Skip to main content

RETRACTED ARTICLE: Tree species diversity predicts aboveground carbon storage through functional diversity and functional dominance in the dry evergreen Afromontane forest of Hararghe highland, Southeast Ethiopia

This article was retracted on 06 September 2023

This article has been updated

Abstract

Background

Regarding the most important ecological challenges, scientists are increasingly debating the relationship between biodiversity and ecosystem function. Despite this, several experimental and theoretical researches have shown inconsistencies in biodiversity and ecosystem function relationships, supporting either the niche complementarity or selection effect hypothesis. The relationship between species diversity, functional diversity, and aboveground biomass carbon was investigated in this study employing standing aboveground carbon (AGC) stock as a proxy measure for ecosystem function. We hypothesized that (i) effects of diversity on AGC can be transmitted through functional diversity and functional dominance; (ii) effects of diversity on AGC would be greater for functional dominance than functional diversity; and (iii) effects of functional diversity and functional dominance on carbon stock varied with metrics and functional traits. Community-weighted means (CWM) of functional traits (wood density, specific leaf area, and maximum plant height) were calculated to assess functional dominance (selection effects). As for functional diversity (complementarity effects), multi-trait functional diversity (selection effects) indices were computed. We tested the first hypothesis using structural equation modeling. For the second hypothesis, the effects of environmental variables such as slope, aspect, and elevation were tested first, and separate linear mixed-effects models were fitted afterward for functional diversity, functional dominance, and the two together.

Results

Results revealed that slope had a significant effect on aboveground carbon storage. Functional diversity and functional dominance were significant predictors of the aboveground carbon storage (22.4%) in the dry evergreen Afromontane forest. The effects of species richness on aboveground carbon storage were mediated by functional diversity and functional dominance of species. This shows that both the selection effects and the niche complementarity are important for aboveground carbon storage prediction. However, the functional diversity effects (niche complementarity) were greater than functional dominance effects (selection effects).

Conclusions

Linking diversity and biodiversity components to aboveground carbon provides better insights into the mechanisms that explain variation in aboveground carbon storage in natural forests, which may help improve the prediction of ecosystem functions.

Background

The relationship between plant diversity and ecosystem functioning is one of the main ecological issues (Ruiz-Jaen and Potvin 2011; Cavanaugh et al. 2014), and some aspects of climate-related effects have been well tested (Durán et al. 2015; Poorter et al. 2015; Wu et al. 2015). Because biomass is an important component of ecosystem function (productivity), it is therefore possible to link the relationship between plant diversity and biomass to the biodiversity component and ecosystem functioning (biomass) (Lasky et al. 2014; Mensah et al. 2016). Productivity is the most common measure of ecosystem function, but given the difficulty in measuring productivity, standing biomass is frequently used as a proxy measure for ecosystem function. Hence, a clear understanding of how diversity and dominance affect ecosystem function is important not only in theory-based forest management but also in helping direct strategies for the conservation and restoration of threatened natural ecosystems (Cavanaugh et al. 2014; Ruiz-Jaen and Potvin 2010; Mensah et al. 2016).

Niche complementarity and selection effect hypotheses are the most common proposed ecological hypothesis used to explain the role of species diversity in ecosystem dynamics, process, and ecosystem functioning (Cavanaugh et al. 2014; Poorter et al. 2015; Mensah et al. 2016). The selection effect hypothesis postulates that a higher probability of occurrence of dominant species or traits (because of competition) would influence the functioning of the ecosystem in a diverse group of species (Grime 1998). The niche complementary effect hypothesis states that increasing diversity would promote a wider range of functional traits and provide opportunities for species to use the resources efficiently, increasing ecosystem function and decreasing competition (Tilman et al. 1997). We usually infer the niche complementary effect hypothesis when explaining higher production of biomass and productivity in highly diverse ecosystems (Mensah et al. 2016). Higher biomass, however, may also result from dominant species with strong resource responses and/or strong ecosystem effects that refer to selection effects (Reich et al. 2001).

The relationship between the components of biodiversity and the function of the ecosystem is incoherent in whether these relationships are positive, negative, or no relationship at all. For example, in the primary Pinus kesiya forest in Yunnan of southwest China, Li et al. (2018) reported a positive relationship between species richness and aboveground biomass. Poorter et al. (2015) found a positive relationship between species diversity and aboveground biomass in tropical forest biomass. Zhang and Chen (2015) found a strong association between diversity and aboveground biomass within a natural temperate spruce and pine forest. Jerzy and Anna (2007) reported negative relationship between species diversity and the storage of biomass in a European pine forest, while other studies have found no relation between aboveground biomass (AGB) and tree species diversity in a forest ecosystem (Whittaker and Heegaard 2003).

In recent decades, many scientists have focused on functional diversity and functional dominance to explain the underlying mechanisms in diversity-carbon relationships (Baraloto et al. 2012; Clark et al. 2012; Ruiz-Jaen and Potvin 2011; Díaz and Cabido 2001; Tilman et al. 1997). A variety of functional traits (functional diversity) among species are assumed to reflect the complementarity effect, and the traits of dominant species (functional dominance) are assumed to reflect the selection effect (Mensah et al. 2016). Some recent forest ecosystem studies support the selection effect hypothesis (Lin et al. 2016; Prado-Junior et al. 2016), while others support the complementarity of niches more (Morin et al. 2011; Wang et al. 2011; Mensah et al. 2016; Sintayehu et al. 2020). Grime (1998) proposed that functional dominance played a more important role in ecosystem functioning when species differed substantially in their abundances. Several scholars have argued that these hypotheses are not mutually exclusive and jointly might determine the effects of diversity on ecosystem functions (Cavanaugh et al. 2014; Lasky et al. 2014), especially through competitive exclusion (selection effects) and complimentary use of resources by competitors (Ali et al. 2016; Poorter et al. 2015; Mensah et al. 2016; Mensah et al. 2018). Understanding whether diversity influences on ecosystem function are most likely mediated through functional diversity than functional dominance, or vice versa, will bring substantial insights into which mechanism is more relevant.

Few studies have addressed the relationships between biodiversity component and ecosystem function in dry tropical forests (Conti and Díaz 2013; Prado-Junior et al. 2016). Using aboveground tree carbon data in a dry evergreen Afromontane forest of Southeast Ethiopia, we examined the relationship between diversity and aboveground carbon stocks through the effects of functional diversity and functional dominance. This study hypothesized that (i) effects of diversity on aboveground carbon stock are transmitted through functional diversity and functional dominance. However, there are notions that diversity and carbon relationships are caused by co-varying environmental factors (Cavanaugh et al. 2014; Mensah et al. 2016). Therefore, elevation, slope, and aspect were considered to be the most important environmental factors in this forest and we tested for their influences on tree aboveground carbon stocks. In addition, while accounting for significant environmental factors effects, we also hypothesized that (ii) effects of diversity on carbon storage would be greater for functional dominance than for functional diversity and (iii) effects of functional diversity and functional dominance on carbon stock varied with metrics and functional traits.

Material and methods

Study area

We conducted the study in the Dindin dry evergreen Afromontane forest on the Hararghe highland, Southeast Ethiopia. The geographical location of the study site lies between 40° 10′ 40″ to 40° 18′ 50″ E and 8° 33′ 0″ to 8° 40′ 40″ N with elevation ranges between 2124 and 3069 m a.s.l. and situated around 336 km southeast of Addis Ababa (Fig. 1). Because of the lack of long-term climatic data for the study site, we used the climate estimator software tool, New LocClim, to produce long-term monthly precipitation and temperature (FAO 2005; Grieser et al. 2006). The mean annual temperature (MAT) and mean annual precipitation (MAP) at the study site estimated to be 25.6 °C and 804 mm/year, respectively. Precipitation at the study site has a bimodal distribution pattern with a long rainy season lasting from June to October and a short rainy season from April to May (Fig. 2). The soils of the study area developed from a wide range of parent materials, including volcanic and mixed limestone and sandstone over a Precambrian basement. Leptosols are the most abundant soil types which are classified as Haplic and Lithic leptosols. Lithic leptosols covers are most extensive on high in mountainous relief hills and parallel ridges and river, gorges having very steep slopes (30–60%) (Tefera et al. 1996; Elias 2016). According to Friis et al. (2010), the vegetation type in the study area is categorized as tropical dry evergreen Afromontane forest, which is characterized by a dry climate. Woody species such as Afrocarpus falcatus, Maesa lanceolata, Allophylus abyssinicus, and Vernonia myriantha are dominant species.

Fig. 1
figure 1

Map of study area and location of sampling sites. The red rectangle represents the study site; the black circles represent individual horizontal transect sampling sites (points). DDAF, Dindin dry evergreen Afromontane forest

Fig. 2
figure 2

The climate diagram of the study area (values generated using New_LocClim). The red and blue lines show the temperature curve and annual precipitation time series, respectively. MAT, mean annual temperature; Max, maximum absolute temperature; Min, minimum absolute temperature; MAP, mean annual precipitation

Vegetation data collection

In 54 plots of 20 × 20 m sizes, a systematic sampling technique was used to collect data on the diversity of plant species along horizontal transect line (Fig. 1). The horizontal transects were located systematically at an elevation interval of 200 m from 2300 to 2900 m (Fig. 1). For each horizontal transect, two sites were selected along with the tree-canopy cover, characterized by the closed and open cover. In the horizontal transect at 2300, the forest is dominated by evergreen conifer Afrocarpus falcatus and at 2500 m, Afrocarpus falcatus-Dombeya aethiopica co-dominate the forest, whereas at 2700 and 2900 m, the forest is dominated by the evergreen conifer Juniperus excesla. The sites within the same horizontal transects were located 200–300 m apart from each other, depending on local conditions. We laid the plots in each horizontal transect out parallel to each other. We adopted this approach for two major reasons: (i) there is a minimum sampling bias as they represent all woody species of the study forest and to capture the natural variation in functional traits along elevation; (ii) the effect of area (the decrease in the area from bottom to the top of a mountain) minimized, as all forest areas included (Berhanu et al. 2016; Jiang et al. 2018). Therefore, this procedure gives a precise indication of species represented and used to achieve better documentation of changes in species distribution patterns along an elevation (Aynekulu et al. 2012; Arellano et al. 2016).

In each plot, all individual tree species were counted, and height and diameter at breast height (dbh) measured. In two perpendicular directions, diameter at breast height (dbh1 and dbh2) was measured with a caliper at 1.3 m above ground level or just above the buttress and we used the average value in biomass estimation. Height and diameter were measured by using Sunto clinometer and clipper, respectively. For this study, we define trees as one-stemmed woody plants with dbh ≥ 5 cm and heights ≥ 3 m. In total, 1210 individual trees of 36 species (31 genera and 27 families) were recorded and measured in the plots (Table S1). The names of the species were identified up to the species level in the field, but in doubtful cases, vouchers were collected and pressed for further identification and confirmation at the Addis Ababa University National Herbarium. Environmental variables such as elevation, slope, aspect, and geographical location of each plot were recorded. The elevation was measured by using a GPS Garmin, whereas the slope and aspect were measured by using a Sunto clinometer.

Estimation of aboveground biomass carbon

Recent allometric equation developed by Chave et al. (2014) can apply in all ecosystems because this equation encompasses important parameters like diameter, height, and wood density. Thus, an allometric equation developed by Chave et al. (2014) was used to estimate the aboveground biomass as given below:

$$ \mathrm{AGB}\ \left(\mathrm{kg}\right)=0.0673\times {\left({\uprho \mathrm{D}}^2\mathrm{H}\right)}^{0.976} $$

where AGB is the aboveground biomass of trees (kg), ρ is the basic wood density (g cm−3), D is the stem diameter at breast height (cm), and H is the height of trees (m).

Aboveground biomass carbon determined by assuming 47% of AGB comprises carbon (IPCC 2006). The total AGB carbon for each plot is calculated by summing up AGB carbon for all species. Carbon stocks were determined for each plot and then extrapolated to tonnes per hectare (Table S2).

Functional diversity and dominance metrics

Taxonomic diversity was used to measure diversity at each plot. Species richness was used to characterize the taxonomic diversity (Magurran 2004). Species richness at plot level refers to the number of different species counted in each plot. To quantify functional diversity and functional dominance, three functional traits that apply to the ecosystem function of interest (i.e., biomass and carbon storage) were considered, because carbon storage strongly depends on wood and foliage structures, in this study, therefore, traits such as wood density (WD), specific leaf area (SLA), and maximum plant height (PHm) were considered. We extracted data on wood density from the Global Wood Density Database (Zanne et al. 2009). In case multiple values were available for a single species, the average wood density was used. Plant functional trait, SLA, and PHm were extracted from the TRY database (Kattge et al. 2011; Maire et al. 2015). When the data for a species were missing, the average genus of wood density and SLA were used.

We estimated functional diversity metrics, functional richness (FRic), functional evenness (FEve), functional divergence (FDiv), and functional dispersion (FDis) (Cavanaugh et al. 2014; Zhang et al. 2014; Finegan et al. 2015; Villéger et al. 2008) for each individual plot (Table S3), using the values of the functional traits by “FDiversity” software (Casanoves et al. 2011). FRic represents the amount of functional space filled by the community (indicates resource use); FEve describes how evenly traits are distributed in functional trait space (indicates effective resource utilization); FDiv relates to how abundance is distributed within the volume of functional trait space occupied by species (i.e., niche differentiation) (Mason et al. 2005; Villéger et al. 2008; Mouchet et al. 2010). Functional dispersion (FDis) stated as the weighted mean distance in the multidimensional trait space of individual species to the weighted centroid of all species (niche complementarity) (Laliberte and Legendre 2010). These diversity indices are multi-trait-based functional diversity metrics that include both the relative weight of each species and the pair-wise functional difference between species.

Functional dominance (single functional trait index) was quantified by estimating the plot level community weight mean (CWM) for each functional trait. CWM is the mean of each species trait value weighted by the relative abundance of that species (Cavanaugh et al. 2014). The CWM for WD, SLA, and PHm estimated at each plot (Table S3), again using the FDiversity software (Casanoves et al. 2010). Multi-trait functional diversity indices were computed after standardization of trait values, while a single functional trait index was calculated without standardization. Standardization was performed due to observed difference in several orders of studied traits magnitude and scale of measurement (Pla et al. 2012).

Statistical analysis

Structural equation modeling

Structural equation modeling (SEM) is a powerful, multivariate statistical model found increasingly in ecological studies to test and evaluate multivariate causal relationships. SEM differs from other modeling techniques as it tests the direct and indirect effects on presupposed causal relationships (Fan et al. 2016). This is important, as we hypothesized that effects of diversity could be mediated through functional diversity and functional dominance. The original motivation of this hypothesis is because of the increased rate of biodiversity loss in the past few decades. To detect the effect of species loss on ecosystem functioning, we need to link measures of biodiversity components to ecosystem functions, such as aboveground biomass storages. Therefore, we tested the indirect and direct effects of diversity (species richness) on aboveground carbon storages. Two separate structural equation models specified representing (a) full mediation: postulating that effects of diversity on AGC are fully mediated through functional diversity and dominance metrics; and (b) partial mediation: stating that there are both direct and indirect diversity effects through functional diversity and functional dominance metrics on aboveground carbon. Because of multiple measures of functional diversity and functional dominance, stepwise selection techniques were used to select the most relevant functional diversity and functional dominance metrics for the aboveground carbon data.

The overall fit of the SEMs model evaluated using the goodness-of-fit index (GFI), a value that should be close to 0.95 or higher (Hu and Bentler 1999), Chi-square (χ2), and standardized root mean residual (SRMR ≤ 0.08). The standardized coefficients were used to make direct comparisons across paths (Grace and Bollen 2005). SEMs fitted in the R statistical software package, using the “sem” function in “lavaan” package, and later, output graphs visualized by using “semPlot” package again in R.

Linear mixed-effects models

Linear mixed-effects models (LMMs) are an important part of statistical models that can be used to analyze correlated data (Galecki and Burzykowski 2013). Before the mixed-effects modeling, the significant environmental factors and species richness effects on aboveground carbon stock were tested. We expect environmental factors to have effects on plant composition, growth, and survival (Mensah et al. 2016; Zhang et al. 2014); for this reason, standing aboveground biomass and carbon stocks could be affected by environmental factors. Here, the focus given to the factors that are determinant and quantifiable in the area, which is topography (i.e., elevation, slope, and aspect) (Geldenhuys 2002). The effects of elevation, slope, and aspect on aboveground carbon stock were tested using simple linear models in R. In case, these variables showed no significant effect on the carbon stock and further analyses performed and multiple linear regressions were used to test their effects on aboveground carbon storage. For both simple and multiple linear models, Shapiro-Wilk tests (W = 0.985, p value = 0.739) were used to check for the normality of the square root transformed aboveground carbon data and of the residuals. Further, Breusch-Pagan tests, value inflation factor (VIF), and Durbin-Watson statistics were used to test for heteroscedasticity, multicollinearity, and autocorrelation between residuals, respectively.

The effect of biodiversity components (i.e., functional diversity and functional dominance) on carbon storage was evaluated by fitting separate linear mixed-effects models (Zuur et al. 2009). Elevation is considered as random factors and each measure of functional diversity (i.e., FRic, FEve, FDiv, and FDis) and functional dominance (i.e., CWM of WD, SLA, and PHm) as fixed effects. Next, mixed-effects models fitted to evaluate the individual effect of each functional diversity and functional dominance metric; the combined effects of functional diversity metrics; the combined effects of functional dominance metrics; and the combined effects of functional diversity and functional dominance metrics. Further, the effect of each diversity component (fixed effects) on aboveground carbon storage was determined using a mixed-effects model (Bates et al. 2015) with “Type III” analysis of variance with Satterthwaite’s method. To run a mixed-effects model, package “lme4” (Bates et al. 2015) was used in R and variables selected by “backward selection” using “cAIC4” package (Saefken et al. 2018) in R. The significant effect of fixed factors was determined using the “lmer” function of the “lmerTest” package (Kuznetsova et al. 2017) also in R. The significance of the random effects determined using likelihood ratio (LR) test, again in the package “lmerTest”. We evaluated the performance of fitted models based on the fit statistics such as the Akaike information criterion (AIC) (Akaike 1974) and the marginal R2, which shows the proportion of variance explained by fixed effects (Nakagawa and Schielzeth 2013).

Results

Diversity effects on AGC mediated through functional diversity and functional dominance

The results of structural equation models fitted to test the mediated effects of diversity (species richness) on AGC, through functional diversity and functional dominance summarized in Table 1 and Fig. 3a. The chi-square value for the “full mediation” model was 24.63 with 14 degrees of freedom and a p value of 0.026, and the standardized root-mean-square residual was 0.07, showing a slightly good fit.

Table 1 Results of the structural equation modeling done to test the effects of species richness on AGC through functional diversity and functional dominance
Fig. 3
figure 3

The best-fit structural equation models (SEMs) for linking species richness, functional diversity, functional dominance, and AGC storage. a Full mediation, b partial mediation. The figures on the lines are the standardized path coefficients. The figures in the box represent the coefficient of determination (R2). Red arrows denote negative effect paths. Black arrows denote positive effect paths. Double-pointed lines represent the residual correlations. We present the significance of each path and model fit statistics in Table 1. AGC, aboveground biomass carbon; FRic, functional richness; FEve, functional evenness; FDis, functional dispersion; FDiv, functional divergence; CWM, community weight mean; WD, wood density; SLA, specific leaf area; PHm, maximum plant height

In the “full mediation” model, species richness revealed a significant positive direct effect on the functional richness (R2 = 0.580; ß = 0.76; p < 0.001; Table 1), but showed no significant positive effect on the AGC (ß = 0.17; p = 0.374; Table 1). There was also a significant positive effect of species richness on functional evenness (R2 = 0.194; ß = −0.44; p < 0.001; Table 1); the latter, however, exhibited a non-significant positive effect on the AGC. Further, species richness showed a positive significant effect on functional dispersion (R2 = 0.157; ß = 0.40; p < 0.001; Table 1), however, non-significant positive effect on AGC was found. There were positive and non-significant effects of species richness on functional divergence (ß = 0.12; P = 0.356; Table 1), the latter of which indicates the negative and slightly significant effect on the AGC (ß = −0.23; p = 0.084; Table 1) at 0.1, which would suggest that the mediated effects of species richness were through functional divergence only. In addition, we found a non-significant correlation between functional richness and functional evenness (ß = 0.29; p = 0.090; Table 1), which would suggest that both functional richness and functional evenness transmitted the mediated effects of species richness.

Out of the three functional dominance metrics, the CWM of SLA and CWM of wood density did not retain any significant path. Only the CWM of maximum plant height showed significant responses to species richness (R2 = 0.07; ß = 0.26; p = 0.040; Table 1), but showed negative and non-significant effect on the AGC (ß = −0.17; p = 0.612; Table 1). The significant residual correlation of CWM of maximum plant height with CWM of wood density (ß = 0.45; p = 0.003); and CWM of SLA (ß = 0.45; p = 0.003; Table 1) confirms that the mediated effects of species richness transmitted through CWM of maximum plant height only.

The partial mediation model fitted by only adding a direct path from species richness to AGC to the “full mediation” model. The second model (partial mediation) had a chi-square value of 19.58 (DF = 12; p = 0.076, Fig. 3b and Table 1), which indicates a good fit and absence of significant deviations between data and the model. There are similarities between the two models in terms of significant and non-significant paths (Table 1). However, the “partial mediation” model showed slightly better fits (GFI = 0.95; R2 = 0.23; SRMR = 0.071) than the “full mediation” model (GFI = 0.93; R2 = 0.14; SRMR = 0.076). We found the significant influence of the added causal path from species richness to AGC, suggesting an existing true direct effect of diversity on the AGC. Both models recommend that the effects of species richness on aboveground carbon mediated through functional diversity and functional dominance, which support both the complementarity hypothesis and selection effect hypothesis.

Effect of environmental variables on aboveground carbon stocks

The effects of environmental factors, notably the slope, were significant, explaining 17.41% of the variation in aboveground carbon (Table 2). However, elevation and aspect showed a non-significant effect on the aboveground carbon (p = 0.168 and 0.701, respectively; Table 2). When accounting for the effects of the elevation, a significant effect of species richness on the aboveground carbon found (ß = 0.78; p = 0.034; Table 2).

Table 2 The effect of elevation, slope, aspect, and species richness on AGC stock

Effects of functional diversity and functional dominance on aboveground carbon (AGC) stock

The separate linear mixed effects models tested the individual effects of functional diversity metrics showed that FRic (ß = 1.02, p = 0.042) and FDis (ß = 16.71, p < 0.001) had a significant positive effect on the AGC (Table 3). Both functional divergence (ß = −10.64, p = 0.485) and functional evenness (ß = 3.82, p = 0.524) had non-significant negative and positive effects on the AGC (Table 3). Regarding the combined effects of functional diversity metrics, we found the significant effects of FDis and FDiv after backward selection for the last model (see Table 3). A significant positive effect of functional dispersion (ß = 21.36; p < 0.001; Table 3) and a significant negative effect of functional divergence therefore showed the effects of functional diversity on AGC (ß = −39.18; p = 0.003; Table 3). Both functional dispersion and functional divergence explained 29% of the variance of AGC.

Table 3 The effect of functional diversity on AGC stock

Out of the three functional dominance metrics used in this study, only CWM of WD showed negative significant effects on the AGC (Table 4). However, CWM of SLA (ß = −0.09; p = 0.250) and CWM of PHm (ß = 0.28; p = 0.124) revealed negative and positive non-significant effects on the AGC, respectively (Table 4). When testing their combined effects on AGC, CWM of WD is not kept by the final model, and the effects of functional dominance were only shown by positive and significant effects of CWM of maximum plant height and negative non-significant effect of CWM of SLA, with 10.1% explained variance (Table 4).

Table 4 The effect of functional dominance on AGC stock

The output of mixed-effects models of functional diversity and functional dominance showed that the marginal R2 (variance explained by fixed factors) in the diversity and AGC relationship was greater for functional diversity (29%) than for functional dominance (10.1%) (Tables 3 and 4), indicating a variation of AGC greater for functional diversity than combined model effects. While analyzing the combined effect of functional diversity and functional dominance on AGC, the result revealed that 22.4% of the deviations of AGC were explained by significant effects of functional dispersion and CWM of wood density (Table 5).

Table 5 The combined effects of functional diversity and functional dominance on AGC stock

Discussion

Species diversity predicts aboveground carbon stocks

In this study, we found significant and positive effect of species richness on aboveground carbon after the effects of environmental drivers (e.g., elevation) accounted for. While this finding is in line with some previous studies that controlled for the effects of environmental factors (Ruiz-Jaen and Potvin 2010; Wu et al. 2015; Mensah et al. 2016), it also supports the positive relationships in diverse natural forests; biomass and carbon stocks increase with increasing species diversity. Various local and global studies on forest ecosystems have observed a positive relationship between species richness and forest biomass or carbon (Cavanaugh et al. 2014; Ruiz-Benito et al. 2014; Wu et al. 2015; Mensah et al. 2016). In addition, studies in subtropical forests (Vance-Chalcraft et al. 2010), spruce dominated forest stands (Wang et al. 2011), collaborative forests in Terai, Nepal (Mandal et al. 2013), and tropical forests (Barrufol et al. 2013) have also reported increases in biomass productivity with increasing diversity.

Increasing species richness could be increase carbon storage potential of forest ecosystems (Ruiz-Jaen and Potvin 2010; Ruiz-Benito et al. 2014). Some other recent evidence showed that a positive effect of species richness on aboveground carbon might be explained through the biotic interactions such as facilitation, where some species could enhance soil fertility (e.g., by fixing nitrogen) for the productivity of other species. This is often used to support the reason that mixed species communities of plantations are more productive than mono-specific stands (Mensah et al. 2016), although maybe increasing species richness increases the probability of inclusion of highly productive and naturally dominant species (Ruiz-Benito et al. 2014; Mensah et al. 2018).

While our result in the dry evergreen Afromontane forest supports the positive species richness-aboveground carbon relationship, it is well known that finding of the inverse effect also exists. For example, a report by Ruiz-Jaen and Potvin (2011) from the natural forest of Barro Colorado Island in Central Panama. Other studies have reported null or negative relationships between aboveground carbon and species diversity in forest ecosystems (Zhang et al. 2014; Whittaker and Heegaard 2003). These discrepancies in findings suggest that the effects of diversity on forest carbon may vary with other factors such as the types, sites, and the succession stages of the forests (Lasky et al. 2014; Wu et al. 2015), and also the specific dimension of the diversity measure used (Vance-Chalcraft et al. 2010; Lasky et al. 2014).

Effects of environmental factors on tree aboveground carbon stock

In this study, we did not find a significant effect of elevation on aboveground carbon stock. In line with our finding, Cavanaugh et al. (2014) and Mensah et al. (2016) reported a non-significant relationship between forest carbon and elevation. However, this finding runs contrary to other previous studies that tested the relationships between elevation and biomass or carbon stock (Ensslin et al. 2015). On the one hand, some scholars found that biomass and carbon stocks could decrease with increasing elevation (Moser et al. 2007). Some previous studies observed a positive correlation between increasing tree carbon and increasing elevation (Zhang et al. 2018; Zhu et al. 2010). This lack of clarity on the relationship between elevation and forest biomass may be partly due to the variation in the elevation range among studies.

Unlike elevation and aspect, slope showed a significant effect, and accounted for 17.41% of aboveground carbon variance, evidencing that variations in carbon stocks can result from topological constraints, particularly differences in slope. We have identified the slope as an important environmental gradient that affects carbon stocks (Chave et al. 2003). This is because they inherently related aboveground carbon to wood and biomass; we can see production and the effect of slope as prior effects of environment on availability of resources (Luizao et al. 2004). For instance, a steeper slope will speed up nutrient and water runoff, as well as favor highly water and nutrient efficient species against others. Taking this into account, tree growth and biomass production might decline in higher slope areas, as results of moisture and nutrient stress (Clark et al. 2010; Durán et al. 2015), whereas flat and gentle slope areas would allow for more water availability, to which plant would be likely to respond positively. The significant effect of slope supports the fact that ecosystem functions and biomass carbon stock are environment-structured (Wu et al. 2015).

Effects of diversity on AGC mediated through functional diversity and functional dominance

In recent decades, interest has increased significantly in determining the multiple biodiversity measures and ecosystem functioning relationships. One of the most examined relationships is that between species richness and productivity (Ruiz-Jaen and Potvin 2010). Using different measures of biodiversity component to provide additional mechanisms motivating the effects of biodiversity on carbon stocks has also gained an increasing interest in recent years (Cadotte et al. 2011; Conti and Díaz 2013; Finegan et al. 2015; Lasky et al. 2014; Ruiz-Benito et al. 2014; Vance-Chalcraft et al. 2010; Ziter et al. 2013). Most of these studies compared the relative effects of species richness and other biodiversity components on aboveground carbon stocks, by assuming that it mediated these effects through functional diversity and functional dominance. We also examined effects of functional diversity and dominance metrics. In this study, the structural equation modeling results confirm this assumption and this, therefore, supports the need to explore beyond species richness to explain the mechanisms that drive the relationship between diversity and productivity. The results further support the hypothesis that neither niche complementarity nor selection effect exclusively affects carbon storage in tropical forests (Cavanaugh et al. 2014; Wu et al. 2017). Therefore, diversity (species richness) predicts carbon stock through effects of functional diversity and functional dominance, because it based these diversity components on specific functional traits, which would reflect functional variations among the species (Díaz and Cabido 2001; Mensah et al. 2016). Indeed, in this study, increased species richness indirectly accounted for differences among species, in terms of the ecological niche and resource use.

Functional diversity effects on aboveground carbon stock

The linear mixed models were used to examine the effects of functional diversity measures on tree aboveground carbon in the Dindin dry evergreen Afromontane forest of Hararghe highland, Southeast Ethiopia. Functional diversity dissected into four relatively independent components: functional richness FRic, evenness FEve, divergence FDiv (Villeger et al. 2008; Mouchet et al. 2010), and dispersion FDis (Laliberte and Legendre 2010). These indices quantify the trait hyper-volume of the community (FRic, FDis) and the distribution of abundance or biomass of the species in this volume (FEve, FDiv). Both these sets of functional characteristics may measure niche complementarity, and therefore, ecosystem processes by functional trait variety. Out of the four functional diversity metrics used in this study, only functional richness and functional dispersion were found to significantly explain the variation in aboveground carbon stock. Some previous evidences for functional diversity effects on aboveground biomass and carbon are consistent with our results. For instance, Yuan et al. (2018) observed a significant effect of functional dispersion on aboveground carbon in temperate mixed forests. Similarly, in tropical rain forests of Bolivia, Brazil, and Costa Rica, Finegan et al. (2015) found functional richness among other functional diversity indices as a significant predictor of aboveground biomass. A study by Ziter et al. (2013) in unmanaged forest fragments in Quebec showed significant and positive relationships between functional dispersion and aboveground carbon. Another study by Finegan et al. (2014) reported significant but negative effects of the functional richness on stand biomass in tropical forests. While in this study, these functional diversity indices have their specific biological meaning, and the positive effect of functional richness on the aboveground carbon might be because of functional richness being positively correlated with species richness (refer SEM output).

The functional trait diversity would increase carbon stock because species with different traits would differ in resource use, and would efficiently use the resources available within the community for higher growth and productivity, showing the importance of niche complementarity effects in facilitating ecosystem processes (Finegan et al. 2015). Unlike functional richness and functional dispersion, functional evenness and functional divergence did not show any relationship with species richness and showed no significant influence on aboveground carbon. According to Laliberte and Legendre (2010), the functional dispersion is the mean distance of each species weighted by its relative abundances to the centroid of all species in a community. Functional dispersion relates to the niche space or volumes of niche space, and functional diversity as measured here could reflect a “niche differentiation” (Carroll et al. 2011). Niche differentiation expressed as the beta niche of species, i.e., differentiating species with different optima between communities across environmental gradients (Silvertown et al. 2006). A greater functional diversity, that is greater value and range of functional traits, would reflect not only the magnitude of “niche differentiation” but also the differences in resource utilization by species, thus promoting diversity effects on ecosystem functioning. This is consistent with Zhu et al. (2016) who observed increasing niche difference contributes to species coexistence and positive diversity effects on biomass productivity.

Functional dominance effects on aboveground carbon stock

Community-weighted mean (CWM) functional trait values, which are community trait values weighted (selection effect) by species abundances (Muscarella and Uriarte 2016; Thakur and Chawla 2019) used to reflect locally optimal strategies of a community and to predict functional dominance effects. CWM functional traits as functional dominance metrics could explain ecological fitness by resource competition ability and environmental filtering (Cornwell and Ackerly 2009; Kraft et al. 2015). Therefore, functional dominance could indicate some aspect of “relative performance or fitness differences” between competitors for limiting resources (Cadotte et al. 2011; Carroll et al. 2011). The dominant functional traits (high wood density and low specific leaf area) in the stressful areas indicate a stress-tolerant life history strategy (Chapman and McEwan 2018). The result showed that functional dominance had a significant effect on tree carbon storage, which indicates the magnitude of relative fitness differences strengthens the influence of diversity on biomass productivity (Carroll et al. 2011). In this study, the functional dominance effects varied with the functional traits. Functional traits are those attributes of an organism or parts of an organism that strongly influence fitness through their effects on the overall structure, function, and diversity of ecosystems (Wieczynski et al. 2019). For instance, CWM of wood density showed a negative and significant effect on aboveground carbon stocks. It is not surprising given that wood density is a potential predictor of tree biomass, which highly correlates with the carbon stock. There is some evidence that CWM of wood density is negatively related to the biomass increment, as being a good predictor of individual tree diameter increments (Finegan et al. 2015). After evaluating biomass stocks in tropical forests, some scholars found that the AGC-wood density relationship varies from negative to null to positive depending on the forest community and forest identity (Baker et al. 2004; Stegen et al. 2009). The finding of the CWM of wood density in this study shows that low wood density species grow faster and are expected to store more biomass. We recommend that conserving and planting low wood density species would likely help to increase the carbon stock.

In the analysis of combined effects of functional dominance metrics, they kept only CWM of specific leaf area and maximum plant height in the last model, with maximum plant height being the significant predictor. This is most likely tree height is a key factor for species-specific or multi-species biomass regressions. Ali et al. (2016) suggests that strong dominance by tall and conservative species, rather than a set of coexisting species with diverse heights and exploitative nature, results in the greatest carbon stock in natural forest ecosystems. Therefore, the positive and significant relationship between CWM of maximum plant height and carbon stocks indicates the potential importance of characteristics of dominant and adult trees for ecosystem functioning and productivity, thus supporting the selection effects hypothesis.

Functional diversity and functional dominance effects aboveground carbon stocks

Functional diversity explained greater variance in aboveground carbon than functional dominance when the amounts of variance explained by the selection effect and niche complementary mechanisms were tested (Tables 3 and 4). Unfortunately, this rejects the second hypothesis of our study and suggests that niche complementarity effects appear to be more important than selection effects. This finding, which is consistent with Mensah et al. (2016) and Sintayehu et al. (2020), supports the notion that functional diversity explains greater variation in aboveground carbon than selection effects. Finegan et al. (2015) and Ruiz-Jaen and Potvin (2011) found that selection effects are more relevant in tropical forests for biomass and carbon storage prediction. The reason for this discrepancy is that in this study, functional dominance metrics (community-weighted mean of a functional trait) were estimated using species relative abundance, but Ruiz-Jaen and Potvin (2011) and Finegan et al. (2015) used species relative basal area and species relative biomass, respectively, as weighting variable in their studies. As noted by Mensah et al. (2016), the strength of the relationship between community-weighted means of functional traits and the ecosystem function of interest could depend on the weighting variables. Therefore, community-weighted mean values of functional trait weighted by biomass or basal area as weighting variables would likely show stronger relation with biomass and carbon than abundance-based community mean values.

In assessing the combined effects of niche and selection (see Table 5), this result supports the assumption that these two hypotheses are not exclusive and may contribute to the functioning of the ecosystem. Previous evidence of complementarity and selection effects on ecosystem function implies that they can contribute to different proportions of ecosystem functions (Fargione et al. 2007). Both complementarity and selection effects promote species coexistence. As reason out by Mensah et al. (2016), these two hypotheses could even be the outcome of interactions of the relative fitness differences and the niche differences, whereby dominant competitors could suppress some species’ populations, to allow effective utilization of the resources. In this study, they mediated the selection effects through specific maximum plant height, which indicates the effect of dominant species and suggests a possible competitive exclusion in terms of utilization of resources (refer to SEM, e.g., sunlight).

Conclusions

This study examined the relationship between diversity and carbon stock in dry evergreen Afromontane forest of Hararghe highland, Southeast Ethiopia, and showed that taxonomic diversity (species richness) predicts carbon storage through functional diversity and functional dominance. Further, the study noted that both the niche complementarity and selection effect hypotheses are important predictors of carbon stock. Carbon stock variations explained by functional diversity (niche complementarity effects) were greater than by functional dominance effects (selection effects). Functional dominance effects strongly transmitted through the CWM of maximum plant height, showing the importance of forest vertical stratification for diversity and carbon relationship. Therefore, complementary effects were induced by complementary light use efficiency of species and trees growing in the understory layer. This study recommends that future research focuses on the relationship between diversity and forest carbon oriented toward a perspective of the forest canopy (or dominant species vs. other species), to contribute additional insights into our understanding of biodiversity and ecosystem function relationship.

Availability of data and materials

The datasets used and/or analyzed during the current study will be available online.

Change history

Abbreviations

AGC:

Aboveground carbon

AGB:

Aboveground biomass

AIC:

Akaike’s information criterion

a.s.l.:

Above sea level

cAIC4:

Conditional Akaike’s information criterion

BP:

Breusch-Pagan tests

CWM:

Community weight mean

dbh:

Diameter at breast height

DW:

Durbin-Watson statistics

FRic:

Functional richness

FEve:

Functional evenness

FDiv:

Functional divergence

FDis:

Functional dispersion

GFI:

Goodness-of-fit index

GPS:

Global positioning system

LMMs:

Linear mixed-effects models

LR:

Likelihood ratio

MAP:

Mean annual precipitation

MAT:

Mean annual temperature

PHm:

Maximum plant height

S :

Species richness

SEM:

Structural equation modeling

SLA:

Specific leaf area

SRMR:

Standardized root mean residual

W :

Shapiro-Wilk tests

WD:

Wood density

VIF:

Value inflation factor

ρ :

Wood density

χ 2 :

Chi-square

References

  • Akaike H (1974) A new look at the statistical model identification. IEEE Trans Automat Contr 19(6):716–723. https://doi.org/10.1109/TAC.1974.1100705

    Article  Google Scholar 

  • Aynekulu E, Aerts R, Moonen P, Denich M, Gebrehiwot K, Vågen T-G, Mekuria W, Boehmer HJ (2012) Altitudinal variation and conservation priorities of vegetation along the Great Rift Valley escarpment, northern Ethiopia. Biodivers Conserv 21(10):2691–2707. https://doi.org/10.1007/s10531-012-0328-9

  • Ali A, Yan E-R, Chen HYH, Chang SX, Zhao Y-T, Yang X-D, Xu M-S (2016) Stand structural diversity rather than species diversity enhances aboveground carbon storage in secondary subtropical forests in Eastern China. Biogeosciences 13(16):4627–4635. https://doi.org/10.5194/bg-13-4627

    Article  CAS  Google Scholar 

  • Arellano G, Cala V, Fuentes A, Cayola L, Jørgensen PM, Macía MJ (2016) A standard protocol for woody plant inventories and soil characterization using temporary 0.1-ha plots in tropical forests. J Trop For Sci 28(4):508–516

  • Baker TR, Phillips OL, Malhi Y, Almeida S, Arroyo L, di Fiore A, Erwin T, Killeen TJ, Laurance SG, Laurance WF, Lewis SL, Lloyd J, Monteagudo A, Neill DA, Patiño S, Pitman NCA, M. Silva JN, Vásquez Martínez R (2004) Variation in wood density determines spatial patterns in Amazonian forest biomass. Glob Chang Biol 10(5):545–562. https://doi.org/10.1111/j.1365-2486.2004.00751.x

    Article  Google Scholar 

  • Baraloto C, Hardy OJ, Paine CET, Dexter KG, Cruaud C, Dunning LT, Gonzalez M‐A, Molino J‐F, Daniel Sabatier, Savolainen V, Chave J (2012) Using functional traits and phylogenetic trees to examine the assembly of tropical tree communities. J Ecol 100(3):690–701

  • Barrufol M, Schmid B, Bruelheide H, Chi XL, Hector A, Ma KP, Michalski S, Tang ZY, Niklaus PA (2013) Biodiversity promotes tree growth during succession in subtropical forest. PLoS ONE 8(12):e81246

  • Bates D, Mächler M, Bolker B, Walker S (2015) Fitting linear mixed-effects models using lme4. J Stat Softw 67(1):1–48. https://doi.org/10.18637/jss.v067.i01

  • Berhanu A, Woldu Z, Demissew S (2016) Elevation patterns of woody taxa richness in the evergreen Afromontane vegetation of Ethiopia. J For Res 28(4):787–793. https://doi.org/10.1007/s11676-016-0350-y

  • Cadotte MW, Carscadden K, Mirotchnick N (2011) Beyond species: functional diversity and the maintenance of ecological processes and services. J Appl Ecol 48(5):1079–1087. https://doi.org/10.1111/j.1365-2664.2011.02048.x

  • Carroll IT, Cardinale B, Nisbet RM (2011) Niche and fitness differences relate the maintenance of diversity to ecosystem function. Ecology 92(5):1157–1165. https://doi.org/10.1890/10-0302.1

  • Casanoves F, Pla L, Di Rienzo JA, Di’az S (2010) FDiversity: a software package for the integrated analysis of functional diversity. Methods Ecol Evol 2:233–237. https://doi.org/10.1111/j.2041-210X.2010.00082.x

    Article  Google Scholar 

  • Casanoves F, Pla L, Di Rienzo JA, Díaz S (2011) FDiversity: a software package for the integrated analysis of functional diversity. Methods Ecol Evol 2:233–237. https://doi.org/10.1111/j.2041-210X.2010.00082.x

  • Cavanaugh KC, Gosnell JS, Davis SL, Ahumada J, Boundja P, Clark DB, Andelman S (2014) Carbon storage in tropical forests correlates with taxonomic diversity and functional dominance on a global scale. Glob Ecol Biogeogr 23(5):563–573. https://doi.org/10.1111/geb.12143

    Article  Google Scholar 

  • Chapman JI, McEwan RW (2018) The role of environmental filtering in structuring Appalachian tree communities: topographic influences on functional diversity are mediated through soil characteristics. Forests 9(1):19. https://doi.org/10.3390/f9010019

  • Chave J, Condit R, Lao S, Caspersen JP, Foster RB, Hubbell SP (2003) Spatial and temporal variation of biomass in a tropical forest: results from a large census plot in Panama. J Ecol 91(2):240–252. https://doi.org/10.1046/j.1365-2745.2003.00757.x

    Article  Google Scholar 

  • Chave J, Goodman RC, Henry M, Martínez-Yrízar A, Mugasha WA, Muller-Landau HC (2014) Improved allometric models to estimate the aboveground biomass of tropical trees. Glob Chang Biol 20(10):3177–3190. https://doi.org/10.1111/gcb.12629

    Article  Google Scholar 

  • Clark CM, Flynn DFB, Butterfield BJ, Reich PB (2012) Testing the link between functional diversity and ecosystem functioning in a minnesota grassland experiment. PLoS ONE 7(12):e52821. https://doi.org/10.1371/journal.pone.0052821

  • Clark DB, Clark DA, Oberbauer SF (2010) Annual wood production in a tropical rain forest in NE Costa Rica linked to climatic variation but not to increasing CO2. Glob Chang Biol 16(2):747–759. https://doi.org/10.1111/j.1365-2486.2009.02004.x

  • Conti G, Díaz S (2013) Plant functional diversity and carbon storage - an empirical test in semi-arid forest ecosystems. J Ecol 101(1):18–28. https://doi.org/10.1111/1365-2745.12012

  • Cornwell WK, Ackerly DD (2009) Community assembly and shifts in plant trait distributions across an environmental gradient in coastal California. Ecol Monogr 79(1):109–126. https://doi.org/10.1890/07-1134.1

    Article  Google Scholar 

  • Díaz S, Cabido M (2001) Vive la différence: plant functional diversity matters to ecosystem processes. Trends Ecol Evol 16(11):646–655

  • Durán SM, Sánchez-Azofeifa GA, Rios RS, Gianoli E (2015) The relative importance of climate, stand variables and liana abundance for carbon storage in tropical forests. Glob Ecol Biogeogr 24(8):939–949. https://doi.org/10.1111/geb.12304

    Article  Google Scholar 

  • Elias E (2016) Soils of the Ethiopian highlands: geomormology and properties. CASCAPE Project, ALTERA, Wageningen University and Research Centre (Wageningen UR), The Netherlands

    Google Scholar 

  • Ensslin A, Rutten G, Pommer U, Zimmermann R, Hemp A, Fischer M (2015) Effects of elevation and land use on the biomass of trees, shrubs, and herbs at Mount Kilimanjaro. Ecosphere 6:45

    Article  Google Scholar 

  • Fan Y, Chen J, Shirkey G, John R, Wu SR, Park H, Shao C (2016) Applications of structural equation modeling (SEM) in ecological studies: an updated review. Ecol Process 5:19. https://doi.org/10.1186/s13717-016-0063-3

  • FAO (2005) New_LocClim: local climate estimator (version 1.10). Environment and natural resources, working paper number 20 (CD-ROM). FAO, Rome

    Google Scholar 

  • Fargione J, Tilman D, Dybzinski R, Lambers JHR, Clark C, Harpole WS et al (2007) From selection to complementarity: shifts in the causes of biodiversity-productivity relationships in a long-term biodiversity experiment. Proc R Soc B: Biol Sci 274(1611):871–876.  https://doi.org/10.1098/rspb.2006.0351

  • Finegan B, Peña-Claros M, de Oliveira A, Ascarrunz N, Bret-Harte MS, Carreño-Rocabado G, Casanoves F, Díaz S, Eguiguren Velepucha P, Fernandez F, Licona JC, Lorenzo L, Salgado Negret B, Vaz M, Poorter L (2014) Does functional trait diversity predict above-ground biomass and productivity of tropical forests? Testing three alternative hypotheses. J Ecol 103(1):191–201. https://doi.org/10.1111/1365-2745.12346

    Article  Google Scholar 

  • Finegan B, Peña-Claros M, de Oliveira A, Ascarrunz N, Bret-Harte MS, Carreño-Rocabado G, Poorter L (2015) Does functional trait diversity predict above- ground biomass and productivity of tropical forests? Testing three alternative hypotheses. J Ecol 103:191–201

  • Friis I, Demissew S, van Breugel P (2010) Atlas of the potential vegetation of Ethiopia, Biologiske Skrifter, vol 58. The Royal Danish Academy of Science and Letters, Copenhagen

    Google Scholar 

  • Galecki A, Burzykowski T (2013) Linear mixed-effects models using R: a step-by-step approach. Springer Texts Statist. https://doi.org/10.1007/978-1-4614-3900-4

  • Geldenhuys CJ (2002) Tropical secondary forest management in Africa: reality and perspectives. Food and Agriculture Organization, Rome

    Google Scholar 

  • Grace BJ, Bollen KA (2005) Interpreting the results from multiple regression and structural equation models. Bull Ecol Soc Am 86:283–295

  • Grieser J, Gommes R, Bernardi M (2006) New LocClim - the local climate estimator of FAO. Geophys Res Abstr 8:2

  • Grime JP (1998) Benefits of plant diversity to ecosystems: immediate, filter, and founder effects. J Ecol 86(6):902–910. https://doi.org/10.1046/j.1365-2745.1998.00306.x

    Article  Google Scholar 

  • Hu L, Bentler PM (1999) Cutoff criteria for fit indexes in covariances structure analysis: convention criteria versus new alternatives. Struct Equation Model Multidisciplinary J 6(1):1–55. https://doi.org/10.1080/10705519909540118

    Article  Google Scholar 

  • IPCC (2006) In: Eggleston HS, Buendia L, Miwa K, Ngara T, Tanabe K (eds) Guidelines for national greenhouse gas inventory prepared by national greenhouse gas inventories programme. IGES, Japan

    Google Scholar 

  • Jerzy S, Anna G (2007) Above-ground standing biomass and tree species diversity in natural stands of Central Europe. J Veg Sci 18:563–570 https://doi.org/10.1111/j.1654-1103.2007.tb02569.x

    Google Scholar 

  • Jiang Z, Ma K, Liu H, Tang Z (2018) A trait-based approach reveals the importance of biotic filter for elevational herb richness pattern. J Biogeogr 45:2288–2298. https://doi.org/10.1111/jbi.13398

  • Kattge J, Diaz S, Lavorel S, Prentice IC, Leadley P, Bönisch G et al (2011) TRY - a global database of plant traits. Glob Chang Biol 17(9):2905–2935. https://doi.org/10.1111/j.1365-2486.2011.02451.x

  • Kraft NJB, Adler PB, Godoy O, James EC, Fuller S, Levin JM (2015) Community phylogenetics and ecosystem functioning: community assembly, coexistence and the environmental filtering metaphor. Funct Ecol. 29(5):592–599. https://doi.org/10.1111/1365-2435.12345

    Article  Google Scholar 

  • Kuznetsova A, Brockhoff PB, Christensen RHB (2017) lmerTest Package: tests in linear mixed effects models. J Stat Softw 82(13):1–26. https://doi.org/10.18637/jss.v082.i13

    Article  Google Scholar 

  • Laliberte E, Legendre P (2010) A distance-based framework for measuring functional diversity from multiple traits. Ecology 91(1):299–305. https://doi.org/10.1890/08-2244.1

  • Lasky JR, Uriarte M, Boukili VK, Erickson DL, John Kress W, Chazdon RL (2014) The relationship between tree biodiversity and biomass dynamics changes with tropical forest succession. Ecol Lett 17(9):1158–1167. https://doi.org/10.1111/ele.12322

    Article  Google Scholar 

  • Li S, Lang X, Liu W, Ou G, Xu H, Su J (2018) The relationship between species richness and aboveground biomass in a primary Pinus kesiya forest of Yunnan, southwestern China. PLoS One 13(1):e0191140 https://doi.org/10.1371/journal.pone.0191140

  • Lin D, Anderson-Teixeira KJ, Lai J, Mi X, Ren H, Ma K (2016) Traits of dominant tree species predict local scale variation in forest aboveground and topsoil carbon stocks. Plant and Soil 409(1-2):435–446. https://doi.org/10.1007/s11104-016-2976-0

  • Luizao RCC, Luizao FJ, Paiva RQ, Monteiro TF, Sousa LS, Kruijt B (2004) Variation of carbon and nitrogen cycling processes along a topographic gradient in a central Amazonian forest. Glob Chang Biol 10(5):592–600. https://doi.org/10.1111/j.1529-8817.2003.00757.x

    Article  Google Scholar 

  • Magurran AE (2004) Measuring biological diversity. Blackwell Science Ltd, Australia

  • Maire V, Wright IJ, Prentice IC, Batjes NH, Bhaskar R, van Bodegom PM et al (2015) Global effects of soil and climate on leaf photosynthetic traits and rates. Glob Ecol Biogeogr 24(6):706–717. https://doi.org/10.1111/geb.12296

    Article  Google Scholar 

  • Mandal RA, Jha PK, Dutta IC, Karmacharya S (2013) Relationship between carbon stock and plant biodiversity in collaborative forests in Terai, Nepal. ISRN Bot:1–7. https://doi.org/10.1155/2013/625767

  • Mason NWH, Mouillot D, Lee WG, Wilson JB (2005) Functional richness, functional evenness and functional divergence: the primary components of functional diversity. Oikos 111(1):112–118. https://doi.org/10.1111/j.0030-1299.2005.13886.x

    Article  Google Scholar 

  • Mensah S, du Toit B, Seifert T (2018) Diversity–biomass relationship across forest layers: implications for niche complementarity and selection effects. Oecologia 187(3):783–795. https://doi.org/10.1007/s00442-018-4144-0

    Article  Google Scholar 

  • Mensah S, Veldtman R, Assogbadjo AE, Glèlè Kakaï R, Seifert T (2016) Tree species diversity promotes aboveground carbon storage through functional diversity and functional dominance. Ecol Evol 6(20):7546–7557. https://doi.org/10.1002/ece3.2525

  • Morin X, Fahse L, Scherer-Lorenzen M, Bugmann H (2011) Tree species richness promotes productivity in temperate forests through strong complementarity between species. Ecol Lett 14(12):1211–1219. https://doi.org/10.1111/j.1461-0248.2011.01691.x

    Article  Google Scholar 

  • Moser G, Hertel D, Leuschner C (2007) Altitudinal change in LAI and stand leaf biomass in tropical montane forests: a transect study in Ecuador and a pan-tropical meta-analysis. Ecosystems 10(6):924–935. https://doi.org/10.1007/s10021-007-9063-6

    Article  Google Scholar 

  • Mouchet MA, Villeger S, Norman WH, Mason NWH, Mouillot D (2010) Functional diversity measures: an overview of their redundancy and their ability to discriminate community assembly rules. Funct Ecol 24(4):867–876. https://doi.org/10.1111/j.1365-2435.2010.01695.x

    Article  Google Scholar 

  • Muscarella R, Uriarte M (2016) Do community-weighted mean functional traits reflect optimal strategies? Proc R Soc B Biol Sci 283(1827):20152434. https://doi.org/10.1098/rspb.2015.2434

  • Nakagawa S, Schielzeth H (2013) A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol Evol 4(2):133–142. https://doi.org/10.1111/j.2041-210x.2012.00261.x

  • Pla L, Casanoves F, Rienzo JD (2012) Quantifying functional biodiversity. Springer, Dordrecht Heidelberg, London, New York. https://doi.org/10.1007/978-94-007-2648-2

    Book  Google Scholar 

  • Poorter L, van der Sande MT, Thompson J, Arets EJMM, Alarcón A, Álvarez-Sánchez J, Ascarrunz N, Balvanera P, Barajas-Guzmán G, Boit A, Bongers F, Carvalho FA, Casanoves F, Cornejo-Tenorio G, Costa FRC, de Castilho CV, Duivenvoorden JF, Dutrieux LP, Enquist BJ, Fernández-Méndez F, Finegan B, Gormley LHL, Healey JR, Hoosbeek MR, Ibarra-Manríquez G, Junqueira AB, Levis C, Licona JC, Lisboa LS, Magnusson WE, Martínez-Ramos M, Martínez-Yrizar A, Martorano LG, Maskell LC, Mazzei L, Meave JA, Mora F, Muñoz R, Nytch C, Pansonato MP, Parr TW, Paz H, Pérez-García EA, Rentería LY, Rodríguez-Velazquez J, Rozendaal DMA, Ruschel AR, Sakschewski B, Salgado-Negret B, Schietti J, Simões M, Sinclair FL, Souza PF, Souza FC, Stropp J, ter Steege H, Swenson NG, Thonicke K, Toledo M, Uriarte M, van der Hout P, Walker P, Zamora N, Peña-Claros M (2015) Diversity enhances carbon storage in tropical forests. Glob Ecol Biogeogr 24(11):1314–1328. https://doi.org/10.1111/geb.12364

  • Prado-Junior JA, Schiavini I, Vale VS, Arantes CS, Van Der Sande MT, Lohbeck M, Poorter L (2016) Conservative species drive biomass productivity in tropical dry forests. J Ecol 104(3):817–827. https://doi.org/10.1111/1365-2745.12543

  • Reich PB, Knops J, Tilman D, Craine J, Ellsworth D, Tjoelker M, Lee T, Wedin D, Naeem S, Bahauddin D, Hendrey G, Jose S, Wrage K, Goth J, Bengston W (2001) Plant diversity enhances ecosystem responses to elevated CO2 and nitrogen deposition. Nature 410(6830):809–810. https://doi.org/10.1038/35071062

    Article  CAS  Google Scholar 

  • Ruiz-Benito P, Gómez-Aparicio L, Paquette A, Messier C, Kattge J, Zavala MA (2014) Diversity increases carbon storage and tree productivity in Spanish forests. Glob Ecol Biogeogr 23(3):311–322. https://doi.org/10.1111/geb.12126

    Article  Google Scholar 

  • Ruiz-Jaen MC, Potvin C (2010) Can we predict carbon stocks in tropical ecosystems from tree diversity? Comparing species and functional diversity in a plantation and a natural forest. New Phytol 189:978–987. https://doi.org/10.1111/j.1469-8137.2010.03501.x

  • Ruiz-Jaen MC, Potvin C (2011) Tree diversity explains variation in ecosystem function in a neotropical forest in Panama. Biotropica 42(6):638–646. https://doi.org/10.1111/j.1744-7429.2010.00631.x

    Article  Google Scholar 

  • Saefken B, Ruegamer D, Kneib T, Greven S (2018) Conditional model selection in mixed-effects models with cAIC4. https://arxiv.org/abs/1803.05664

    Google Scholar 

  • Silvertown J, Dodd M, Gowing D, Lawson C, McConway K (2006) Phylogeny and the hierarchical organization of plant diversity. Ecology 87(sp7):S39–S49. https://doi.org/10.1890/0012-9658(2006)87[39:pathoo]2.0.co;2

  • Sintayehu DW, Belayneh A, Dechassa N (2020) Aboveground carbon stock is related to land cover and woody species diversity in tropical ecosystems of Eastern Ethiopia. Ecol Process 9:37. https://doi.org/10.1186/s13717-020-00237-6

  • Stegen JC, Swenson NG, Valencia R, Enquist BJ, Thompson J (2009) Above-ground forest biomass is not consistently related to wood density in tropical forests. Glob Ecol Biogeogr 18(5):617–625. https://doi.org/10.1111/j.1466-8238.2009.00471.x

    Article  Google Scholar 

  • Tefera M, Chernet T, Haro W (1996) Explanation of the geological map of Ethiopia: scale 1: 2000,000, vol 3. Ethiopian Institute of Geological Surveys Bull, Addis Ababa

  • Thakur D, Chawla A (2019) Functional diversity along elevational gradients in the high-altitude vegetation of the western Himalaya. Biodivers Conserv 28:1977–1996. https://doi.org/10.1007/s10531-019-01728-5

  • Tilman D, Knops J, Wedin D, Reich P, Ritchie M, Siemann E (1997) The influence of functional diversity and composition on ecosystem processes. Science 277(5330):1300–1302. https://doi.org/10.1126/science.277.5330.1300

    Article  CAS  Google Scholar 

  • Vance-Chalcraft HD, Willig MR, Cox SB, Lugo AE, Scatena FN (2010) Relationship between aboveground biomass and multiple measures of biodiversity in subtropical forest of Puerto Rico. Biotropica 42(3):290–299. https://doi.org/10.1111/j.1744-7429.2009.00600.x

    Article  Google Scholar 

  • Villéger S, Mason NWH, Mouillot D (2008) New multidimensional functional diversity indices for a multifaceted framework in functional ecology. Ecology 89:2290–2301

  • Wang W, Lei X, Ma Z, Kneeshaw DD, Peng C (2011) Positive relationship between aboveground carbon stocks and structural diversity in spruce-dominated forest stands in New Brunswick, Canada. For Sci 57(2):506–515

  • Whittaker R, Heegaard E (2003) What is the observed relationship between species richness and productivity? Ecology 84(2):3384–3390. https://doi.org/10.1007/s00442-003-1343-z

  • Wieczynski DJ, Boyle B, Buzzard V, Duran SM, Henderson AN, Hulshof CM, Kerkhoff AJ, McCarthy MC, Michaletz ST, Swenson NG, Asner GP, Bentley LP, Enquist BJ, Savage VM (2019) Climate shapes and shifts functional biodiversity in forests worldwide. PNAS 116(2):587–592. https://doi.org/10.1073/pnas.1813723116

    Article  CAS  Google Scholar 

  • Wu H, Xiang W, Fang X, Lei P, Ouyang S, Deng X (2017) Tree functional types simplify forest carbon stock estimates induced by carbon concentration variations among species in a subtropical area. Sci Rep 7:4992. https://doi.org/10.1038/s41598-017-05306-z

    Article  CAS  Google Scholar 

  • Wu J, Zhang Q, Li A, Liang C (2015) Historical landscape dynamics of Inner Mongolia: patterns, drivers, and impacts. Landsc Ecol 30:1579–1598

  • Yuan Z, Wang S, Ali A, Gazol A, Ruiz-Benito P, Wang X, Lin F, Ye J, Hao Z, Loreau M (2018) Aboveground carbon storage is driven by functional trait composition and stand structural attributes rather than biodiversity in temperate mixed forests recovering from disturbances. Ann For Sci 75(3):67. https://doi.org/10.1007/s13595-018-0745-3

    Article  Google Scholar 

  • Zanne AE, Lopez-Gonzalez G, Coomes DA, Ilic J, Jansen S, Lewis SL, Miller RB, Swenson NG, Wiemann MC, Chave J (2009) Global wood density database. Dryad. Identifier: http://hdl.handle.net/10255/dryad.235.

    Google Scholar 

  • Zhang J-T, Li M, Nie E (2014) Pattern of functional diversity along an altitudinal gradient in the Baihua Mountain Reserve of Beijing, China. Braz J Bot 37(1):37–45. https://doi.org/10.1007/s40415-014-0047-4

  • Zhang QP, Wang J, Gu HL, Zhang Z-G, Wang Q (2018) Effects of continuous slope gradient on the dominance characteristics of plant functional groups and plant diversity in alpine meadows. Sustainability 10(12):4805. https://doi.org/10.3390/su10124805

  • Zhang Y, Chen HYH (2015) Individual size inequality links forest diversity and above-ground biomass. J Ecol 103:1245–1252. https://doi.org/10.1111/1365-2745.12425

    Article  Google Scholar 

  • Zhu B, Wang X, Fang J, Piao S, Shen H, Zhao S, Peng C (2010) Altitudinal changes in carbon storage of temperate forests on Mt Changbai, Northeast China. J Plant Res 123(4):439–452. https://doi.org/10.1007/s10265-009-0301-1

    Article  Google Scholar 

  • Zhu J, Jiang L, Zhang Y (2016) Relationships between functional diversity and aboveground biomass production in the Northern Tibetan alpine grasslands. Sci Rep 6(1):34105. https://doi.org/10.1038/srep34105

    Article  CAS  Google Scholar 

  • Ziter C, Bennett EM, Gonzalez A (2013) Functional diversity and management mediate aboveground carbon stocks in small forest fragments. Ecosphere 4(7):85 https://doi.org/10.1890/ES13-00135.1

    Article  Google Scholar 

  • Zuur AF, Ieno EN, Walker N, Saveliev AA, Smith GM (2009) Mixed effects models and extensions in ecology with R. Springer New York, New York. https://doi.org/10.1007/978-0-387-87458-6

    Book  Google Scholar 

Download references

Acknowledgements

We appreciate the collaboration of the Oromia Forest and Wildlife Enterprise, Dindin district workers for their help during the fieldwork in the Dindin natural forest. Besides, we thank Mr. Mekonnen Beyene (Oda Bultum University, Ethiopia) for his collaboration in collecting vegetation data from all the plots of this study. They also acknowledged the Oda Bultum University for offering a postgraduate study opportunity for the first author. In addition, we thank Dr. Motuma Tolera (Hawassa University, Ethiopia) and Dr. Dong Gill Kim (Hawassa University, Ethiopia) for their valuable suggestions for this paper.

Funding

We obtained funding for this research from Oda Bultum University, Ethiopia.

Author information

Authors and Affiliations

Authors

Contributions

Mengistu Teshome and Zebene Asfaw conceived the presented idea and developed the theory and analytical methods. Mengistu Teshome also planned and carried out the measurements. Mengistu Teshome, Zebene Asfaw, and Muktar Mohammed discussed the methodologies and results. Mengistu Teshome wrote the manuscript with help from Zebene Asfaw and Muktar Mohammed. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Mengistu Teshome Wondimu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

We have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This article has been retracted. Please see the retraction notice for more detail:https://doi.org/10.1186/s13717-023-00458-5

Supplementary Information

Additional file 1: Table S1.

List of tree species identified from Dindin dry evergreen Afromontane forest of Hararghe highland, southeast Ethiopia. Table S2. Estimated aboveground carbon stock for each plot in Dindin dry evergreen Afromontane forest of Hararghe highland, southeast Ethiopia. Table S3. Estimated values of functional diversity and functional dominance indices per plot in Dindin dry evergreen Afromontane forest of Hararghe highland, southeast Ethiopia.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wondimu, M.T., Nigussie, Z.A. & Yusuf, M.M. RETRACTED ARTICLE: Tree species diversity predicts aboveground carbon storage through functional diversity and functional dominance in the dry evergreen Afromontane forest of Hararghe highland, Southeast Ethiopia. Ecol Process 10, 47 (2021). https://doi.org/10.1186/s13717-021-00322-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13717-021-00322-4

Keywords