sites for large ungulates in Lake Mburo National Park, Uganda. J Zool Lond 267:97–102 Nutting WL, Haverty MI, LaFage JP (1987) Physical and chemical alteration of soil by two subterranean termite species in Sonoran Desert grassland. J Arid Env 12:233–239 Nyamphene KW (1986) The use of termite mounds in Zimbabwe peasant agriculture. Trop Agric 63:191–192 Okwakol MJN (2000) Changes in termite (Isoptera) communities due to the clearance and cultivation of tropical forest in Uganda. Afr J Ecol 38:1–7 Owen-Smith RN (1988) Megaherbivores: the influence of very large body size on ecology. Cambridge University Press, Cambridge, UK Palgrave MC (2002) Trees of Southern Africa, 3rd edn. Struik Publ, Cape Town, South Africa Pe´rez-Garcı´a EA, Meave JA (2006) Coexistence and divergence of savanna and tropical dry forest in southern Mexico. J Biogeog 33:438–447 Phillips S, Namaganda M, Lye KA (2003) 115 Ugandan grasses. Makerere herbarium handbook no. 1. Makerere University, Uganda Pomeroy DE (1983) Some effects of the mound-building termites on the soils of a semi-arid area of Kenya. J Soil Sci 34:555–570 Rannestad OT, Danielsen T, Moe SR, Stokke S (2006) Adjacent pastoral areas support higher densities of wild ungulates than the Lake Mburo National Park in Uganda. J Trop Ecol 22:675–683 Sabiiti EN, Wein RW (1987) Fire and Acacia seeds: a hypothesis of colonization success. J Ecol 74:937–946 Sharam G, Sinclair ARE, Turkington R (2006) Establishment of broad-leaved thickets in Serengeti, Tanzania: the influence of fire, browsers, grass competition, and elephants. Biotropica 38:599–605 Tilman D (1988) Plant strategies and the dynamics and structure of plant communities. Monographs in population biology. Princeton University Press, Princeton, New Jersey, USA Tropin G (1983) Flore du Rwanda. Spermatophytes, vol 1–3. Musee Royal de L’Afrique Centrale, Tervuren, Belgique Watson JP (1969) Water movement in two termite mounds in Rhodesia. J Ecol 57:441–451

123

40 Watson JP (1976) The composition of mounds of the termite Macrotermes falciger (Gersta¨cker) on soil derived from granite in three rainfall zones of Rhodesia. J Soil Sci 27:495–503 Watson JP (1977) The use of mounds of the termite Macrotermes falciger (Gersta¨cker) as a soil amendment. J Soil Sci 28:664–672 Whittaker RH (1977) Evolution of species diversity in land communities. In: Hecht MK, Steere WS, Wallace B (eds) Evolutionary biology, vol 10. Plenum Press, New York, USA, pp 1–67

123

Plant Ecol (2009) 202:31–40 Wild H (1952) The vegetation of southern Rhodesian termitaria. Rhod Agric J 49:280–292 Wood TG (1978) Food and feeding habits of termites. In: Brian MV (ed) Production ecology of ants and termites. Cambridge University Press, Cambridge, UK, pp 55–90 Wood TG, Sands WA (1978) The role of termites in ecosystems. In: Brian MV (ed) Production ecology of ants and termites. Cambridge University Press, Cambridge, UK, pp 245–292