You are on page 1of 9

Ecological Indicators 111 (2020) 106025

Contents lists available at ScienceDirect

Ecological Indicators
journal homepage: www.elsevier.com/locate/ecolind

Biomass and soil carbon along altitudinal gradients in temperate Cedrus T


deodara forests in Central Himalaya, India: Implications for climate change
mitigation

Mehraj A. Sheikha, Munesh Kumara, N.P. Todariaa, Rajiv Pandeyb,
a
Department of Forestry and Natural Resources, HNB Garhwal University, Srinagar Garhwal, Uttarakhand, India
b
Forest Research Institute, Dehradun, India

A R T I C LE I N FO A B S T R A C T

Keywords: Carbon inventories are urgently needed for understanding climate dynamics and implementing climate miti-
Climate response gation strategies, but such data is scarce in for forest ecosystem of Himalaya. Therefore, the present study focuses
Carbon sequestration to supplement the existing information by estimating carbon stocks in temperate forests dominated by Cedrus
Forest litter deodara along the altitudinal gradients in the Central Himalaya. Three altitudes i.e., lower (1750 m), middle
GHG removal
(1900 m) and upper (2050 m) of the stand was considered to understand and estimate biomass and soil carbon
REDD+
storage potential of the forests based on the standard protocol. The results showed that the soil carbon stock
(SOC) decreased significantly with increasing soil depths and altitudes. Litterfall production in the forests varied
adversely with altitude. Above and below-ground biomass carbon stock also decreased along with altitude. The
study observed that the total carbon stock (soil, trees and forest floor) of the Cedrus deodara forest in different
altitudes was 395.4 t ha−1 (lower altitude), 321.6 t ha−1 (middle altitude) and 282.5 t ha−1 (upper altitude).
The estimates of the Cedrus forests would provide guidelines for estimating carbon for forest-based mitigation
activities in the Himalayan region.

1. Introduction offsetting capabilities of the forests. The precise estimation of forest


carbon is also required to understand the role of the forest for mitiga-
The concentration of atmospheric CO2 has accelerated upward tive actions to achieve the National Determined Commitment (NDC)
during the past few decades with an annual rate of 1.91 parts per such as REDD+ (Kishwan et al., 2012); and evaluating the role of
million (ppm) due to various anthropogenic activities, leading to in- forests for their ecosystems services to the poor (Pandey, 2009). The
crease in the atmospheric CO2 concentration from 317 ppm in 1958 to estimates of forest carbon are also required for evaluating forest de-
410 ppm in 2018 (NOAA, 2017). The change in CO2 is instrumental for gradation, as anthropogenic extractions from forests result in the de-
changing the temperature and has large implications on the social, gradation of forests (Malik et al., 2016). For example, carbon emissions
economical and ecological systems. Therefore, there is a need to reduce from world forests due to deforestation and forest degradation have
the emissions leading to limit the increase in global temperature below been reduced by 1 GT i.e. from 3.9 GT to 2.9 GT during 2011–2015
2 °C in respect to pre-industrial level (i.e., measured from 1750) (IPCC, (FAO, 2015).
2014). Forests are important for global carbon cycle because they se- In contrast, precise estimates for the carbon in specific forests are
quester carbon from the atmosphere, and also emit carbon resulting largely not available due to lack of resources and lack of experts across
from loss of biomass (Houghton et al., 2009; Kishwan et al., 2012). the globe. Besides this, the volume of work for the estimation of the
The carbon storage by the forests contributes to emission reduction forest carbon for larger forest patch such as at sub-national, national
by offsetting the GHG emissions by other sources (Houghton 2005; and global level are huge and require large number of experts
Houghton et al., 2009). For example, the mitigative role of India’s (Houghton et al., 2009; Sharma et al., 2010b). Moreover, the changes in
forests was estimated to 9.3% at the year 2000; and 4.9% in 2020 for the structure and composition of forests along with differences in in-
the annual emissions of the country (Kishwan et al., 2012). Precise dividual tree sizes and shapes restrict the precise evaluation of forests
information of forest carbon stock is important for understanding the carbon of a larger patch. The assessment of sequestration potential of


Corresponding author at: Forest Research Institute, Dehradun, Uttarakhand, India.
E-mail addresses: mehrajshkh@gmail.com (M.A. Sheikh), rajivfri@yahoo.com (R. Pandey).

https://doi.org/10.1016/j.ecolind.2019.106025
Received 13 December 2018; Received in revised form 12 October 2019; Accepted 16 December 2019
1470-160X/ © 2019 Elsevier Ltd. All rights reserved.
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

forests were attempted in different parcels of forests across the globe for region. The geo-coordinates of the three sites were N 30o08′00.52″ and
various purposes such as at a global scale (Houghton, 2005; Houghton E 078o46′32.38″ (Site-I), N 30o09′43.46″ and E 078o45′22.44″ (Site-II),
et al., 2009); forest type basis in India (Pandey et al., 2011); tree N 30o09′35.29″ and E 078o45′19.79″ (Site-III) with an altitude of
dominance basis in Himalaya (Kumar and Pandey, 2017); plantation 1750 masl, 1900 masl and 2050 masl, respectively. The approximate
basis at sub-national scale in India (Gupta and Pandey, 2008); habitat age of the forest is above 100 years, as informed by the villagers during
basis in England (Alonso et al., 2012). Moreover, several attempts have the survey. Three random selected sample plots were laid in each stand
also been made for carbon estimation for the country (Haripriya, 2000; following Hairiah et al. (2001). The size of the plot was 20 × 100 m for
Chhabra and Dadhwal, 2004). However, for forestry-based mitigation measurement of trees having stem diameter > 30 cm at breast height
actions, these studies were challenged due to the lack of precise sta- for estimation of form factors for each species. A subplot of size
tistics for carbon stock for the specific land parcel leading to undermine (5 m × 40 m) was laid for evaluation of trees of stem diameter between
the mitigative role of forests at one side and restrict the planning ap- 5 and 30 cm and tree density (ha−1) was estimated for the two classes
proach for the address of climate change impacts on the other (IPCC, of trees in each plot. Within these plots, smaller square subplots of
2014). Therefore, precise estimation of carbon stock at either stand 0.5 m2 size were laid for collection of soil samples. Litter traps were laid
level or species level or other required land parcels would be imperative within these smaller subplots for collection of litter. For analysis of soil,
and would lead to supplementing the reduction of overall complexities 6 samples each at four depths (6 samples × 4 depths) were collected
of mitigation options (Wutzler et al., 2007; UNFCCC, 2015). Moreover, from each sample sub-plot at four different soil depths i.e., 0–0.3, 0–0.5,
understanding the change in carbon stocks in the forests would further 0–1.0, 0–1.5 m. Therefore, from each stand 72 samples were considered
assist the forest manager to regulate the rules and regulation for ef- for estimation of soil organic carbon (SOC) consisting of six samples at
fective management of forests focussing to deliverables of services to four depths from the 3 sample plots. Soil samples were collected by
the humans. metalcore sampling cylinder (core method).
The present study was motivated by the fact that the precise esti-
mation of carbon stock is essentially required for various purposes in- 2.3. Analysis of physico-chemical properties of soil
cluding incentivization for mitigation actions. This study was con-
sidered to address the gap by providing an estimate of carbon for Cedrus The collected soil samples were brought to the laboratory following
tree dominant forests in the Himalaya region. The study attempts to standard protocol and labeled for further processing. The samples were
estimate carbon stock in the moist (Cedrus deodara Loud.) Deodar forest air-dried grounded in a cyclotec-1030 mill and passed through a 2 mm
(Type – 12/C1C) (Champion and Seth, 1968) of the Himalayan region in sieve (Jackson, 1967). The fractions > 2 mm and < 2 mm size was
India. Cedrus deodara, usually found in pure stands, occurs at altitudinal stored separately and the fine fractions were used for further analysis.
range from 1750 to 2150 m asl and typically gregarious. Major as-
sociates of the Cedrus deodara are Abies pindrow, Picea smithiana and
Pinus wallichiana (Luna, 2005). The usual girth of tree ranges between 2.3.1. Bulk density
2.4 and 3.6 m and height up to of 35 m (Luna, 2005). The tree is used Collected soil samples were oven-dried at 100 ± 5 °C for 48 h and
for several purposes including timber for building. Keeping this in view weighed. The bulk density was estimated using a standard core method
along with the wide distribution and importance of Cedrus deodara (Wilde et al., 1964) using the following formula.
forests, it is important to estimate biomass and soil carbon storage ca- Bulk density(g/cm3) = Mass of dry soil(g)/Volume of dry soil(cm3)
pacity for predicting carbon storage in Himalayan forests. The present
study attempts to estimate carbon stock of Cedrus deodara forests as
well as the seasonal contribution of the forest floor to carbon stock in 2.3.2. Coarse fragment
the forest. Percent coarse fragments (> 2 mm size) in soils were estimated to
evaluate the soil mass. The soil samples were kept in the 2 mm sieve,
2. Materials and methods and running water was supplied through the sample. Soil particles <
2 mm were allowed to wash away with running water. The remaining
2.1. Focus of study fractions in the sieve were dried and weighted for estimation of per-
centage coarse fragment.
Uttrakhand with a geographic area of 53,484 km2 lies between the
latitude of 28° 42́ and 31° 28́ N and longitude of 77° 35́ and 81° 50́ E,
with altitudinal ranges between 400 and 3500 m. amsl (Fig. 1). The 2.3.3. Moisture content (%)
forests in the state are spread in 65% of the total geographical area The moisture content of the soil samples was estimated as per the
comprising sub-tropical Chir pine coniferous and mixed deciduous standard method (Misra, 1968
woody of the temperate zone with tree species such as Juglans regia, Moisture content(%) = [(Fresh weight − Dry weight)/Fresh weight] × 100
Picea smithiana, Cedrus deodara and Pinus gerardiana. The study focuses
on the hilly part of the state i.e., Garhwal Himalaya. The vegetation and
soils of Garhwal Himalaya vary according to aspect, altitude and cli- 2.3.4. Estimation of soil carbon stock in soil
mate. The hilly terrain of the region contains densely forested slopes Walkley and Black, Krishan et al. (2009) an analytical technique
containing generally clayey to sandy loams sedentary soil with podzo- was applied to estimate soil organic carbon (SOC) by weight to volume
lization. The climate of the study region is temperate with very cold in (Walkley and Black, 1934). The soil carbon stock (SOC) was estimated
winter and pleasant in summer. The temperature of the region ranges by Batjes (1996) approach. For an individual profile with k depths, the
between −2 °C to 20 °C in January and 24 °C to 36 °C in June and July total organic carbon by volume is estimated by the following formula:
(Kumari et al. 2018) with an average annual rainfall of 2180 mm
k
generally commencing from mid-June and extending till mid-Sep-
tember with relative humidity between 54% and 63%.
Qi = ∑ ρi Pi Di (1 − Si)
i=1

2.2. Sampling for data collection where Q i is the total amount of organic carbon (in Mg m−2) over depth,
ρi is the bulk density (Mg m−3) of layer i, Pi is the proportion of organic
Three forest stands of Cedrus deodara located at three different al- carbon (g C g−1) in layer i, Di is the thickness of ith layer (m), and Si is
titudes at north aspect were selected at random for sampling from the the volume of the fraction of fragments > 2 mm (Batjes, 1996).

2
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

Fig. 1. Location map of the study area.

2.4. Biomass estimation 2.4.2. Branch biomass


The total number of branches irrespective of size was counted on
2.4.1. Aboveground biomass each of the sample trees and categorized on the basis of basal diameter
The diameter at breast height (dbh) was measured by tree calliper into three groups, viz < 6 cm, 6–10 cm and > 10 cm. Fresh weight of
and height (h) through Ravi’s multimeter. The form factor (F) of a tree two randomly selected branches from each group were recorded, se-
was estimated with Spiegel relaskope by using formula F = 2*h1/3h parately. The dry weight of branches was estimated by using equation
(Pressler, 1895; Bitlerlich, 1984) where h1 is the height at half dbh. (Chidumaya, 1990) as; Bdwi = Bfwi/(1 + Medbi) where Bdwi is the oven-
Volume (V) was estimated with V = F × h × g, where g the basal area dry weight of branches, Bfwi the fresh weight of branches and Medbi the
g = π (dbh/2)2 (Pressler, 1895) as exemplified by Koul and Panwar moisture content of branches on a dry weight basis. Total branch bio-
(2008). Biomass of wood was estimated by multiplying specific gravity mass per sample tree was determined as; Bbt = n1 bw1 + n2 bw2 + n3
with volume. Specific gravity was estimated based on stem core by the bw3 where Bbt is the branch biomass per tree, ni the no of branches in
maximum moisture method (Koul and Panwar, 2008). the ith branch group (i = 1(1)3), wi is the weight of the one branch in
the group (Koul and Panwar, 2008).

3
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

2.4.3. Leaf biomass (120–150 cm) (Table 1). The depth-wise analysis revealed that carbon
Leaves from five randomly selected branches of individual trees stock increased with soil depth from 0 to 30 cm to 0–150 cm in all the
were collected from five randomly selected trees per quadrate. The three altitudes i.e., 104.8 ± 13.11 to 174.3 ± 23.69 t ha−1 (lower
leaves were weighed and oven-dried separately to a constant weight at altitude), 105.3 ± 15.34 to 173.0 ± 11.87 t ha−1 (middle altitude)
80 ± 5 °C. The average leaf biomass was estimated by multiplying the and 89.7 ± 20.93 to 139.3 ± 12.70 t ha−1 (upper altitude) (Table 1).
average biomass of the leaves per branch with the numberof branches
in a single tree and the number of trees in a quadrate (Chidumaya,
1990; Koul and Panwar, 2008). 3.3. Above and below ground biomass and carbon

2.4.4. Belowground biomass In lower altitude, six tree species i.e., Cedrus deodara, Cupressus
Below ground biomass density (BGBD) (fine and coarse roots) was torulosa, Pinus roxburghii, Quercus leucotrichophora, Quercus floribunda
estimated for each species using the allometric equation by Cairns et al. and Myrica esculenta were observed with decreasing density and basal
(1997). area (Table 2). The maximum mean diameter was observed for Cedrus
deodara followed by Pinus roxburghii and Cupressus torulosa with bole
BGBD = exp{ −1.059 + 0.884 × ln(AGBD) + 0.284}
volume of 563.6, 2.8 and 126.3 m3 ha−1, respectively (Table 2). Total
where AGBD is above ground biomass density. AGB and biomass in different tree components i.e., bole, branch, twig
and foliage is reported in Table 3. Total above-ground tree biomass was
2.4.5. Understorey and Herbaceous biomass highest (307.3 t ha−1) in Cedrus deodara followed by Cupressus torulosa
Herbaceous biomass was estimated by collecting all the plants from (70.5 t ha−1), Pinus roxburghii (2.3 t ha−1), Quercus leucotrichophora
a randomly laid sub-quadrate of size 1 m × 1 m in the sample plot. All (0.3 t ha−1), Quercus floribunda (0.2 t ha−1) and Myrica esculenta
collected plants were oven-dried 65 ± 5 °C to a constant weight. Shrub (0.04 t ha−1) (Table 3). The BGB for each tree was also estimated and
biomass was also estimated by applying a similar procedure. reported in Table 3. The total biomass (AG and BG) for each tree species
as Cedrus deodara, Cupressus torulosa, Pinus roxburghii, Quercus leuco-
2.4.6. Litter biomass trichophora, Quercus floribunda and Myrica esculenta were 387.0, 92.2,
All vegetation < 5 cm dbh were harvested from the considered 3.4, 0.5, 0.3 and 0.07 t ha−1, respectively. The distribution of carbon
sample plot. Litter was collected in litter traps on sample plots and mass stock (t ha−1) allocation in different components of each species is
of the total fresh sample (in g m−2) was measured. A composite fresh reported in Table 4. The species wise, AG and BG carbon stock were
sub-sample (~300 g) were taken and weighted after thorough mixing 138.3 and 35.8 t ha−1 (Cedrus deodara), 31.7 and 9.8 t ha−1 (Cupressus
for subsequent oven drying at 80 °C for conversion to dry weight and for torulosa), 1.1 and 0.5 t ha−1 (Pinus roxburghii), 0.12 and 0.1 t ha−1
further analysis (Hairiah et al., 2001). (Quercus leucotrichophora), 0.1 and 0.04 t ha−1 (Quercus floribunda) and
0.02 t ha−1 and 0.01 t ha−1 (Myrica esculenta). The maximum carbon
Total dry weight(kg m2) stock (AG and BG) was observed for Cedrus deodara followed by Cu-
Total fresh weight(kg) × Subsample dry weight (g) pressus torulosa (Table 4).
=
Subsample fresh weight(g) × Sampled area(m2) In middle altitude, the associated species of Cedrus deodara were
Cupressus torulosa, Rhododendron arboreum, Quercus leucotrichophora
and Quercus floribunda. The highest density and basal area were ob-
2.4.7. Estimation of carbon stock in biomass served for Cedrus deodara followed by Cupressus torulosa, Rhododendron
Biomass contains carbon from 45 to 50% of dry matter (Chan, 1982; arboreum, Quercus leucotrichophora and Quercus floribunda (Table 2). In
Kishwan et al., 2012; Pandey et al., 2014). In the present study, the this site, Cedrus deodara was having maximum mean diameter and
total amount of carbon stocked was 0.45 times of dry biomass as sug- height followed by Cupressus torulosa, Rhododendron arboreum, Quercus
gested by Woomer (1999). floribunda and Quercus leucotrichophora (Table 2). Total AGB with a
distribution of biomass in different components i.e. bole, branch, twig
3. Results and foliage were estimated and reported in Table 3. The highest total
AGB (208.7 t ha−1) and BGB (56.7 t ha−1) were observed for Cedrus
The collected data were statistically analyzed for various para- deodara with a total biomass of 265.4 t ha−1 followed by Cupressus
meters for Cedrus deodara stand. The details of descriptive statistics of torulosa, Rhododendron arboreum, Quercus leucotrichophora and Quercus
different parameters and carbon for various observed tree species as floribunda (Table 3). The maximum carbon stock (AG and BG) was
well as forest floor in the stand are discussed below in details. observed in Cedrus deodara followed by Cupressus torulosa, Rhododen-
dron arboreum, Quercus floribunda and Quercus leucotrichophora
3.1. Physico-chemical properties of soil (Table 4).
In upper altitude, the highest density and basal area were observed
The diverse trend in soil moisture was observed across the three for Cedrus deodara with 258 individual ha−1 and 34.1 m2 ha−1, re-
altitudes with a maximum at the higher elevation. In lower altitude, soil spectively (Table 2). The maximum diameter was found in Cedrus
moisture was maximum at the surface i.e. 0–30 cm and minimum at the deodara followed by Cupressus torulosa, Quercus leucotrichophora, Rho-
120–150 cm. Bulk density (BD) and carbon fraction (CF) was minimum dodendron arboreum and Quercus floribunda (Table 2). Total AGB accu-
at 0–30 cm and increased with depth (Table 1). In middle altitude, the mulation and its allocation to different tree components i.e., bole,
moisture content reduced from 24.3 ± 5.50% (0–30 cm) to branch, twig and foliage are reported in Table 3. AGB was highest in
13.3 ± 2.19% (120–150 cm). However, BD and CF were observed to Cedrus deodara for each component i.e., bole, branch, twig and foliage
increase with increasing depth (Table 1). In upper altitude, moisture and lowest in Quercus floribunda for the same components. The AGB and
percent was decreased with increasing soil depth, however, BD and CF BGB for Cedrus deodara were 177.2 and 49.3 t ha−1 with a total biomass
increased with increasing soil depth (Table 1). of 226.5 t ha−1, followed by Cupressus torulosa, Quercus leuco-
trichophora, Rhododendron arboreum and Quercus floribunda (Table 3).
3.2. Soil carbon Maximum total (AG and BG) carbon stock (101.9 t ha−1) was found in
Cedrus deodara followed by Cupressus torulosa, Quercus leucotrichophora,
In all the three altitudes (lower, middle and upper), SOC and carbon Rhododendron arboreum and Quercus floribunda respectively (Table 4).
stock were decreased from the top layer (0–30 cm) to bottom

4
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

Table 1
Physical properties of soil in Cedrus deodara stands of the study site.
Site (Altitude) Depth (cm) Moisture (%) BD (g cm3) CF (%) SOC (%) SOC stock (t ha−1)

Site – I (Lower altitude) 0–30 19.0 ± 13.41 1.2 ± 0.06 13.1 ± 2.46 3.4 ± 0.42 104.8 ± 13.11
0–60 15.9 ± 8.04 1.2 ± 0.06 12.3 ± 1.20 2.2 ± 0. 37 138.3 ± 23.39
0–100 6.7 ± 2.25 1.2 ± 0.07 14.3 ± 3.65 1.5 ± 0.34 157.4 ± 35.37
0–150 6.3 ± 1.79 1.3 ± 0.08 16.0 ± 4.54 1.1 ± 0.15 174.3 ± 23.69

Site – II (Middle altitude) 0–30 24.3 ± 5.50 1.2 ± 0.16 3.4 ± 3.14 3.1 ± 0.44 105.3 ± 15.34
0–60 16.5 ± 3.62 1.2 ± 0.06 3.2 ± 0.28 2.2 ± 0.72 154.1 ± 51.93
0–100 14.2 ± 0.95 1.3 ± 0.07 6.9 ± 6.39 1.3 ± 0.24 152.7 ± 28.02
0–150 13.3 ± 2.19 1.3 ± 0.15 9.9 ± 7.97 1.0 ± 0.07 173.0 ± 11.87

Site – III (Upper altitude) 0–30 30.9 ± 0.84 1.1 ± 0.06 7.0 ± 1.21 2.8 ± 0.66 89.7 ± 20.93
0–60 15.6 ± 4.69 1.2 ± 0.10 6.3 ± 0.99 1.9 ± 0.62 128.5 ± 42.37
0–100 10.8 ± 4.36 1.3 ± 0.09 8.5 ± 3.87 1.1 ± 0.06 128.8 ± 7.29
0–150 8.8 ± 1.40 1.3 ± 0.18 8.2 ± 3.22 0.8 ± 0.07 139.3 ± 12.70
Significance * ns ns ** *
CD 9.45 0.44 13.47

*Significance (p < 0.0 5); **Significance (p < 0.01); and ns - non-significance (p < 0.05).

Table 2
Mean diameter, height, density, basal area, volume and the specific gravity of wood for prominent tree species of different stands.
Site and Altitude Species Mean diameter Mean height (m) Form Average trees Density Basal area Bole volume Specific
(cm) factor sampled (individual ha−1) (m2 ha−1) (m3 ha−1) gravity
(2000 m2)

Site – I (Lower Cedrus deodara 47.8 ± 12.95 30.1 ± 3.98 0.346 52 365 53.8 563.6 0.437
altitude) Cupressus torulosa 38.0 ± 17.31 24.5 ± 8.50 0.345 12.33 181.5 10.3 126.3 0.433
Pinus roxburghii 42.3 ± 00 35.0 ± 00 0.337 0.33 1.5 0.24 2.8 0.632
Quercus 5.1 ± 2.90 7.0 ± 1.41 0.247 0.66 33.5 0.11 0.2 0.865
leucotrichophora
Quercus floribunda 6.3 ± 0.94 4.8 ± 0.35 0.281 0.66 33.5 0.10 0.1 0.60
Myrica esculenta 5.5 ± 00 3.0 ± 0.00 0.222 0.33 16.5 0.03 0.03 0.737

Site – II (Middle Cedrus deodara 43.8 ± 11.20 28.1 ± 2.86 0.337 45.66 318 40.3 384.8 0.437
altitude) Cupressus torulosa 36.0 ± 10.30 24.6 ± 5.83 0.344 10 95 6.4 55.1 0.433
Rhododendron 28.0 ± 13.41 11.4 ± 3.21 0.338 1.33 53 2.2 8.8 0.628
arboreum
Quercus 14.1 ± 00 10.0 ± 00 0.256 0.33 33 0.3 0.9 0.865
leucotrichophora
Quercus floribunda 16.5 ± 1.97 9.0 ± 5.66 0.261 0.66 16.5 0.4 1.5 0.60

Site – III (Upper Cedrus deodara 44.8 ± 11.64 28.0 ± 2.09 0.338 36.66 258 34.1 326.3 0.437
altitude) Cupressus torulosa 43.2 ± 18.85 23.0 ± 8.81 0.338 4.66 68 4.5 53.7 0.433
Rhododendron 25.4 ± 10.14 12.0 ± 2.43 0.3 4 106.5 4.8 14.1 0.628
arboreum
Quercus 28.1 ± 4.22 15.0 ± 1.17 0.306 2 55 3.0 11.6 0.865
leucotrichophora
Quercus floribunda 25.0 ± 27.28 4.5 ± 2.12 0.444 0.33 18 0.3 0.02 0.60
Significance ** ns ns ** **
CD 6.26 0.97 24.19

**Significance (p < 0.01) and ns non-significance (p < 0.05); Average values of trees (n = 3plots each of 2000 m2); Form factor is mean of trees of three different
plots in each altitude, estimated based on field survey data.

3.4. Litter and understory biomass the season. The understory biomass was estimated of
1.3 ± 0.66 t ha−1 comprising a total forest floor biomass of
In lower altitude, the litter production was estimated for different 8.6 ± 2.07 t ha−1 (Table 6). The total forest floor carbon was
seasons i.e., summer, rainy and winter. The maximum litter production 3.8 ± 0.93 t ha−1 in middle altitude (Table 6). In upper altitude, the
was in summer season followed by rainy and winter seasons (Table 5) biomass of litter production was 3.1 ± 1.48 (summer season),
with an annual litter production of 7.6 ± 1.11 t ha−1 (Table 6). The 1.4 ± 0.38 (rainy season) and 1.2 ± 0.62 (winter season) with an
litter production differs among seasons significantly (p < 0.001) annual production of 5.7 ± 1.04 t ha−1 (Table 5). The understory
(Table 5). The understory biomass was estimated to be biomass was estimated 0.7 ± 0.09 t ha−1 which makes a total forest
0.5 ± 0.18 t ha−1 with a total forest floor biomass of floor biomass of 6.3 ± 1.13 t ha−1 and a total forest floor carbon of
8.1 ± 1.29 t ha−1 (Table 6). The distribution of total forest floor 2.8 ± 0.51 t ha−1 (Table 6).
carbon (litter carbon and underground biomass carbon) was
3.4 ± 0.50 t ha−1 and 0.2 ± 0.08 t ha−1 with a total forest floor
carbon of 3.6 ± 0.58 t ha−1 (Table 6). In middle altitude, the max- 4. Discussion
imum litter production was recorded 3.9 ± 1.32 in summer season
followed by 2.2 ± 0.79 and 1.2 ± 0.85 for rainy and winter seasons, 4.1. Soil carbon in relation to depth and altitude
respectively with an annual production of 7.2 ± 1.41 t ha−1 (Table 5).
The carbon stock of litter also followed a similar trend with respect to The vertical distributions of SOC in all the sites were in decreasing
order with increasing soil depth (Table 1). The higher SOC content in

5
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

Table 3
Biomass (t ha−1) allocation in different components of different species in different stands.
Site (Altitude) Species AGB (t Biomass allocation (t ha−1) (Mean ± SE) BGB (t TB (t
ha−1) ha−1) ha−1)
Bole Branch Twig Foliage

Site – I (Lower Cedrus deodara 307.3 246.3 ± 104.40 38.4 ± 15.73 (12.5) 11.5 ± 4.05 (3.7) 11.1 ± 4.65 (3.6) 79.6 387.0
altitude) (80.2)
Cupressus torulosa 70.5 54.7 ± 27.60 (77.6) 9.3 ± 3.97 (13.2) 2.8 ± 0.61 (4.0) 3.7 ± 1.62 (5.3) 21.7 92.2
Pinus roxburghii 2.3 1.7 ± 3.03 (73.9) 0.3 ± 0.61 (13.0) 0.1 ± 0.22 (4.4) 0.1 ± 0.19 (4.4) 1.1 3.4
Quercus 0.3 0.2 ± 0.30 (58.8) 0.1 ± 0.16 (29.4) 0.03 ± 0.06 (8.8) 0.01 ± 0.02 (2.9) 0.2 0.5
leucotrichophora
Quercus floribunda 0.2 0.1 ± 0.15 (62.5) 0.04 ± 0.06 (25.0) 0.01 ± 0.02 (6.3) 0.01 ± 0.01 (6.3) 0.1 0.2
Myrica esculenta 0.04 0.02 ± 0.04 (57.1) 0.01 ± 0.02(28.6) 0.003 ± 0.01 (8.6) 0.002 ± 0.00 (5.7) 0.03 0.07

Site – II (Middle Cedrus deodara 208.6 168.2 ± 45.36 (80.6) 23.9 ± 4.98(11.5) 7.9 ± 1.12 (3.8) 8.7 ± 1.40 (4.1) 56.7 265.4
altitude) Cupressus torulosa 30.6 23.8 ± 9.72 (77.8) 3.9 ± 1.48(12.8) 1.4 ± 0.53 (4.6) 1.5 ± 0.65 (4.9) 10.3 41.0
Rhododendron 8.3 5.5 ± 4.16 (66.3) 1.4 ± 1.02 (16.9) 0.7 ± 0.46 (8.4) 0.7 ± 0.45 (8.4) 3.3 11.6
arboreum
Quercus 1.1 0.6 ± 1.04 (54.6) 0.3 ± 0.55 (27.3) 0.1 ± 0.20 (9.1) 0.1 ± 0.16 (9.1) 0.6 1.7
leucotrichophora
Quercus floribunda 1.4 0.9 ± 0.89 (64.3) 0.3 ± 0.28 (21.4) 0.1 ± 0.09 (7.1) 0.1 ± 0.09 (7.1) 0.7 2.1

Site – III (Upper Cedrus deodara 177.2 142.6 ± 28.62 (80.5) 19.2 ± 1.94 (10.8) 8.1 ± 2.24 (4.6) 7.3 ± 1.77 (4.1) 49.3 226.5
altitude) Cupressus torulosa 30.4 23.3 ± 20.86 (76.4) 4.0 ± 3.84 (13.1) 1.7 ± 1.55(5.6) 1.5 ± 1.41 (4.9) 10.4 40.77
Rhododendron 13.3 8.9 ± 6.20 (66.9) 2.1 ± 1.78 (15.8) 1.3 ± 0.85(9.8) 1.0 ± 0.86 (7.5) 4.9 18.2
arboreum
Quercus 19.4 10.1 ± 1.70 (52.1) 6.0 ± 1.22(30.9) 2.2 ± 0.52 (11.3) 1.2 ± 0.51 (5.7) 7.0 26.4
leucotrichophora
Quercus floribunda 0.03 0.01 ± 0.03 (41.7) 0.01 ± 0.01 (41.7) 0.002 ± 0.00 (8.3) 0.002 ± 0.00 (8.3) 0.02 0.1
Significance ns *** *** ** * ** **
CD – 59.72 8.57 2.488 2.753 2.196 13.54

*Significance (p < 0.0 5); **Significance (p < 0.01); ***Significance (p < 0.001) and ns non-significance (p < 0.05); (AGB: Above ground biomass; BGB: Below
ground biomass; TB: Total biomass), Value within parenthesis is proportion of AGB.

the top layers of soil may be due to rapid decomposition of forest litter soil layer contains most of the root tips and has porous and nutritious
and mixing with top surface materials. Similar results were also re- soil (Samina et al., 1999; Persson, 1980). Soil BD increases with soil
ported for tropical soils by Dinakaran and Krishnayya (2008), and for depths and therefore affects the penetration of root and subsequently
temperate soil by Jobbágy and Jackson (2000) and Sheikh et al. (2009). the root biomass (Nambiar and Sands, 1992). SOC is also influenced by
The decrease in the SOC content with increasing soil depth may be inter and intra interaction of various biophysical factors with vegeta-
attributed to favorable environment for topsoil and a decrease in bio- tion types (Lal, 2005) though sufficient information is not available to
logical activity with increasing soil depth (Jobbágy and Jackson, 2000). understand the variation in size and residence time of different frac-
In the present study, SOC proportion was higher in comparison to be- tions of SOC (Dinakaran and Krishnayya, 2008). In the present study,
lowground biomass of trees. It may be due to the higher microbial ac- the SOC ranged from 0.8 to 3.4% across the depths, which is lower than
tivities in the upper soil layers (0–20 cm). Higher microbial activities at that the observed SOC ranges between 3.48 and 3.64 % for Cedrus
upper soil layers leading toaccelerating decomposition are reported by deodara forests at Central Himalaya by Joshi et al. (1991). The differ-
Singh et al. (1990), Upadhyay and Singh (1989). Moreover, the upper ences in SOC may be attributed to the surface layer soil SOC. Further,

Table 4
Carbon stock (t ha−1) allocation in different components and species in different stand.
Site (Altitude) Species Biomass carbon allocation (t ha−1) AGC (t ha−1) BGC (t ha−1) TC (t ha−1)

Bole Branch Twig Foliage

Site – I (lower altitude) Cedrus deodara 110.8 17.3 5.2 5.0 138.3 35.8 174.1
Cupressus torulosa 24.6 4.2 1.3 1.7 31.7 9.8 41.5
Pinus roxburghii 0.8 0.2 0.1 0.1 1.0 0.5 1.6
Quercus leucotrichophora 0.1 0.04 0.01 0.01 0.12 0.1 0.2
Quercus floribunda 0.04 0.02 0.01 0.004 0.05 0.04 0.1
Myrica esculenta 0.01 0.004 0.002 0.001 0.02 0.01 0.03

Site – II (Middle altitude) Cedrus deodara 75.7 10.7 3.5 3.9 93.9 25.5 119.4
Cupressus torulosa 10.7 1.8 0.6 0.7 13.8 4.7 18.5
Rhododendron arboretum 2.5 0.6 0.3 0.3 3.7 1.5 5.2
Quercus leucotrichophora 0.3 0.1 0.1 0.04 0.5 0.3 0.8
Quercus floribunda 0.4 0.1 0.04 0.04 0.6 0.3 0.9

Site – III (Upper altitude) Cedrus deodara 64.2 8.6 3.6 3.3 79.7 22.2 101.9
Cupressus torulosa 10.5 1.8 0.7 0.66 13.7 4.7 18.4
Rhododendron arboretum 4.0 0.9 0.6 0.5 6.0 2.2 8.2
Quercus leucotrichophora 4.5 2.7 1.0 0.6 8.7 3.1 11.9
Quercus floribunda 0.01 0.003 0.001 0.001 0.01 0.02 0.02
Significance * * ns ns ** ** **
CD 4.36 0.71 0.19 0.10 5.11 0.99 6.09

ns
*Significance (p < 0.05); **Significant (p < 0.01) and non-significance (p < 0.05).

6
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

Table 5
Season wise litter production (t ha−1) in different stands.
Site (Altitude) Litter production (t ha−1) (Mean ± SE) Significance CD

Summer Rainy Winter

Site – I (Lower altitude) 3.8 ± 0.93 1.9 ± 0.73 1.8 ± 0.7 *** 0.95
Site – II (Middle altitude) 3.9 ± 1.32 2.2 ± 0.79 1.2 ± 0.85
Site – III (Upper altitude) 3.1 ± 1.48 1.4 ± 0.38 1.2 ± 0.62

***Significance (p < 0.001).

the cumulative SOC stock up to 150 cm ranged from 139.3 ± 12.70 4.2.2. Biomass and carbon stock in relation to Regional, national and global
(upper altitude) to 174.3 ± 23.69 t ha−1 (lower altitude), which is level
lower than the national average content of soil organic carbon as The mean value of AGB in the present study was 290.3 t ha−1,
182.94 t C ha−1 (Jha et al., 2003). which is higher than the reported (141.20 t ha−1) of Haripriya (2000)
The difference of SOC across the depth and altitude was significant for Cedrus deodara forests of India. The above-ground biomass was
(p < 0.001). The decrease in SOC with increasing altitude may be 137–245 t ha−1 and 239.8 t ha−1 respectively, as reported by Tiwari
attributed to low plant density at higher altitude. Moreover, the basal and Singh (1987), Sheikh et al. (2011) for Central Himalayan forests,
area also decreased from 64.59 to 49.6 m2 ha−1 with increasing alti- which are also much lower than the present value. The differences may
tude. The mean diameter (cm), basal area (m2 ha−1) and volume (m3 be attributed to the sampling, location besides the characteristics of the
ha−1) were also significantly different across the different altitudes. forests. However, these values are comparable to the biomass ranged of
However, mean height and density were not significantly different with 200–600 t ha−1 generally found in the mature forests of the world
altitudes. It has been reported that tree density and basal area decreases (Whittaker, 1975), but are higher than the reported (133–202 t ha−1)
with altitude for Himalayan forests (Gairola et al., 2008). The decline in for the majority of the conifer forests of the temperate regions (Reichle,
vegetation with increasing altitude results in less accumulation of litter 1981).
leading to low inputs for accumulation organic carbon in soils (Sevgi The AGC stock in the present study was 130.6 t ha−1 and higher
and Tecimen, 2008). The variation in SOC along altitude may be at- than the reported values (70.60 t ha−1) of Haripriya (2000) for Cedrus
tributed to the spatial variation of soil bulk density (Li et al., 2010) or deodara forests of Himalayas. The AGC reported by Singh et al. (1985)
better stabilization of SOC at lower altitudes (Sheikh et al., 2009) or for poor (highly disturbed) and (less disturbed) forests of Central Hi-
variation in above ground litter inputs (Miao et al., 2019). malayas are 35–75.20 t ha−1 and 75.20–131.5 t ha−1 respectively and
lower than the present study, however, within the range of undisturbed
forests of Central Himalayas i. e., 131.5–225.6 t ha−1. The total carbon
4.2. Above-ground biomass and carbon values of the present study (167.5 t ha−1) were much higher than the
national average reported by Sheikh et al. (2011) and Chhabra et al.
4.2.1. Biomass and carbon allocation in different components (2002) with values of 42.92 and 61.06 t ha−1 respectively. However,
Allocation of the biomass of Cedrus deodara to the various compo- these values are lower than that reported by Sheikh et al. (2011) for
nents namely, bole, branch, twig and foliage were compared with other Cedrus deodara forests of Binsar and Tarkishwar in Garhwal Himalayas
studies of the temperate forests of Central Himalayas. The biomass al- which were 245.31 t ha−1. The higher carbon stock in the present study
location percent in bole was higher than reported values for Central may be mainly because of the high density and basal area as well as
Himalayan Forests (Rana et al., 1989). There was a decrease in bole, matured forests besides less disturbed, as observed during the survey.
branch, twig and foliage biomass with increasing altitude in species Old-growth forests are considered to be carbon neutral because the
across all the stands. The difference in bole and branch biomass was amount of carbon uptake by photosynthesis is equal to the release
significant (p < 0.001) with increasing altitude. Twig biomass and through the plant and soil respiration (Odum, 1969). However, in-
foliage biomass differed across the altitude. The total biomass and creasing evidence show that old-growth, undisturbed forests can still
carbon allocation with altitude were also significantly different. The accumulate carbon even as net primary productivity (NPP) declines
BGB and BGC were also significant (p < 0.01) with altitude. The es- (Zhou et al., 2006).
timates of BGB and BGC were based on Cairns et al. (1997) due to
unavailability of the estimates for the species and the forest patch.
Therefore, for a broader perspective, it was used however, the estimates 4.2.3. Biomass and carbon in relation with altitude
may be imprecise. Cairns et al (1997) was considered with the view that A decreasing trend of biomass and carbon stock was observed with
the study contains the data from some Himalayan studies focussing to respect to altitude in the study. The decrease in biomass may be due to
Cedrus deodar forests such as Garkoti and Singh (1995) for higher ele- decrease in density (631.5, 515.5 and 505.5 individual ha−1 in lower
vations and Singh et al. (1994) for a different altitudinal range of Indian altitude, middle altitude and upper altitude) and basal area (64.6, 49.6
Himalaya. and 46.6 m2 ha−1) in lower altitude, middle altitude and upper altitude

Table 6
Total forest floor (litter and below ground) biomass (t ha−1) and forest floor carbon (t ha−1).
Site (Altitude) Forest floor biomass (t ha−1) (Mean ± SE) Forest floor carbon (t ha−1) (Mean ± SE)

Litter production (t ha−1 yr−1) Understory biomass (t ha−1) TFFB (t ha−1) Litter (t ha−1) Understory carbon (t ha−1) TFFC (t ha−1)

Site – I (Lower altitude) 7.6 ± 1.11 0.5 ± 0.18 8.1 ± 1.29 3.4 ± 0.50 0.2 ± 0.08 3.6 ± 0.58
Site – II (Middle altitude) 7.2 ± 1.41 1.3 ± 0.66 8.4 ± 2.07 3.3 ± 0.63 0.6 ± 0.30 3.8 ± 0.93
Site – III (Upper altitude) 5.7 ± 1.04 0.7 ± 0.09 6.3 ± 1.13 2.6 ± 0.47 0.3 ± 0.04 2.8 ± 0.51
Significance * ns ns * ns ns
CD 1.36 0.59

TFFB: Total forest floor biomass; TFFC: Total forest floor carbon; *Significance (p < 0.0 5); ns
non-significance (p < 0.05).

7
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

of trees with increasing altitude. The decrease in biomass with an in- may provide precise estimates of altitude-wise biomass expansion
crease in altitude may also be attributed to a decrease of mean tree factor (BEF) and can be used as a reference in future studies specific to
height and mean diameter with altitude. Sangarun et al. (2007), has Himalayan conifer forests. Moreover, this study provides an estimate of
also reported a decrease of height with an increase in altitude. In- overall soil C stock up to a depth of 1.5 m in contrast to many previous
creasing elevation may affect tree growth rates and stand structure studies in the Central Himalayan region, which has been restricted only
because of reduced air and soil temperatures often increased rainfall, up to a depth of 30 or 60 cm. The present evaluation would facilitate for
and alterations in nutrient availability and soil chemistry (Korner, the soil carbon estimation upto 1.5 m depth leading to addition for
2007). efficient analysis of overall soil carbon in Himalayan conifer forests.
These estimates of carbon may be useful for a variety of purposes
4.3. Litter production related to carbon sequestration in forests and facilitate to achieve the
climate change responses i.e., mitigation and adaptation. The result is
Litterfall production decreased with altitude (Table 6) as also re- more precise as the study is based on primary data with sufficient
ported by Zhou et al. (2006). The significant negative relationships samples collected from a wide distributional range of Cedrus deodara
between elevations and reproductive litterfall production suggest that forests along altitude gradient and therefore is an update of the carbon
small changes in temperature may result in significant changes in re- estimate. These updated estimates can be used for assessing forestry
productive allocation. Increased allocation to reproduction may have contribution for emission reduction for the region as well as act as the
negative impacts on other functions such as root and leaf growth by altitude-wise default for the Himalayan region for the other Cedrus
depleting stores of carbon and nutrients (Chapin et al. 1990). If climate deodara forests. The estimate would also facilitate improved trade-off
change does cause montane forests to move to higher elevations as for forests between altitude wise carbon sequestrations and associated
predicted by many researchers, they may experience significant de- offsetting of polluting gases. The biomass and soil C estimate would also
creases in reproductive allocation which could decrease their pro- add value to forest-based incentive mechanism for climate change mi-
ductivity in the future (Saxe et al. 2001). The litterfall production varies tigation actions specially REDD + program. The precise estimate of
from 5.7 ± 1.04 to 7.6 ± 1.11 t ha−1 yr−1 which is within the range carbon would encourage better participation of the communities in the
of montane temperate forests, tropical dry deciduous forests, mangrove forest-based mitigation action as the total carbon by forests would be
and swamp, temperate evergreen, temperate dry evergreen, sub- measured more due to additional depth of soil carbon.
tropical-Montane (Chhabra and Dadhwal, 2004), however these values This study also provides specific information about the soil carbon
are much lower than tropical rain forests and other plantations. The being captured by the forests at different altitude. Henceforth it will
variation may be attributed to variation in the composition of species, add to the existing knowledge about the tree species-specific informa-
age and climate as these factors influence the litterfall production tion for more precise estimation at different altitudinal gradient of soil
Sharma et al. (2010a). The difference in litterfall production might be carbon and thus facilitate the use of Tier III approach of carbon esti-
due to variation in basal area or climate in between these sites. In all the mation. For estimating the net benefit with carbon trading under dif-
forest sites, maximum litter production was in summer season followed ferent scenarios for the region, this study may act as a baseline i.e.,
by rainy and winter seasons (Table 5). The seasonal pattern of litterfall reference point and thus supplement information for valuing for all
may be attributed to differences in climatic factors, such as temperature incremental carbon. The present study can be used as a reference point
and moisture and intrinsic genetic factors (Jamaludheen and Kumar, in the future, if REDD + projects would be implemented in Himalayan
1999). In the present study, the peak litterfall was observed in the forests. Moreover, the quantification would also add value to forest
summer season (March-June). The peak rate of litter falls of Cedrus management for comparing the values of the Cedrus deodara forests
deodara in summer season has also been reported in the adjoining Ku- with other forests, leading to better and species based forest manage-
moun Division of the Himalaya followed by rainy and winter seasons, ment.
which are similar to our study (Joshi et al., 1991). Average seasonal Author contribution Statement
(summer, rainy and winter) pattern of litter production (t ha−1 yr−1) MAS (Ph.D. Scholar), MK (Ph.D. Supervisor), NPT (Ph.D. Co-su-
among three different sites was significant (p < 0.05). However, the pervisor) have designed the study; MAS collected, compiled and ana-
total forest floor biomass and carbon were not significant. This may be lyzed the data, and developed a draft of the manuscript. MK, NPT and
attributed to the overall similar production of forest floor biomass along RP read the data, analysis and improved the draft and finally, all au-
with the altitude. thors approved the draft for submission.
Climate change is altering the location of species all over the planet
and with global warming, species are shifting towards the poles and up Declaration of Competing Interest
elevation to maintain their existence (Guo et al., 2018). Moreover;
warming due to climate change in the Himalaya leads to changes in The authors declare that they have no known competing financial
diversity and distribution of plant communities (Shrestha et al., 2012) interests or personal relationships that could have appeared to influ-
and assist in the upward displacement of the plant species (Telwala ence the work reported in this paper.
et al., 2013) and species range shifts to higher elevations. A shift of
23–998 m in species’ upper elevation limit with a mean upward dis- References
placement rate of 27.53–622.04 m/decade is reported with changes in
the plant assemblages and community structure in the Himalaya during I. Alonso K. Weston R. Gregg M. Morecroft Carbon storage by habitat - Review of the
last century (Telwala et al., 2013). The changes in species composition evidence of the impacts of management decisions and condition on carbon stores and
sources Natural England Research Reports, Number 2012 NERR043.
and structure are instrumental for changing in the biomass of the Batjes, N.H., 1996. Total carbon and nitrogen in the soils of the world. Eur. J. Soil Sci. 47,
community and therefore would affect the overall carbon flux of the 151–163.
community. Bitlerlich, W., 1984. The Relaskop Idea Slough: Commonwealth Agricultural Bureause.
Farnham Royal, England.
Cairns, M.A.S., Brown, E.H., Baumgardner, G.A., 1997. Root biomass allocation in the
5. Conclusion world‟s upland forests. Oecologia 111 (1), 1–11.
Champion, H.G., Seth, S.K., 1968. The Forest Types of India. Govt. of India, New Delhi.
Chan, Y.H., 1982. Storage and release of organic carbon in peninsular Malaysia. Int. J.
Out of ten species with highest growing stock in Indian forests; six Environ. Stud. 18, 211–222.
species i.e., Pinus roxburghii, Abies pindrow, Quercus semecarpifolia, Chapin, F.S., Schulze, E., Mooney, H.A., 1990. The ecology and economics of storage in
Cedrus deodara, Pinus excels and Abies smithiana are found in Indian plants. Annu. Rev. Ecol. Syst. 21, 423–447.
Chhabra, A., Dadhwal, V.K., 2004. Assessment of major pools and fluxes of carbon in
mountains (FSI, 2009). These component-wise estimates of the forests

8
M.A. Sheikh, et al. Ecological Indicators 111 (2020) 106025

Indian forests. Clim. Change 64, 341–360. subsoil on root development, water uptake and growth of Radiata pine. Tree Physiol.
Chhabra, A., Palria, S., Dadhwal, V.K., 2002. Growing stock based forest biomass estimate 10, 20.
for India. Biomass Bioenergy 22, 187–194. NOAA, 2017. National Oceanic and Atmospheric Administration. Atmospheric CO2 at
Chidumaya, E.N., 1990. Aboveground woody biomass structure and productivity in a Mauna Loa observatory [http://WWW.CO2NOW.org].
Zambezian woodland. For. Ecol. Manage. 36, 33–46. Odum, E.P., 1969. The strategy of ecosystem development. Science 164, 262–270.
Dinakaran, J., Krishnayya, N.S.R., 2008. Variation in type of vegetal cover and hetero- Pandey, R., 2009. Forest resource utilization by tribal community of Jaunsar. Indian For.
geneity of soil organic carbon in affecting sink capacity of tropical soils. Curr. Sci. 135 (5), 436–441.
94, 9. Pandey, R., Rawat, G.S., Kishwan, J., 2011. Changes in distributions of carbon in various
FAO, 2015. FAO assessment of forests and carbon stocks, 1990–2015: Reduced overall forest types of India from 1995–2005. Silva Lusitana 19 (1), 41–54.
emissions, but increased degradation. FAO Forestry Paper I4470E/1/03.15. Rome. Pandey, R., Hom, S.K., Harrison, S., Yadav, V.K., 2014. Mitigation potential of important
FSI, 2009. India State of Forest Report. Government of India, Dehra Dun, Forest Survey of farm and forest trees: a potentiality for clean development mechanism afforestation
India, Ministry of Environment and Forests. reforestation (CDMA R) project and reducing emissions from deforestation and de-
Gairola, S., Rawal, R.S., Todaria, N.P., 2008. Forest vegetation patterns along an altitu- gradation, along with conservation and enhancement of carbon stocks (REDD+).
dinal gradient in Sub-alpine zone of West Himalaya India. Afr. J. Plant Sci. 2 (6), Mitig Adapt Strateg Glob Change; DOI 10.1007/s11027-014-9591-2.
42–48. Persson, H., 1980. Spatial distribution of fine root growth mortality and decomposition in
Garkoti, S.C., Singh, S.P., 1995. Variation in net primary productivity and biomass of young Scots pine stand in Central Sweden. Oikos 34, 77–87.
forests in the high mountains of Central Himalaya. J. Veg. Sci. 6, 23–28. Pressler, M., 1895. Das Gesetz der Stambildung Leipzig.153.
Guo, F., Lenoir, J., Bonebrake, T.C., 2018. Land-use change interacts with climate to Rana, B.S., Singh, S.P., Singh, R.P., 1989. Biomass and net primary productivity in central
determine elevational species redistribution. Nat. Commun. 9 (1). https://doi.org/10. Himalayan forest along an altitudinal gradient. For. Ecol. Manage. 27, 199–218.
1038/s41467-018-03786-9. Reichle, D.E., 1981. Dynamic Properties of Forest Ecosystem. Cambridge Univ. Press,
Gupta, M.K., Pandey, R., 2008. Soil organic carbon pool under different plantations in Cambridge, pp. 657.
some districts of Uttrakhand and Haryana. Indian J. For. 31 (3), 369–374. Samina, U., Singh, S.P., Rawat, Y.S., 1999. Fine root productivity and turnover in two
Hairiah, K., Sitompul, S.M., Noordwijk, M., Palm, 2001. Methodology for sampling evergreen central Himalayan forests. Ann. Bot. 84, 87–94.
carbon stocks above and below ground. ASB Lecture Notes 4B. International Centre Sangarun, P.W., Srisang, K., Jaroensutasinee, M., 2007. Cloud Forest Characteristics of
for Research in Agro forestry, Indonesia, Published in http://www.icraf.cgiar.org/ Khao Nan. World Academy of Science, Engineering and Technology, Thailand,
sea. pp. 32.
Haripriya, G.S., 2000. Estimates of biomass in Indian forests. Biomass Bioenergy 19, Saxe, H., Cannell, G., Johnson, O., Ryan, M., Vourlitis, G., 2001. Tree and forest func-
245–258. tioning in response to global warming. New Phytol. 149, 369–400.
Houghton, R.A., Hall, F., Goetz, S.J., 2009. Importance of biomass in the global carbon Sevgi, O., Tecimen, H.B., 2008. Changes in Austrian Pine forest floor properties in relation
cycle. J. Geophys. Res. 114, G00E03. https://doi.org/10.1029/2009JG000935. with altitude in mountainous areas. J. For. Sci. 54, 306–313.
Houghton, R.A., 2005. Aboveground forest biomass and the global carbon balance. Global Sharma, C.M., Baduni, N.P., Gairola, S., Ghildiyal, S.K., Suyal, S., 2010a. The effect of
Change Biol. 11, 945–958. slope aspects on the forest composition, community structure and soil nutrient status
IPCC, 2014. Summary for policymakers. In: Stocker, T.F., Qin, D., Plattner, G.K., Tignor, of some major natural temperate forest types of Garhwal Himalaya. J. For. Res. 21
M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M. (Eds.), (3), 331–337.
“Climate change 2013: The Physical Science Basis”, Contribution of Working Group I Sharma, C.M., Baduni, N.P., Gairola, S., Ghildiyal, S.K., Suyal, S., 2010b. Tree diversity
to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. and carbon stocks of some major forest types of Garhwal Himalaya, India. For. Ecol.
Cambridge University Press, Cambridge. Manage. 260 (12), 2170–2179.
Jackson, M.L., 1967. Soil Chemical analysis. Prentice-Hall of India, Pvt. Ltd., New Delhi, Sheikh, M.A., Kumar, M., Bussmann, R.W., 2009. Altitudinal variation in soil organic
pp. 498. carbon stock in coniferous subtropical and broadleaf temperate forests in Garhwal
Jamaludheen, V., Kumar, B.M., 1999. Litter of multipurpose trees in Kerala, India: var- Himalaya. Carbon Balance Manage. 4, 1–6.
iations in the amount, quality, decay rates and release of nutrients. For. Ecol. Manage. Sheikh, M.A., Kumar, M., Bussman, R.W., Todaria, N.P., 2011. Forest carbon stocks and
115, 1–11. fluxes in physiographic zones of India. Carbon Balance Manage. 6, 15. http://www.
Jha, M.N., Gupta, M.K., Saxena, A., Kumar, R., 2003. Soil organic carbon store in different cbmjournal.com/content/6/1/15.
forests in India. Indian Forester 129 (6), 714–724. Shrestha, U.B., Gautam, S., Bawa, K.S., 2012. Widespread climate change in the
Jobbágy, E.G., Jackson, R.B., 2000. The vertical distribution of soil carbon and its relation Himalayas and associated changes in local ecosystems. PLoS ONE 7 (5), e36741.
to climate and vegetation. Ecol. Appl. 10, 423–436. Singh, J.S., Tiwari, A.K., Saxena, A.K., 1985. Himalayan forests: a net source of carbon to
Joshi, M., Mer, G.S., Singh, S.P., Rawat, Y.S., 1991. Seasonal pattern of total soil re- the atmosphere. Environ. Conserv. 12, 67–69.
spiration in undisturbed and disturbed ecosystems of Central Himalaya. Biol. Fertil. Singh, S.P., Adhikari, B.S., Zobel, D.B., 1994. Biomass, productivity, leaf longevity, and
Soils 11, 267–272. forest structure in the central Himalaya. Ecol. Monogr. 64, 401–421.
Kishwan, J., Pandey, R., Dadhwal, V.K., 2012. Emission removal capability of India’s Singh, S.P., Pande, K., Upadhyay, V.P., Singh, J.S., 1990. Fungal communities associated
forest and tree cover. Small Sc. For. 11 (1), 61–72. with the decomposition of a common leaf litter (Quercus leucotrichophora A. Camus)
Korner, C., 2007. The use of ‘altitude’ in ecological research. Trends Ecol. Evol. 22, along an elevational transect in the central Himalaya. Biol. Fertil. Soils 9, 245–251.
569–574. Telwala, Y., Brook, B.W., Manish, K., Pandit, M.K., 2013. Climate-induced Elevational
Koul, D.N., Panwar, P., 2008. Prioritizing land-management options for carbon seques- range shifts and increase in plant species richness in a Himalayan biodiversity epi-
tration potential. Curr. Sci. 95, 658–663. centre. PLoS ONE 8 (2), e57103. https://doi.org/10.1371/journal.pone.0057103.
Krishan, G., Srivastav, S.K., Kumar, S., Saha, S.K., Dadhwal, V.K., 2009. Quantifying the Tiwari, A.K., Singh, J.S., 1987. Analysis of forest land-use and vegetation in a part of
underestimation of soil organic carbon by the Walkley and Black technique – Central Himalaya, using aerial photographs. Environ. Conserv. 14, 233–244.
Examples from Himalayan and Central Indian soils. Curr. Sci. 96 (8), 25. UNFCCC, 2015. Measurements for Estimation of Carbon Stocks in Afforestation and
Kumar, M., Pandey, R., 2017. Carbon stock loss of Chir pine forest through tree felling in Reforestation Project Activities under the Clean Development Mechanism: A Field
Lower Himalaya. Environ. Risk Assess. Remediat. 1 (1), 19–21. Manual. UNFCCC, Germany.
Kumari, S., Mehta, J.P., Shafi, Snobar, Dhiman, P., 2018. Vegetational analysis of woody Upadhyay, V.P., Singh, J.S., 1989. Nitrogen release pattern in Nainital hills, India. Indian
vegetation in burnt and unburnt forest communities of Pauri Garhwal Himalaya. For. 115, 320–326.
Plant Arch. 18 (1), 135–143. Walkley, A., Black, I.A., 1934. An examination of the Degtjareff method for determining
Lal, R., 2005. Forest soils and carbon sequestration. For. Ecol. Manage. 220, 242–258. soil organic matter, and a proposed modification of the chromic acid titration
Li, P., Wang, Q., Endo, T., Zhao, X., Kakubari, Y., 2010. Soil organic carbon stock is method. Soil Sci. 34, 29–38.
closely related to aboveground vegetation properties in cold-temperate mountainous Whittaker, R.H., 1975. Community and Ecosystems, second ed. MacMillan, N.Y.
forests. Geoderma 154, 407–415. Wilde, S.A., Voigt, G.K., Iyer, J.G., 1964. Soil and Plant Analysis of Tree Culture. Oxford
Luna, R.K., 2005. Plantation Trees. International Book Distributors, Dehrudun. Publishing House, Calcutta, India.
Malik, J., Bhatt, A.B., Pandey, R., 2016. Anthropogenic disturbances and their impact on Woomer, P.L., 1999. Impact of cultivation of carbon fluxes in woody savannas of southern
vegetation in western Himalaya India. J. Mt. Sci. 13 (1), 69–82. https://doi.org/10. Africa. Water Air Soil Pollut. 70, 403–412.
1007/s11629-015-3533-7. Wutzler, T., Profft, I., Mund, M., 2007. Quantifying tree biomass carbon stocks, their
Miao, R., Ma, J., Liu, Y., Liu, Y., Yang, Z., Guo, M., 2019. Variability of aboveground litter changes and uncertainties using routine stand taxation inventory data. Silva Fennica
inputs alters soil carbon and nitrogen in a coniferous-broadleaf mixed forest of 45 (3), 359–377.
central China. Forests 10 (2), 188–211. Zhou, G.Y., Liu, S.G., Li, Z.A., Zhang, D.Q., Tang, X.L., Zhou, C.Y., Yan, J.H., Mo, J.M.,
Misra, R., 1968. Ecology Workbook. Oxford and IBH Publishing, Calcutta. 2006. Old-growth forests can accumulate carbon in soils. Science 314, 1417.
Nambiar, E.K.S., Sands, R., 1992. Effect of compaction and simulated root channels in the

You might also like