ArticlePDF Available

Agricultural Expansion and Intensification in the Foothills of Mount Kenya: A Landscape Perspective

MDPI
Remote Sensing
Authors:
  • Center for training and Integrated Research In ASAL Development

Abstract and Figures

This study spatially assesses, quantifies, and visualizes the agricultural expansion and land use intensification in the northwestern foothills of Mount Kenya over the last 30 years: processes triggered by population growth, and, more recently, by large-scale commercial investments. We made use of Google Earth Engine to access the USGS Landsat data archive and to generate cloud-free seasonal composites. These enabled us to accurately differentiate between rainfed and irrigated cropland, which was important for assessing agricultural intensification. We developed three land cover and land use classifications using the random forest classifier, and assessed land cover and land use change by creating cross-tabulation matrices for the intervals from 1987 to 2002, 2002 to 2016, and 1987 to 2016 and calculating the net change. We then applied a landscape mosaic approach to each classification to identify landscape types categorized by land use intensity. We discuss the impacts of landscape changes on natural habitats, biodiversity, and water. Kappa accuracies for the three classifications lay between 78.3% and 82.1%. Our study confirms that rainfed and irrigated cropland expanded at the expense of natural habitats, including protected areas. Agricultural expansion took place mainly in the 1980s and 1990s, whereas agricultural intensification largely happened after 2000. Since then, not only large-scale producers, but also many smallholders have begun to practice irrigated farming. The spatial pattern of agricultural expansion and intensification in the study area is defined by water availability. Agricultural intensification and the expansion of horticulture agribusinesses increase pressure on water. Furthermore, the observed changes have heightened pressure on pasture and idle land due to the decrease in natural and agropastoral landscapes. Conflicts between pastoralists, smallholder farmers, large-scale ranches, and wildlife might further increase, particularly during the dry seasons and in years of extreme drought.
Content may be subject to copyright.
remote sensing
Article
Agricultural Expansion and Intensification in the
Foothills of Mount Kenya: A Landscape Perspective
Sandra Eckert 1, *ID , Boniface Kiteme 2, Evanson Njuguna 2and Julie Gwendolin Zaehringer 1
1Centre for Development and Environment, University of Bern, Hallerstrasse 10, 3012 Bern, Switzerland;
julie.zaehringer@cde.unibe.ch
2Centre for Training and Integrated Research in ASAL Development, 10400 Nanyuki, Kenya;
b.kiteme@africaonline.co.ke (B.K.); e.njuguna@cetrad.org (E.N.)
*Correspondence: sandra.eckert@cde.unibe.ch; Tel.: +41-31-631-5439
Received: 3 July 2017; Accepted: 28 July 2017; Published: 31 July 2017
Abstract:
This study spatially assesses, quantifies, and visualizes the agricultural expansion and land
use intensification in the northwestern foothills of Mount Kenya over the last 30 years: processes
triggered by population growth, and, more recently, by large-scale commercial investments. We made
use of Google Earth Engine to access the USGS Landsat data archive and to generate cloud-free
seasonal composites. These enabled us to accurately differentiate between rainfed and irrigated
cropland, which was important for assessing agricultural intensification. We developed three land
cover and land use classifications using the random forest classifier, and assessed land cover and land
use change by creating cross-tabulation matrices for the intervals from 1987 to 2002, 2002 to 2016,
and 1987 to 2016 and calculating the net change. We then applied a landscape mosaic approach to each
classification to identify landscape types categorized by land use intensity. We discuss the impacts
of landscape changes on natural habitats, biodiversity, and water. Kappa accuracies for the three
classifications lay between 78.3% and 82.1%. Our study confirms that rainfed and irrigated cropland
expanded at the expense of natural habitats, including protected areas. Agricultural expansion took
place mainly in the 1980s and 1990s, whereas agricultural intensification largely happened after
2000. Since then, not only large-scale producers, but also many smallholders have begun to practice
irrigated farming. The spatial pattern of agricultural expansion and intensification in the study
area is defined by water availability. Agricultural intensification and the expansion of horticulture
agribusinesses increase pressure on water. Furthermore, the observed changes have heightened
pressure on pasture and idle land due to the decrease in natural and agropastoral landscapes. Conflicts
between pastoralists, smallholder farmers, large-scale ranches, and wildlife might further increase,
particularly during the dry seasons and in years of extreme drought.
Keywords:
land cover and land use change; landscape change; agricultural intensification;
greenhouse cultivation; environmental impacts; remote sensing; Kenya
1. Introduction
Changes in land cover and land use worldwide have reached an unprecedented pace, magnitude,
and spatial extent [
1
5
]. The changes affect and contribute to local, regional, and global aspects of
earth system functioning [
1
3
,
6
]. In African landscapes, pastoralism, shifting cultivation, permanent or
semi-permanent agriculture, and agroforestry have altered the environment to a point that the present
landscape is the product of human-induced changes as much as natural variation in vegetation [
7
].
Over the last decades, anthropogenic impacts and competition over land have become issues of
major concern [
8
,
9
]. Many areas in Africa are experiencing substantial human population growth,
and, as a result, a shift away from extensive pastoral livestock-keeping to subsistence-oriented
Remote Sens. 2017,9, 784; doi:10.3390/rs9080784 www.mdpi.com/journal/remotesensing
Remote Sens. 2017,9, 784 2 of 20
small-scale farming and agroforestry [
10
]. In addition, the last 10–15 years have also seen agricultural
intensification. Triggered mainly by large-scale agricultural investments, it has become an additional
important driver of land cover and land use change [8,10,11].
In Kenya, horticulture has grown faster over the last decade than any other industry in the
agricultural sector [
12
,
13
]. In the Mount Kenya region, substantial agricultural expansion driven by
population growth, in combination with recurring cycles of drought, has since the late 1990s been
causing conflicts both between upstream and downstream water users and between pastoralists and
farmers [
12
,
14
]. The more recent expansion of horticultural agribusinesses in the same area is further
increasing pressure on limited natural resources, particularly water and land [1517].
The livelihood systems of small-scale farmers in this region have been studied intensively for more
than 20 years [
18
23
]. However, to date, there is only little spatially explicit quantitative information on
agricultural expansion and land use intensification in the region, if any at all. It can only be estimated
to which land covers’ cost, and to what extent, land cover and land use changes related to agricultural
expansion and intensification have shaped the Mount Kenya region’s landscapes.
A growing body of literature focuses on the assessment of irrigated and rainfed cropland and
related changes using multi-temporal or multi-sensor data [
24
26
]. However, many studies focus on
the global or national scale, using multi-temporal MODIS satellite data that are too coarse to capture
small-scale irrigated plots or regional changes [
27
29
]. More recently, scholars have begun to explore
phenological profiles from Landsat time series, which has led to more accurate and more detailed
cropland identification and change assessments [
30
32
]. Some studies have successfully assessed
agricultural intensification in areas where it manifested itself in a change of land cover [
33
], a greater
number of cropping cycles [
34
], or an increase in center-pivot irrigation areas [
35
]. However, most of
these studies focus on pixel-level land cover and land use changes [
32
,
36
] rather than landscape-level
changes towards more intense land use systems.
Land change science offers a strong conceptual framework to analyze transitions in land use
systems dominated by smallholders [
37
]. It seeks to understand the dynamics of land cover and land
use as a coupled human-environmental system [
38
]. These dynamics lead to distinct spatial land cover
and land use patterns, creating mosaics of landscapes [
39
]. Landscape analysis is therefore considered
a suitable approach for monitoring these distinct land cover and land use patterns, and transitions
between them [
2
,
40
]. Geographical information sciences and remote sensing provide powerful tools to
undertake such research [37,41,42].
Against this background, the aim of this study is to spatially assess and quantify agricultural
expansion and land use intensification in the northwestern foothills of Mount Kenya over the last
30 years. Remote sensing-based land cover and land use classifications for three points in time
provided the basis for extracting land use intensification. We present a landscape mosaic approach
that enabled us, based on specific combinations of land cover and land use classes, to identify
different landscape types categorized by land use intensity. Further, we differentiated landscape
types according to the presence of woody biomass (e.g., trees, shrubs, and bushes). Focusing on
the landscape scale has enabled us to assess and visualize changes in landscape type—including
agricultural intensification—and to discuss their impacts on natural habitats, biodiversity, and water.
2. Materials and Methods
2.1. Study Area
The study area of 249,147 ha lies in the northwestern foothills of Mount Kenya, within the upper
Ewaso Ng’iro basin, and includes parts of Laikipia, Meru, and Nyeri counties. The upper Ewaso
Ng’iro basin encompasses steep ecological gradients, as it drops from 5199 m above sea level at the
summit of the mountain to an average altitude of 1500 m above sea level in the northwest. Climatic
conditions range from semi-humid (1000–1500 mm of rainfall annually) near Mount Kenya in the east
Remote Sens. 2017,9, 784 3 of 20
to semi-arid (400–900 mm rainfall) and arid (about 350 mm rainfall) towards the west [
43
]. The farther
away from Mount Kenya, the dryer the conditions.
There are two distinct rainy seasons per year. The long rains last from mid-March to mid-June
(sowing and planting time). They are followed by a dry season from June to September or early
October (harvesting time). A second, much shorter rainy season occurs in November. The two
rainy seasons largely determine the cropping calendar in this semi-arid, water-scarce area, as most
small-scale farmers rely on the seasonal rains. The rains are unreliable and unpredictable in terms
of onset, duration, and termination [
44
]. This variability impacts greatly on all natural resources
and particularly on water, which continues to become scarcer. The major river systems in the area
show a significant decline in water even though there has been no significant change in the rainfall
regime [
18
]. The growing number of water abstractions for irrigation, livestock, and domestic purposes
has intensified competition for this scarce resource [45].
The study area has experienced substantial land use conversions since the beginning of the 20th
century. Traditionally, the area was inhabited by the Maasai. At that time, much of the study area was
covered with fire-modified acacia bushland and grassland [
19
], which the Maasai used as pasture.
Subsequently, several major land use and socio-economic transitions occurred [
14
]. The dominant
land use changed from pastoralism to extensive large-scale farming and ranching, which was reserved
for European settlers [
46
]. After Kenya’s independence, the land was distributed to immigrating
small-scale subsistence farmers, shifting the dominant land use from extensive ranching to small-scale
mixed farming [
14
], and leading to an increase in the population of Laikipia County from 60,000 in
1960 to over 400,000 in 2009 [
47
]. The past 15 to 20 years, finally, have seen the development of a highly
technologized, export-oriented horticulture sector practicing greenhouse and high-input vegetable and
flower production [
48
]. Today, the area’s very fertile soils are used both by these high-input, large-scale
commercial farms and by smallholders. The geospatial pattern of the different land use systems is
largely determined by water availability. Closer to Mount Kenya, a dense population of smallholders
practices small-scale farming; more recently, large-scale horticulture and greenhouse farms have been
established in the area. In the drier areas farther away from Mount Kenya, land use is dominated by
agropastoralism and pastoralism, interspersed with wildlife reserves and tourist lodges. The study
area is depicted in Figure 1.
Remote Sens. 2017, 9, 784 3 of 20
There are two distinct rainy seasons per year. The long rains last from mid-March to mid-June
(sowing and planting time). They are followed by a dry season from June to September or early
October (harvesting time). A second, much shorter rainy season occurs in November. The two rainy
seasons largely determine the cropping calendar in this semi-arid, water-scarce area, as most small-
scale farmers rely on the seasonal rains. The rains are unreliable and unpredictable in terms of onset,
duration, and termination [44]. This variability impacts greatly on all natural resources and
particularly on water, which continues to become scarcer. The major river systems in the area show
a significant decline in water even though there has been no significant change in the rainfall regime
[18]. The growing number of water abstractions for irrigation, livestock, and domestic purposes has
intensified competition for this scarce resource [45].
The study area has experienced substantial land use conversions since the beginning of the 20th
century. Traditionally, the area was inhabited by the Maasai. At that time, much of the study area
was covered with fire-modified acacia bushland and grassland [19], which the Maasai used as
pasture. Subsequently, several major land use and socio-economic transitions occurred [14]. The
dominant land use changed from pastoralism to extensive large-scale farming and ranching, which
was reserved for European settlers [46]. After Kenya’s independence, the land was distributed to
immigrating small-scale subsistence farmers, shifting the dominant land use from extensive ranching
to small-scale mixed farming [14], and leading to an increase in the population of Laikipia County
from 60,000 in 1960 to over 400,000 in 2009 [47]. The past 15 to 20 years, finally, have seen the
development of a highly technologized, export-oriented horticulture sector practicing greenhouse
and high-input vegetable and flower production [48]. Today, the area’s very fertile soils are used both
by these high-input, large-scale commercial farms and by smallholders. The geospatial pattern of the
different land use systems is largely determined by water availability. Closer to Mount Kenya, a
dense population of smallholders practices small-scale farming; more recently, large-scale
horticulture and greenhouse farms have been established in the area. In the drier areas farther away
from Mount Kenya, land use is dominated by agropastoralism and pastoralism, interspersed with
wildlife reserves and tourist lodges. The study area is depicted in Figure 1.
Figure 1. Overview of the study area located in the foothills of Mount Kenya (map projection of
detailed study area map: UTM 37S).
Figure 1.
Overview of the study area located in the foothills of Mount Kenya (map projection of
detailed study area map: UTM 37S).
Remote Sens. 2017,9, 784 4 of 20
2.2. Assessing Land Cover and Land Use Change since the Late 1980s
2.2.1. Satellite Data Preprocessing
For this study, we queried the USGS Landsat data archives for Landsat 5, Landsat 7, and Landsat
8 scenes using the Google Earth Engine (GEE) cloud computing environment. We worked with the
Landsat Ecosystem Disturbance Adaptive Processing System (LEDAPS) surface reflectance products
of Landsat 5 TM, Landsat 7 ETM+, and Landsat 8 OLI provided by USGS. These products are already
geometrically coregistered, orthorectified, and atmospherically corrected. They are provided together
with a cloud mask and a quality assessment (QA) band.
We generated three different image collections representing the situation in 1987, 2002, and 2016.
In order to obtain cloud-free seasonal composites of surface reflectance we had to include several years
of imagery for each point in time. For 1987, we used imagery acquired between 1984 and 1987; for 2002,
we used data acquired between 1999 and 2002; and for 2016 we used data acquired between 2014
and 2016. Figure S1 in the Supplementary Materials provides an overview of the numbers of scenes
that were available for each month and year, indicating which ones were used for this study. For all
scenes acquired within these three time periods, we calculated the Normalized Difference Vegetation
Index (NDVI). Then, we used the LEDAPS QA band to remove clouded pixels, resulting in a stack
of cloud-free pixel values for each pixel location, which we then reduced to a monthly composite by
choosing the median pixel value for each pixel and for each month. The use of the median pixel value
ensures that outlier values (e.g., due to cloud shadows or clouds that were not previously removed
by the QA band) are excluded. This was done for each optical Landsat band as well as the NDVI.
Based on a visual check, we finally downloaded those monthly composites for the three time periods
that (1) contained no or few no-data areas (due to masked clouded pixels), and (2) did not contain
any cloud shadows. This resulted in three raster data stacks that contained three to four monthly
composites representing the study area in one of the two dry seasons and at least one of the two wet
seasons. Such seasonal composites representing key phenological time windows can be helpful for
separating certain land cover and land use classes in a reliable way [49,50].
2.2.2. Land Cover and Land Use Classification and Change Analysis
The field reference data required for training and validating the classifier were collected during a
field visit in February 2016. Additional reference data were digitized in Google Earth, which offered
high-resolution data captured in 2001, 2003, 2013, and 2016. For the 1984–1987 composite, reference data
were collected by inspecting the Landsat composites and delineating representative samples together
with fellow researchers who have been working in the study area and following its development since
the early 1980s, and have therefore acquired a great deal of local expertise [14,22,44,46,51].
We defined land cover and land use classes that reflect the natural vegetation cover in the
study area, as well as ones that reflect land covers and land uses that developed with increasing
human activity in the study area (Table 1). Further, we distinguished “rainfed cropland” from
“irrigated cropland”, and included a “greenhouses” land use class to identify and assess agricultural
intensification in the study area.
Table 1. Land cover and land use classes characterizing the landscapes in the study area.
Land Cover and Land Use Class Description
Bare land Bare soil including dirt roads, rock outcrops, and sand
Cropland
Rainfed cropland Plots of varying size covered with crops or ploughed
Irrigated cropland Plots showing a high amount of green vegetation cover during dry seasons
Savannah grassland Grassland interspersed with bushes, shrubs, and trees at a low to medium density,
and fallows with a vegetation cover
Bush- and shrubland Areas with a bush and shrub cover of medium to high density and an understory
that is bare or covered with grass or dry matter
Remote Sens. 2017,9, 784 5 of 20
Table 1. Cont.
Land Cover and Land Use Class Description
Forest Natural or plantation forests, including riparian forests, and very densely grown
bush- and shrubland with a high amount of green vegetation
Waterbodies Small and shallow natural waterbodies and larger artificial reservoirs for irrigation
Settlements Settlements, large buildings, tarmac
Greenhouses Glass or plastic greenhouses
The three data stacks representing typical dry- and wet-season conditions in 1987, 2002, and 2016
were split up into homogeneous subsets of predominantly natural habitats and predominantly cropland
areas in order to avoid misclassifications between bush- and shrubland and rainfed cropland, which
have very similar spectral characteristics.
All subsets were classified using random forest (RF), an ensemble method for supervised
classification, and regression trees (CART), first developed by Breiman in [
52
]. RF is a high-performance
machine learning algorithm based on an ensemble of decision trees. Even slight variations in training
data cause CARTs to differ significantly in their structure. This characteristic of CARTs can be combined
with bootstrap aggregation and random feature selection to create independent predictors [
53
]. It has
many benefits compared to traditional classifiers [
52
,
54
,
55
]. RF is relatively insensitive to the number
and multi-collinearity of input data, and makes no assumptions about distributions. Furthermore, it has
been shown to provide reliable and stable classification results, outperforming other classifiers [
56
59
].
In our classification, we used 1000 trees for the RF model, and the number of selected features was set
as the square root of all features. The Gini coefficient served as the impurity criterion. The accuracy
of the resulting land cover and land use classifications was assessed using 40% (1987), 51% (2002),
and 57% (2016) of the collected field reference data as independent validation points. We calculated
overall accuracy, class-wise user ’s and producer’s accuracies, as well as kappa and F1 accuracies [
60
].
The overall accuracies for 1987, 2002, and 2016 range between 83.8% and 86.7%. The kappa accuracies
lie between 78.3% and 82.1%, and the average F1 accuracies between 83.6% and 87.8%. Detailed class
accuracies are indicated in Table S1 in the Supplementary Materials.
We assessed land cover and land use change for each pixel of the study area by creating
cross-tabulation matrices for the intervals from 1987 to 2002, from 2002 to 2016, and from 1987 to 2016,
and by calculating net change. Furthermore, when analyzing class changes, we accounted for the
spatial extent of the various land cover and land use classes, as this makes it possible to differentiate
between random changes and systematic change processes [61,62].
2.2.3. A Landscape Mosaic Approach to Capture the Intensification of Agricultural Land Use
Land cover and land use change maps are of limited suitability for assessing land use intensification
processes and landscape changes. This is because certain landscape types, as well as certain
intensification processes and landscape changes, are hard to capture spatially by remote sensing.
Doing so requires analyzing changes in the combinations of land covers and land uses in a certain
context area. To address this problem, Messerli et al. [
63
] introduced an approach that interprets
a pixel’s land cover by taking into account human-environment interactions and the condition of
neighboring pixels. The approach was successfully applied to distinguish shifting cultivation systems
from permanent land use systems in Laos and Madagascar [
64
,
65
]. Arvor et al. [
40
] developed a similar
approach to map and analyze the soybean agricultural frontier in Mato Grosso, Brazil. In the present
study, we adapted the approach to the different land use systems found in Kenya in order to identify
land use intensification and landscape changes. The following paragraphs describe how we proceeded.
First, we defined the size of the context area to be considered as a square of 2
×
2 km. This is large
enough to contain the typical combinations of land cover and land use classes that define the different
land use systems of varying land use intensity (e.g., greenhouses with outdoor irrigated cropland
and irrigation water reservoirs). Once the size was defined, we used focal statistics to calculate class
Remote Sens. 2017,9, 784 6 of 20
percentages per context area. Then, we developed a matrix of landscape types based on five categories
of agricultural land use intensity and three categories of woody biomass cover, as well as a decision
rule set for each category.
As a guideline for categorization, we used the conceptual model of frontier dynamics proposed
by DeFries et al. [
2
]. The five agricultural land use intensity categories were defined based on (1) the
share of natural land cover classes in the context area, and (2) the relation between rainfed cropland,
irrigated cropland, and greenhouses. This resulted in the following intensity categories:
Natural landscape: Natural vegetation cover classes (bare land, savannah grassland, bush- and
shrubland, and forest) cover more than 80% of the context area (I1)
Agropastoralism: Savannah grassland covers a greater share of the context area than rainfed
cropland, but cropland covers at least 5% of the context area (I2)
Rainfed farming: Rainfed cropland and irrigated cropland together cover more than 20% of the
context area, but the share of rainfed cropland is larger than that of irrigated cropland (I3)
Irrigated farming: Rainfed cropland and irrigated cropland together cover more than 20% of the
context area, but the share of irrigated cropland is larger than that of rainfed cropland (I4)
Large-scale commercial farming: Greenhouses and waterbodies together cover more than 3% of
the context area (I5)
The four woody biomass categories were defined based on the amount and type of woody
biomass present in the context area. This categorization makes it possible to capture change in forest
and woodland areas, which is closely linked to biodiversity [
66
] as well as to agroforestry land use
systems, in which tree crops play an important role. Agroforestry is considered to be an important
sustainable farming system in Africa [67].
High forest cover: Forest covers at least 20% of the context area (FO)
Mostly bush- and shrubland: Bush- and shrubland and savannah grassland together cover a
larger share of the context area than forest (BS)
Little woody biomass: Bare land covers at least 20% of the context area (BA)
No woody biomass: All areas that do not fall in one of the above categories, e.g., monoculture
cropland (NO)
The resulting matrix of landscape types is presented in Figure 2.
Remote Sens. 2017, 9, 784 6 of 20
on five categories of agricultural land use intensity and three categories of woody biomass cover, as
well as a decision rule set for each category.
As a guideline for categorization, we used the conceptual model of frontier dynamics proposed
by DeFries et al. [2]. The five agricultural land use intensity categories were defined based on (1) the
share of natural land cover classes in the context area, and (2) the relation between rainfed cropland,
irrigated cropland, and greenhouses. This resulted in the following intensity categories:
Natural landscape: Natural vegetation cover classes (bare land, savannah grassland, bush- and
shrubland, and forest) cover more than 80% of the context area (I1)
Agropastoralism: Savannah grassland covers a greater share of the context area than rainfed
cropland, but cropland covers at least 5% of the context area (I2)
Rainfed farming: Rainfed cropland and irrigated cropland together cover more than 20% of the
context area, but the share of rainfed cropland is larger than that of irrigated cropland (I3)
Irrigated farming: Rainfed cropland and irrigated cropland together cover more than 20% of the
context area, but the share of irrigated cropland is larger than that of rainfed cropland (I4)
Large-scale commercial farming: Greenhouses and waterbodies together cover more than 3% of
the context area (I5)
The four woody biomass categories were defined based on the amount and type of woody
biomass present in the context area. This categorization makes it possible to capture change in forest
and woodland areas, which is closely linked to biodiversity [66] as well as to agroforestry land use
systems, in which tree crops play an important role. Agroforestry is considered to be an important
sustainable farming system in Africa [67].
High forest cover: Forest covers at least 20% of the context area (FO)
Mostly bush- and shrubland: Bush- and shrubland and savannah grassland together cover a
larger share of the context area than forest (BS)
Little woody biomass: Bare land covers at least 20% of the context area (BA)
No woody biomass: All areas that do not fall in one of the above categories, e.g., monoculture
cropland (NO)
The resulting matrix of landscape types is presented in Figure 2.
Figure 2. Landscape types categorized by land use intensity and woody biomass cover (adapted from
Zaehringer et al. [65]).
The five binary maps representing the five intensity categories are largely complementary, but
some overlaps nonetheless exist. For this reason, we combined them according to their intensity
levels, with the map showing the most intensive category at the top and the one showing the least
Figure 2.
Landscape types categorized by land use intensity and woody biomass cover (adapted from
Zaehringer et al. [65]).
The five binary maps representing the five intensity categories are largely complementary,
but some overlaps nonetheless exist. For this reason, we combined them according to their intensity
levels, with the map showing the most intensive category at the top and the one showing the least
Remote Sens. 2017,9, 784 7 of 20
intensive one at the bottom. The four binary maps representing woody biomass were processed
analogously, with the map showing the category with the most woody biomass at the top and the one
showing the category with the least woody biomass at the bottom. The resulting two categorical maps
were then combined to produce a landscape mosaic map. This procedure was repeated for each of the
three points in time.
The resulting three landscape mosaic maps depict aggregated land use information at the
landscape level for the three points in time studied. This makes it possible to detect changes that affect
entire land use systems, as well as to assess intensification processes [
65
]. In order to analyze landscape
change and quantify the intensification of agricultural land use, we cross-tabulated the landscape
mosaic statistics for the three points in time.
The landscape mosaic maps were evaluated based on visual inspection and local expertise.
A quantitative assessment at the landscape level is not feasible based on the available field reference
data. Furthermore, single pixel classification errors in the land cover maps have little influence on the
landscape-level mosaic maps.
3. Results
3.1. Land Cover and Land Use Changes in the Study Area
Table 2shows the net changes in area that occurred in each land cover and land use class over
the last 30 years, split up into two time intervals. A look at the changes over the entire 30 years
shows strikingly that massive changes between 1987 and 2016 occurred mainly in four classes: while
savannah grassland and bush- and shrubland decreased by 46,105 ha and 11,837 ha, respectively,
cropland (rainfed and irrigated) increased by 47,752 ha. This change is also clearly visible in Figure 3.
A look at the two partial intervals shows that the transition to cropland first occurred on the
northwestern slopes and foothills closer to Mount Kenya, followed by areas farther away from the
mountain and areas near rivers. Smaller amounts of savannah grassland were converted to forest,
bush- and shrubland, and irrigated cropland. While the increase in rainfed cropland is much smaller
during the second interval (0.85%), irrigated cropland continues to increase at a similar rate as during
the first interval (3.87%). The conversion to irrigated cropland mainly happened near Mount Kenya,
near greenhouses or irrigation water reservoirs, or along rivers, where riparian forests and wetlands
were converted to irrigated or rainfed cropland.
A number of forest plantations were established during the first interval, increasing the forest
area by 8509 ha. During the second interval, the total forest area shrank again (by 605 ha) due to the
harvesting of trees from plantations and the conversion of riparian and other small forest patches to
irrigated and rainfed cropland. Class changes between savannah grassland, bush- and shrubland,
and forest were also observed. These can partly be attributed to natural succession and tree harvesting,
but in part they are also caused by variability in green vegetation cover between the three points in
time investigated due to rainfall variability.
Table 2.
Net area changes for each land cover and land use class within the study area in hectares (ha)
and percent (%) of the total area analyzed between 1987 and 2002, between 2002 and 2016, and across
the entire time interval from 1987 to 2016.
Net Changes 1987–2002 2002–2016 1987–2016
ha % ha % ha %
Rainfed cropland 28,740 11.6 2079 0.8 29,438 11.9
Bare land 228 0.1 1026 0.4 3351 1.4
Waterbodies 24 0.0 73 0.0 97 0.0
Irrigated cropland 8882 3.6 9515 3.9 18,315 7.4
Savannah grassland 41,023 16.6 6030 2.5 46,105
18.7
Forest 8509 3.4 605 0.2 7816 3.2
Settlements 306 0.1 31 0.0 322 0.1
Greenhouses 21 0.0 604 0.2 624 0.3
Bush- and shrubland 5689 2.3 6693 2.7 11,837 4.8
Remote Sens. 2017,9, 784 8 of 20
Remote Sens. 2017, 9, 784 8 of 20
Figure 3. Land cover and land use classification maps of the study area around 1987, 2002, and 2016
(map projection: UTM 37S).
Settlements grew considerably during the first interval, while waterbodies and greenhouses
expanded mostly during the second interval. Between 1987 and 2016, greenhouses increased by 624
ha and waterbodies by 97 ha. The appearance of greenhouses surrounded by small water reservoirs
can be observed particularly along the main road connecting Nairobi with the towns of Naro Moru,
Nanyuki, and Timau. Figure 4A shows this impressive development east of Timau, while Figure 4B
shows the conversion of savannah grassland, wetlands, as well as riparian and other small forests to
rainfed and irrigated cropland northwest of Naro Moru, along the Naro Moru and Burguret rivers.
Areas cultivated by large-scale investors are mostly used for highly technologized commercial
floriculture or horticulture. Floriculture farms are typically characterized by greenhouses and water
storage reservoirs, while horticulture farms are generally surrounded by irrigated cropland.
Areas classified as greenhouses in 2016 were formerly either used as rainfed cropland (297 ha)
or converted from savannah grassland (239 ha), bush- and shrubland (37 ha), and forest (25 ha).
Detailed land cover and land use class change matrices for the two time intervals are provided in
Tables S2 and S3 in the Supplementary Materials.
Figure 3.
Land cover and land use classification maps of the study area around 1987, 2002, and 2016
(map projection: UTM 37S).
Settlements grew considerably during the first interval, while waterbodies and greenhouses
expanded mostly during the second interval. Between 1987 and 2016, greenhouses increased by 624 ha
and waterbodies by 97 ha. The appearance of greenhouses surrounded by small water reservoirs
can be observed particularly along the main road connecting Nairobi with the towns of Naro Moru,
Nanyuki, and Timau. Figure 4A shows this impressive development east of Timau, while Figure 4B
shows the conversion of savannah grassland, wetlands, as well as riparian and other small forests to
rainfed and irrigated cropland northwest of Naro Moru, along the Naro Moru and Burguret rivers.
Areas cultivated by large-scale investors are mostly used for highly technologized commercial
floriculture or horticulture. Floriculture farms are typically characterized by greenhouses and water
storage reservoirs, while horticulture farms are generally surrounded by irrigated cropland.
Areas classified as greenhouses in 2016 were formerly either used as rainfed cropland (297 ha) or
converted from savannah grassland (239 ha), bush- and shrubland (37 ha), and forest (25 ha). Detailed
land cover and land use class change matrices for the two time intervals are provided in Tables S2
and S3 in the Supplementary Materials.
Remote Sens. 2017,9, 784 9 of 20
Remote Sens. 2017, 9, 784 9 of 20
Figure 4. Two subsets of the land cover and land use classifications for 1987, 2002, and 2016 showing
(A) the development of greenhouses and irrigation water reservoirs east of Timau, and (B) the
conversion of savannah grassland, forests mostly along rivers, and wetlands northwest of Naro Moru,
along the Naro Moru and Burguret rivers to rainfed and irrigated cropland (map projection: UTM
37S).
3.2. Landscape Changes in the Study Area
The availability of aggregated land use information at the landscape level makes it possible to
detect changes that affect entire land use systems, as well as to assess and visualize processes of
cropland expansion and use intensification (Figure 5).
In 1987, the study area was dominated by natural vegetation and agropastoral land use systems,
which together covered around 75% of the study area. In 2016, this was no longer the case, with these
landscape types now covering only 45% of study area (Figures 6 and 7). This reduction happened at
the expense of landscapes characterized by natural vegetation (forest, bush- and shrubland,
grassland; 22%). Such areas were converted, mostly during the first interval, in the vicinity of Mount
Kenya, where rainfall in those years was comparably reliable; during the second interval, the change
affected natural landscapes in the lower, more arid plains farther away from Mount Kenya but near
rivers and wetlands. Most of these landscapes were converted to smallholder farmland (23%)
dominated by rainfed, low-input agriculture (BS-I3, BA-I3). One exception is an area northeast of
Timau, where large-scale cereal cultivation in monoculture (NO-I3) was practiced already before
1987. Here, there was little change over the past 30 years, except in some plots that were given to
smallholder farmers practicing multicropping and agroforestry, resulting in a conversion from NO-
Figure 4.
Two subsets of the land cover and land use classifications for 1987, 2002, and 2016 showing
(
A
) the development of greenhouses and irrigation water reservoirs east of Timau, and (
B
) the
conversion of savannah grassland, forests mostly along rivers, and wetlands northwest of Naro
Moru, along the Naro Moru and Burguret rivers to rainfed and irrigated cropland (map projection:
UTM 37S).
3.2. Landscape Changes in the Study Area
The availability of aggregated land use information at the landscape level makes it possible to
detect changes that affect entire land use systems, as well as to assess and visualize processes of
cropland expansion and use intensification (Figure 5).
In 1987, the study area was dominated by natural vegetation and agropastoral land use systems,
which together covered around 75% of the study area. In 2016, this was no longer the case, with these
landscape types now covering only 45% of study area (Figures 6and 7). This reduction happened at
the expense of landscapes characterized by natural vegetation (forest, bush- and shrubland, grassland;
22%). Such areas were converted, mostly during the first interval, in the vicinity of Mount Kenya,
where rainfall in those years was comparably reliable; during the second interval, the change affected
natural landscapes in the lower, more arid plains farther away from Mount Kenya but near rivers
and wetlands. Most of these landscapes were converted to smallholder farmland (23%) dominated
by rainfed, low-input agriculture (BS-I3, BA-I3). One exception is an area northeast of Timau, where
large-scale cereal cultivation in monoculture (NO-I3) was practiced already before 1987. Here, there
was little change over the past 30 years, except in some plots that were given to smallholder farmers
Remote Sens. 2017,9, 784 10 of 20
practicing multicropping and agroforestry, resulting in a conversion from NO-I3 to FO-I3 or FO-I4.
Throughout the study area, an increase in tree cover—mostly due to the adoption of agroforestry
systems—occurred particularly in areas with sufficient water availability, that is, in the semi-humid
areas near Mount Kenya or along perennial rivers. Since 2002, agroforestry has expanded along rivers,
where riparian forests and wetlands were converted mainly to irrigated cropland. Agroforestry is also
increasingly found at higher elevations near the boundaries of the Mount Kenya National Park and
National Forest, where water is more readily available.
Remote Sens. 2017, 9, 784 10 of 20
I3 to FO-I3 or FO-I4. Throughout the study area, an increase in tree cover—mostly due to the adoption
of agroforestry systems—occurred particularly in areas with sufficient water availability, that is, in
the semi-humid areas near Mount Kenya or along perennial rivers. Since 2002, agroforestry has
expanded along rivers, where riparian forests and wetlands were converted mainly to irrigated
cropland. Agroforestry is also increasingly found at higher elevations near the boundaries of the
Mount Kenya National Park and National Forest, where water is more readily available.
Figure 5. Landscape mosaic maps for 1987, 2002, and 2016, overlaid with waterways (map projection:
UTM 37S). The map for 2016 additionally shows different types of protected or managed areas and
the road network. Land use intensity increases with the color gradient changing from dark green
towards dark pink. The darker and more intense the green and olive colors, the more woody biomass
there is. The more intense the orange, red, and pink colors, the less woody biomass there is and the
more intensively the land is used.
Figure 5.
Landscape mosaic maps for 1987, 2002, and 2016, overlaid with waterways (map projection:
UTM 37S). The map for 2016 additionally shows different types of protected or managed areas and the
road network. Land use intensity increases with the color gradient changing from dark green towards
dark pink. The darker and more intense the green and olive colors, the more woody biomass there
is. The more intense the orange, red, and pink colors, the less woody biomass there is and the more
intensively the land is used.
Remote Sens. 2017,9, 784 11 of 20
Remote Sens. 2017, 9, 784 11 of 20
Figure 6. (A) Shares of landscape types in the total study area in percent for 1987, 2002, and 2016, and
(B) gains and losses in each landscape type expressed as percentages of the total study area for the
three intervals from 1987 to 2002, from 2002 to 2016, and from 1987 to 2016. FO = high forest cover, BS
= mostly bush- and shrubland, BA = little woody biomass, NO = no woody biomass.
Figure 6.
(
A
) Shares of landscape types in the total study area in percent for 1987, 2002, and 2016,
and (
B
) gains and losses in each landscape type expressed as percentages of the total study area for the
three intervals from 1987 to 2002, from 2002 to 2016, and from 1987 to 2016. FO = high forest cover,
BS = mostly bush- and shrubland, BA = little woody biomass, NO = no woody biomass.
Remote Sens. 2017,9, 784 12 of 20
Remote Sens. 2017, 9, 784 12 of 20
Figure 7. Sankey plot showing changes from one landscape type to another between 1987, 2002, and
2016. The colors and labels follow the scheme shown in Figure 2.
Mixed agropastoral landscape types remained fairly stable in size, but shifted farther away from
Mount Kenya towards more arid areas. Irrigated farming was still rare in 1987 (<1% of study area),
but became more prominent and appeared on the landscape mosaic map in 2002 (2.5%); in 2016,
finally, it covered 4% of the study area. As mentioned above, it is mainly practiced by horticulture
farms along the main road from Nanyuki to Timau, as well as by small-scale farmers, who mostly
practice it in conjunction with agroforestry and have reliable access to irrigation water (FO-I4 and BS-
I4). Another feature that first appears in the 2002 landscape mosaic map is the horticulture and
floriculture hotspot area (FO-I4 and FO-I5) between Nanyuki and Timau. At that time, there were no
greenhouses in the study area, and flowers were grown in open fields. However, greenhouses are
needed for the production of high-value, high-quality flowers [68], which the floriculture sector
started investing in around 2002 [69,70]. In 2016, landscapes characterized by such high-intensity
agriculture already covered 3% of the study area. Most of them are located along rivers and along or
only a short distance away from the main road.
3.3. Changes in Tree Cover and Woody Biomass
Landscapes with a high amount of woody biomass (FO and BS) experienced an increase between
1987 and 2002, followed by a slight decrease until 2016 (see Figure 8). The increase is mainly due to
an expansion of riparian forests, and, even more so, of bush- and shrubland west to northwest of
Naro Moru. By 2016, the area covered by these landscapes shrank again; the establishment of several
tree plantations east of Nanyuki was unable to compensate for the loss of natural, forest-dominated
landscapes that occurred, for example, north of Timau as well as south and southwest of Naro Moru,
where large tracts of natural landscape were converted to rainfed farmland (clearly visible in the 2016
landscape mosaic map). Landscapes with little woody biomass have increased slightly since 1987,
but at a low level. Landscapes without woody biomass show an overall increase since 1987, but have
been on the decline since 2002.
Figure 7.
Sankey plot showing changes from one landscape type to another between 1987, 2002,
and 2016. The colors and labels follow the scheme shown in Figure 2.
Mixed agropastoral landscape types remained fairly stable in size, but shifted farther away from
Mount Kenya towards more arid areas. Irrigated farming was still rare in 1987 (<1% of study area),
but became more prominent and appeared on the landscape mosaic map in 2002 (2.5%); in 2016, finally,
it covered 4% of the study area. As mentioned above, it is mainly practiced by horticulture farms along
the main road from Nanyuki to Timau, as well as by small-scale farmers, who mostly practice it in
conjunction with agroforestry and have reliable access to irrigation water (FO-I4 and BS-I4). Another
feature that first appears in the 2002 landscape mosaic map is the horticulture and floriculture hotspot
area (FO-I4 and FO-I5) between Nanyuki and Timau. At that time, there were no greenhouses in
the study area, and flowers were grown in open fields. However, greenhouses are needed for the
production of high-value, high-quality flowers [
68
], which the floriculture sector started investing
in around 2002 [
69
,
70
]. In 2016, landscapes characterized by such high-intensity agriculture already
covered 3% of the study area. Most of them are located along rivers and along or only a short distance
away from the main road.
3.3. Changes in Tree Cover and Woody Biomass
Landscapes with a high amount of woody biomass (FO and BS) experienced an increase between
1987 and 2002, followed by a slight decrease until 2016 (see Figure 8). The increase is mainly due to
an expansion of riparian forests, and, even more so, of bush- and shrubland west to northwest of
Naro Moru. By 2016, the area covered by these landscapes shrank again; the establishment of several
tree plantations east of Nanyuki was unable to compensate for the loss of natural, forest-dominated
landscapes that occurred, for example, north of Timau as well as south and southwest of Naro Moru,
where large tracts of natural landscape were converted to rainfed farmland (clearly visible in the 2016
landscape mosaic map). Landscapes with little woody biomass have increased slightly since 1987,
but at a low level. Landscapes without woody biomass show an overall increase since 1987, but have
been on the decline since 2002.
Remote Sens. 2017,9, 784 13 of 20
Figure 8.
Area (ha) covered by each of the four woody biomass categories in 1987, 2002, and 2016.
FO = high forest cover, BS = mostly bush- and shrubland, BA = little woody biomass, NO = no
woody biomass.
3.4. Landscape Changes in Protected Areas
Most larger natural landscapes that have remained stable in the study area are protected areas
where land uses other than conservation are restricted. The most prominent protected area in the
study area is Mount Kenya National Park and National Forest, which has experienced only minor
changes along its boundaries. These changes are either from one natural landscape category to another
or from landscapes dominated by forest or bush- and shrubland to landscapes dominated by rainfed
and irrigated agroforestry. In contrast, the small Lusoi Forest Reserve in the south of the study area
experienced substantial change. In 1987, it was surrounded by a vast area of natural bush- and
shrubland. By 2016, most of this landscape disappeared, and about half of the reserve was converted
into a landscape dominated by rainfed farming, albeit still surrounded by a large area of bush- and
shrubland. In the west, partially outside the study area, lies the Ol Pejeta Conservancy, which today
encompasses about 360 km
2
. A cattle ranch in colonial times, it became a wildlife reserve in the
1980s and gradually increased in size to its current extent. Although its natural landscape has largely
remained stable, a notable change occurred in the southern part of the reserve, which was converted
to large-scale monoculture wheat fields after 1987. This change is clearly visible in the landscape
mosaic maps for 2002 and 2016. The 2016 map further shows that land use intensified along the rivers
forming the borders of the Ol Pejeta Conservancy. Besides these protected areas, the study area also
contains two private ranches: Solio Ranch in the south, dedicated to conserving rhinoceros, and Borana
Ranch and Conservancy in the north. On Borana Ranch, the natural landscape still covers a large
area, whereas in the southeastern part of Solio Ranch it was recently converted to rainfed smallholder
farmland, which is clearly visible on the landscape mosaic map for 2016.
4. Discussion
Important prerequisites for this study were the successful and accurate mapping of the most
important natural vegetation covers, the differentiation between rainfed and irrigated cropland, and the
Remote Sens. 2017,9, 784 14 of 20
identification of greenhouses and irrigation water reservoirs. All three could only be achieved by
using multi-seasonal satellite data composites, making subsets of predominantly natural habitats and
predominantly cropland areas, and verifying the mapping results through careful visual comparison
with Google Earth imagery. We would like to note that settlement areas might be under-represented in
our land cover and land use maps, as most individual buildings and small settlements in the study area
are too small to be captured clearly in the Landsat data. Furthermore, variations in the shares of natural
vegetation covers and bare soil, as well as of rainfed cropland and irrigation cropland, between the
three points in time investigated might partially be due to interannual variations in seasonal rainfall.
Thus, the observed increase in irrigated cropland might be slightly overestimated. The same applies
to the increase in rainfed cropland at the cost of savannah grassland, a small portion of which may
have been caused by misclassifications between the two classes due to the very heterogeneous and
small-scale pattern of cropland and grazing plots. We estimate these errors to be within one percent of
net change.
We would further like to note that the accuracies of the three pixel-based land cover and land use
change analyses are only as good as the multiplied accuracies of each individual land cover and land
use classification [
71
]. However, single pixel classification errors in these classification maps have little
influence on the landscape-level mosaic maps.
Depending on how agricultural intensification manifests itself, a pixel-level assessment may
make sense. This is the case, for example, when natural vegetation cover changes to cropland [
72
],
when plot size increases [
73
], when irrigated cropland increases [
35
], or when the number of cropping
cycles increases [
34
]. However, this study shows that when agricultural intensification manifests itself
in land use system changes, an assessment at the landscape level may be better suited to capture
these processes. Based on the idea that certain landscape types can only be captured by analyzing
the combination of land covers and land uses in a certain context area, we defined landscape type
categories and developed a decision rule set for each category that enabled us to differentiate landscape
types by land use intensity and the amount of woody biomass present. We defined the categories
based on the landscape compositions observed in the field. The category definitions might need to
be adjusted if the approach is transferred to a different agroecological zone composed of different
landscapes and land use systems.
The results presented in this paper show impressively how much, where, and at the expense of
which landscapes cropland has expanded and land use has intensified since 1987. The transformation
from extensive agropastoral use to rainfed smallholder farming mainly happened between 1987 and
2002, but it still continues wherever land, and, more importantly, water is available. Between 2002 and
2016, we can observe a shift away from purely rainfed to increasingly irrigated smallholder farming,
as well as the development of high-input commercial vegetable and flower farming, which is practiced
in open fields and increasingly also in greenhouses. Commercial growers require year-round irrigation
to supply export markets regardless of the season. Indeed, commercial water use is highest precisely
when the least water is available: one reason being that European demand for vegetables peaks during
the study area’s driest period. The same is true of flowers, with Valentine’s Day falling in February,
the driest month in the study area. Increasing commercial use strains water resources during times of
scarcity and sets the stage for conflicts among different water users [
48
,
51
]. To avoid such conflicts
while further increasing production, most horticulture farms in the study area have begun to reduce
their dependence on local river water. Over the last decade, they have invested in the construction
of irrigation water reservoirs and boreholes. This explains the increase in surface water by 97 ha
between 1987 and 2016. Furthermore, most commercial farms harvest rainwater from the greenhouse
roofs and apply drip irrigation. Nonetheless, it remains unknown whether the intensification of
water use will have long-term impacts on the groundwater table, be it in the highlands or in the
Merti aquifer downstream. Furthermore, intensive small- and large-scale farming is impairing the
water’s quality, as it affects the physico-chemical properties of surface waters [
16
,
74
76
]. Muriithi
and Yu [
16
] took water samples in the study area and examined several physico-chemical parameters.
Remote Sens. 2017,9, 784 15 of 20
They found high pollutant concentrations in areas with intensive small-scale farming and large-scale
commercial horticulture.
Our results regarding land use system changes further show that natural habitats in the study
area have decreased in the last 30 years. Some of the protected areas—which are important habitats for
Kenya’s wildlife—are continuously shrinking. The southern part of the Ol Pejeta Conservancy was
converted into large-scale monoculture wheat fields. In the case of Solio Ranch, the Government of
Kenya purchased the southeastern part of the rhinoceros reserve in 2007 to resettle landless small-scale
farmers, leading to a transformation of the natural habitat into rainfed farmland [
77
]. The farmers
had previously been evicted from their original homes in the Mount Kenya and Aberdare forests
due to concerns about adverse environmental impacts of the area’s overpopulation, and since then
had had to live as squatters in the surrounding towns. Bond [
15
] found that, besides water, the local
population perceives land and pasture as the main natural resources related to conflicts in Laikipia
County. The potential for conflict between pastoralists, smallholder farmers, large-scale ranches,
and wildlife is particularly prevalent during the dry seasons, when migrating pastoralists move into
the area in search of pasture, water, and idle land to feed their livestock [15].
The massive reduction in natural vegetation and the intensification of agricultural land use
also have consequences for biodiversity and cause land degradation. The replacement of bush- and
shrubland, grassland, and forests with crops reduces plant species diversity, although mixed cropping
and agroforestry conserve native plant species better than single-crop farming [
10
]. In formerly less
forested areas, such as the area northeast of Timau that used to consist mainly of large monoculture
wheat farms, mixed cropping and agroforestry has led to an increase in trees, which may also have led
to an increase in biodiversity by attracting new species of birds [
10
]. However, wherever farming is
intensified and habitat diversity is reduced, biodiversity also declines. Land degradation is affected
in a similar way: soil erosion increases with decreasing vegetation cover, and the productivity of
agricultural land declines unless manure or other fertilizers are added. The main drivers of land
degradation besides soil erosion are the depletion of organic matter (soil organic carbon), degradation
of the soil structure, and a decline in the availability of major nutrients (N, P, K) and micro elements in
the soil. Furthermore, toxicity may increase due to acidification and salinization, especially in land
used for irrigated farming [10].
In sum, agricultural intensification and the expansion of horticulture agribusinesses further
increase pressure on the study area’s limited natural resources, potentially aggravating the condition
of the environment and the situation of disadvantaged land users. These developments certainly offer
new economic opportunities, but they also come with constraints for sustainable
development [17,78]
.
Therefore, it is important to ensure that agricultural production in Africa is intensified in a
sustainable manner.
5. Conclusions
This study provides spatially explicit information about the expansion of farming and its more
recent intensification in the northwestern foothills of Mount Kenya. We analyzed how these changes
affect the landscapes in the study area. We made use of Google Earth Engine to access the USGS
Landsat data archive and to generate cloud-free seasonal composites. These seasonal composites then
enabled us to differentiate between rainfed and irrigated cropland with considerably high accuracy
(F1 class values ranging between 73.5% and 84.7% for the two classes). By applying a landscape
mosaic approach to the land cover and land use classifications, we were able to derive landscape types
categorized by land use intensity and the amount of woody biomass present. The results and their
analysis led us to the following conclusions:
Rainfed and irrigated cropland expanded by 47,752 ha, mainly at the expense of savannah
grassland, bush- and shrubland, and forest, which showed overall losses of 46,105 ha, 11,837 ha,
and 605 ha, respectively. This amounts to a 30% decrease in natural habitats in the study area
over the last 30 years. The conversion to rainfed cropland mainly happened between 1987 and
Remote Sens. 2017,9, 784 16 of 20
2002, although it continued on after that at a much lower level. The intensity of agricultural
land use began to increase between 2002 and 2016, as further humid forest, bush- and shrubland,
and grassland areas along rivers, as well as rainfed cropland areas were converted into irrigated
cropland. Not only large-scale producers, but also many smallholders have begun to practice
irrigated farming. In addition, the area has seen a rapid development of high-input commercial
greenhouse horticulture farming (since 2002, greenhouses increased by 604 ha and irrigation
water reservoirs by 73 ha).
Natural wildlife habitats continue to shrink. Agricultural expansion and intensification affects
not only non-protected areas, but also private ranches and wildlife reserves as well as small
forest reserves in the study area. However, Mount Kenya National Park and National Forest
remained fairly stable. The overall forested area has decreased only slightly thanks to a number
of afforestation projects near the boundary of Mount Kenya National Park and National Forest.
The massive reduction in natural habitats and the intensification of agriculture have diverse
impacts on biodiversity. While the observed reduction in natural habitats has reduced biodiversity
at the regional level, the observed increase in agroforestry farming has increased it locally.
The changes also affect land degradation. Potential future consequences of agricultural
intensification include soil erosion and—unless fertilizer is applied—a decline in the soil’s organic
matter, degradation of the soil structure, and a reduction in major nutrients.
Water availability defines the spatial pattern of agricultural expansion and intensification in the
study area. Water has always been a scarce resource in the region. Agricultural intensification
and the expansion of horticulture agribusinesses further increase pressure on this limited natural
resource. Furthermore, the observed changes have also heightened pressure on pasture and
idle land due to the decrease in natural and agropastoral landscapes. As a result, conflicts
between pastoralists, smallholder farmers, large-scale ranches, and wildlife might further increase,
particularly during the dry seasons and in years of extreme drought.
Spatially explicit information on agricultural expansion and intensification is highly relevant to
understand patterns of land use change and their impacts on the environment and human well-being.
Particularly in developing countries, which often lack such spatial information and where land use
is undergoing substantial change, such up-to-date spatial information can support policymakers,
land use planners, and land users in achieving sustainable agricultural intensification.
Supplementary Materials:
The following are available online at www.mdpi.com/2072-4292/9/8/784/s1,
Figure S1: Number of Landsat surface reflectance products sorted by year, month, and season that were available
in Google Earth Engine. Collections highlighted in green were chosen for the analysis. Table S1: Overall and
class-wise accuracies for 1987 (A), 2002 (B), and 2016 (C). Table S2: Matrix of land cover and land use class changes
between 1987 and 2002. Table S3: Matrix of land cover and land use class changes between 2002 and 2016.
Acknowledgments:
The research for this publication was conducted as part of the BELMONT Forum and
FACCE–JPI project “African Food, Agriculture, Land and Natural Resource Dynamics, in the context of global
agro-food-energy system changes (AFGROLAND)” (Grant Number: 40FA40_160405). The project is funded by
the Swiss National Science Foundation, the French National Research Agency, and the South African National
Research Foundation.
Author Contributions:
Sandra Eckert and Julie Gwendolin Zaehringer conceived and designed the experiments;
Sandra Eckert processed and analyzed the data and wrote the paper; Boniface Kiteme and Evanson Njuguna
contributed to the collection of field reference data and to the writing of the paper; and Julie Gwendolin Zaehringer
contributed to the writing of the paper as well.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Lambin, E.F.; Turner, B.L.; Geist, H.J.; Agbola, S.B.; Angelsen, A.; Bruce, J.W.; Coomes, O.T.; Dirzo, R.;
Fischer, G.; Folke, C.; et al. The causes of land-use and land-cover change: Moving beyond the myths.
Glob. Environ. Chang. 2001,11, 261–269. [CrossRef]
Remote Sens. 2017,9, 784 17 of 20
2.
DeFries, R.S.; Foley, J.A.; Asner, G.P. Land-use choices: Balancing human needs and ecosystem function.
Front. Ecol. Environ. 2004,2, 249–257. [CrossRef]
3.
Lambin, E.F.; Meyfroidt, P. Global land use change, economic globalization, and the looming land scarcity.
Proc. Natl. Acad. Sci. USA 2011,108, 3465–3472. [CrossRef] [PubMed]
4.
Mustard, J.F.; Defries, R.S.; Fisher, T.; Moran, E. Land-use and land-cover change pathways and impacts.
In Land Change Science; Springer: Dordrecht, The Netherlands, 2012; pp. 411–429.
5.
Hansen, M.C.; Potapov, P.V.; Moore, R.; Hancher, M.; Turubanova, S.A.; Tyukavina, A.; Thau, D.;
Stehman, S.V.; Goetz, S.J.; Loveland, T.R.; et al. High-Resolution Global Maps of 21st-Century Forest
Cover Change. Science 2013,342, 850–853. [CrossRef] [PubMed]
6.
Foley, J.A.; DeFries, R.; Asner, G.P.; Barford, C.; Bonan, G.; Carpenter, S.R.; Chapin, F.S.; Coe, M.T.; Daily, G.C.;
Gibbs, H.K.; et al. Global Consequences of Land Use. Science 2005,309, 570–574. [CrossRef] [PubMed]
7.
Bongers, F.; Tennigkeit, T. Degraded Forests in Eastern Africa: Management and Restoration; Routledge: London,
UK, 2010.
8.
Brink, A.B.; Bodart, C.; Brodsky, L.; Defourney, P.; Ernst, C.; Donney, F.; Lupi, A.; Tuckova, K. Anthropogenic
pressure in East Africa—Monitoring 20 years of land cover changes by means of medium resolution satellite
data. Int. J. Appl. Earth Obs. Geoinform. 2014,28, 60–69. [CrossRef]
9.
Niamir-Fuller, M.; Kerven, C.; Reid, R.; Milner-Gulland, E. Co-existence of wildlife and pastoralism on
extensive rangelands: Competition or compatibility? Pastor. Res. Policy Pract. 2012,2, 1–14. [CrossRef]
10.
Maitima, J.M.; Mugatha, S.M.; Reid, R.S.; Gachimbi, L.N.; Majule, A.; Lyaruu, H.; Pomery, D.; Mathai, S.;
Mugisha, S. The linkages between land use change, land degradation and biodiversity across East Africa.
Afr. J. Environ. Sci. Technol. 2009,3, 310–325.
11.
Cotula, L.; International Institute for Environment and Development; Food and Agriculture Organization of
the United Nations; International Fund for Agricultural Development. Land Grab or Development Opportunity?
Agricultural Investment and International Land Deals in Africa; IIED: London, UK; FAO: Rome, Italy; IFAD:
Rome, Italy, 2009.
12.
Notter, B.; MacMillan, L.; Viviroli, D.; Weingartner, R.; Liniger, H.-P. Impacts of environmental change on
water resources in the Mt. Kenya region. J. Hydrol. 2007,343, 266–278. [CrossRef]
13.
Ministry of Agriculture, Livestock and Fisheries, Government of Kenya, National Horticultural Policy.
Available online: http://extwprlegs1.fao.org/docs/pdf/ken147935.pdf (accessed on 3 July 2017).
14.
Wiesmann, U. Sustainable Regional Development in Rural Africa: Conceptual Framework and Case Studies from
Kenya; Geographica Bernensia: Bern, Switzerland, 1998.
15.
Bond, J. Conflict, Development and Security at the Agro–Pastoral–Wildlife Nexus: A Case of Laikipia County,
Kenya. J. Dev. Stud. 2014,50, 991–1008. [CrossRef]
16.
Muriithi, F.K.; Yu, D. Understanding the Impact of Intensive Horticulture Land-Use Practices on Surface
Water Quality in Central Kenya. Environments 2015,2, 521–545. [CrossRef]
17.
Ulrich, A. Export-Oriented Horticultural Production in Laikipia, Kenya: Assessing the Implications for Rural
Livelihoods. Sustainability 2014,6, 336–347. [CrossRef]
18.
Ulrich, A.; Ifejika Speranza, C.; Roden, P.; Kiteme, B.; Wiesmann, U.; Nüsser, M. Small-scale farming in
semi-arid areas: Livelihood dynamics between 1997 and 2010 in Laikipia, Kenya. J. Rural Stud.
2012
,28,
241–251. [CrossRef]
19.
Roden, P.; Bergmann, C.; Ulrich, A.; Nüsser, M. Tracing divergent livelihood pathways in the drylands:
A perspective on two spatially proximate locations in Laikipia County, Kenya. J. Arid Environ.
2016
,124,
239–248. [CrossRef]
20.
McCord, P.F.; Cox, M.; Schmitt-Harsh, M.; Evans, T. Crop diversification as a smallholder livelihood strategy
within semi-arid agricultural systems near Mount Kenya. Land Use Policy 2015,42, 738–750. [CrossRef]
21.
Ogalleh, S.; Vogl, C.; Eitzinger, J.; Hauser, M. Local Perceptions and Responses to Climate Change and
Variability: The Case of Laikipia District, Kenya. Sustainability 2012,4, 3302–3325. [CrossRef]
22.
Wiesmann, U. Socioeconomic Viewpoints on Highland-Lowland Systems: A Case Study on the Northwest
Side of Mount Kenya. Mt. Res. Dev. 1992,12, 375–381. [CrossRef]
23.
Kohler, T.; Wiesmann, U.M. Kleinbäuerliche Besiedlung und die Ausprägung von Haushaltsstrategien am
Mount Kenya. In Welt der Alpe—Gebirge der Welt; Jeanneret, F., Wastl-Walter, D., Wiesmann, U., Schwyn, M.,
Eds.; Haupt Verlag: Bern, Switzerland, 2003; pp. 185–196.
Remote Sens. 2017,9, 784 18 of 20
24.
Atzberger, C. Advances in Remote Sensing of Agriculture: Context Description, Existing Operational
Monitoring Systems and Major Information Needs. Remote Sens. 2013,5, 949–981. [CrossRef]
25.
Ozdogan, M.; Yang, Y.; Allez, G.; Cervantes, C. Remote sensing of irrigated agriculture: Opportunities and
challenges. Remote Sens. 2010,2, 2274–2304. [CrossRef]
26.
Thenkabail, P.S.; Wu, Z. An automated cropland classification algorithm (ACCA) for Tajikistan by combining
Landsat, MODIS, and secondary data. Remote Sens. 2012,4, 2890–2918. [CrossRef]
27.
Estel, S.; Kuemmerle, T.; Levers, C.; Baumann, M.; Hostert, P. Mapping cropland-use intensity across Europe
using MODIS NDVI time series. Environ. Res. Lett. 2016,11. [CrossRef]
28.
Gumma, M.K.; Thenkabail, P.S.; Teluguntla, P.; Rao, M.N.; Mohammed, I.A.; Whitbread, A.M. Mapping
rice-fallow cropland areas for short-season grain legumes intensification in South Asia using MODIS 250 m
time-series data. Int. J. Digit. Earth 2016,9, 981–1003. [CrossRef]
29.
Teluguntla, P.; Thenkabail, P.S.; Xiong, J.; Gumma, M.K.; Congalton, R.G.; Oliphant, A.; Poehnelt, J.; Yadav, K.;
Rao, M.; Massey, R. Spectral matching techniques (SMTs) and automated cropland classification algorithms
(ACCAs) for mapping croplands of Australia using MODIS 250-m time-series (2000–2015) data. Int. J.
Digit. Earth 2016,10, 1–34. [CrossRef]
30.
Müller, H.; Rufin, P.; Griffiths, P.; Barros Siqueira, A.J.; Hostert, P. Mining dense Landsat time series for
separating cropland and pasture in a heterogeneous Brazilian savanna landscape. Remote Sens. Environ.
2015,156, 490–499. [CrossRef]
31.
Jin, N.; Tao, B.; Ren, W.; Feng, M.; Sun, R.; He, L.; Zhuang, W.; Yu, Q. Mapping Irrigated and Rainfed Wheat
Areas Using Multi-Temporal Satellite Data. Remote Sens. 2016,8, 207. [CrossRef]
32.
Knauer, K.; Gessner, U.; Fensholt, R.; Forkuor, G.; Kuenzer, C. Monitoring agricultural expansion in Burkina
Faso over 14 years with 30 m resolution time series: the role of population growth and implications for the
environment. Remote Sens. 2017,9, 132. [CrossRef]
33.
Morton, D.C.; DeFries, R.S.; Shimabukuro, Y.E.; Anderson, L.O.; Arai, E.; del Bon Espirito-Santo, F.;
Freitas, R.; Morisette, J. Cropland expansion changes deforestation dynamics in the southern Brazilian
Amazon. Proc. Natl. Acad. Sci. USA 2006,103, 14637–14641. [CrossRef] [PubMed]
34.
Kontgis, C.; Schneider, A.; Ozdogan, M. Mapping rice paddy extent and intensification in the Vietnamese
Mekong River Delta with dense time stacks of Landsat data. Remote Sens. Environ.
2015
,169, 255–269.
[CrossRef]
35.
Maron, M.; Fitzsimons, J.A. Agricultural intensification and loss of matrix habitat over 23 years in the West
Wimmera, south-eastern Australia. Biol. Conserv. 2007,135, 587–593. [CrossRef]
36.
Edlinger, J.; Conrad, C.; Lamers, J.; Khasankhanova, G.; Koellner, T. Reconstructing the spatio-temporal
development of irrigation systems in Uzbekistan using Landsat time series. Remote Sens.
2012
,4, 3972–3994.
[CrossRef]
37.
Turner, B.L.; Lambin, E.F.; Reenberg, A. The emergence of land change science for global environmental
change and sustainability. Proc. Natl. Acad. Sci. USA 2007,104, 20666–20671. [CrossRef] [PubMed]
38.
Reenberg, A. Land systems research in Denmark: Background and perspectives. Geogr. Tidsskr. Dan. J. Geogr.
2006,106, 1–6. [CrossRef]
39.
Dubreuil, V.; Laques, A.-É.; Nédélec, V.; Arvor, D.; Gurgel, H. Paysages et fronts pionniers amazoniens sous
le regard des satellites: L’exemple du Mato Grosso. L’Espace Géographique 2008,37, 57–74. [CrossRef]
40.
Arvor, D.; Dubreuil, V.; Simões, M.; Bégué, A. Mapping and spatial analysis of the soybean agricultural
frontier in Mato Grosso, Brazil, using remote sensing data. GeoJournal 2013,78, 833–850. [CrossRef]
41.
Gutman, G.; Janetos, A.C.; Justice, C.O.; Moran, E.F.; Mustard, J.F.; Rindfuss, R.R.; Skole, D.; Turner, B.L., II;
Cochrane, M.A. Land Change Science: Observing, Monitoring and Understanding Trajectories of Change on the
Earth’s Surface; Springer: Dordrecht, The Netherlands, 2004.
42.
Kuemmerle, T.; Erb, K.; Meyfroidt, P.; Müller, D.; Verburg, P.H.; Estel, S.; Haberl, H.; Hostert, P.; Jepsen, M.R.;
Kastner, T.; et al. Challenges and opportunities in mapping land use intensity globally. Curr. Opin. Environ.
Sustain. 2013,5, 484–493. [CrossRef] [PubMed]
43.
Berger, P. Rainfall and Agroclimatology of the Laikipia Plateau, Kenya; Geographica Bernensia: Berne, Switzerland,
1989.
44.
Liniger, H.; Gikonyo, J.; Kiteme, B.; Wiesmann, U. Assessing and Managing Scarce Tropical Mountain Water
Resources: The Case of Mount Kenya and the Semiarid Upper Ewaso Ng’iro Basin. Mt. Res. Dev.
2005
,25,
163–173. [CrossRef]
Remote Sens. 2017,9, 784 19 of 20
45.
Aeschbacher, J.; Liniger, H.; Weingartner, R. River water shortage in a highland-lowland system: A case
study of the impacts of water abstraction in the Mount Kenya region. Mt. Res. Dev.
2005
,25, 155–162.
[CrossRef]
46.
Kohler, T. Land Use in Transition: Aspects and Problems of Small Scale Farming in a New Environment: The Example
of Laikipia District, Kenya; Geographica Bernensia: Bern, Switzerland, 1987.
47.
KNBS—Kenya National Bureau of Statistics. Available online: https://www.knbs.or.ke/ (accessed on
6 June 2017).
48.
Lanari, N.; Liniger, H.; Kiteme, B.P. Commercial Horticulture in Kenya: Adapting to Water Scarcity; University of
Bern: Bern, Switzerland, 2016.
49.
Griffiths, P.; Müller, D.; Kuemmerle, T.; Hostert, P. Agricultural land change in the Carpathian ecoregion
after the breakdown of socialism and expansion of the European Union. Environ. Res. Lett.
2013
,8, 1–12.
[CrossRef]
50.
Griffiths, P.; Kuemmerle, T.; Baumann, M.; Radeloff, V.C.; Abrudan, I.V.; Lieskovsky, J.; Munteanu, C.;
Ostapowicz, K.; Hostert, P. Forest disturbances, forest recovery, and changes in forest types across the
Carpathian ecoregion from 1985 to 2010 based on Landsat image composites. Remote Sens. Environ.
2014
,
151, 72–88. [CrossRef]
51.
Kiteme, B.P.; Liniger, H.; Notter, B.; Wiesmann, U.; Kohler, T. Dimensions of Global Change in African
Mountains: The Example of Mount Kenya. In International Human Dimensions Programme on Global
Environmental Change: Mountain Regions: Laboratories for Adaptation; Rechkemmer, A., Ed.; IHDP: Bonn,
Germany, 2008; Volume 2, pp. 18–23.
52. Breiman, L. Random Forests. Mach. Learn. 2001,45, 5–32. [CrossRef]
53.
Jakimow, B.; Oldenburg, C.; Rabe, A.; Waske, B.; van der Linden, S.; Hostert, P. Manual for Application:
ImageRF (1.1). Available online: http://dev.geo.hu-berlin.de/enmap-box/documentation/applications/
imageRF/help/imageRF_Manual.pdf (accessed on 31 July 2017).
54.
Friedman, J.; Hastie, T.; Tibshirani, R. The Elements of Statistical Learning; Springer: New York, NY, USA, 2001.
55.
Gislason, P.O.; Benediktsson, J.A.; Sveinsson, J.R. Random forests for land cover classification. Pattern
Recognit. Lett. 2006,27, 294–300. [CrossRef]
56.
Belgiu, M.; Drăgu¸t, L. Random forest in remote sensing: A review of applications and future directions.
ISPRS J. Photogramm. Remote Sens. 2016,114, 24–31. [CrossRef]
57.
Karlson, M.; Ostwald, M.; Reese, H.; Sanou, J.; Tankoano, B.; Mattsson, E. Mapping tree canopy cover and
aboveground biomass in Sudano-Sahelian woodlands using Landsat 8 and random forest. Remote Sens.
2015
,
7, 10017–10041. [CrossRef]
58.
Li, X.; Chen, W.; Cheng, X.; Wang, L. A comparison of machine learning algorithms for mapping of complex
surface-mined and agricultural landscapes using ZiYuan-3 stereo satellite imagery. Remote Sens.
2016
,8, 514.
[CrossRef]
59.
Rodriguez-Galiano, V.F.; Ghimire, B.; Rogan, J.; Chica-Olmo, M.; Rigol-Sanchez, J.P. An assessment of the
effectiveness of a random forest classifier for land-cover classification. ISPRS J. Photogramm. Remote Sens.
2012,67, 93–104. [CrossRef]
60.
Congalton, R.G.; Green, K. Assessing the Accuracy of Remotely Sensed Data: Principles and Practices; CRC Press:
Boca Raton, LA, USA, 2008.
61.
Pontius, R.G.; Huffaker, D.; Denman, K. Useful techniques of validation for spatially explicit land-change
models. Ecol. Model. 2004,179, 445–461. [CrossRef]
62.
Aldwaik, S.Z.; Pontius, R.G. Intensity analysis to unify measurements of size and stationarity of land changes
by interval, category, and transition. Landsc. Urban Plan. 2012,106, 103–114. [CrossRef]
63.
Messerli, P.; Heinimann, A.; Epprecht, M. Finding homogeneity in heterogeneity—A new approach to
quantifying landscape mosaics developed for the Lao PDR. Hum. Ecol.
2009
,37, 291–304. [CrossRef]
[PubMed]
64.
Hett, C.; Castella, J.-C.; Heinimann, A.; Messerli, P.; Pfund, J.-L. A landscape mosaics approach for
characterizing swidden systems from a REDD+ perspective. Appl. Geogr. 2012,32, 608–618. [CrossRef]
65.
Zaehringer, J.G.; Hett, C.; Ramamonjisoa, B.; Messerli, P. Beyond deforestation monitoring in conservation
hotspots: Analysing landscape mosaic dynamics in north-eastern Madagascar. Appl. Geogr.
2016
,68, 9–19.
[CrossRef]
Remote Sens. 2017,9, 784 20 of 20
66.
Cumming, D.H.; Fenton, M.B.; Rautenbach, I.L.; Taylor, R.D.; Cumming, G.S.; Cumming, M.S.; Dunlop, J.M.;
Ford, A.G.; Hovorka, M.D.; Johnston, D.S. Elephants, woodlands and biodiversity in southern Africa. S. Afr.
J. Sci. 1997,93, 231–236.
67.
Mbow, C.; Van Noordwijk, M.; Luedeling, E.; Neufeldt, H.; Minang, P.A.; Kowero, G. Agroforestry solutions
to address food security and climate change challenges in Africa. Curr. Opin. Environ. Sustain.
2014
,6, 61–67.
[CrossRef]
68.
Gebreeyesus, M.; Sonobe, T. Global Value Chains and Market Formation Process in Emerging Export Activity:
Evidence from Ethiopian Flower Industry. J. Dev. Stud. 2012,48, 335–348. [CrossRef]
69.
Xia, Y.; Deng, X.; Zhou, P.; Shima, K.; da Silva, J.A.T. The World Floriculture Industry: Dynamics of
production and markets. In Floriculture, Ornamental and Plant Biotechnology: Advances and Topical Issues.
Available online: http://www.globalsciencebooks.info/Books/images/FOPBVolume5Outline.pdf (accessed
on 31 July 2017).
70.
Whitaker, M.; Kolavalli, S. Floriculture in Kenya. In Technology, Adaptation, and Exports—How Some Developing
Countries Got it Right; Chandra, V., Ed.; The World Bank: Washington, DC, USA, 2006; pp. 335–367.
71.
Verburg, P.H.; Neumann, K.; Nol, L. Challenges in using land use and land cover data for global change
studies. Glob. Chang. Biol. 2011,17, 974–989. [CrossRef]
72.
Galford, G.L.; Mustard, J.F.; Melillo, J.; Gendrin, A.; Cerri, C.C.; Cerri, C.E. Wavelet analysis of MODIS time
series to detect expansion and intensification of row-crop agriculture in Brazil. Remote Sens. Environ.
2008
,
112, 576–587. [CrossRef]
73.
White, E.V.; Roy, D.P. A contemporary decennial examination of changing agricultural field sizes using
Landsat time series data. Geo Geogr. Environ. 2015,2, 33–54. [CrossRef] [PubMed]
74.
Kibichii, S.; Shivoga, W.A.; Muchiri, M.; Miller, S.N. Macroinvertebrate assemblages along a land-use
gradient in the upper River Njoro watershed of Lake Nakuru drainage basin, Kenya. Lakes Reserv. Res.
Manag. 2007,12, 107–117. [CrossRef]
75.
Otiang’a-Owiti, G.E.; Oswe, I.A. Human impact on lake ecosystems: The case of Lake Naivasha, Kenya.
Afr. J. Aquat. Sci. 2007,32, 79–88. [CrossRef]
76.
Nyakundi, W.O.; Magoma, G.; Ochora, J.; Nyende, A.B. Survey of pesticide use and application patterns
among farmers: A case study from selected horticultural farms in Rift Valley and Central provinces, Kenya.
In Proceedings of the 2010 JKUAT Scientific Technological and Industrialization Conference, Nairobi, Kenya,
17–19 November 2010; pp. 618–630.
77.
Meersohn Meinecke, S.M.; Pandey, A.; Thant, A.M.; Madalinska, A.; Sørensen, C.; Wilson, M.
Crop Production and Livelihood Strategies in a Resettlement Village in Kenya. Available online:
http://www.tropentag.de/2012/abstracts/full/371.pdf (accessed on 31 July 2017).
78.
Zaehringer, J.G.; Wambugu, G.; Kiteme, B.P.; Eckert, S. Large-scale agricultural investments on the western
slopes of Mount Kenya: Perceived impacts on small-scale farmers’ land use and the environment. Ecol. Soc.
2017, submitted.
©
2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... Because the natural landscapes of Mt Kenya are increasingly isolated and surrounded by land converted to rain-fed cropland and non-native forest plantations (including mixed agriculture small plot shamba farming, timber plantations, and livestock grazing) (Eckert et al., 2017), a secondary objective of our project is to establish which wildlife species can occupy disturbed habitat outside of protected areas. This study sheds light on elevational patterns in large mammal diversity, a largely missing component of elevational patterns across taxa, which tends to be dominated by work on smaller and less mobile taxa ( Figure 1). ...
... Below an elevation of about 2200-2300 m the natural forest cover of Mt Kenya region has been almost completely lost and is replaced by agricultural fields and exotic timber plantations (Cupressus Hagenia-Hypericum forest (3220-3285 m) (Table 1) (Hedberg, 1951). (Bussman, 1994;Eckert et al., 2017). In all stages this habitat type is under intensive land use and occupation, including the corresponding wildlife trapping/hunting. ...
... Above 2401 m, the National Park management area is strictly off limits to consumptive extraction or livestock use ( Figure 2). Given that the conversion of natural habitat into agricultural habitat has rapidly increased over the last 40 years (Eckert et al., 2017), there is a sharp distinction between protected habitats at elevations above 2401 m in along the Burguret Trail and mostly disturbed habitat at lower elevations. ...
Article
Full-text available
Conservation areas encompassing elevation gradients are biodiversity hotspots because they contain a wide range of habitat types in a relatively small space. Studies of biodiversity patterns along elevation gradients, mostly on small mammal or bird species, have documented a peak in diversity at mid elevations. Here, we report on a field study of medium and large mammals to examine the impact of elevation, habitat type, and gross primary productivity on community structure. Species richness was observed using a camera trap transect with 219 sites situated across different habitat types from 2329 to 4657 m above the sea level on the western slope of Mt Kenya, the second highest mountain in Africa. We found that the lowest elevation natural habitats had the highest species richness and relative abundance and that both metrics decreased steadily as elevation increased, paralleling changes in gross primary productivity, and supporting the energy richness hypothesis. We found no evidence for the mid‐domain effect on species diversity. The lowest elevation degraded Agro‐Forestry lands adjacent to the National Park had high activity of domestic animals and reduced diversity and abundance of native species. The biggest difference in community structure was between protected and unprotected areas, followed by more subtle stepwise differences between habitats at different elevations. Large carnivore species remained relatively consistent but dominant herbivore species shifted along the elevation gradient. There was some habitat specialization and turnover in species, such that the elevation gradient predicts a high diversity of species, demonstrating the high conservation return for protecting mountain ecosystems for biodiversity conservation.
... Land use/land cover classifications were used to determine the change in land use/land cover classes during the investigation period. Using cross-tabulation [46][47][48], changes in areas between the different classes were determined and mapped for three periods, namely for changes between the years 2002 and 2013, between 2013 and 2022, and between 2002 and 2022. Annual land use/land cover change has been assessed [15,49] and percentage changes for each land use/land cover type over time were calculated [47,[50][51][52][53]. ...
Article
Full-text available
Since 2002, numerous sustainable land management (SLM) interventions have been implemented in Ethiopia, such as agroforestry, area closure, forage development, gully rehabilitation, and conservation agriculture. In addition, watershed-based developments contributed comprehensively to a better use of existing natural resources. This study determined the impact of Integrated Watershed Management (IWM) on land use/land cover for the Yesir watershed in Northern Ethiopia. Supervised image classification algorithms were applied to a time series of Landsat 5 (2002) and Landsat 8 (2013 and 2022) images to produce land use/land cover maps. A Geographic Information System was applied to analyze and map changes in land use/land cover for settlements, agricultural land, grazing land, and land covered with other vegetation. In focus group discussions, the time series maps were analyzed and compared with the integrated watershed management practices to analyze their impacts. The results document that integrated watershed management practices have contributed to a significant change in land use/land cover in the study area over the past 20 years. The quantitative analysis of land use/land cover between the years 2002 and 2022 only revealed a downward trend in agricultural land. Considering the value of the Normalized Difference Vegetation Index (NDVI) as a biophysical feature for the increase of green mass, this indicator also documents an improvement in land use/land cover with regard to sustainable land management and consequently poverty alleviation.
... The ongoing removal of riparian forests is likely to speeden up and intensify drought impacts, and reduces the resilience of riverine ecosystems, including its 2065 human and nonhuman inhabitants, to recover from droughts and prepare for the next. Also in mountaineous forests, in the Mt Kenya region, human activities in the form of agricultural intensification and the expansion of horticulture agribusinesses have increased pressure on water resources, pasture, and idle land because of a shrinkage in area of natural and agro-pastoral landscapes (Eckert et al., 2017). ...
Preprint
Full-text available
Droughts are often long lasting phenomena, without a distinct start or end, and with impacts cascading across sectors and systems, creating long-term legacies. Nevertheless, our current perception and management of droughts and their impacts is often event-based, which can limit the effective assessment of drought risks and reduction of drought impacts. Here, we advocate for changing this perspective and viewing drought as a hydro-eco-social continuum. We take a systems theory perspective and focus on how “memory” causes feedback and interactions between parts of the interconnected systems at different time scales. We first discuss the characteristics of the drought continuum with a focus on the hydrological, ecological, and social systems separately; and then study the system of systems. Our analysis is based on a review of the literature and a study of five cases: Chile, the Colorado River Basin in the US, Northeast Brazil, Kenya, and the Rhine River Basin in Northwest Europe. We find that the memories of past dry and wet periods, carried by both bio-physical (e.g. groundwater, vegetation) and social systems (e.g. people, governance), influence how future drought risk manifests. We identify four archetypes of drought dynamics: Impact & recovery; Slow resilience-building; Gradual collapse; and High resilience, big shock. The interactions between the hydrological, ecological and social systems result in systems shifting between these types, which plays out differently in the five case studies. We call for more research on drought pre-conditions and recovery in different systems, on dynamics cascading between systems and triggering system changes, and on dynamic vulnerability and maladaptation. Additionally, we argue for more continuous monitoring of drought hazards and impacts, modelling tools that better incorporate memories and adaptation responses, and management strategies that increase social and institutional memory to better deal with the complex hydro-eco-social drought continuum and identify effective pathways to adaptation.
... However, rather than universally supporting agricultural intensification, more consideration needs to be given to the specific local food system, especially local foods that could potentially be lost under intensification strategies (Ickowitz et al., 2019). In fact, agricultural intensification and expansion can increase pressure on water (Eckert et al., 2017), GHG emissions (Olén et al., 2021;Van Loon et al., 2019) and biodiversity (Shaver et al., 2015;Zabel et al., 2019). ...
Article
Full-text available
Bridging the gap between the micro and the macro scale in modelling food security to inform context-specific regionalised policies remains a major scientific challenge. A better understanding of the relations between global and local drivers impacting local food self-sufficiency (LFSS) is essential. We applied to the whole Mediterranean environmental area (Southern and Northern) a modelling framework for structural estimates (PLS-PM) using qualitative and quantitative methods to combine local-level information from field surveys and participatory workshops with global-level data. Our findings show that farmland expansion and intensification spatially disconnected from urban consumption areas do not appear to foster LFSS. On the other hand, public policies appear key to enhancing LFSS in the Mediterranean area if appropriate to the particular regional context. We outline how this multi-level modelling methodology can contribute to a place-based approach by informing context-specific regionalised policies aimed at food security.
... The GEE cloud platform stores open-source datasets such as Landsat and Sentinel on a global scale, allowing users to retrieve data for visualization and computation according to their needs (Tamiminia et al. 2020). With its stored open-source datasets and efficient computational capabilities, GEE has become one of the major cloud platforms for remote sensing classification today (Eckert et al. 2017;Liu et al. 2020a, b;Phalke et al. 2020;Wang et al. 2020;Zhang et al. 2020), and has been used in many situations (Dong et al. 2016;Noel et al. 2017). Of course, apart from the GEE cloud platform, there exist many other cloud platforms. ...
Article
Full-text available
The unique geographical location of the land-sea transition makes the ecological environment of the Yellow River Delta very fragile and vulnerable to human activities. As one of the characteristics of anthropogenic activities, monitoring the spatiotemporal changes of impervious surface area (ISA) is of great significance to the protection of the ecological environment in the Yellow River Delta (YRD). Based on the Landsat historical images and computing resources provided by Google Earth Engine (GEE), an ISA mapping method was developed through combining spectral, texture features and random forest algorithm, and subsequently was applied to generate the spatiotemporal distribution data of ISA of the YRD for 1992, 1998, 2004, 2010, 2016 and 2021. The experimental results demonstrated that the proposed method achieved satisfactory accuracy, with an average overall accuracy of 92.23% and an average Kappa coefficient of 0.9090. Through further time-series analysis of ISA, it found that the area of ISA in the YRD increased from the initial 394.87 km² to 1081.74 km² during study periods, and the annual growth rate broke through new highs, ranging from the initial 1.01 km²/year to 67.87 km²/year. According to the research results, urban development activities in the region should be strictly restricted in order to protect the ecological environment of the YRD.
... The UN (United Nations) FAO (Food and Agriculture Organization) Africover project (Di Gregorio, 2005) which was specifically dedicated to crop classification, generated the most recent and detailed land cover map for Central-Eastern African countries in the year 2000s. Since then, population growth and major policy changes have caused the increase in use of land for agropastoralism and systematic expansion of cropland area (Eckert et al., 2017, Kennedy et al., 2018. On the other hand, certain provinces of the region have been deeply affected by the loss of agriculture due to climatic change and human activities (Negussie et al., 2011, Roba & Oba, 2013, Nguru & Rono, 2013, Mganga et al., 2015, Balogh et al., 2016. ...
Article
The SBAM (Satellite Based Agricultural Monitoring) project, funded by the Italian Space Agency aims at: developing a validated satellite imagery based method for estimating and updating the agricultural areas in the region of Central-Africa; implementing an automated process chain capable of providing periodical agricultural land cover maps of the area of interest and, possibly, an estimate of the crop yield. The project aims at filling the gap existing in the availability of high spatial resolution maps of the agricultural areas of Kenya. A high spatial resolution land cover map of Central-Eastern Africa including Kenya was compiled in the year 2000 in the framework of the Africover project using Landsat images acquired, mostly, in 1995. We investigated the use of phenological information in supporting the use of remotely sensed images for crop classification and monitoring based on Landsat 8 and, in the near future, Sentinel 2 imagery. Phenological information on crop condition was collected using time series of NDVI (Normalized Difference Vegetation Index) based on Landsat 8 images. Kenyan countryside is mainly characterized by a high number of fragmented small and medium size farmlands that dramatically increase the difficulty in classification; 30 m spatial resolution images are not enough for a proper classification of such areas. So, a pan-sharpening FIHS (Fast Intensity Hue Saturation) technique was implemented to increase image resolution from 30 m to 15 m. Ground test sites were selected, searching for agricultural vegetated areas from which phenological information was extracted. Therefore, the classification of agricultural areas is based on crop phenology, vegetation index behaviour retrieved from a time series of satellite images and on AEZ (Agro Ecological Zones) information made available by FAO (FAO, 1996) for the area of interest. This paper presents the results of the proposed classification procedure in comparison with land cover maps produced in the past years by other projects. The results refer to the Nakuru County and they were validated using field campaigns data. It showed a satisfactory overall accuracy of 92.66 % which is a significant improvement with respect to previous land cover maps.
... (64-bit) with 2.0× vertical exaggeration to show topographic relief (Google Earth/DigitalGlobe, 2021). and recent decades (Castro, 1991b;Fanstone, 2016), but with some stable forestlines and reforestation areas (Gathaara, 1999;Hansen et al., 2013;Eckert et al., 2017). At present on Mount Kenya, fires occur most frequently in the Ericaceous vegetation, along trail routes, along protected area boundaries and on the drier leeward northwestern flank (Fig. 2) Henry et al., 2019b). ...
Article
Full-text available
Past forest fire events and fire frequencies are reconstructed with sediment–charcoal records at lake catchment spatial scales. Few quantitative palaeofire analyses exist in tropical montane forests, where fire return intervals are long (decadal and centennial scales) because of the infrequency of fire weather and fuel conditions. Fire return intervals are a key characteristic of fire regimes and changing fire frequencies rapidly alter land cover compositions and vegetation structure. Charcoal records from small lakes with relatively small catchments covered with dense forest provide an opportunity to reconstruct low‐frequency, high‐severity fires through a time series decomposition approach to identify charcoal peaks above a varying background rate as a proxy for palaeofire events. The sediment core from Rumuiku wetland on Mount Kenya, equatorial eastern Africa, accumulated a nearly linear age–depth model and provided a high temporal resolution (10 years cm–1) sieved charcoal count record (>125 µm). Pollen analysis showed a significant change in montane forest assemblage occurred at 21 200 cal a bp from a montane forest with abundant Podocarpus and Juniperus to a forest with more abundant Hagenia. This change in forest altered the vegetation composition and structure with concomitant changes to the fire regime. Forest biomass in the Hagenia forests decreased and it is likely that fire activity qualitatively changed toward lower intensity and lower severity fires. The quantitative fire event reconstruction focuses on the interval from 27 000 to 16 500 cal a bp and the older montane forest that experienced higher severity fires from 27 000 to 21 200 cal a bp, which reconstructed a temporally heterogeneous fire regime with fire return intervals that ranged from 30–430 years and a mean of 120 years (median 160 years) in the catchment. These are the first estimates of fire return intervals of mountain forests in eastern Africa. We then explore the potential for further comparative research and incremental research contributions to improve quantitative and qualitative palaeofire research in tropical forest ecosystems. We discuss the potential to use these types of data for characterizing variables of fire regimes prior to ostensibly significant modification by anthropogenic activity as well as during the recent past as human land use pressures increased within Afromontane forests.
... Similarly, Maua et al. (2022) and Ondiek et al. (2020) observed that the rich wetland resources of the Nzoia and Lake Victoria drainage systems experienced rampant anthropogenic exploitation within the period under our study. Studies in Kenya have also attributed cropland expansion to agricultural extensification practices (Eckert et al. 2017;Mwangi et al. 2018). At the SSA scale, cropland expansion in the late twentieth century resulted from rising economic activities, improved technology, declining soil fertility, and climate change variability (Jellason et al. 2021). ...
Article
Full-text available
Population growth and increasing demand for agricultural production continue to drive global cropland expansions. These expansions lead to the overexploitation of fragile ecosystems, propagating land degradation, and the loss of natural diversity. This study aimed to identify the factors driving land use/land cover changes (LULCCs) and subsequent cropland expansion in Trans Nzoia County in Kenya. Landsat images were used to characterize the temporal LULCCs in 30 years and to derive cropland expansions using change detection. Logistic regression (LR), boosted regression trees (BRTs), and evidence belief functions (EBFs) were used to model the potential drivers of cropland expansion. The candidate variables included proximity and biophysical, climatic, and socioeconomic factors. The results showed that croplands replaced other natural land covers, expanding by 38% between 1990 and 2020. The expansion in croplands has been at the expense of forestland, wetland, and grassland losses, which declined in coverage by 33%, 71%, and 50%, respectively. All the models predicted elevation, proximity to rivers, and soil pH as the critical drivers of cropland expansion. Cropland expansions dominated areas bordering the Mt. Elgon forest and Cherangany hills ecosystems. The results further revealed that the logistic regression model achieved the highest accuracy, with an area under the curve (AUC) of 0.96. In contrast, EBF and the BRT models depicted AUC values of 0.86 and 0.77, respectively. The findings exemplify the relationships between different potential drivers of cropland expansion and contribute to developing appropriate strategies that balance food production and environmental conservation.
... To date, most of the pollinator-based research comes from Europe and North America, while significant data gaps occur for Asian and African regions that are currently experiencing an intense agricultural and industrial development (Timberlake and Morgan, 2018). Specifically, in Sub-Saharan Africa, the land use intensification through urban and agricultural expansion is increasing as fast as the population growth (Eckert et al., 2017;Sulemana et al., 2019). In Sub-Saharan countries, agriculture represents the main source of family sustainment (Stein et al., 2017) with about 80% of the population relying on subsistence farming in Tanzania . ...
Thesis
Full-text available
One of the pronounced global challenges facing ecologists is how to feed the current growing human population while sustaining biodiversity and ecosystem services. To shed light on this, I investigated the impact of human land use on bee diversity and plant-pollinator interactions in Tanzania Savannah ecosystems. The thesis comprises the following chapters: Chapter I: General Introduction This chapter provides the background information including the study objectives and hypotheses. It highlights the ecological importance of bees and the main threats facing bee pollinators with a focus on two land-use practices namely livestock grazing and agriculture. It also highlights the diversity and global distribution of bees. It further introduces the tropical savannah ecosystem, its climate, and vegetation characteristics and explains spectacular megafauna species of the system that form centers of wildlife tourism and inadequacy knowledge on pollinators diversity of the system. Finally, this chapter describes the study methodology including, the description of the study area, study design, and data collection. Chapter II: Positive effects of low livestock grazing intensity on East African bee assemblages mediated by increases in floral resources The impact of livestock grazing intensity on bee assemblage has been subjected to research over decades. Moreover, most of these studies have been conducted in temperate Europe and America leaving the huge tropical savannah of East Africa less studied. Using sweep netting and pan traps, a total of 183 species (from 2,691 individuals) representing 55 genera and five families were collected from 24 study sites representing three levels of livestock grazing intensity in savannah ecosystem of northern Tanzania. Results have shown that moderate livestock grazing slightly increased bee species richness. However, high livestock grazing intensity led to a strong decline. Besides, results revealed a unimodal distribution pattern of bee species richness and mean annual temperature. It was also found that the effect of livestock grazing and environmental temperature on bee species richness was mediated by a positive effect of moderate grazing on floral resource richness. The study, therefore, reveals that bee communities of the African savannah zone may benefit from low levels of livestock grazing as this favors the growth of flowering plant species. A high level of livestock grazing intensity will cause significant species losses, an effect that may increase with climatic warming. Chapter III: Agricultural intensification with seasonal fallow land promotes high bee diversity in Afrotropical drylands This study investigated the impact of local agriculture intensification on bee diversity in the Afro tropical drylands of northern Tanzania. Using sweep netting and pan traps, a total of 219 species (from 3,428 individuals) representing 58 genera and six families were collected from 24 study sites (distributed from 702 to 1708 m. asl) representing three levels of agriculture intensity spanning an extensive gradient of mean annual temperature. Results showed that bee species richness increased with agricultural intensity and with increasing temperature. However, the effects of agriculture intensity and temperature on bee species richness were mediated by the positive effects of agriculture and temperature on floral resource richness used by bee pollinators. Moreover, results showed that variation of bee body sizes increases with agricultural intensification, “that effect”, however, diminished in environments with higher temperatures. This study reveals that bee assemblages in Afrotropical drylands benefit from agriculture intensification in the way it is currently practiced. Further intensification, including year-round irrigated crop monocultures and extensive use of agrochemicals, is likely to exert a negative impact on bee diversity and pollination services, as reported in temperate regions. Moreover, several bee species were restricted to natural savannah habitats. Therefore, to conserve bee communities in Afro tropical drylands and guarantee pollination services, a mixture of savannah and agriculture, with long periods of fallow land should be maintained. Chapter IV: Impact of land use intensification and local features on plants and pollinators in Sub-Saharan smallholder farms For the first time in the region, this study explores the impact of land-use intensification on plants and pollinators in Sub-Saharan smallholder farms. The study complemented field surveys of bees with a modern DNA metabarcoding approach to characterize the foraged plants and thus built networks describing plant-pollinator interactions at the individual insect level. This information was coupled with quantitative traits of landscape composition and floral availability surrounding each farm. The study found that pollinator richness decreased with increasing impervious and agricultural cover in the landscape, whereas the flower density at each farm correlated with pollinator richness. The intensification of agricultural land use and urbanization correlated with a higher foraging niche overlap among pollinators due to the convergence of individuals' flower-visiting strategies. Furthermore, within farms, the higher availability of floral resources drove lower niche overlap among individuals, greater abundance of flower visitors shaped higher generalization at the networks level (H2I), possibly due to increased competition. These mechanistic understandings leading to individuals’ foraging niche overlap and generalism at the network level, could imply stability of interactions and the pollination ecosystem service. The integrative survey proved that plant-pollinator systems are largely affected by land use intensification and by local factors in smallholder farms of Sub-Saharan Africa. Thus, policies promoting nature-based solutions, among which the introduction of more pollinator-friendly practices by smallholder farmers, could be effective in mitigating the intensification of both urban and rural landscapes in this region, as well as in similar Sub-Saharan contexts. Chapter V: A synopsis of the Bee occurrence data of northern Tanzania This study represents a synopsis of the bee occurrence data of northern Tanzania obtained from a survey in the Kilimanjaro, Arusha, and Manyara regions. Bees were sampled using two standardized methods, sweep netting and colored pan traps. The study summed up 953 species occurrences of 45 species belonging to 20 genera and four families (Halictidae, Apidae, Megachilidae, and andrenidae) A. This study serves as the baseline information in understanding the diversity and distribution of bees in the northern parts of the country. Understanding the richness and distribution of bees is a critical step in devising robust conservation and monitoring strategies for their populations since limited taxonomic information of the existing and unidentified bee species makes their conservation haphazard. Chapter VI: General discussion In general, findings obtained in these studies suggest that livestock grazing and agriculture intensification affects bee assemblages and floral resources used by bee pollinators. Results have shown that moderate livestock grazing intensity may be important in preserving bee diversity. However, high level of livestock grazing intensity may result in a strong decline in bee species richness and abundance. Moreover, findings indicate that agriculture intensification with seasonal fallow lands supports high floral resource richness promoting high bee diversity in Afrotropical drylands. Nonetheless, natural savannahs were found to contain unique bee species. Therefore, agriculture intensification with seasonal fallow should go in hand with conserving remnant savannah in the landscapes to increase bee diversity and ensure pollination services. Likewise, findings suggest that increasing urbanization and agriculture cover at the landscape level reduce plant and pollinator biodiversity with negative impacts on their complex interactions with plants. Conversely, local scale availability of floral resources has shown the positive effects in buffering pollinators decline and mitigating all detrimental effects induced by land-use intensification. Moreover, findings suggest that the impact of human land use (livestock grazing and agriculture) do not act in isolation but synergistically interacts with climatic factors such as mean annual temperature, MAT. The impact of MAT on bee species richness in grazing gradient showed to be more detrimental than in agriculture habitats. This could probably be explained by the remaining vegetation cover following anthropogenic disturbance. Meaning that the remaining vegetation cover in the agricultural gradient probably absorbs the solar radiations hence reducing detrimental effect of mean annual temperature on bee species richness. This one is not the case in grazing gradient since the impact of livestock grazing is severe, leaving the bare land with no vegetation cover. Finally, our findings conclude that understanding the interplay of multiple anthropogenic activities and their interaction with MAT as a consequence of ongoing climate change is necessary for mitigating their potential consequences on bee assemblages and the provision of ecosystem services. Moreover, future increases in livestock grazing and agriculture intensification (including year-round crop irrigated monocultures and excessive use of agrochemicals) may lead to undesirable consequences such as species loss and impair the provision of pollination services.
Conference Paper
Full-text available
This Sustainable Land Use and Natural Resource Management (SLUSE) research project investigates the reconstruction of agricultural livelihoods in a young resettlement community known as Village 3 in a semi-arid area in the central highlands of Kenya. A majority of the villagers are formerly displaced people who were granted land as part of a resettlement scheme initiated by the Kenyan Government in 2007, after some twenty years of living as squatters. In order to better understand the constraints and needs of the resettled community as they begin to construct their new agriculture based livelihoods, the paper explores the importance of crop production compared to other livelihood activities, and the factors influencing farmers’ decision-making. A combination of qualitative and quantitative methods drawing from both social and natural science disciplines were employed to establish the livelihood activities the villagers are engaged in and how they regard the contribution of these activities to their livelihood strategies. Furthermore the study explored the existing crop cultivation practices and the role of culture, knowledge, property rights, climate, soil fertility, water availability, and access to markets, credit and inputs on the famers’ choice of crop production system. The results from the questionnaire survey, several interviews and a number of PRAs show that agriculture is the most important activity in Solio Village 3, although livestock, casual labour and small businesses are also prominent. Farmers are faced with constraints to their crop production, most notably the difficult climatic conditions, lack of financial capital, and lack of knowledge about the most suitable crops to grow. These factors influence their choice of crop production system, and they now focus on drought resistant crops such as beans, although they still try to cultivate traditional crops like maize. Most farmers are engaged in at least one other livelihood activity than agricultural production, a diversification strategy aimed at increasing their livelihood security.
Article
Full-text available
Burkina Faso ranges amongst the fastest growing countries in the world with an annual population growth rate of more than three percent. This trend has consequences for food security since agricultural productivity is still on a comparatively low level in Burkina Faso. In order to compensate for the low productivity, the agricultural areas are expanding quickly. The mapping and monitoring of this expansion is difficult, even on the basis of remote sensing imagery, since the extensive farming practices and frequent cloud coverage in the area make the delineation of cultivated land from other land cover and land use types a challenging task. However, as the rapidly increasing population could have considerable effects on the natural resources and on the regional development of the country, methods for improved mapping of LULCC (land use and land cover change) are needed. For this study, we applied the newly developed ESTARFM (Enhanced Spatial and Temporal Adaptive Reflectance Fusion Model) framework to generate high temporal (8-day) and high spatial (30 m) resolution NDVI time series for all of Burkina Faso for the years 2001, 2007, and 2014. For this purpose, more than 500 Landsat scenes and 3000 MODIS scenes were processed with this automated framework. The generated ESTARFM NDVI time series enabled extraction of per-pixel phenological features that all together served as input for the delineation of agricultural areas via random forest classification at 30 m spatial resolution for entire Burkina Faso and the three years. For training and validation, a randomly sampled reference dataset was generated from Google Earth images and based on expert knowledge. The overall accuracies of 92% (2001), 91% (2007), and 91% (2014) indicate the well-functioning of the applied methodology. The results show an expansion of agricultural area of 91% between 2001 and 2014 to a total of 116,900 km 2. While rainfed agricultural areas account for the major part of this trend, irrigated areas and plantations also increased considerably, primarily promoted by specific development projects. This expansion goes in line with the rapid population growth in most provinces of Burkina Faso where land was still available for an expansion of agricultural area. The analysis of agricultural encroachment into protected areas and their surroundings highlights the increased human pressure on these areas and the challenges of environmental protection for the future.
Article
Full-text available
Mapping croplands, including fallow areas, are an important measure to determine the quantity of food that is produced, where they are produced, and when they are produced (e.g. seasonality). Furthermore, croplands are known as water guzzlers by consuming anywhere between 70% and 90% of all human water use globally. Given these facts and the increase in global population to nearly 10 billion by the year 2050, the need for routine, rapid, and automated cropland mapping year-after-year and/or season-after-season is of great importance. The overarching goal of this study was to generate standard and routine cropland products, year-after-year, over very large areas through the use of two novel methods: (a) quantitative spectral matching techniques (QSMTs) applied at continental level and (b) rule-based Automated Cropland Classification Algorithm (ACCA) with the ability to hind-cast, now-cast, and future-cast. Australia was chosen for the study given its extensive croplands, rich history of agriculture, and yet nonexistent routine yearly generated cropland products using multi-temporal remote sensing. This research produced three distinct cropland products using Moderate Resolution Imaging Spectroradiometer (MODIS) 250-m normalized difference vegetation index 16-day composite time-series data for 16 years: 2000 through 2015. The products consisted of: (1) cropland extent/areas versus cropland fallow areas, (2) irrigated versus rainfed croplands, and (3) cropping intensities: single, double, and continuous cropping. An accurate reference cropland product (RCP) for the year 2014 (RCP2014) produced using QSMT was used as a knowledge base to train and develop the ACCA algorithm that was then applied to the MODIS time-series data for the years 2000–2015. A comparison between the ACCA-derived cropland products (ACPs) for the year 2014 (ACP2014) versus RCP2014 provided an overall agreement of 89.4% (kappa = 0.814) with six classes: (a) producer’s accuracies varying between 72% and 90% and (b) user’s accuracies varying between 79% and 90%. ACPs for the individual years 2000–2013 and 2015 (ACP2000–ACP2013, ACP2015) showed very strong similarities with several other studies. The extent and vigor of the Australian croplands versus cropland fallows were accurately captured by the ACCA algorithm for the years 2000–2015, thus highlighting the value of the study in food security analysis. The ACCA algorithm and the cropland products are released through http://croplands.org/app/map and http://geography.wr.usgs.gov/science/croplands/algorithms/australia_250m.html
Article
Full-text available
Land cover mapping (LCM) in complex surface-mined and agricultural landscapes could contribute greatly to regulating mine exploitation and protecting mine geo-environments. However, there are some special and spectrally similar land covers in these landscapes which increase the difficulty in LCM when employing high spatial resolution images. There is currently no research on these mixed complex landscapes. The present study focused on LCM in such a mixed complex landscape located in Wuhan City, China. A procedure combining ZiYuan-3 (ZY-3) stereo satellite imagery, the feature selection (FS) method, and machine learning algorithms (MLAs) (random forest, RF; support vector machine, SVM; artificial neural network, ANN) was proposed and first examined for both LCM of surface-mined and agricultural landscapes (MSMAL) and classification of surface-mined land (CSML), respectively. The mean and standard deviation filters of spectral bands and topographic features derived from ZY-3 stereo images were newly introduced. Comparisons of three MLAs, including their sensitivities to FS and whether FS resulted in significant influences, were conducted for the first time in the present study. The following conclusions are drawn. Textures were of little use, and the novel features contributed to improve classification accuracy. Regarding the influence of FS: FS substantially reduced feature set (by 68% for MSMAL and 87% for CSML), and often improved classification accuracies (with an average value of 4.48% for MSMAL using three MLAs, and 11.39% for CSML using RF and SVM); FS showed statistically significant improvements except for ANN-based MSMAL; SVM was most sensitive to FS, followed by ANN and RF. Regarding comparisons of MLAs: for MSMAL based on feature subset, RF achieved the greatest overall accuracy of 77.57%, followed by SVM and ANN; for CSML, SVM had the highest accuracies (87.34%), followed by RF and ANN; based on the feature subsets, significant differences were observed for MSMAL and CSML using any pair of MLAs. In general, the proposed approach can contribute to LCM in complex surface-mined and agricultural landscapes.
Research
Full-text available
Like other countries in East Africa, Kenya faces seasonal water shortages that make it important to use and distribute water in an optimal way. One of Kenya’s biggest water users is its growing commercial horticulture sector, which exports fruits, vegetables, and especially flowers to Europe and elsewhere. Economically, the sector is a big success: it is Kenya’s second largest foreign exchange earner and a major employer. In 2014, for example, the horticulture sector contributed EUR 1.7 billion to the economy, with 42% of the profits coming from exports. Neighbouring countries in East Africa have sought to emulate this model. But its economic benefits must be weighed carefully against relevant social and environmental risks, including competition over precious water.
Article
Full-text available
The goal of this study was to map rainfed and irrigated rice-fallow cropland areas across South Asia, using MODIS 250 m time-series data and identify where the farming system may be intensified by the inclusion of a short-season crop during the fallow period. Rice-fallow cropland areas are those areas where rice is grown during the kharif growing season (June–October), followed by a fallow during the rabi season (November–February). These cropland areas are not suitable for growing rabi-season rice due to their high water needs, but are suitable for a short -season (≤3 months), low water-consuming grain legumes such as chickpea (Cicer arietinum L.), black gram, green gram, and lentils. Intensification (double-cropping) in this manner can improve smallholder farmer’s incomes and soil health via rich nitrogen-fixation legume crops as well as address food security challenges of ballooning populations without having to expand croplands. Several grain legumes, primarily chickpea, are increasingly grown across Asia as a source of income for smallholder farmers and at the same time providing rich and cheap source of protein that can improve the nutritional quality of diets in the region. The suitability of rainfed and irrigated rice-fallow croplands for grain legume cultivation across South Asia were defined by these identifiers: (a) rice crop is grown during the primary (kharif) crop growing season or during the north-west monsoon season (June–October); (b) same croplands are left fallow during the second (rabi) season or during the south-east monsoon season (November–February); and (c) ability to support low water-consuming, short-growing season (≤3 months) grain legumes (chickpea, black gram, green gram, and lentils) during rabi season. Existing irrigated or rainfed crops such as rice or wheat that were grown during kharif were not considered suitable for growing during the rabi season, because the moisture/water demand of these crops is too high. The study established cropland classes based on the every 16-day 250 m normalized difference vegetation index (NDVI) time series for one year (June 2010–May 2011) of Moderate Resolution Imaging Spectroradiometer (MODIS) data, using spectral matching techniques (SMTs), and extensive field knowledge. Map accuracy was evaluated based on independent ground survey data as well as compared with available sub-national level statistics. The producers’ and users’ accuracies of the cropland fallow classes were between 75% and 82%. The overall accuracy and the kappa coefficient estimated for rice classes were 82% and 0.79, respectively. The analysis estimated approximately 22.3 Mha of suitable rice-fallow areas in South Asia, with 88.3% in India, 0.5% in Pakistan, 1.1% in Sri Lanka, 8.7% in Bangladesh, 1.4% in Nepal, and 0.02% in Bhutan. Decision-makers can target these areas for sustainable intensification of short-duration grain legumes.
Article
Full-text available
Irrigation is crucial to agriculture in arid and semi-arid areas and significantly contributes to crop development, food diversity and the sustainability of agro-ecosystems. For a specific crop, the separation of its irrigated and rainfed areas is difficult, because their phenology is similar and therefore less distinguishable, especially when there are phenology shifts due to various factors, such as elevation and latitude. In this study, we present a simple, but robust method to map irrigated and rainfed wheat areas in a semi-arid region of China. We used the Normalized Difference Vegetation Index (NDVI) at a 30 × 30 m spatial resolution derived from the Chinese HJ-1A/B (HuanJing(HJ) means environment in Chinese) satellite to create a time series spanning the whole growth period of wheat from September 2010 to July 2011. The maximum NDVI and time-integrated NDVI (TIN) that usually exhibit significant differences between irrigated and rainfed wheat were selected to establish a classification model using a support vector machine (SVM) algorithm. The overall accuracy of the Google-Earth testing samples was 96.0%, indicating that the classification results are accurate. The estimated irrigated-to-rainfed ratio was 4.4:5.6, close to the estimates provided by the agricultural sector in Shanxi Province. Our results illustrate that the SVM classification model can effectively avoid empirical thresholds in supervised classification and realistically capture the magnitude and spatial patterns of rainfed and irrigated wheat areas. The approach in this study can be applied to map irrigated/rainfed areas in other regions when field observational data are available.
Article
Full-text available
Global agricultural production will likely need to increase in the future due to population growth, changing diets, and the rising importance of bioenergy. Intensifying already existing cropland is often considered more sustainable than converting more natural areas. Unfortunately, our understanding of cropping patterns and intensity is weak, especially at broad geographic scales. We characterized and mapped cropping systems in Europe, a region containing diverse cropping systems, using four indicators: (a) cropping frequency (number of cropped years), (b) multi-cropping (number of harvests per year), (c) fallow cycles, and (d) crop duration ratio (actual time under crops) based on the MODIS Normalized Difference Vegetation Index (NDVI) time series from 2000 to 2012. Second, we used these cropping indicators and self-organizing maps to identify typical cropping systems. The resulting six clusters correspond well with other indicators of agricultural intensity (e.g., nitrogen input, yields) and reveal substantial differences in cropping intensity across Europe. Cropping intensity was highest in Germany, Poland, and the eastern European Black Earth regions, characterized by high cropping frequency, multi-cropping and a high crop duration ratio. Contrarily, we found lowest cropping intensity in eastern Europe outside the Black Earth region, characterized by longer fallow cycles. Our approach highlights how satellite image time series can help to characterize spatial patterns in cropping intensity—information that is rarely surveyed on the ground and commonly not included in agricultural statistics: our clustering approach also shows a way forward to reduce complexity when measuring multiple indicators. The four cropping indicators we used could become part of continental-scale agricultural monitoring in order to identify target regions for sustainable intensification, where trade-offs between intensification and the environmental should be explored.
Article
Common understanding of the causes of land-use and land-cover change is dominated by simplifications which, in turn, underlie many environment-development policies. This article tracks some of the major myths on driving forces of land-cover change and proposes alternative pathways of change that are better supported by case study evidence. Cases reviewed support the conclusion that neither population nor poverty alone constitute the sole and major underlying causes of land-cover change worldwide. Rather, peoples’ responses to economic opportunities, as mediated by institutional factors, drive land-cover changes. Opportunities and
Article
A random forest (RF) classifier is an ensemble classifier that produces multiple decision trees, using a randomly selected subset of training samples and variables. This classifier has become popular within the remote sensing community due to the accuracy of its classifications. The overall objective of this work was to review the utilization of RF classifier in remote sensing. This review has revealed that RF classifier can successfully handle high data dimensionality and multicolinearity, being both fast and insensitive to overfitting. It is, however, sensitive to the sampling design. The variable importance (VI) measurement provided by the RF classifier has been extensively exploited in different scenarios, for example to reduce the number of dimensions of hyperspectral data, to identify the most relevant multisource remote sensing and geographic data, and to select the most suitable season to classify particular target classes. Further investigations are required into less commonly exploited uses of this classifier, such as for sample proximity analysis to detect and remove outliers in the training samples.