Bush encroachment

Last updated

View of bush encroached land at the Waterberg Plateau Park in Otjozondjupa Region, Namibia JLP 2323-2.jpg
View of bush encroached land at the Waterberg Plateau Park in Otjozondjupa Region, Namibia

Bush encroachment (also shrub encroachment, woody encroachment, bush thickening, woody plant proliferation) is a natural phenomenon characterised by the increase in density of woody plants, bushes and shrubs, at the expense of the herbaceous layer, grasses and forbs. It predominantly occurs in grasslands, savannas and woodlands and can cause biome shifts from open grasslands and savannas to closed woodlands. The term bush encroachment refers to the expansion of native plants and not the spread of alien invasive species. It is thus defined by plant density, not species. Bush encroachment is often considered an ecological regime shift and can be a symptom of land degradation.

Contents

Its causes include land use intensification, such as high grazing pressure and the suppression of wildfires. Climate change is found to be an accelerating factor for woody encroachment. The impact of bush encroachment is highly context specific. It is often found to have severe negative consequences on key ecosystem services, especially biodiversity, animal habitat, land productivity and groundwater recharge. Across rangelands, woody encroachment has led to significant declines of productivity, threatening the livelihoods of affected land users. The phenomenon is observed across different ecosystems and with different characteristics and intensities globally.

Various countries actively counter woody encroachment, through adapted grassland management practices and targeted bush thinning. In some cases, areas affected by woody encroachment are classified as carbon sinks and form part of national greenhouse gas inventories. However, carbon sequestration effects of woody encroachment are highly context specific and still insufficiently researched.

Causes

Woody encroachment is assumed to have its origins at the beginning of Holocene and the start of warming, with tropical species expanding their ranges away from the equator into more temperate regions. But it has occurred at unparalleled rates since the mid 19th century. [1] [2] Among earliest published notions of bush encroachment are publications of R. Staples in 1945, [3] O. West in 1947 [4] and Heinrich Walter in 1954. [5]

Various factors have been found to contribute to the process of bush encroachment. A general distinction can be made between bush encroachment due to land intensification and bush encroachment after land abandonment. Literature further suggests that the causes of woody encroachment differ significantly between wet and dry savanna. [6] With regard to its causes, bush encroachment is distinctly different from alien plant invasion, which is caused by the spread of deliberately or accidentally introduced species. [7]

Land abandonment

Where land is abandoned, the rapid spread of native bush plants is often observed. This is for example the case in former forest areas in the Alps that have been converted to agricultural land and later abandoned. In Southern Europe encroachment is thus linked to rural exodus. [8]

Land intensification

Driver of woody encroachment can change over time. While overgrazing has in the past frequently been found to be a main driver of woody encroachment, it is observed that woody encroachment continues in the respective areas even after grazing reduced or even ceases. [9]

Global drivers

While changes in land management are often seen as the main driver of woody encroachment, some studies suggest that global drivers increase woody vegetation regardless of land management practices. [22]

Species

A wide range of different woody species have been identified as encroacher species globally. As opposed to invasive species, encroacher species are indigenous to the respective ecosystem and their classification as encroachers depends on how strongly they grow into landscapes that they previously did not dominate. Comparisons of encroaching and non-encroaching vachellia species found that encroaching species have a higher acquisition and competition for resources. Their canopy architecture is different and only encroaching tree species reduce the productivity of perennial vegetation. [33]

Impact on Ecosystem Services

Woody encroachment constitutes a shift in plant composition with far-reaching impact on the affected ecosystems. While it is commonly identified as a form of land degradation, with severe negative consequences for various ecosystem services, such as biodiversity, groundwater recharge, carbon storage capacity and herbivore carrying capacity, this link is not universal. Impacts are dependent on species, scale and environmental context factors and shrub encroachment can have significant positive impacts on ecosystem services as well. [34] While woody plant encroachment is not generally synonymous with degradation, it is found to contribute to degradation of arid ecosystems. [35] There is a need for ecosystem-specific assessments and responses to woody encroachment. [36] Generally, the following context factors determine the ecological impact of woody encroachment: [37]

Woody encroachment is often seen as a form of land degradation and an expression of desertification. Due to its ambiguous role of contributing to greening and desertification, it has been termed "green desertification". [42] However, the link to desertification is not universal. During woody encroachment the herbaceous cover in the intercanopy zones typically remains intact, while during desertification these zones degrade and turn into bare soil devoid of organic matter. [43] For example, in the Mediterranean region shrub establishment can contribute to the reversal of ongoing desertification. [44]

Biodiversity

Bush expands at the direct expense of other plant species, potentially reducing plant diversity and animal habitats. [45] These effects are context specific, a meta analysis of 43 publications of the time period 1978 to 2016 found that bush encroachment has distinct negative effects on species richness and total abundance in Africa, especially on mammals and herpetofauna, but positive effects in North America. [46] However, in context specific analyses also in Northern America negative effects are observed. For example, piñon-juniper encroachment threatens up to 350 sagebrush-associated plant and animal species in the USA. [47] A study of 30 years of woody encroachment in Brazil found a significant decline of species richness by 27%. [48] Shrub encroachment may result in increase vertebrate species abundance and richness. However, these encroached habitats and their species assemblages may become more sensitive to droughts. [49] [50]

Evidence of biodiversity losses include the following:

Groundwater recharge and soil moisture

Water balance Surface water cycle.svg
Water balance

Woody plant encroachment is frequently linked to reduced groundwater recharge, based on evidence that bushes consume significantly more rainwater than grasses and encroachment alters water streamflow. [73] The downward movement of water is hindered by increased root density and depth. [74] [75] [76] [77] The impact on groundwater recharge differs between sandstone bedrocks and karst regions as well as between deep and shallow soils. [74] Although this is strongly context dependent, bush control can be an effective method for the improvement of groundwater recharge. [78] Applied research, assessing the water availability after brush removal, was conducted in Texas USA, resulting in an increase in water availability in all cases. [79] Studies in the United States further find that dense encroachment with Juniperus virginiana is capable of transpiring nearly all rainfall, thus altering groundwater recharge significantly. [80] An exception is shrub encroachment on slopes, where groundwater recharge can increase under encroachment. [20] [81]

While there is general consensus that bush encroachment has an ecohydrological impact, concrete experience with changes in groundwater recharge is however largely based on anecdotal evidence or regionally and temporally limited research projects. [82] Moreover, there is limited understanding how hydrological cycles through woody encroachment affects carbon influx and efflux, with both carbon gains and losses possible. [73]

Besides groundwater recharge, woody encroachment increases tree transpiration and evaporation of soil moisture, due to increased canopy cover. [83]

Carbon sequestration

Against the background of global efforts to mitigate climate change, the carbon sequestration and storage capacity of natural ecosystems receives increasing attention. Grasslands constitute 40% of Earth's natural vegetation [84] and hold a considerable amount of the global Soil Organic Carbon. Shifts in plant species composition and ecosystem structure, especially through woody encroachment, lead to significant uncertainty in predicting carbon cycling in grasslands. [85] [86] The impact of bush control on the carbon sequestration and storage capacity of the respective ecosystems is an important management consideration.

Research on the changes to carbon sequestration under bush encroachment and bush control is still insufficient. [87] [88] The Intergovernmental Panel on Climate Change (IPCC) states that bush encroachment generally leads to increased aboveground woody carbon, while belowground carbon changes depend on annual rainfall and soil type. The panel further elaborates that a global assessment of the net change in carbon stocks due to woody plant encroachment has not been conducted yet. [49] Factors relevant for comparisons of carbon sequestration potentials between encroached and non-encroached grasslands, include the following: above-ground net primary production (ANPP), below-ground net primary production (BNPP), photosynthesis rates, plant respiration rates, plant litter decomposition rates, soil microbacterial activity.

The impact of woody encroachment on soil organic carbon is found to be dependent on rainfall, with soil organic carbon increasing in dry ecosystems and decreasing in mesic ecosystems under encroachment. [94] [88] In wet environments, grasslands have more soil carbon than shrublands and woodlands. Under shrub encroachment, the losses in soil carbon can be sufficient to offset the gains of above-ground carbon gains. [95] [96] [97] [98] [99] Degradation of grasslands has in some areas led to the loss of up to 40% of the ecosystem's soil organic carbon. [90] An important factor is that under bush encroachment the increased photosynthetic potential is largely offset by increased plant respiration and respective carbon losses. [100]
Soil organic carbon changes need to be viewed at landscape level, as there are differences between under canopy and inter canopy processes. When a landscape becomes increasingly encroached and the remaining open grassland patches are overgrazed as a result, soil organic carbon may decrease. [101] [34] In South Africa, bush encroachment was found to slow decomposition rates of litter, which took twice the time to decay under bush encroachment compared to open savannas. This suggests a significant impact of woody encroachment on the soil organic carbon balance. [102] In pastoral lands of Ethiopia, bush encroachment was found to have little to now positive effect on soil organic carbon and woody encroachment restriction was the most effective way to maintain soil organic carbon. [103] In the United States, substantial soil organic carbon sequestration was observed in deeper portions of the soil, following woody encroachment. [104]
A meta-analysis of 142 studies found that shrub encroachment alters soil organic carbon (0–50 cm), with changes ranging between -50 and 300 percent. Soil organic carbon increased under the following conditions: semi arid and humid regions, encroachment by leguminous shrubs as opposed to non-legumes, sandy soils as opposed to clay soils. The study further concludes that shrub encroachment has a mainly positive effect on top-soil organic carbon content, with significant variations among climate, soil and shrub types. [105] There is a lack of standardised methodologies to assess the effect of woody encroachmetn on soil organic carbon. [88]

Land productivity

Bush encroachment directly impacts land productivity, as widely documented in the context of animal carrying capacity. In the Southern African country Namibia it is assumed that agricultural carrying capacity of rangelands has declined by two-thirds due to bush encroachment. In East Africa there is evidence that an increase of bush cover of 10 percent reduced grazing by 7 percent, with land becoming unusable as rangeland when the bush cover reaches 90 per cent. [114] [115] In Northern America, each 1 percent of increase in woody cover implies a reduction of 0.6 to 1.6 cattle per 100 hectares. [116]

Also touristic potential of land is found to decline in areas with heavy bush encroachment, with visitors shifting to less encroached areas and better visibility of wildlife. [117]

Rural livelihoods

While the ecological effects of woody encroachment are multifold and vary depending on encroachment density and context factors, woody encroachment is often considered to have a negative impact on rural livelihoods. In Africa 21% of the population depend on rangeland resources. Woody encroachment typically leads to an increase in less palatable woody species at the expense of palatable grasses. This reduces the resources available to pastoral communities and rangeland based agriculture at large. [118] Woody encroachment has negative consequences on livelihoods especially arid areas, [37] which support a third of the world population's livelihoods. [119] [120]

Others

In the United States, woody encroachment has been linked to the spread of tick-borne pathogens and respective disease risk for humans and animals. [121] In the Arctic tundra, shrub encroachment can reduce cloudiness and contribute to a raise in temperature. [122] In Northern America, significant increases in temperature and rainfall were linked to woody encroachment, amounting to values up to 214mm and 0.68 °C respectively. This is caused by a decrease in surface albedo. [123]

Targeted bush control in combination with the protection of larger trees is found to improve scavenging that regulates disease processes, alters species distributions, and influences nutrient cycling. [124]

Quantification and monitoring

There is no static definition of what is considered woody encroachment, especially when encroachment of indigenous plants occurs. While it is simple to determine vegetation trends (e.g. an increase in woody plants over time), it is more complex to determine thresholds beyond which an area is to be considered as encroached. Various definitions as well as quantification and mapping methods have been developed.

In Southern Africa, the BECVOL method (Biomass Estimates from Canopy Volume) finds frequent application. It determines Evapotranspiration Tree Equivalents (ETTE) per selected area. This data is used for comparison against climatic factors, importantly annual rainfall, to determine whether the respective areas have a higher number of woody plants than considered sustainable. [45]

Remote sensing imagery is frequently used to determine land the extend of woody encroachment. Shortcomings of this methodology include difficulties to distinguish species and the inability to detect small shrubs. [125] Moreover, UAV-based multispectral data and Lidar data are frequently used to quantify woody encroachment. [126] The probability of bush encroachment for the African continent has been mapped using GIS data and the variables precipitation, soil moisture and cattle density. [127]

Rephotography is found to be an effective tool for the monitoring of vegetation change, including woody encroachment [128] and forms the basis of various encroachment assessments. [31]

In most affected ecosystems, knowledge of historical land cover is limited to the availability of photographic evidence or written records. Methods to overcome this knowledge gap include the assessment of pollen records. In a recent application, vegetation cover of the past 130 years in a bush encroached area in Namibia was established. [129]

Bush control

Goats can function as a natural measure against bush encroachment or the re-establishment of seedlings after bush thinning. Boer goat444.jpg
Goats can function as a natural measure against bush encroachment or the re-establishment of seedlings after bush thinning.

Bush control refers to the active management of the density of woody species in grasslands. Although woody encroachment in many instances is a direct consequence of unsustainable management practices, it is unlikely that the introduction of more sustainable practices alone (e.g. the management of fire and grazing regimes) will achieve to restore already degraded areas. Encroached grasslands can constitute a stable state, meaning that without intervention the vegetation will not return to its previous composition. [130] Responsive measures, such as mechanical removal, are needed to restore a different balance between woody and herbaceous plants. [131] Once a high woody plant density is established, woody plants contribute to the soil seed bank more than grasses [132] and the lack of grasses presents less fuel for fires, reducing their intensity. [14] This perpetuates woody encroachment and necessitates intervention, if the encroached state is undesirable for the functions and use of the respective ecosystems. Most interventions constitute a selective thinning of bush densities, although in some contexts also repeat clear-cutting has shown to effectively restore diversity of typical savanna species. [133] In decision making on which woody species to thin out and which to retain, structural and functional traits of the species play a key role. [134]

Types of interventions

The term bush control, or brush management, refers to actions that are targeted at controlling the density and composition of bushes and shrubs in a given area. Such measures either serve to reduce risks associated with bush encroachment, such as wildfires, or to rehabilitate the affected ecosystems. It is widely accepted that encroaching indigenous woody plants are to be reduced in numbers, but not eradicated. This is critical as these plants provide important functions in the respective ecosystems, e.g. they serve as habitat for animals. [135] [136] Efforts to counter bush encroachment fall into the scientific field of restoration ecology and are primarily guided by ecological parameters, followed by economic indicators. Three different categories of measures can be distinguished:

Control methods

Fire fighter administering prescribed fire as management tool to remove woody encroachment near Mt. Adams, Washington, US USFWS Prescribed Fire at Conboy Lake NWR (22489393517).jpg
Fire fighter administering prescribed fire as management tool to remove woody encroachment near Mt. Adams, Washington, US

Natural bush control

Among others through the introduction of browsers, such as Boer goats, [139] administering controlled fires, [13] [140] [141] [142] or rewilding ecosystems with historic herbivory fauna. [143] [144]

Fire was found to be especially effective in reducing bush densities, when coupled with the natural event of droughts [145] or the intentional introduction of browsers. [146] [147] Fires have the advantage that they consume the seeds of woody plants in the grass layer before germination, therefore reducing the grasslands sensitivity to encroachment. [148] Prerequisite for successful bush control through fire is sufficient fuel load, thus fires have a higher effectiveness in areas where sufficient grass is available. Furthermore, fires must be administered regularly to address re-growth. Bush control through fire is found to be more effective when applying a range of fire intensities over time. [149] The relation between prescribed fire and tree mortality, is subject of ongoing research. [150] Research further suggests that the success rate of prescribed fires differs depending on the season during which it is applied. [151]

There is evidence that some rural farming communities have used small ruminants, like goats, to prevent bush encroachment for decades. [152]

Also targeted grazing systems can function as management tool for biodiversity conservation in grasslands. This is subject of ongoing research. [153]

Chemical bush control

Wood densities are frequently controlled through the application of herbicides, in particular arboricides. Frequently applied herbicides are based on the active ingredients tebuthiuron, ethidimuron, bromacil and picloram. [154] In East Africa, first comprehensive experiments on the effectiveness of such bush control date back to 1958–1960. [155]

Mechanical bush control

Cutting or harvesting of bushes and shrubs with manual or mechanised equipment. Mechanical cutting of woody plants is followed by stem-burning, fire or browsing to suppress re-growth. [156] Some studies find that mechanical bush control is more sustainable than controlled fires, because burning leads to deeper soil degradation and faster recovering of shrubs. [157] Bush that is mechanically harvested is often burnt on piles, [158] but can also serve as feedstock for value addition.

Challenges

Literature emphasizes that a restoration of bush encroached areas to a desired previous non-encroached state is difficult to achieve and the recovery of key-ecosystem may be short-lived or not occur. Intervention methods and technologies must be context specific to achieve their intended outcome. [159] [1] [160] Current efforts of selective plant removal are found to have slowed or halted woody encroachment in respective areas, but are sometimes found to be outpaced by continuing encroachment. [161] [162]

When bush thinning is implemented in isolation, without follow-up measures, grassland may not be rehabilitated. This is because such once-off treatments typically target small areas at a time and they leave plant seeds behind enabling rapid re-establishment of bushes. A combination of preventative measures, addressing the causes of bush encroachment, and responsive measures, rehabilitating affected ecosystems, can overcome bush encroachment in the long-run. [148] [163] [164]

In grassland conservation efforts, the implementation of measures across networks of private lands, instead of individual farms, remains a key challenge. [161] Due to the high cost of chemical or mechanical removal of woody species, such interventions are often implemented on a small scale, i.e. a few hectares at a time. This differs from natural control processes before human land use, e.g. widespread fires and vegetation pressure by free roaming wildlife. As a result, the interventions often have limited impact on the continued dispersal and spread of woody plants. [141]

Countering woody encroachment can be costly and largely depends on the financial capacity of land users. Linking bush control to the concept of Payment for ecosystem services (PES) has been explored in some countries. [165]

Relation to climate change mitigation and adaptation

Amount of carbon stored in Earth's various terrestrial ecosystems, in gigatonnes Carbon stored in ecosystems.png
Amount of carbon stored in Earth's various terrestrial ecosystems, in gigatonnes

Consideration in GHG inventories

Given scientific uncertainties, it varies widely how countries factor woody encroachment and the control thereof into their national Greenhouse Gas Inventories. In early carbon sink quantifications, woody encroachment was found to account for as much as 22% to 40% of the regional carbon sink in the USA, [167] [168] while it is considered a key uncertainty in the US carbon balance [169] [170] and the sink capacity is found to decrease when encroachment has reached its maximum extent. [171] Also in Australia woody encroachment constitutes a high proportion of the national carbon account. [172] [173] In South Africa, woody encroachment was estimated to have added around 21.000 Gg CO2 to the national carbon sink, [174] while it is has been highlighted that especially the loss of grass roots leads to losses of below-ground carbon, which is not fully compensated by gains of above-ground carbon. [175]

It is suggested that the classification of encroached grasslands and savannas as carbon sinks may often be incorrect, underestimating soil organic carbon losses. [176] [93]

Beyond difficulties to conclusively quantify the changes in carbon storage, promoting carbon storage through woody encroachment can constitute a trade-off, as it may reduce biodiversity of savanna endemics [177] [48] and core ecosystem services, like land productivity and water availability.

Grassland conservation can make a significant contribution to global carbon sequestration targets, but compared to sequestration potential in forestry and agriculture, this is still insufficiently explored and implemented. [178]

Bush Control as adaptation measure

Some countries, for example South Africa, acknowledge inconclusive evidence on the emissions effect of bush thinning, but strongly promote it as a means of climate change adaptation. [179] Geographic selection of intervention areas, targeting areas that are at an early stage of encroachment, can minimise above-ground carbon losses and therewith minimise the possible trade-off between mitigation and adaptation. [180] The Intergovernmental Panel on Climate Change (IPCC) reflects on this trade-off: "This variable relationship between the level of encroachment, carbon stocks, biodiversity, provision of water and pastoral value can present a conundrum to policymakers, especially when considering the goals of three Rio Conventions: UNFCCC, UNCCD and UNCBD. Clearing intense bush encroachment may improve species diversity, rangeland productivity, the provision of water and decrease desertification, thereby contributing to the goals of the UNCBD and UNCCD as well as the adaptation aims of the UNFCCC. However, it would lead to the release of biomass carbon stocks into the atmosphere and potentially conflict with the mitigation aims of the UNFCCC." The IPPC further lists bush control as relevant measure under ecosystem-based adaptation and community-based adaptation. [49]

Grassland conservation versus afforestation

With afforestation having gained popularity as a measure to create or enhance carbon sinks and thereby mitigate global climate change, there are calls to more carefully select suitable ecosystems. Conservation efforts increasingly target grasslands, savannas and open-canopy woodlands, recognising their importance for biodiversity and ecosystem services. Accepting woody encroachment or the invasion of alien woody species as a measure to mitigate climate change, can have severe negative consequences for the respective ecosystems. [181] [182] It is found that grasslands are frequently misidentified as degraded forests and targeted by afforestation efforts. [183] According to an analysis of areas identified to have forest restoration potential by the World Resources Institute, this includes up to 900 million hectares grasslands. [184] In Africa alone, 100 million hectares of grasslands are found to be at risk by misdirected afforestation efforts. Among the areas mapped as degraded forests are the Serengeti and Kruger National Parks, which have not been forested for several million years. [185] The Intergovernmental Panel on Climate Change (IPCC) states that mitigation action, such as reforestation or afforestation, can encroach on land needed for agricultural adaptation and therewith threaten food security, livelihoods and ecosystem functions. [32]

Global extent

Depiction of terrestrial biomes around the world Vegetation.png
Depiction of terrestrial biomes around the world

Woody encroachment occurs on all continents in a variety of ecosystems. Its causes, extent and response measures differ and are highly context specific. [186] [187] Ecosystems affected by woody encroachment include closed shrublands, open shrublands, woody savannas, savannas, and grasslands. It can occur not only in tropical and subtropical climates, but also in temperate areas. [35]

In sub-Saharan Africa, woody vegetation cover has increased by 8% during the past three decades, mainly through woody encroachment. Overall, 750 million hectares of non-forest biomes experienced significant net gains in woody plant cover, which is more than three times the area that experienced net losses of woody vegetation. [188] In around 249 million hectares of African rangelands long-term climate change was found to be the key driver of vegetation change. [118] In Southern Africa, woody encroachment has been identified as the main factor of greening, i.e. of the increase in vegetation cover detected through remote sensing. [189] [190]

In Southern Europe an estimated 8 percent of land area has transitioned from grazing land to woody vegetation between 1950 and 2010. [191]

In the Eurasian Steppe, the largest grassland globally, climate change linked bush encroachment has been found to occur at around 1% per decade. [192]

Affected ecoregions

Northern Europe

Woody encroachment is common in the Alpine tundra of Norway and Sweden [193] [194] Also in the Coastal meadows of Estonia bush encroachment is observed, resulting from land abandonment. [195] In Ireland and Denmark, dry grasslands are affected by woody encroachment. In Ireland extensive low input farming helps to prevent further encroachment by Blackthorn and Hazel , while high density stands are actively thinned out. [196]

Central Europe and European Alps

Areas that formerly were forests require continuous maintenance to avoid bush encroachment. When active land cultivation ends, fallow land is the result and gradual spread of shrubs and bushes can follow. Animal species once native to Central Europe effectively countered this natural process. These include herbivores such as European bison, auerochs (extinct), red deer and feral horse. Grassland and heath are considered to require protection due to their biodiversity as well as to preserve cultural landscapes. Bush encroachment is therefore frequently countered with selective removal of woody biomass or through the seasonal or year-round introduction of grazing animal species, such as sheep, goats, heck cattle or horses. Bush encroachment occurs in the Alps, where structural change in agriculture leads to the abandonment of land. Alnus viridis is the most widely distributed shrub species in the sub-alpine zone and is found to severely impair species richness and beta diversity when encroaching grassland. [197] Woody encroachment in the alpine tundra is associated with aboveground carbon storage and a slowdown of the biogeochemical cycle. [198] 70 percent of cultivated land in the Eastern Alps are affected by woody encroachment. [199] Also in Hungary bush encroachment is linked to the abandonment of formerly cultivated land. Moderate encroachment is found to have no negative impact on biodiversity and suppression of woody plants is considered an effective restoration approach. [200]

Mediterranean Basin

The Mediterranean region is widely reported to be affected by bush encroachment, which is often a transition into the establishment of trees in former grasslands. [201] This is found to have negative effects on biodiversity and to magnify climate and related droughts. [202] Further, it adversely affects soil organic matter. [203] At the same time encroaching shrubs are also found to have a positive effect, reversing the desertification process. [204] [44] Areas experiencing woody encroachment have more extended droughts and higher usage of deep water and this is expected to increase under future climate scenarios. [205] In the Spanish Pyrenees, woody encroachment is connected to land abandonment [206] and affects around 80 percent of cultivated land. [207] [208]

North American grasslands

North American grasslands have been found to be affected by woody plant encroachment. Documentation of shrub encroachment caused by fire exclusion was documented as early as 1968. [209]

United States of America

In the United States, affected ecosystems include the Chihuahuan Desert, the Sonoran Desert, the northern and southern Rocky Mountains, the sagebrush steppe, as well as the Southern and Central Great Plains. Poor grazing management and fire suppression are among the documented causes. [49] [210] Woody plant expansion is considered one of the greatest contemporary threats to mesic grasslands of the central United States. [131] Woody encroachment is estimated to lead to a loss of 75% of potential grass biomass in the Great Plains. [211] In the western US, woody plants have increased on around 44 million hectares since 1999. [212] Among encroaching species is piñon-juniper which mostly encroaches in shrubland adjacent to wooded areas. Up to 350 sagebrush-associated plant and animal species are threatened as a result. In the northern Great Basin piñon-juniper has encroached 0.45 million hectares since 2001 alone. [47] The rate at which grassland is lost to woody encroachment is found to equal the rate of conversion of grassland to agricultural land. [148] Also the tundra ecosystems of Colorado and Alaska areaffected by the rapid expansion of woody shrubs. [213] [214]

Negative impacts on forage production and an interrelation with carbon sequestration are documented. [109] At the same time in the semiarid karst savanna of Texas, USA, woody plant encroachment has been found to improve soil infiltrability and therewith groundwater recharge. [215] Over a period of 69 years, woody encroachment in Texas has increase aboveground carbon stocks by 32%. [216] Bird population decline as a result of woody encroachment has been identified as a critical conservation concern, [63] with bird populations found to have decreased by nearly two-thirds over the last half-century. [64]

Through government funded conservation programmes, shrubs and trees are thinned out systematically in affected ecosystems. This is found to revive habitat for birds and improve other ecosystem services. [217] There is evidence that selective thinning with post-treatment has successfully reversed the effects of conifer encroachment in studied areas. [163] At the same time study areas in Nebraska, where Juniperus virginiana encroachment was treated with fire, showed that woody cover stayed low and stable for 8-10 years after fire treatment, but rapid re-encroachment then followed. [218]

Asian temperate savanna and steppe

China

Temperate savanna-like ecosystems in Northern China are found to be affected by shrub encroachment, linked to unsustainable grazing and climate change. [219] In the Inner Mongolia steppe shrub encroaches steppe. [220]

India

Grassland in India Panna grassland IMG 20161127 102539353 HDR.jpg
Grassland in India

Semi-arid Banni grasslands of western India are found to be affected by bush encroachment, with affects both species composition and behaviour of nocturnal rodents. [221]

Australian lowland woodlands

In Australia woody encroachment is observed across all lowland grassy woodland as well as semi-arid floodplain wetlands and coastal ecosystems, with substantial implications for biodiversity conservation and ecosystem services. [222] [223]

Latin American grasslands

Argentina

In the Gran Chaco intense shrub encroachment has detrimental impact on livestock economies, especially in the Formosa Province. Livestock pressure and the lack of wildfires have been main causes. [224]

Brazil

Map of the Cerrado ecoregion in Brazil as delineated by the World Wide Fund for Nature Ecoregion NT0704.svg
Map of the Cerrado ecoregion in Brazil as delineated by the World Wide Fund for Nature

Wide-ranging woody encroachment is found in the Cerrado, a savannah ecosystem in central Brazil. Studies found that 19% of its area, approximately 17 million hectares, show significant bush encroachment. Among the researched causes are fire suppression and land use abandonment. [225] Fire suppression is linked to Brazil's conservation policy that aims to deforestation in the Amazon, but achieves the limitation of fires also in the Cerrado. [226] This ecological change is linked to the disturbance of ecohydrological processes. [227] In some areas of the Cerrado, open grassland and wetlands has largely disappeared. [228] A contributing factor to the loss of the natural Cerrado savanna ecosystem is the planting of monocultures, such as pine, for wood production. When pine is removed and plantations abandoned, areas turn into low-diversity forests lacking savanna species. [229] Also in the highland grassland of Southern Brazil, bush encroachment caused by land management changes is seen as a significant threat biodiversity, human wellbeing and cultural heritage in grassland ecosystems. [19] [230]

Nicaragua

In Nicaragua Vachellia pennatula is known to encroach due to land intensification as well as land abandonment. [231]

Eastern African grasslands

Across Eastern Africa, including protected areas, woody encroachment has been noted as a challenge. [232] It has first been documented in the 1970s, with scientists indicating that woody encroachment is the rule rather than the exception in East Africa. [233]

Ethiopia

Flowers of the Prosopis juliflora plant that is a common invasive species in Ethiopia and other countries Starr 070404-6610 Prosopis juliflora.jpg
Flowers of the Prosopis juliflora plant that is a common invasive species in Ethiopia and other countries

Grasslands in the Borana Zone in southern Ethiopia are found to be effected by bush encroachment, specifically by Senegalia mellifera, Vachellia reficiens and Vachellia oerfota. [234] [235] Woody plants constitute 52% of vegetation cover. [236] This negatively affects species richness and diversity of plant species. [237] Experiments have shown the effectiveness of bush control of different woody species by cutting and stem-burning, cutting with fire-browse combination, cutting and fire as well as cutting and browsing. Post-management techniques were effective in sustaining savanna ecology. [156] In the Bale lowlands, woody encroachment is found to have increased by 546% between 1990 and 2020, transforming grassland into bushland. [238]

Woody encroachment has been found to reduce herbage yield and therewith rangeland productivity. [239] Under woody encroachment, less meat and milk is produced per head of cattle, which challenges traditional pastoral diets. [240] [241]

Also the invasive species Prosopis juliflora has expanded rapidly since it was introduced in the 1970s, with direct negative consequences on pasture availability and therewith agriculture. Prosopis is native to Central America and was introduced in an attempt to halt land degradation and provide a source of firewood and animal fodder, but has since then encroached into various ecosystems and become a main driver of degradation. [242] The Afar Region is most severely affected. The wood of the invasive species is commonly used as household fuel in the form of firewood and charcoal. [243] [244] [245] [246]

Shrub encroachment in forest areas of Ethiopia, such as the Desa’a Forest, reduces carbon stocks. [247]

Kenya

In Kenya, woody encroachment has been identified as a main type of land-cover change in grasslands, reducing the grazing availability for pastoralists. [248] Studied areas show an increase of woodland by 39% and a decrease of grassland by 74%, with Vachellia reficiens and Vachellia nubica as a dominant species. Observed causes include overgrazing, suppression of wildfires, the reduction of rain as well as the introduction of bush seeds through smallstock [249] [250] Older studies had suggested that an increase in bush cover by 10% reduces grazing by 7%, and grazing is eliminated completely by 90% bush cover. [115] Also Euclea divinorum is a dominant encroaching species. [251] Adaptation strategies include the integration of browsers into the livestock mix, for example goats and camels. [252] [253] In areas where Acacia mellifera encroaches, manual bush thinning during the late dry season combined with reseeding of native grasses and soil conservation measures, proved to be an effective restoration measure with 34% improvement in perennial grass cover. [254]

In the Baringo County of Kenya, up to 30% of grasslands have disappeared due to the invasion of Prosopis juliflora . [90] [255] Clearing Prosopisjuliflora to restore grasslands can increase soil organic carbon and generate value through carbon credit schemes. [256]

Tanzania

In Tanzania woody encroachment has been studied in the savanna ecosystem of the Maswa Game Reserve, with detected shrub growth rates of up to 2.6% per annum. Vachellia drepanolobium is dominant species. [257]

Uganda

Bush encroachment in Uganda is found to have negative impacts on livestock farming. In selected study areas farm income was twice as high on farm that implemented bush control, compared to farms with high bush densities. [258] [259]

West African Guinean and Sudanian savannas

Map of the Guinean forest-savanna mosaic ecoregion Ecoregion AT0707.svg
Map of the Guinean forest-savanna mosaic ecoregion

Bush encroachment is observed across several West African countries, especially in the Guinean savanna and the Sudanian savanna.

Ivory Coast

In Ivory Coast late dry season fires were found to reduce bush encroachment in the Guinean savanna. [260]

Cameroon

In Cameroon, among the regions affected by bush encroachment is Adamawa Region, near the Nigerian border. It has been labelled "pastoral thickets" due to the suspected relation to livestock grazing pressure. [261]

Central African Republic

In the 1960s pastoral land in the Central African Republic was mapped and bush encroachment attributed to livestock pressure as well as reduced fire intensity. [261]

Southern African Savanna

Namibia

Dichrostachys cinerea, a common encroacher species in Namibia Dichrostachys cinerea-IMG 9469.jpg
Dichrostachys cinerea, a common encroacher species in Namibia

Bush encroachment is estimated to affect up to 45 million hectares of savanna in Namibia. As a result, agricultural productivity in Namibia has declined by two-thirds throughout the past decades. The phenomenon affects both commercial and communal farming in Namibia, mostly the central, eastern and north-eastern regions. [262] [263] Common encroacher species include Dichrostachys cinerea, which is most dominant in areas with higher precipitation. [264]

The government of Namibia has recognised bush encroachment as a key challenge for the national economy and food safety. In its current National Development Plan 5, it stipulates that bush shall be thinned on a total of 82.200 hectares per annum. [265] The reduction of bush encroachment on 1.9 million hectares until 2040 is one of Namibia's primary Land Degradation Neutrality Targets under the UNCCD framework. [266] The Government of Namibia pursues a value addition strategy, promoting the sustainable use of bush biomass, which in turn is expected to finance bush harvesting operations. Existing value chains include wood briquettes for household use, woodchips for thermal and electrical energy generation (currently used at Ohorongo Cement factory and at Namibia Breweries Limited), export charcoal, biochar as soil enhancer and animal feed supplement, animal feed, flooring and decking material, predominantly using the invasive species Prosopis, wood carvings, firewood and construction material, i.e. wood composite material. [267]

Increasingly, the encroaching bush is seen as a resource for a biomass industry. Economic assessments were conducted to quantify and value various key ecosystem services and land use options that are threatened by bush encroachment. The evaluation was part of the Economics of Land Degradation (ELD) Initiative, a global initiative established in 2011 by the United Nations Convention to Combat Desertification, the German Federal Ministry for Economic Cooperation and Development, and the European Commission. Based on a national study, cost-benefit analysis suggests a programme of bush control to generate an estimated and aggregated potential net benefit of around N$48.0 billion (USD 3.8 billion) (2015 prices, discounted) over 25 years when compared with a scenario of no bush thinning. This implies a net benefit of around N$2 billion (USD 0.2 billion) (2015 prices, discounted) per annum in the initial round of 25 years. [268]

Namibia has a well-established charcoal sector, which currently comprises approximately 1,200 producers, which employ a total of 10,000 workers. Most producers are farmers, who venture into charcoal production as a means to combat bush encroachment on their land. However, increasingly small enterprises also venture into charcoal making. As per national forestry regulations, charcoal can only be produced from encroaching species. In practice, it however proves difficult to ensure full compliance with these regulations, as the charcoal production is highly decentralised and the inspection capacities of the Directorate of Forestry are low. Voluntary FSC certification has sharply increased in recent years, due to respective demand in many off-take countries, such as the United Kingdom, France and Germany. Due to exclusive use of encroacher bush for charcoal production, rendering the value chain free from deforestation, Namibian charcoal has been dubbed the "greenest charcoal" in an international comparison. [269] In 2016 the Namibia Charcoal Association (NCA) emerged as a legal entity through a restructuring process of the Namibia Charcoal Producers Association, previously attached to Namibia Agricultural Union. It is a non-profit entity and the official industry representation, currently representing an estimated two-thirds of all charcoal producers in the country.

Namibia Biomass Industry Group is a non-profit association under Section 21 of the Companies Act (Act 28 of 2004) of Namibia, founded in 2016. It functions as the umbrella representative body of the emerging bush-based biomass sector in the country with voluntary paid membership. The core objectives, as enlisted in the Articles of Association, include to develop market opportunities for biomass from harvested encroacher bush as well as to address industry bottlenecks, such as skills shortages and research and development needs. The De-bushing Advisory Service is a division of the association, mandated with the dissemination of knowledge on the topics of bush encroachment, bush control and biomass use. Services are provided upon inquiry and are considered a public service and therefore not charged. According to its websites, services include technical advice on bush control and biomass use, environmental advice, the strengthening of existing agricultural outreach services and linkage with service providers. [270] [271]

In 2019, the three Namibian farmers' unions (NNFU, NAU/NLU, NECFU) together with the Ministry of Agriculture, Water and Forestry published a best strategy document called "Reviving Namibia's Livestock Industry". [272] The document states that the Namibian livestock industry is in decline due to the loss of palatable perennial grasses and the increase in bush encroachment. Namibia's rangelands show higher levels of bare ground, lower levels of herbaceous cover, lower perennial grass cover, and higher bush densities over large areas. Bush thickening leads to direct competition for moisture with desirable forage species and detrimentally influences the health of the soil. The best practice document identifies tried and tested practices of both emerging and established farmers from communal and title deed farms. These practices include the Split Ranch Approach, several Holistic Management approaches and the Mara Fodder Bank Approach. Other best practices include bush thinning, landscape re-hydration and fodder production. The unions state that there is major scope for enhancing livelihoods through the sustainable use of bush products. In addition, increased profitability and productivity of the sector will have a major impact on the 70% of the Namibian population that depends directly or indirectly on the rangeland resource for their economic well-being and food security.

Both the Forestry Act and the Environmental Act of Namibia govern bush control. Special harvesting permits as well as Environmental Clearance Certificates are applicable to all bush harvesting activities. Responsible Authority is the Ministry of Environment, Forestry and Tourism. Effective April 2020 the Forest Stewardship Council introduced a national Namibian FSC standard (National Forest Stewardship Standard) that is closely aligned to the global FSC certification standard, but takes into consideration context specific parameters, such as bush encroachment. [273] In early 2020, the total land area certified under the FSC standard for the purpose of bush thinning and biomass processing was reported to amount to 1.6 million hectares. [274]

Botswana

Healthy savanna landscape in Botswana Late afternoon in the Okavango savanna.jpg
Healthy savanna landscape in Botswana

Bush encroachment in Botswana has been documented from as far back as 1971. [275] [276] Around 3.7 million hectares of land in Botswana is affected by bush encroachment, that is over 6% of the total land area. Encroaching species include Acacia tortilis, Acacia erubescens, Acacia mellifera , Dichrostachys cinerea , Grewia flava, and Terminalia sericea . [277] Ecological surveys found bush encroachment affecting both communal grazing areas and private farmland, with particular prevalence in semi-arid ecosystems. [278] [279] Encroachment is considered a key form of land degradation, mainly because of the country's significant dependence on agricultural productivity. [165] In selected areas, charcoal production has been introduced as a measure to reduce bush densities. [280] [281] [275] [282]

South Africa

In South Africa bush encroachment entails the abundance of indigenous woody vegetation in grassland and savanna biomes. [128] These biomes make up 27.9% and 32.5% of the land surface area. About 7.3 million hectares are directly affected by bush encroachment, impacting rural communities socio-economically. [283] [284] Common encroaching species include Vachellia karoo, Senegalia mellifera, Dichrostachys cinera, Rhus undulata and Rhigozum trichotomum. [285]

Through the public works and conservation programme Working for Water, the government of South Africa allocated approximately 100 million USD per annum for the management of native encroaching species. [286] Land users in South Africa commonly combat woody encroachment through clear felling, burning, intensive browsing or chemical control in the form of herbicide application. [285] Studies have found a positive effect of bush thinning on grass biomass production over short periods of time. [287]

The Kruger National Park is largely affected by bush encroachment, which highlights that global drivers cause encroachment also outside typical rangeland settings. [288]

Lesotho

In 1998, around 16% of Lesotho's rangelands where estimated to be affected by woody encroachment, linked to grazing pressure. [289] Encroaching species include Leucosidea sericea and Chrysocoma and a negative impact of water catchment areas is suspected. [290]

Eswatini

Studies in the Lowveld savannas of Eswatini confirm different heavy woody plant encroachment, especially by Dichrostachys cinerea, among other factors related to grazing pressure. In selected study areas the shrub encroachment increased from 2% in 1947 to 31% in 1990. In some affected areas, frequent fires, coupled with drought, reduced bush densities over time. [145] [291]

Zambia

Woody encroachment has been recorded in southern Zambia. Between 1986 and 2010 woody cover increased from 26% to 45% in Kafue Flats and Lochinvar National Park. A common encroacher species is Dichrostachys cinerea. [292]

Zimbabwe

There is evidence of bush encroachment in Zimbabwe, among others by Vachellia karroo. [293] Document notions of woody encroachment in Zimbabwe and its impact on land use date back to 1945. [3] [4]

Other ecoregions

There is evidence of woody encroachment by Acacia leata, Acacia mellifera, Acacia polyacantha, Acacia senegal and Vachellia seyal in Sudan. [294]

Reference map

The following map displays the countries that are addressed in this article, i.e. countries that feature ecosystems with woody encroachment.

See also

Related Research Articles

Forest Dense collection of trees covering a relatively large area

A forest is an area of land dominated by trees. Hundreds of definitions of forest are used throughout the world, incorporating factors such as tree density, tree height, land use, legal standing and ecological function. The Food and Agriculture Organization defines a forest as land spanning more than 0.5 hectares with trees higher than 5 meters and a canopy cover of more than 10 percent, or trees able to reach these thresholds in situ. It does not include land that is predominantly under agricultural or urban land use. Using this definition FRA 2020 found that forests covered 4.06 billion hectares or approximately 31 percent of the global land area in 2020.

Overgrazing When plants are grazed for extended periods without sufficient recovery time

Overgrazing occurs when plants are exposed to intensive grazing for extended periods of time, or without sufficient recovery periods. It can be caused by either livestock in poorly managed agricultural applications, game reserves, or nature reserves. It can also be caused by immobile, travel restricted populations of native or non-native wild animals.

Grassland Area with vegetation dominated by grasses

Grasslands are areas where the vegetation is dominated by grasses (Poaceae). However, sedge (Cyperaceae) and rush (Juncaceae) can also be found along with variable proportions of legumes, like clover, and other herbs. Grasslands occur naturally on all continents except Antarctica and are found in most ecoregions of the Earth. Furthermore, grasslands are one of the largest biomes on earth and dominate the landscape worldwide. There are different types of grasslands: natural grasslands, semi-natural grasslands, and agricultural grasslands. They cover 31–43% of the Earth's land area.

Savanna Mixed woodland-grassland ecosystem

A savanna or savannah is a mixed woodland-grassland ecosystem characterised by the trees being sufficiently widely spaced so that the canopy does not close. The open canopy allows sufficient light to reach the ground to support an unbroken herbaceous layer consisting primarily of grasses.

Mesquite Several species of leguminous trees

Mesquite is a common name for several plants in the genus Prosopis, which contains over 40 species of small leguminous trees. They are native to dry areas in the Americas. They have extremely long roots to seek water from very far under ground. As a legume, mesquites are one of the few sources of fixed nitrogen in the desert habitat. These trees bloom from spring to summer. They often produce fruits known as "pods". Prosopis spp. are able to grow up to 8 metres (26 ft) tall, depending on site and climate. They are deciduous and depending on location and rainfall can have either deep or shallow roots. Prosopis is considered long-lived because of the low mortality rate after the dicotyledonous stage and juveniles are also able to survive in conditions with low light and drought. The Cahuilla indigenous people of western North America were known to eat the seeds of mesquite.

Etosha National Park National park of Namibia

Etosha National Park is a national park in northwestern Namibia and one of the largest national parks in Africa. It was proclaimed a game reserve in March 1907 in Ordinance 88 by the Governor of German South West Africa, Friedrich von Lindequist. It was designated as Wildschutzgebiet in 1958, and was elevated to the status of a national park in 1967 by an act of parliament of the Republic of South Africa. It spans an area of 22,270 km2 (8,600 sq mi) and gets its name from the large Etosha pan which is almost entirely within the park. With an area of 4,760 km2 (1,840 sq mi), the Etosha pan covers 23% of the total area of the national park. The area is home to hundreds of species of mammals, birds and reptiles, including several threatened and endangered species such as the black rhinoceros.

Meadow Open habitat vegetated primarily by non-woody plants

A meadow is an open habitat, or field, vegetated by grasses, herbs, and other non-woody plants. Trees or shrubs may sparsely populate meadows, as long as these areas maintain an open character. Meadows may be naturally occurring or artificially created from cleared shrub or woodland. They can occur naturally under favourable conditions, but they are often maintained by humans for the production of hay, fodder, or livestock. Meadow habitats, as a group, are characterized as "semi-natural grasslands", meaning that they are largely composed of species native to the region, with only limited human intervention.

<i>Bromus tectorum</i> Species of grass

Bromus tectorum, known as downy brome, drooping brome or cheatgrass, is a winter annual grass native to Europe, southwestern Asia, and northern Africa, but has become invasive in many other areas. It now is present in most of Europe, southern Russia, Japan, South Africa, Australia, New Zealand, Iceland, Greenland, North America and western Central Asia. In the eastern US B. tectorum is common along roadsides and as a crop weed, but usually does not dominate an ecosystem. It has become a dominant species in the Intermountain West and parts of Canada, and displays especially invasive behavior in the sagebrush steppe ecosystems where it has been listed as noxious weed. B. tectorum often enters the site in an area that has been disturbed, and then quickly expands into the surrounding area through its rapid growth and prolific seed production.

Rangeland Biomes which can be grazed by animals or livestock (grasslands, woodlands, prairies, etc)

Rangelands are grasslands, shrublands, woodlands, wetlands, and deserts that are grazed by domestic livestock or wild animals. Types of rangelands include tallgrass and shortgrass prairies, desert grasslands and shrublands, woodlands, savannas, chaparrals, steppes, and tundras. Rangelands do not include forests lacking grazable understory vegetation, barren desert, farmland, or land covered by solid rock, concrete and/or glaciers.

Agroforestry Land use management system

Agroforestry is a land use management system in which trees or shrubs are grown around or among crops or pastureland. This diversification of the farming system initiates an agroecological succession, like that in natural ecosystems, and so starts a chain of events that enhance the functionality and sustainability of the farming system. Trees also produce a wide range of useful and marketable products from fruits/nuts, medicines, wood products, etc. This intentional combination of agriculture and forestry has multiple benefits, such as greatly enhanced yields from staple food crops, enhanced farmer livelihoods from income generation, increased biodiversity, improved soil structure and health, reduced erosion, and carbon sequestration. Agroforestry practices are highly beneficial in the tropics, especially in subsistence smallholdings in sub-Saharan Africa and have been found to be beneficial in Europe and the United States.

Fire ecology Study of fire in ecosystems

Fire ecology is a scientific discipline concerned with natural processes involving fire in an ecosystem and the ecological effects, the interactions between fire and the abiotic and biotic components of an ecosystem, and the role as an ecosystem process. Many ecosystems, particularly prairie, savanna, chaparral and coniferous forests, have evolved with fire as an essential contributor to habitat vitality and renewal. Many plant species in fire-affected environments require fire to germinate, establish, or to reproduce. Wildfire suppression not only eliminates these species, but also the animals that depend upon them.

Afforestation Establishment of trees where there were none previously

Afforestation is the establishment of a forest or stand of trees (forestation) in an area where there was no previous tree cover. Many government and non-governmental organizations directly engage in afforestation programs to create forests and increase carbon capture. Afforestation is an increasingly sought-after method to fight climate concerns, as it is known to increase the soil quality and organic carbon levels into the soil, avoiding desertification.

<i>Atriplex nummularia</i> Species of plant

Atriplex nummularia is a species of saltbush from the family Amaranthaceae and is a large woody shrub known commonly as oldman saltbush. A. nummularia is native to Australia and occurs in each of the mainland states, thriving in arid and semi-arid inland regions.

Agriculture in Namibia contributes around 5% of the national Gross Domestic Product though 25% to 40% of Namibians depend on subsistence agriculture and herding. Primary products included livestock and meat products, crop farming and forestry. Only 2% of Namibia's land receives sufficient rainfall to grow crops. As all inland rivers are ephemeral, irrigation is only possible in the valleys of the border rivers Oranje, Kunene, and Okavango, and also at the Hardap Irrigation Scheme.

<i>Vachellia reficiens</i> Species of legume

Vachellia reficiens, commonly known as red-bark acacia, red thorn, false umbrella tree, or false umbrella thorn, is a deciduous tree or shrub of the pea family (Fabaceae) native to southern Africa, often growing in an upside-down cone shape and with a relatively flat crown.

Montane ecosystems Ecosystems found in mountains

Montane ecosystems are found on the slopes of mountains. The alpine climate in these regions strongly affects the ecosystem because temperatures fall as elevation increases, causing the ecosystem to stratify. This stratification is a crucial factor in shaping plant community, biodiversity, metabolic processes and ecosystem dynamics for montane ecosystems. Dense montane forests are common at moderate elevations, due to moderate temperatures and high rainfall. At higher elevations, the climate is harsher, with lower temperatures and higher winds, preventing the growth of trees and causing the plant community to transition to montane grasslands, shrublands or alpine tundra. Due to the unique climate conditions of montane ecosystems, they contain increased numbers of endemic species. Montane ecosystems also exhibit variation in ecosystem services, which include carbon storage and water supply.

Holistic management (agriculture)

Holistic Management in agriculture is an approach to managing resources that was originally developed by Allan Savory. Holistic Management is a registered trademark of Holistic Management International.

Grassland degradation, also called vegetation or steppe degradation is a biotic disturbance in which grass struggles to grow or can no longer exist on a piece of land due to causes such as overgrazing, burrowing of small mammals, and climate change. Since the 1970s, it has been noticed to affects plains and plateaus of alpine meadows or grasslands, most notably being in the Philippines and in the Tibetan and Inner Mongolian region of China where 2460 km2 of grassland is degraded each year. Across the globe it is estimated that 23% of the land is degraded. It takes years and sometimes even decades, depending on what is happening to that piece of land, for a grassland to become degraded. The process is slow and gradual but so is restoring degraded grassland. Initially only patches of grass appear to die and appear brown in nature; but the degradation process, if not addressed, can spread to many acres of land. As a result, the frequency of landslides and dust storms may increase. The degraded land's less fertile ground cannot yield crops nor can animals graze in these fields. With a dramatic decrease in plant diversity in this ecosystem, more carbon and nitrogen may be released into the atmosphere. These results can have serious effects on humans such as displacing herders from their community; a decrease in vegetables, fruit, and meat that are regularly acquired from these fields; and a catalyzing effect on global warming.

Carbon farming

Carbon farming is a name for a variety of agricultural methods aimed at sequestering atmospheric carbon into the soil and in crop roots, wood and leaves. The aim of carbon farming is to increase the rate at which carbon is sequestered into soil and plant material with the goal of creating a net loss of carbon from the atmosphere. Increasing a soil's organic matter content can aid plant growth, increase total carbon content, improve soil water retention capacity and reduce fertilizer use. As of 2016, variants of carbon farming reached hundreds of millions of hectares globally, of the nearly 5 billion hectares (1.2×1010 acres) of world farmland. In addition to agricultural activities, forests management is also a tool that is used in carbon farming. The practice of carbon farming is often done by individual land owners who are given incentive to use and to integrate methods that will sequester carbon through policies created by governments. Carbon farming methods will typically have a cost, meaning farmers and land-owners typically need a way in which they can profit from the use of carbon farming and different governments will have different programs.

<i>Grewia flava</i> Species of plant in the genus Grewia

Grewia flava, the brandy bush, wild currant, velvet raisin, or raisin tree, is a species of flowering plant in the family Malvaceae, native to southern Africa. A common shrub species, it is spreading into grasslands due to human rangeland management practices, and increasing rainfall. The berries are sweet and edible, but have little flesh and so are typically collected to ferment into alcoholic beverages. The desert truffle Kalaharituber pfeilii is often found in association with its roots.

References

  1. 1 2 Van Auken, O.W. (July 2009). "Causes and consequences of woody plant encroachment into western North American grasslands". Journal of Environmental Management. 90 (10): 2931–2942. doi:10.1016/j.jenvman.2009.04.023. PMID   19501450.
  2. Archer, Steve; Boutton, Thomas W.; Hibbard, K.A. (2001), "Trees in Grasslands", Global Biogeochemical Cycles in the Climate System, Elsevier, pp. 115–137, doi:10.1016/b978-012631260-7/50011-x, ISBN   978-0-12-631260-7 , retrieved 3 July 2021
  3. 1 2 Staples, R.R. 1945. Veld Burning. Rhodesian Agricultural Journal 42, 44-52.
  4. 1 2 West, O. (1947). "Thorn bush encroachment in relation to the management of veld grazing". Rhodesian Agricultural Journal. 44: 488–497. OCLC   709537921.
  5. 1 2 Walter, H. (1954). "Die Verbuschung, eine Erscheinung der subtropischen Savannengebiete, und ihre ökologischen Ursachen". Vegetatio Acta Geobot (in German). 5: 6–10. doi:10.1007/BF00299544. S2CID   12772783.
  6. Devine, Aisling P.; McDonald, Robbie A.; Quaife, Tristan; Maclean, Ilya M. D. (2017). "Determinants of woody encroachment and cover in African savannas". Oecologia. 183 (4): 939–951. Bibcode:2017Oecol.183..939D. doi:10.1007/s00442-017-3807-6. ISSN   0029-8549. PMC   5348564 . PMID   28116524.
  7. Trollope, W.S.W.; Trollope, Lynne A.; Bosch, O.J.H. (March 1990). "Veld and pasture management terminology in southern Africa". Journal of the Grassland Society of Southern Africa. 7 (1): 52–61. doi:10.1080/02566702.1990.9648205. ISSN   0256-6702.
  8. Moreira, Francisco; Viedma, Olga; Arianoutsou, Margarita; Curt, Thomas; Koutsias, Nikos; Rigolot, Eric; Barbati, Anna; Corona, Piermaria; Vaz, Pedro; Xanthopoulos, Gavriil; Mouillot, Florent (2011). "Landscape – wildfire interactions in southern Europe: Implications for landscape management". Journal of Environmental Management. 92 (10): 2389–2402. doi:10.1016/j.jenvman.2011.06.028. hdl:10400.5/16228. PMID   21741757.
  9. Zinnert, Julie C.; Nippert, Jesse B.; Rudgers, Jennifer A.; Pennings, Steven C.; González, Grizelle; Alber, Merryl; Baer, Sara G.; Blair, John M.; Burd, Adrian; Collins, Scott L.; Craft, Christopher (May 2021). "State changes: insights from the U.S. Long Term Ecological Research Network". Ecosphere. 12 (5). doi:10.1002/ecs2.3433. ISSN   2150-8925. S2CID   235484735.
  10. Jeltsch, Florian; Milton, Suzanne J.; Dean, W. R. J.; Rooyen, Noel Van (1997). "Analysing Shrub Encroachment in the Southern Kalahari: A Grid-Based Modelling Approach". The Journal of Applied Ecology. 34 (6): 1497. doi:10.2307/2405265. JSTOR   2405265.
  11. Brown, Joel R.; Archer, Steve (1999). "Shrub invasion of grassland: recruitment is continuous and not regulated by herbaceous biomass or density". Ecology. 80 (7): 2385–2396. doi:10.1890/0012-9658(1999)080[2385:SIOGRI]2.0.CO;2. hdl:1969.1/182279. ISSN   0012-9658.[ permanent dead link ]
  12. O'Connor, Tim G; Puttick, James R; Hoffman, M Timm (4 May 2014). "Bush encroachment in southern Africa: changes and causes". African Journal of Range & Forage Science. 31 (2): 67–88. doi:10.2989/10220119.2014.939996. ISSN   1022-0119. S2CID   81059843.
  13. 1 2 Trollope, W.S.W. (1980). "Controlling bush encroachment with fire in the savanna areas of South Africa". Proceedings of the Annual Congresses of the Grassland Society of Southern Africa. 15 (1): 173–177. doi:10.1080/00725560.1980.9648907. ISSN   0072-5560.
  14. 1 2 Van Langevelde, Frank; Van De Vijver, Claudius A. D. M.; Kumar, Lalit; Van De Koppel, Johan; De Ridder, Nico; Van Andel, Jelte; Skidmore, Andrew K.; Hearne, John W.; Stroosnijder, Leo; Bond, William J.; Prins, Herbert H. T. (2003). "Effects of Fire and Herbivory on the Stability of Savanna Ecosystems". Ecology. 84 (2): 337–350. doi:10.1890/0012-9658(2003)084[0337:EOFAHO]2.0.CO;2. hdl:20.500.11755/3d42107b-dbca-4edd-8f47-4405a2531e16. ISSN   0012-9658.
  15. Archibald, Sally; Roy, David P.; van Wilgen, Brian W.; Scholes, Robert J. (March 2009). "What limits fire? An examination of drivers of burnt area in Southern Africa". Global Change Biology. 15 (3): 613–630. Bibcode:2009GCBio..15..613A. doi:10.1111/j.1365-2486.2008.01754.x.
  16. Staver, C.; Archibald, S.; Levin, S.A. (2011). "The Global Extent and Determinants of Savanna and Forest as Alternative Biome States". Science. 334 (6053): 230–232. Bibcode:2011Sci...334..230S. doi:10.1126/science.1210465. PMID   21998389. S2CID   11100977.
  17. Lehmann, Caroline E. R.; Archibald, Sally A.; Hoffmann, William A.; Bond, William J. (2011). "Deciphering the distribution of the savanna biome". New Phytologist. 191 (1): 197–209. doi:10.1111/j.1469-8137.2011.03689.x. PMID   21463328.
  18. Ratajczak, Zak; Nippert, Jesse B.; Briggs, John M.; Blair, John M. (2014). Sala, Osvaldo (ed.). "Fire dynamics distinguish grasslands, shrublands and woodlands as alternative attractors in the Central Great Plains of North America". Journal of Ecology. 102 (6): 1374–1385. doi:10.1111/1365-2745.12311.
  19. 1 2 Sühs, Rafael Barbizan; Giehl, Eduardo Luís Hettwer; Peroni, Nivaldo (December 2020). "Preventing traditional management can cause grassland loss within 30 years in southern Brazil". Scientific Reports. 10 (1): 783. Bibcode:2020NatSR..10..783S. doi:10.1038/s41598-020-57564-z. ISSN   2045-2322. PMC   6972928 . PMID   31964935.
  20. 1 2 Schreiner-McGraw, Adam P.; Vivoni, Enrique R.; Ajami, Hoori; Sala, Osvaldo E.; Throop, Heather L.; Peters, Debra P. C. (December 2020). "Woody Plant Encroachment has a Larger Impact than Climate Change on Dryland Water Budgets". Scientific Reports. 10 (1): 8112. Bibcode:2020NatSR..10.8112S. doi:10.1038/s41598-020-65094-x. ISSN   2045-2322. PMC   7229153 . PMID   32415221.
  21. Skarpe, Christina (December 1990). "Shrub Layer Dynamics Under Different Herbivore Densities in an Arid Savanna, Botswana". The Journal of Applied Ecology. 27 (3): 873–885. doi:10.2307/2404383. JSTOR   2404383.
  22. Wigley, Benjamin J.; Bond, William J.; Hoffman, M. Timm (March 2010). "Thicket expansion in a South African savanna under divergent land use: local vs. global drivers?". Global Change Biology. 16 (3): 964–976. Bibcode:2010GCBio..16..964W. doi:10.1111/j.1365-2486.2009.02030.x.
  23. Bond, W. J.; Midgley, G. F.; Woodward, F. I. (2003). "The importance of low atmospheric CO 2 and fire in promoting the spread of grasslands and savannas". Global Change Biology. 9 (7): 973–982. Bibcode:2003GCBio...9..973B. doi:10.1046/j.1365-2486.2003.00577.x via Wiley.
  24. Tabares, Ximena; Zimmermann, Heike; Dietze, Elisabeth; Ratzmann, Gregor; Belz, Lukas; Vieth‐Hillebrand, Andrea; Dupont, Lydie; Wilkes, Heinz; Mapani, Benjamin; Herzschuh, Ulrike (January 2020). "Vegetation state changes in the course of shrub encroachment in an African savanna since about 1850 CE and their potential drivers". Ecology and Evolution. 10 (2): 962–979. doi: 10.1002/ece3.5955 . PMC   6988543 . PMID   32015858.
  25. Luvuno, Linda; Biggs, Reinette; Stevens, Nicola; Esler, Karen (2018). "Woody Encroachment as a Social-Ecological Regime Shift". Sustainability. 10 (7): 2221. doi: 10.3390/su10072221 .
  26. Kumar, Dushyant; Pfeiffer, Mirjam; Gaillard, Camille; Langan, Liam; Scheiter, Simon (2 June 2020). "Climate change and elevated CO<sub>2</sub> favor forest over savanna under different future scenarios in South Asia". Biogeosciences. 18 (9): 2957–2979. doi:10.5194/bg-2020-169.
  27. Archer SR; Davies K.W; Fulbright T.E; McDaniel K.C; Wilcox B.P.; Predick K.I (2011). "Brush management as a rangeland conservation strategy: a critical evaluation". Conservation benefits of rangeland practices: assessment, recommendations, and knowledge gaps. Allen Press. ISBN   978-0984949908.
  28. García Criado, M; Myers‐Smith, IH; Bjorkman, AD; Lehmann, CER; Stevens, N. (May 2020). "Woody plant encroachment intensifies under climate change across tundra and savanna biomes" (PDF). Glob. Ecol. Biogeogr. 29 (5): 925–943. doi:10.1111/geb.13072. hdl:20.500.11820/cd2cc523-9683-4a09-a6e0-53b354932bf9.
  29. Ncisana, Lusanda; Mkhize, Ntuthuko R; Scogings, Peter F (9 May 2021). "Warming promotes growth of seedlings of a woody encroacher in grassland dominated by C 4 species". African Journal of Range & Forage Science: 1–9. doi:10.2989/10220119.2021.1913762. ISSN   1022-0119. S2CID   236563738.
  30. Stevens, Nicola; Erasmus, B. F. N.; Archibald, S.; Bond, W. J. (19 September 2016). "Woody encroachment over 70 years in South African savannahs: overgrazing, global change or extinction aftershock?". Philosophical Transactions of the Royal Society B: Biological Sciences. 371 (1703): 20150437. doi:10.1098/rstb.2015.0437. ISSN   0962-8436. PMC   4978877 . PMID   27502384.
  31. 1 2 Ward, David; Hoffman, M Timm; Collocott, Sarah J (4 May 2014). "A century of woody plant encroachment in the dry Kimberley savanna of South Africa". African Journal of Range & Forage Science. 31 (2): 107–121. doi:10.2989/10220119.2014.914974. ISSN   1022-0119. S2CID   85329588.
  32. 1 2 IPCC, 2018: Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty [Masson-Delmotte, V., P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P.R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T. Maycock, M. Tignor, and T. Waterfield (eds.)]. In Press.
  33. Bora, Zinabu; Wang, Yongdong; Xu, Xinwen; Angassa, Ayana; You, Yuan (July 2021). "Effects comparison of co-occurring Vachellia tree species on understory herbaceous vegetation biomass and soil nutrient: Case of semi-arid savanna grasslands in southern Ethiopia". Journal of Arid Environments. 190: 104527. Bibcode:2021JArEn.190j4527B. doi:10.1016/j.jaridenv.2021.104527.
  34. 1 2 Eldridge DJ, Soliveres S (2014). "Are shrubs really a sign of declining ecosystem function? Disentangling the myths and truths of woody encroachment in Australia". Australian Journal of Botany. 62 (7): 594–608. doi:10.1071/BT14137 via CSIRO.
  35. 1 2 Deng, Yuanhong; Li, Xiaoyan; Shi, Fangzhong; Hu, Xia; Gillespie, Thomas (31 August 2021). "Woody plant encroachment enhanced global vegetation greening and ecosystem water‐use efficiency". Global Ecology and Biogeography: geb.13386. doi:10.1111/geb.13386. ISSN   1466-822X.
  36. 1 2 3 Eldridge, David J.; Bowker, Matthew A.; Maestre, Fernando T.; Roger, Erin; Reynolds, James F.; Whitford, Walter G. (2011). "Impacts of shrub encroachment on ecosystem structure and functioning: towards a global synthesis". Ecology Letters. 14 (7): 709–722. doi:10.1111/j.1461-0248.2011.01630.x. ISSN   1461-0248. PMC   3563963 . PMID   21592276.
  37. 1 2 Maestre, Fernando T.; Eldridge, David J.; Soliveres, Santiago; Kéfi, Sonia; Delgado-Baquerizo, Manuel; Bowker, Matthew A.; García-Palacios, Pablo; Gaitán, Juan; Gallardo, Antonio; Lázaro, Roberto; Berdugo, Miguel (November 2016). "Structure and Functioning of Dryland Ecosystems in a Changing World". Annual Review of Ecology, Evolution, and Systematics. 47 (1): 215–237. doi:10.1146/annurev-ecolsys-121415-032311. ISSN   1543-592X. PMC   5321561 . PMID   28239303.
  38. Eldridge, David J.; Soliveres, Santiago; Bowker, Matthew A.; Val, James (4 June 2013). "Grazing dampens the positive effects of shrub encroachment on ecosystem functions in a semi-arid woodland". Journal of Applied Ecology. 50 (4): 1028–1038. doi:10.1111/1365-2664.12105. ISSN   0021-8901.
  39. 1 2 Soliveres, Santiago; Maestre, Fernando T.; Eldridge, David J.; Delgado-Baquerizo, Manuel; Quero, José Luis; Bowker, Matthew A.; Gallardo, Antonio (December 2014). "Plant diversity and ecosystem multifunctionality peak at intermediate levels of woody cover in global drylands: Woody dominance and ecosystem functioning". Global Ecology and Biogeography. 23 (12): 1408–1416. doi:10.1111/geb.12215. PMC   4407977 . PMID   25914607.
  40. Riginos, Corinna; Grace, James B.; Augustine, David J.; Young, Truman P. (November 2009). "Local versus landscape-scale effects of savanna trees on grasses". Journal of Ecology. 97 (6): 1337–1345. doi:10.1111/j.1365-2745.2009.01563.x. ISSN   0022-0477.
  41. 1 2 3 Knapp, Alan K.; Briggs, John M.; Collins, Scott L.; Archer, Steven R.; Bret-Harte, M. Syndonia; Ewers, Brent E.; Peters, Debra P.; Young, Donald R.; Shaver, Gaius R.; Pendall, Elise; Cleary, Meagan B. (2008). "Shrub encroachment in North American grasslands: shifts in growth form dominance rapidly alters control of ecosystem carbon inputs: SHRUB ENCROACHMENT INTO GRASSLANDS ALTERS CARBON INPUTS". Global Change Biology. 14 (3): 615–623. doi:10.1111/j.1365-2486.2007.01512.x.
  42. Conant, F. (1982). Thorns paired, sharply recurved: Cultural controls and rangeland quality in East Africa. In Spooner, B., and Mann, H. (eds.), Desertification and Development; Dryland Ecology in Social Perspective. Academic Press, London.
  43. Asner, Gregory P.; Elmore, Andrew J.; Olander, Lydia P.; Martin, Roberta E.; Harris, A. Thomas (21 November 2004). "Grazing Systems, Ecosystem Responses, and Global Change". Annual Review of Environment and Resources. 29 (1): 261–299. doi:10.1146/annurev.energy.29.062403.102142. ISSN   1543-5938.
  44. 1 2 Maestre, Fernando T.; Bowker, Matthew A.; Puche, María D.; Belén Hinojosa, M.; Martínez, Isabel; García-Palacios, Pablo; Castillo, Andrea P.; Soliveres, Santiago; Luzuriaga, Arántzazu L.; Sánchez, Ana M.; Carreira, José A. (September 2009). "Shrub encroachment can reverse desertification in semi-arid Mediterranean grasslands". Ecology Letters. 12 (9): 930–941. doi:10.1111/j.1461-0248.2009.01352.x. PMID   19638041.
  45. 1 2 Smit, G.N. (2005). "Tree thinning as an option to increase herbaceous yield of an encroached semi-arid savanna in South Africa". BMC Ecol. 5: 4. doi: 10.1186/1472-6785-5-4 . PMC   1164409 . PMID   15921528.
  46. Stanton RA Jr; Boone Iv WW; Soto-Shoender J; Fletcher RJ Jr; Blaum N; McCleery RA (2018). "Shrub encroachment and vertebrate diversity: a global meta-analysis". Glob. Ecol. Biogeogr. 27 (3): 368–379. doi:10.1111/geb.12675.
  47. 1 2 3 "Cutting Trees Gives Sage-Grouse Populations a Boost, Scientists Find". Audubon. 10 June 2021. Retrieved 19 June 2021.
  48. 1 2 3 Abreu RC, Hoffmann WA, Vasconcelos HL, Pilon NA, Rossatto DR, Durigan G (2017). "The biodiversity cost of carbon sequestration in tropical savanna". Science Advances. 3: e1701284 (8): e1701284. Bibcode:2017SciA....3E1284A. doi: 10.1126/sciadv.1701284 . PMC   5576881 . PMID   28875172.
  49. 1 2 3 4 5 IPCC, 2019: Climate Change and Land: an IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems, P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pörtner, D. C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J. Petzold, J. Portugal Pereira, P. Vyas, E. Huntley, K. Kissick, M. Belkacemi, J. Malley, (eds.). In press.
  50. Schooley, Robert L.; Bestelmeyer, Brandon T.; Campanella, Andrea (July 2018). "Shrub encroachment, productivity pulses, and core-transient dynamics of Chihuahuan Desert rodents". Ecosphere. 9 (7): e02330. doi:10.1002/ecs2.2330.
  51. Mogashoa, R.; Dlamini, P.; Gxasheka, M. (2020). "Grass species richness decreases along a woody plant encroachment gradient in a semi-arid savanna grassland, South Africa". Landscape Ecol. 36 (2): 617–636. doi:10.1007/s10980-020-01150-1. S2CID   228882177.
  52. Ratajczak, Z.; Nippert, J.; Collins, S. (2012). "Woody encroachment decreases diversity across North American grasslands and savannas". Ecology. 93 (4): 697–703. doi: 10.1890/11-1199.1 . PMID   22690619.
  53. Barbara I. Bleho; Christie L. Borkowsky; Melissa A. Grantham; Cary D. Hamel (2021). "A 20 y Analysis of Weather and Management Effects on a Small White Lady's-slipper (Cypripedium candidum) Population in Manitoba". The American Midland Naturalist. 185 (1): 32–48. doi:10.1637/0003-0031-185.1.32 (inactive 31 May 2021).CS1 maint: DOI inactive as of May 2021 (link)
  54. She, W.; Bai, Y.; Zhang, Y. (2021). "Nitrogen-enhanced herbaceous competition threatens woody species persistence in a desert ecosystem". Plant Soil. 460 (1–2): 333–345. doi:10.1007/s11104-020-04810-y. S2CID   231590340.
  55. Smit, Izak P. J.; Prins, Herbert H. T. (17 September 2015). Crowther, Mathew S. (ed.). "Predicting the Effects of Woody Encroachment on Mammal Communities, Grazing Biomass and Fire Frequency in African Savannas". PLOS ONE. 10 (9): e0137857. Bibcode:2015PLoSO..1037857S. doi: 10.1371/journal.pone.0137857 . ISSN   1932-6203. PMC   4574768 . PMID   26379249.
  56. Nghikembua, Matti T.; Marker, Laurie L.; Brewer, Bruce; Mehtätalo, Lauri; Appiah, Mark; Pappinen, Ari (1 October 2020). "Response of wildlife to bush thinning on the north central freehold farmlands of Namibia". Forest Ecology and Management. 473: 118330. doi:10.1016/j.foreco.2020.118330.
  57. Misher, Chetan; Vanak, Abi Tamim (15 March 2021). "Occupancy and diet of the Indian desert fox Vulpes vulpes pusilla in a Prosopis juliflora invaded semi-arid grassland". Wildlife Biology. 2021 (1). doi:10.2981/wlb.00781. ISSN   0909-6396. S2CID   233685264.
  58. Chen, Anping; Reperant, Leslie; Fischhoff, Ilya R.; Rubenstein, Daniel I. (2021). "Increased vigilance of plains zebras (Equus quagga) in response to more bush coverage in a Kenyan savanna". Climate Change Ecology. 1: 100001. doi:10.1016/j.ecochg.2021.100001.
  59. Meik, Jesse M; Jeo, Richard M; Mendelson III, Joseph R; Jenks, Kate E (2002). "Effects of bush encroachment on an assemblage of diurnal lizard species in central Namibia". Biological Conservation. 106 (1): 29–36. doi:10.1016/s0006-3207(01)00226-9. ISSN   0006-3207.
  60. Andersen, Erik M.; Steidl, Robert J. (2019). "Woody plant encroachment restructures bird communities in semiarid grasslands". Biological Conservation. 240: 108276. doi:10.1016/j.biocon.2019.108276.
  61. Bakker, K. K. 2003. A synthesis of the effect of woody vegetation on grassland nesting birds. Proceedings of the South Dakota Academy of Science 82:233–236.
  62. Schultz P. 2007. Does bush encroachment impact foraging success of the critically endangered Namibian population of the Cape Vulture Gyps coprotheres? MSc thesis, University of Cape Town, South Africa.
  63. 1 2 Jane E. Austin; Deborah A. Buhl (2021). "Breeding Bird Occurrence Across a Gradient of Graminoid- to Shrub-Dominated Fens and Fire Histories". The American Midland Naturalist. 185 (1): 77–109. doi:10.1637/0003-0031-185.1.77 (inactive 31 May 2021).CS1 maint: DOI inactive as of May 2021 (link)
  64. 1 2 Rosenberg, Kenneth V.; Dokter, Adriaan M.; Blancher, Peter J.; Sauer, John R.; Smith, Adam C.; Smith, Paul A.; Stanton, Jessica C.; Panjabi, Arvind; Helft, Laura; Parr, Michael; Marra, Peter P. (4 October 2019). "Decline of the North American avifauna". Science. 366 (6461): 120–124. Bibcode:2019Sci...366..120R. doi:10.1126/science.aaw1313. ISSN   0036-8075. PMID   31604313. S2CID   203719982.
  65. Hofmeyr SD, Symes CT, Underhill LG (2014). "Secretarybird Sagittarius serpentarius Population Trends and Ecology: Insights from South African Citizen Science Data". PLOS ONE. 9(5) e96772 (5): e96772. Bibcode:2014PLoSO...996772H. doi: 10.1371/journal.pone.0096772 . PMC   4016007 . PMID   24816839.
  66. Lautenbach, J. M.; R. T. Plumb; S. G. Robinson; C. A. Hagen; D. A. Haukos; J. C. Pitman (2017). "Lesser Prairie-Chicken Avoidance of Trees in a Grassland Landscape". Rangeland Ecology & Management. 70: 78–86. doi: 10.1016/j.rama.2016.07.008 .
  67. "Endangered Species Act listing proposed for lesser prairie-chicken". www.agri-pulse.com. Retrieved 19 June 2021.
  68. Mahamued, B.; Donald, P.; Collar, N.; Marsden, S.; Ndang'Ang'A, P.; Wondafrash, M.; Lloyd, H. (2021). "Rangeland loss and population decline of the critically endangered Liben Lark Heteromirafra archeri in southern Ethiopia" (PDF). Bird Conservation International. 1–14: 1–14. doi:10.1017/S0959270920000696. S2CID   234250627.
  69. Spottiswoode, C.N.; Wondafrash, M.; Gabremichael, M.N.; Abebe, Y.D.; Mwangi, M.A.K.; Collar, N.J.; Dolman, P.M. (2009). "Rangeland degradation is poised to cause Africa's first recorded avian extinction". Animal Conservation. 12 (3): 249–257. doi:10.1111/j.1469-1795.2009.00246.x.
  70. 1 2 Murray, Darrel B.; Muir, James P.; Miller, Michael S.; Erxleben, Devin R.; Mote, Kevin D. (2021). "Effective Management Practices for Increasing Native Plant Diversity on Mesquite Savanna-Texas Wintergrass-Dominated Rangelands". Rangeland Ecology & Management. 75: 161–169. doi:10.1016/j.rama.2021.01.001. S2CID   232105321.
  71. Sirami, C.; Monadjem, A. (2012). "Changes in bird communities in Swaziland savannas between 1998 and 2008 owing to shrub encroachment". Diversity and Distributions. 18 (4): 390–400. doi: 10.1111/j.1472-4642.2011.00810.x .
  72. Ubach A, Páramo F, Gutiérrez C, Stefanescu C (2020). "Vegetation encroachment drives changes in the composition of butterfly assemblages and species loss in Mediterranean ecosystems". Insect Conservation and Diversity. 13 (2): 151–161. doi:10.1111/icad.12397. S2CID   213753973.
  73. 1 2 Huxman, Travis E.; Wilcox, Bradford P.; Breshears, David D.; Scott, Russell L.; Snyder, Keirith A.; Small, Eric E.; Hultine, Kevin; Pockman, William T.; Jackson, Robert B. (2005). "Ecohydrological Implications of Woody Plant Encroachment". Ecology. 86 (2): 308–319. doi:10.1890/03-0583. hdl:1969.1/179270. JSTOR   3450949.
  74. 1 2 Acharya, Bharat; Kharel, Gehendra; Zou, Chris; Wilcox, Bradford; Halihan, Todd (17 October 2018). "Woody Plant Encroachment Impacts on Groundwater Recharge: A Review". Water. 10 (10): 1466. doi: 10.3390/w10101466 . ISSN   2073-4441.
  75. Zou, Chris; Twidwell, Dirac; Bielski, Christine; Fogarty, Dillon; Mittelstet, Aaron; Starks, Patrick; Will, Rodney; Zhong, Yu; Acharya, Bharat (1 December 2018). "Impact of Eastern Redcedar Proliferation on Water Resources in the Great Plains USA—Current State of Knowledge". Water. 10 (12): 1768. doi: 10.3390/w10121768 . ISSN   2073-4441.
  76. Sandvig, Renee M.; Phillips, Fred M. (August 2006). "Ecohydrological controls on soil moisture fluxes in arid to semiarid vadose zones: Ecohydrology of Arid Vadose Zones". Water Resources Research. 42 (8). doi:10.1029/2005WR004644.
  77. Seyfried, M. S.; Schwinning, S.; Walvoord, M. A.; Pockman, W. T.; Newman, B. D.; Jackson, R. B.; Phillips, F. M. (February 2005). "Ecohydrological Control of Deep Drainage in Arid and Semiarid Regions". Ecology. 86 (2): 277–287. doi:10.1890/03-0568. ISSN   0012-9658.
  78. Fan, Y.; Li, X.-Y.; Li, L.; Wei, J.-Q.; Shi, F.-Z.; Yao, H.-Y.; Liu, L. (2018). "Plant Harvesting Impacts on Soil Water Patterns and Phenology for Shrub-encroached Grassland". Water. 10 (6): 736. doi: 10.3390/w10060736 .
  79. Institute., Texas Agricultural Experiment Station. Texas State Soil and Water Conservation Board. Blackland Research and Extension Center. United States. Natural Resources Conservation Service. Texas Water Resources (2000). Brush management/water yield feasibility studies for four watersheds in Texas. Texas Water Resources Institute. OCLC   385192401.
  80. Caterina, Giulia L.; Will, Rodney E.; Turton, Donald J.; Wilson, Duncan S.; Zou, Chris B. (November 2013). "Water use of Juniperus virginiana trees encroached into mesic prairies in Oklahoma, USA: JUNIPERUS VIRGINIANA WATER USE IN MESIC PRAIRIE". Ecohydrology. 7 (4): 1124–1134. doi:10.1002/eco.1444.
  81. "Shrub encroachment on grasslands can increase groundwater recharge". UC Riverside News. Retrieved 19 June 2021.
  82. W. Rosenthal; W. Dugas; S. Bednarz; T. Dybala; R. Muttiah (2002). "Simulation of Brush Removal within Eight Watersheds in Texas". 2002 Chicago, IL July 28–31, 2002. St. Joseph, MI: American Society of Agricultural and Biological Engineers. doi:10.13031/2013.10415.
  83. Zhang, L.; Dawes, W. R.; Walker, G. R. (March 2001). "Response of mean annual evapotranspiration to vegetation changes at catchment scale". Water Resources Research. 37 (3): 701–708. Bibcode:2001WRR....37..701Z. doi:10.1029/2000WR900325.
  84. Ramankutty, Navin; Evan, Amato T.; Monfreda, Chad; Foley, Jonathan A. (2008). "Farming the planet: 1. Geographic distribution of global agricultural lands in the year 2000: GLOBAL AGRICULTURAL LANDS IN 2000". Global Biogeochemical Cycles. 22 (1). doi:10.1029/2007GB002952.
  85. Pendall, E.; Bachelet, D.; Conant, R. T.; El Masri, B.; Flanagan, L. B.; Knapp, A. K.; Liu, J.; Liu, S.; Schaeffer, S. M. (2018). Cavallaro, N.; Shrestha, G.; Birdsey, R.; Mayes, M. A.; Najjar, R.; Reed, S.; Romero-Lankao, P.; Zhu, Z. (eds.). "Chapter 10: Grasslands. Second State of the Carbon Cycle Report". doi:10.7930/soccr2.2018.ch10.Cite journal requires |journal= (help)
  86. Houghton, R. A. (2003). "Why are estimates of the terrestrial carbon balance so different?: CARBON BALANCE OF TERRESTRIAL ECOSYSTEMS". Global Change Biology. 9 (4): 500–509. doi:10.1046/j.1365-2486.2003.00620.x.
  87. Sankey, Temuulen; Shrestha, Rupesh; Sankey, Joel B.; Hardegree, Stuart; Strand, Eva (2013). "Lidar-derived estimate and uncertainty of carbon sink in successional phases of woody encroachment". Journal of Geophysical Research: Biogeosciences. 118 (3): 1144–1155. Bibcode:2013JGRG..118.1144S. doi:10.1002/jgrg.20088.
  88. 1 2 3 Naikwade, Pratap (16 September 2021). "Changes in Soil Carbon Sequestration during Woody Plant Encroachment in Arid Ecosystems". Plantae Scientia. 4 (4–5): 266–276. doi:10.32439/ps.v4i4-5.266-276. ISSN   2581-589X.
  89. Barger, Nichole N.; Archer, Steven R.; Campbell, John L.; Huang, Cho-ying; Morton, Jeffery A.; Knapp, Alan K. (10 August 2011). "Woody plant proliferation in North American drylands: A synthesis of impacts on ecosystem carbon balance". Journal of Geophysical Research. 116 (G4): G00K07. Bibcode:2011JGRG..116.0K07B. doi:10.1029/2010JG001506. ISSN   0148-0227.
  90. 1 2 3 Mbaabu, Purity Rima; Olago, Daniel; Gichaba, Maina; Eckert, Sandra; Eschen, René; Oriaso, Silas; Choge, Simon Kosgei; Linders, Theo Edmund Werner; Schaffner, Urs (2020). "Restoration of degraded grasslands, but not invasion by Prosopis juliflora, avoids trade-offs between climate change mitigation and other ecosystem services". Scientific Reports. 10 (1): 20391. doi:10.1038/s41598-020-77126-7. ISSN   2045-2322. PMC   7686326 . PMID   33235254.
  91. Pinno, Bradley D.; Wilson, Scott D. (2011). "Ecosystem carbon changes with woody encroachment of grassland in the northern Great Plains". Écoscience. 18 (2): 157–163. doi:10.2980/18-2-3412. ISSN   1195-6860. S2CID   86413227.
  92. Wigley, Benjamin J.; Augustine, David J.; Coetsee, Corli; Ratnam, Jayashree; Sankaran, Mahesh (May 2020). "Grasses continue to trump trees at soil carbon sequestration following herbivore exclusion in a semiarid African savanna". Ecology. 101 (5): e03008. doi:10.1002/ecy.3008. ISSN   0012-9658. PMID   32027378.
  93. 1 2 3 Liu, Yun-Hua; Cheng, Jun-Hui; Schmid, Bernhard; Tang, Li-Song; Sheng, Jian-Dong (1 April 2020). Zhang, Wen-Hao (ed.). "Woody plant encroachment may decrease plant carbon storage in grasslands under future drier conditions". Journal of Plant Ecology. 13 (2): 213–223. doi:10.1093/jpe/rtaa003. ISSN   1752-993X.
  94. Mureva, Admore; Ward, David; Pillay, Tiffany; Chivenge, Pauline; Cramer, Michael (2018). "Soil Organic Carbon Increases in Semi-Arid Regions while it Decreases in Humid Regions Due to Woody-Plant Encroachment of Grasslands in South Africa". Scientific Reports. 8 (1): 15506. Bibcode:2018NatSR...815506M. doi:10.1038/s41598-018-33701-7. ISSN   2045-2322. PMC   6195563 . PMID   30341313.
  95. 1 2 Barger, N. N., Archer, S. R., Campbell, J. L., Huang, C., Morton, J. A., and Knapp, A. K. (2011). "Woody plant proliferation in North American drylands: A synthesis of impacts on ecosystem carbon balance". J. Geophys. Res. 116 G00K07 (G4): G00K07. Bibcode:2011JGRG..116.0K07B. doi: 10.1029/2010JG001506 .CS1 maint: multiple names: authors list (link)
  96. Jackson RB, Banner JL, Jobbágy EG, Pockman WT, Wall DH (2002). "Ecosystem carbon loss with woody plant invasion of grasslands". Nature. 418 (6898): 623–626. Bibcode:2002Natur.418..623J. doi:10.1038/nature00910. PMID   12167857. S2CID   14566976.CS1 maint: multiple names: authors list (link)
  97. Goodale, Christine L.; Davidson, Eric A. (2002). "Uncertain sinks in the shrubs". Nature. 418 (6898): 593–594. doi:10.1038/418593a. ISSN   0028-0836. PMID   12167839. S2CID   4428502.
  98. Duke University (2002). "Trees Encroaching Grasslands May Lock Up Less Carbon Than Predicted". ScienceDaily. Retrieved 6 February 2021.
  99. Jackson, Robert B.; Banner, Jay L.; Jobbágy, Esteban G.; Pockman, William T.; Wall, Diana H. (2002). "Ecosystem carbon loss with woody plant invasion of grasslands". Nature. 418 (6898): 623–626. Bibcode:2002Natur.418..623J. doi:10.1038/nature00910. ISSN   0028-0836. PMID   12167857. S2CID   14566976.
  100. Scott, Russell L.; Huxman, Travis E.; Williams, David G.; Goodrich, David C. (2006). "Ecohydrological impacts of woody-plant encroachment: seasonal patterns of water and carbon dioxide exchange within a semiarid riparian environment". Global Change Biology. 12 (2): 311–324. Bibcode:2006GCBio..12..311S. doi:10.1111/j.1365-2486.2005.01093.x.
  101. Abril A, Barttfeld P & Bucher EH (2005). "The effect of fire and overgrazing disturbances on soil carbon balance in the Dry Chaco forest". Forest Ecology and Management. 206 (1–3): 399–405. doi:10.1016/j.foreco.2004.11.014 via ScienceDirect.
  102. Leitner, Monica; Davies, Andrew B.; Parr, Catherine L.; Eggleton, Paul; Robertson, Mark P. (2018). "Woody encroachment slows decomposition and termite activity in an African savanna". Global Change Biology. 24 (6): 2597–2606. Bibcode:2018GCBio..24.2597L. doi:10.1111/gcb.14118. hdl:2263/64671. PMID   29516645. S2CID   3722515.
  103. Yusuf, Hasen M.; Treydte, Anna C.; Sauerborn, Jauchim (13 October 2015). Balestrini, Raffaella (ed.). "Managing Semi-Arid Rangelands for Carbon Storage: Grazing and Woody Encroachment Effects on Soil Carbon and Nitrogen". PLOS ONE. 10 (10): e0109063. Bibcode:2015PLoSO..1009063Y. doi: 10.1371/journal.pone.0109063 . ISSN   1932-6203. PMC   4603954 . PMID   26461478.
  104. Zhou, Yong; Boutton, Thomas W.; Wu, X. Ben (2017). McCulley, Rebecca (ed.). "Soil carbon response to woody plant encroachment: importance of spatial heterogeneity and deep soil storage". Journal of Ecology. 105 (6): 1738–1749. doi:10.1111/1365-2745.12770.
  105. Li, He; Shen, Haihua; Chen, Leiyi; Liu, Taoyu; Hu, Huifeng; Zhao, Xia; Zhou, Luhong; Zhang, Pujin; Fang, Jingyun (2016). "Effects of shrub encroachment on soil organic carbon in global grasslands". Scientific Reports. 6 (1): 28974. Bibcode:2016NatSR...628974L. doi:10.1038/srep28974. ISSN   2045-2322. PMC   4937411 . PMID   27388145.
  106. Terrer, C.; Phillips, R. P.; Hungate, B. A.; Rosende, J.; Pett-Ridge, J.; Craig, M. E.; van Groenigen, K. J.; Keenan, T. F.; Sulman, B. N.; Stocker, B. D.; Reich, P. B. (25 March 2021). "A trade-off between plant and soil carbon storage under elevated CO2". Nature. 591 (7851): 599–603. Bibcode:2021Natur.591..599T. doi:10.1038/s41586-021-03306-8. ISSN   0028-0836. PMID   33762765. S2CID   232355402.
  107. Schlesinger, William H.; Pilmanis, Adrienne M. (1998). "Plant-soil interactions in deserts". Biogeochemistry. 42 (1/2): 169–187. doi:10.1023/A:1005939924434. S2CID   93294785.
  108. Puttock, Alan; Dungait, Jennifer A. J.; Macleod, Christopher J. A.; Bol, Roland; Brazier, Richard E. (December 2014). "Woody plant encroachment into grasslands leads to accelerated erosion of previously stable organic carbon from dryland soils". Journal of Geophysical Research: Biogeosciences. 119 (12): 2345–2357. Bibcode:2014JGRG..119.2345P. doi:10.1002/2014JG002635. ISSN   2169-8953.
  109. 1 2 3 Archer, Steven R.; Andersen, Erik M.; Predick, Katharine I.; Schwinning, Susanne; Steidl, Robert J.; Woods, Steven R. (2017), Briske, David D. (ed.), "Woody Plant Encroachment: Causes and Consequences", Rangeland Systems, Cham: Springer International Publishing, pp. 25–84, doi:10.1007/978-3-319-46709-2_2, ISBN   978-3-319-46707-8 , retrieved 8 March 2021
  110. Petrie, M. D.; Collins, S. L.; Swann, A. M.; Ford, P. L.; Litvak, M.E. (2015). "Grassland to shrubland state transitions enhance carbon sequestration in the northern Chihuahuan Desert". Global Change Biology. 21 (3): 1226–1235. Bibcode:2015GCBio..21.1226P. doi:10.1111/gcb.12743. ISSN   1354-1013. PMID   25266205.
  111. Throop, Heather L.; Munson, Seth; Hornslein, Nicole; McClaran, Mitchel P. (22 July 2021). "Shrub influence on soil carbon and nitrogen in a semi-arid grassland is mediated by precipitation and largely insensitive to livestock grazing". Arid Land Research and Management: 1–20. doi:10.1080/15324982.2021.1952660. ISSN   1532-4982.
  112. Scott, Russell L.; Biederman, Joel A.; Hamerlynck, Erik P.; Barron‐Gafford, Greg A. (2015). "The carbon balance pivot point of southwestern U.S. semiarid ecosystems: Insights from the 21st century drought". Journal of Geophysical Research: Biogeosciences. 120 (12): 2612–2624. Bibcode:2015JGRG..120.2612S. doi:10.1002/2015JG003181. ISSN   2169-8953.
  113. Clemmensen, Karina Engelbrecht; Durling, Mikael Brandström; Michelsen, Anders; Hallin, Sara; Finlay, Roger D.; Lindahl, Björn D. (June 2021). Liu, Lingli (ed.). "A tipping point in carbon storage when forest expands into tundra is related to mycorrhizal recycling of nitrogen". Ecology Letters. 24 (6): 1193–1204. doi:10.1111/ele.13735. ISSN   1461-023X. PMID   33754469. S2CID   232323007.
  114. Oba G, Post E, Syvertsen PO, Stenseth NC (2000). "Bush cover and range condition assessments in relation to landscape and grazing in southern Ethiopia". Landscape Ecology. 15 (6): 535–546. doi:10.1023/A:1008106625096. S2CID   21986173.
  115. 1 2 van., Wijngaarden, Willem (November 1985). Elephants, trees, grass, grazers : relationships between climate, soils, vegetation and large herbivores in a semi-arid savanna ecosystem (Tsavo, Kenya). International Institute for Aerospace Survey and Earth Sciences. ISBN   90-6164-048-2. OCLC   870274791.
  116. Anadón, J. D.; Sala, O. E.; Turner, B. L.; Bennett, E. M. (2 September 2014). "Effect of woody-plant encroachment on livestock production in North and South America". Proceedings of the National Academy of Sciences. 111 (35): 12948–12953. Bibcode:2014PNAS..11112948A. doi: 10.1073/pnas.1320585111 . ISSN   0027-8424. PMC   4156688 . PMID   25136084.
  117. Gray, E.F.; Bond, W.J. (2013). "Will woody plant encroachment impact the visitor experience and economy of conservation areas?". Koedoe. 55 (1). Art. #1106. doi: 10.4102/koedoe.v55i1.1106 .
  118. 1 2 D'Adamo, Francesco; Ogutu, Booker; Brandt, Martin; Schurgers, Guy; Dash, Jadunandan (July 2021). "Climatic and non-climatic vegetation cover changes in the rangelands of Africa". Global and Planetary Change. 202: 103516. Bibcode:2021GPC...20203516D. doi:10.1016/j.gloplacha.2021.103516.
  119. Yu, Peng; Qiuying, Zhang; Yuanzhan, Chen; Ning, Xu; Yunfeng, Qiao; Chao, Tian; Hirwa, Hubert; Diop, Salif; Guisse, Aliou; Fadong, Li (12 May 2021). "Resilience, Adaptability, and Regime Shifts Thinking: A Perspective of Dryland Socio-Ecology System". Journal of Resources and Ecology. 12 (3). doi:10.5814/j.issn.1674-764x.2021.03.007. ISSN   1674-764X. S2CID   234474418.
  120. The Earth as transformed by human action : global and regional changes in the biosphere over the past 300 years. B. L. Turner. Cambridge: Cambridge University Press with Clark University. 1990. ISBN   0-521-36357-8. OCLC   20294746.CS1 maint: others (link)
  121. Noden, Bruce H.; Tanner, Evan P.; Polo, John A.; Fuhlendorf, Sam D. (June 2021), Invasive woody plants as foci of tick-borne pathogens: eastern redcedar in the southern Great Plains, Journal of Vector Ecology, 46 (1), 12–18
  122. Cho, Mee-Hyun; Yang, Ah-Ryeon; Baek, Eun-Hyuk; Kang, Sarah M.; Jeong, Su-Jong; Kim, Jin Young; Kim, Baek-Min (May 2018). "Vegetation-cloud feedbacks to future vegetation changes in the Arctic regions". Climate Dynamics. 50 (9–10): 3745–3755. doi:10.1007/s00382-017-3840-5. ISSN   0930-7575. S2CID   54037132.
  123. Ge, Jianjun; Zou, Chris (August 2013). "Impacts of woody plant encroachment on regional climate in the southern Great Plains of the United States: Woody Encroachment and Climate". Journal of Geophysical Research: Atmospheres. 118 (16): 9093–9104. doi:10.1002/jgrd.50634.
  124. Lima, K. A.; Stevens, N.; Wisely, S. M.; Fletcher, R. J.; Monadjeme, A.; Austim, J. D.; Mahlabae, T.; McCleery, R. A. (2021). "Landscape heterogeneity and woody encroachment decrease mesocarnivore scavenging in a savanna agro-ecosystem". Rangeland Ecology and Management. 78: 104–111. doi:10.1016/j.rama.2021.06.003. ISSN   1550-7424.
  125. Goslee, S.C; Havstad, K.M; Peters, D.P.C; Rango, A; Schlesinger, W.H (2003). "High-resolution images reveal rate and pattern of shrub encroachment over six decades in New Mexico, U.S.A." Journal of Arid Environments. 54 (4): 755–767. Bibcode:2003JArEn..54..755G. doi:10.1006/jare.2002.1103.
  126. Zhao, Yujin; Liu, Xiaoliang; Wang, Yang; Zheng, Zhaoju; Zheng, Shuxia; Zhao, Dan; Bai, Yongfei (September 2021). "UAV-based individual shrub aboveground biomass estimation calibrated against terrestrial LiDAR in a shrub-encroached grassland". International Journal of Applied Earth Observation and Geoinformation. 101: 102358. Bibcode:2021IJAEO.10102358Z. doi:10.1016/j.jag.2021.102358.
  127. Graw, Valerie; Oldenburg, Carsten; Dubovyk, Olena (2016). "Bush Encroachment Mapping for Africa: Multi-Scale Analysis with Remote Sensing and GIS". SSRN Electronic Journal. doi:10.2139/ssrn.2807811. ISSN   1556-5068.
  128. 1 2 Hottman, M.T.; O'Connor, T.G. (1999). "Vegetation change over 40 years in the Weenen/Muden area, KwaZulu-Natal: evidence from photo-panoramas". African Journal of Range & Forage Science. 16 (2–3): 71–88. doi:10.2989/10220119909485721. ISSN   1022-0119.
  129. Tabares, Ximena; Ratzmann, Gregor; Kruse, Stefan; Theuerkauf, Martin; Mapani, Benjamin; Herzschuh, Ulrike (25 March 2021). "Relative pollen productivity estimates of savanna taxa from southern Africa and their application to reconstruct shrub encroachment during the last century". The Holocene. 31 (7): 095968362110031. Bibcode:2021Holoc..31.1100T. doi:10.1177/09596836211003193. ISSN   0959-6836. S2CID   233680350.
  130. Hao, Guang; Yang, Nan; Dong, Ke; Xu, Yujuan; Ding, Xinfeng; Shi, Xinjian; Chen, Lei; Wang, Jinlong; Zhao, Nianxi; Gao, Yubao (10 May 2021). "Shrub‐encroached grassland as an alternative stable state in semiarid steppe regions: Evidence from community stability and assembly". Land Degradation & Development. 32 (10): 3142–3153. doi:10.1002/ldr.3975. ISSN   1085-3278. S2CID   235543749.
  131. 1 2 Briggs, John M.; Knapp, Alan K.; Blair, John M.; Heisler, Jana L.; Hoch, Greg A.; Lett, Michelle S.; McCARRON, James K. (2005). "An Ecosystem in Transition: Causes and Consequences of the Conversion of Mesic Grassland to Shrubland". BioScience. 55 (3): 243. doi:10.1641/0006-3568(2005)055[0243:AEITCA]2.0.CO;2. ISSN   0006-3568.
  132. Miaojun Ma, Scott L Collins, Zak Ratajczak, Guozhen Du (2021). "Soil Seed Banks, Alternative Stable State Theory, and Ecosystem Resilience". BioScience. 71 (7): 697–707. doi:10.1093/biosci/biab011. ISSN   0006-3568.CS1 maint: multiple names: authors list (link)
  133. Giles, André L.; Flores, Bernardo M.; Rezende, Andréia Alves; Weiser, Veridiana de Lara; Cavassan, Osmar (August 2021). "Thirty years of clear-cutting maintain diversity and functional composition of woody-encroached Neotropical savannas". Forest Ecology and Management. 494: 119356. doi:10.1016/j.foreco.2021.119356.
  134. Eldridge, David J.; Ding, Jingyi (March 2021). "Remove or retain: ecosystem effects of woody encroachment and removal are linked to plant structural and functional traits". New Phytologist. 229 (5): 2637–2646. doi:10.1111/nph.17045. ISSN   0028-646X. PMID   33118178.
  135. Kambongi, T; Heyns, L; Rodenwoldt, D; Edwards, S (8 February 2021). "A description of daytime resting sites used by brown hyaenas (Parahyaena brunnea) from a high-density, enclosed population in north-central Namibia". Namibian Journal of Environment. 5.CS1 maint: multiple names: authors list (link)
  136. Choi, D; Fish, A.; Moorman, C; DePerno, C: Schillaci, J (2021). "Breeding-season Survival, Home-range Size, and Habitat Selection of Female Bachman's Sparrows". Southeastern Naturalist. 1: 105–116.CS1 maint: multiple names: authors list (link)
  137. O'Connor, T G; Kuyler, P; Kirkman, K P; Corcoran, B (11 August 2010). "Which grazing management practices are most appropriate for maintaining biodiversity in South African grassland?". African Journal of Range & Forage Science. 27 (2): 67–76. doi:10.2989/10220119.2010.502646. ISSN   1022-0119. S2CID   84555081.
  138. Webb, Nicholas P.; Stokes, Christopher J.; Marshall, Nadine A. (October 2013). "Integrating biophysical and socio-economic evaluations to improve the efficacy of adaptation assessments for agriculture". Global Environmental Change. 23 (5): 1164–1177. doi:10.1016/j.gloenvcha.2013.04.007.
  139. Grande, D. (2013). "Endozoochorus seed dispersal by goats: recovery, germinability and emergence of five Mediterranean shrub species". Spanish Journal of Agricultural Research. 11 (2): 347–355. doi: 10.5424/sjar/2013112-3673 .
  140. I. P. J. Smit; G. P. Asner; N. Govender; N. R. Vaughn; B. W. van Wilgen (2016). "An examination of the potential efficacy of high-intensity fires for reversing woody encroachment in savannas". Journal of Applied Ecology. 53 (5): 1623–1633. doi: 10.1111/1365-2664.12738 .
  141. 1 2 Twidwell, D.; Fuhlendorf, S.D.; Taylor, C.A. Jr; Rogers, W.E. (2013). "Refining thresholds in coupled fire-vegetation models to improve management of encroaching woody plants in grasslands". J. Appl. Ecol. 50 (3): 603–613. doi: 10.1111/1365-2664.12063 .
  142. Fuhlendorf, Samuel D.; Engle, David M.; Kerby, Jay; Hamilton, Robert (2009). "Pyric Herbivory: Rewilding Landscapes through the Recoupling of Fire and Grazing". Conservation Biology. 23 (3): 588–598. doi:10.1111/j.1523-1739.2008.01139.x. ISSN   0888-8892. JSTOR   29738775. PMID   19183203.
  143. Hempson, Gareth P.; Archibald, Sally; Bond, William J. (8 December 2017). "The consequences of replacing wildlife with livestock in Africa". Scientific Reports. 7 (1): 17196. Bibcode:2017NatSR...717196H. doi:10.1038/s41598-017-17348-4. ISSN   2045-2322. PMC   5722938 . PMID   29222494.
  144. Venter, Zander S.; Hawkins, Heidi-Jayne; Cramer, Michael D. (2017). "Implications of historical interactions between herbivory and fire for rangeland management in African savannas". Ecosphere. 8 (10): e01946. doi:10.1002/ecs2.1946. ISSN   2150-8925.
  145. 1 2 Roques, K.G.; O'Connor, T.G.; Watkinson, A.R. (2001). "Dynamics of shrub encroachment in an African savanna: relative influences of fire, herbivory, rainfall and density dependence: Dynamics and causes of shrub encroachment". Journal of Applied Ecology. 38 (2): 268–280. doi:10.1046/j.1365-2664.2001.00567.x.
  146. Trollope, W.S.W. (1974). "Role of fire in preventing bush encroachment in the Eastern Cape". Proceedings of the Annual Congresses of the Grassland Society of Southern Africa. 9 (1): 67–72. doi:10.1080/00725560.1974.9648722. ISSN   0072-5560.
  147. Wedel, Emily R.; Nippert, Jesse B.; Hartnett, David C. (6 July 2021). "Fire and browsing interact to alter intra-clonal stem dynamics of an encroaching shrub in tallgrass prairie". Oecologia. 196 (4): 1039–1048. doi:10.1007/s00442-021-04980-1. ISSN   0029-8549. PMID   34228246.
  148. 1 2 3 D. Twidwell, D. Fogarty. "A guide to reducing risk and vulnerability to woody encroachment in rangelands" (PDF). University of Nebraska-Lincoln.
  149. Bielski, Christine H.; Scholtz, Rheinhardt; Donovan, Victoria M.; Allen, Craig R.; Twidwell, Dirac (August 2021). "Overcoming an "irreversible" threshold: A 15-year fire experiment". Journal of Environmental Management. 291: 112550. doi:10.1016/j.jenvman.2021.112550. PMID   33965707.
  150. Nippert, Jesse B.; Telleria, Lizeth; Blackmore, Pamela; Taylor, Jeffrey H.; O'Connor, Rory C. (September 2021). "Is a Prescribed Fire Sufficient to Slow the Spread of Woody Plants in an Infrequently Burned Grassland? A Case Study in Tallgrass Prairie". Rangeland Ecology & Management. 78: 79–89. doi:10.1016/j.rama.2021.05.007.
  151. Novak, Erin N.; Bertelsen, Michelle; Davis, Dick; Grobert, Devin M.; Lyons, Kelly G.; Martina, Jason P.; McCaw, W. Matt; O'Toole, Matthew; Veldman, Joseph W. (September 2021). "Season of prescribed fire determines grassland restoration outcomes after fire exclusion and overgrazing". Ecosphere. 12 (9). doi:10.1002/ecs2.3730. ISSN   2150-8925.
  152. H., Jacobs, Alan (1980). Pastoral Maasai and tropical rural development. Agricultural development in Africa: issues of public policy. New York: Praeger. pp. 275–300. OCLC   772636262.
  153. Rodrigo Baggio, Gerhard E. Overbeck, Giselda Durigan, Valério D. Pillar (June 2021). "To graze or not to graze: A core question for conservation and sustainable use of grassy ecosystems in Brazil". Perspectives in Ecology and Conservation. 19 (3): 256–266. doi:10.1016/j.pecon.2021.06.002. ISSN   2530-0644.CS1 maint: multiple names: authors list (link)
  154. Smit, G.N., C.G.F. Ritcher, and A.J. Aucamp. 1999. Bush encroachment: An approach to understanding and managing the problem. In Veld management in South Africa, ed. N.M. Tainton. Pietermaritzburg: University of Natal Press.
  155. Pratt, D. J. (April 1971). "Bush-Control Studies in the Drier Areas of Kenya. VI. Effects of Fenuron (3-Phenyl-1,1-Dimethylurea)". The Journal of Applied Ecology. 8 (1): 239–245. doi:10.2307/2402141. JSTOR   2402141.
  156. 1 2 Hare, Malicha Loje; Xu, Xinwen; Wang, Yongdong; Gedda, Abule Ibro (December 2020). "The effects of bush control methods on encroaching woody plants in terms of die-off and survival in Borana rangelands, southern Ethiopia". Pastoralism. 10 (1): 16. doi:10.1186/s13570-020-00171-4. ISSN   2041-7136. S2CID   220881346.
  157. Alados, C.L.; Saiz, H.; Nuche, P.; Gartzia, M.; Komac, B.; De Frutos, Á.; Pueyo, Y. (4 September 2019). "Clearing vs. burning for restoring Pyrenean grasslands after shrub encroachment". Cuadernos de Investigación Geográfica. 45 (2): 441. doi:10.18172/cig.3589. ISSN   1697-9540.
  158. Albrecht, Matthew A.; Dell, Noah D.; Engelhardt, Megan J.; Reid, J. Leighton; Saxton, Michael L.; Trager, James C.; Waldman, Claire; Long, Quinn G. (3 September 2021). "Recovery of herb‐layer vegetation and soil properties after pile burning in a Midwestern oak woodland". Restoration Ecology. doi:10.1111/rec.13547. ISSN   1061-2971.
  159. Archer, S.; Predick, K. (2014). "An ecosystem services perspective on brush management: research priorities for competing land-use objectives". Journal of Ecology. 102 (6): 1394–1407. doi: 10.1111/1365-2745.12314 .
  160. Scholtz, Rheinhardt; Fuhlendorf, Samuel D.; Uden, Daniel R.; Allred, Brady W.; Jones, Matthew O.; Naugle, David E.; Twidwell, Dirac (July 2021). "Challenges of Brush Management Treatment Effectiveness in Southern Great Plains, United States". Rangeland Ecology & Management. 77: 57–65. doi:10.1016/j.rama.2021.03.007. S2CID   234820208.
  161. 1 2 Fogarty DT, Roberts CP, Uden DR, Donovan VM, Allen CR, Naugle DE, Jones MO, Allred BW, Twidwell D (2020). "Woody Plant Encroachment and the Sustainability of Priority Conservation Areas". Sustainability. 12 (20): 8321. doi: 10.3390/su12208321 .
  162. Van Wilgen, B.W.; Forsyth, G.G.; Le Maitre, D.C.; Wannenburgh, A.; Kotzé, J.D.F.; van den Berg, E.; Henderson, L. (2012). "An assessment of the effectiveness of a large, national-scale invasive alien plant control strategy in South Africa". Biol. Conserv. 148: 28–38. doi:10.1016/j.biocon.2011.12.035. hdl:10019.1/113015.
  163. 1 2 Halpern, Charles B.; Antos, Joseph A. (2021). "Rates, patterns, and drivers of tree reinvasion 15 years after large-scale meadow-restoration treatments". Restoration Ecology. n/a (n/a): e13377. doi:10.1111/rec.13377. ISSN   1526-100X. S2CID   233367081.
  164. Nghikembua, Matti T; Marker, Laurie L; Brewer, Bruce; Leinonen, Arvo; Mehtätalo, Lauri; Appiah, Mark; Pappinen, Ari (27 March 2021). "Restoration thinning reduces bush encroachment on freehold farmlands in north-central Namibia". Forestry: An International Journal of Forest Research. 94 (4): cpab009. doi:10.1093/forestry/cpab009. ISSN   0015-752X.
  165. 1 2 Reed, M.S.; Stringer, L.C.; Dougill, A.J.; Perkins, J.S.; Atlhopheng, J.R.; Mulale, K.; Favretto, N. (March 2015). "Reorienting land degradation towards sustainable land management: Linking sustainable livelihoods with ecosystem services in rangeland systems". Journal of Environmental Management. 151: 472–485. doi:10.1016/j.jenvman.2014.11.010. PMID   25617787.
  166. Kayler, Z., Janowiak, M., Swanston, C. (2017). "The Global Carbon Cycle". Considering Forest and Grassland Carbon in Land Management. General Technical Report WTO-GTR-95. 95. United States Department of Agriculture, Forest Service. pp. 3–9. doi:10.2737/WO-GTR-95.CS1 maint: multiple names: authors list (link)
  167. Pacala, S. W. (22 June 2001). "Consistent Land- and Atmosphere-Based U.S. Carbon Sink Estimates". Science. 292 (5525): 2316–2320. Bibcode:2001Sci...292.2316P. doi:10.1126/science.1057320. PMID   11423659. S2CID   31060636.
  168. Boutton, T.W.; Liao, J.D.; Filley, T.R.; Archer, S.R. (26 October 2015), Lal, Rattan; Follett, Ronald F. (eds.), "Belowground Carbon Storage and Dynamics Accompanying Woody Plant Encroachment in a Subtropical Savanna", SSSA Special Publications, Madison, WI, USA: American Society of Agronomy and Soil Science Society of America, pp. 181–205, doi:10.2136/sssaspecpub57.2ed.c12, ISBN   978-0-89118-859-9 , retrieved 7 March 2021
  169. Houghton, R. A. (23 July 1999). "The U.S. Carbon Budget: Contributions from Land-Use Change". Science. 285 (5427): 574–578. doi:10.1126/science.285.5427.574. PMID   10417385.
  170. Thijs, Ann (2014). Biotic and abiotic controls on carbon dynamics in a Central Texas encroaching savanna (Thesis).
  171. Hurtt, G. C.; Pacala, S. W.; Moorcroft, P. R.; Caspersen, J.; Shevliakova, E.; Houghton, R. A.; Moore, B. (5 February 2002). "Projecting the future of the U.S. carbon sink". Proceedings of the National Academy of Sciences. 99 (3): 1389–1394. Bibcode:2002PNAS...99.1389H. doi: 10.1073/pnas.012249999 . ISSN   0027-8424. PMC   122200 . PMID   11830663.
  172. Burrows, W. H.; Henry, B. K.; Back, P. V.; Hoffmann, M. B.; Tait, L. J.; Anderson, E. R.; Menke, N.; Danaher, T.; Carter, J. O.; . McKeon, G. M (1 August 2002). "Growth and carbon stock change in eucalypt woodlands in northeast Australia: ecological and greenhouse sink implications: GROWTH and CARBON STOCK CHANGE IN EUCALYPT WOODLANDS". Global Change Biology. 8 (8): 769–784. doi:10.1046/j.1365-2486.2002.00515.x.
  173. Kelley, D I; Harrison, S P (1 October 2014). "Enhanced Australian carbon sink despite increased wildfire during the 21st century". Environmental Research Letters. 9 (10): 104015. Bibcode:2014ERL.....9j4015K. doi:10.1088/1748-9326/9/10/104015. ISSN   1748-9326.
  174. Thompson, M. (2018). "South African National Land-Cover 2018 Report & Accuracy Assessment". Department of Environment, Forestry and Fisheries South Africa.
  175. Coetsee, Corli; Gray, Emma F.; Wakeling, Julia; Wigley, Benjamin J.; Bond, William J. (5 December 2012). "Low gains in ecosystem carbon with woody plant encroachment in a South African savanna". Journal of Tropical Ecology. 29 (1): 49–60. doi:10.1017/s0266467412000697. ISSN   0266-4674. S2CID   85575373.
  176. Jackson, Robert B.; Banner, Jay L.; Jobbágy, Esteban G.; Pockman, William T.; Wall, Diana H. (2002). "Ecosystem carbon loss with woody plant invasion of grasslands". Nature. 418 (6898): 623–626. Bibcode:2002Natur.418..623J. doi:10.1038/nature00910. ISSN   0028-0836. PMID   12167857. S2CID   14566976.
  177. Pellegrini, Adam F.A.; Socolar, Jacob B.; Elsen, Paul R.; Giam, Xingli (2016). "Trade-offs between savanna woody plant diversity and carbon storage in the Brazilian Cerrado". Global Change Biology. 22 (10): 3373–3382. Bibcode:2016GCBio..22.3373P. doi:10.1111/gcb.13259. PMID   26919289.
  178. T., Conant, Richard (2010). Challenges and opportunities for carbon sequestration in grassland systems : a technical report on grassland management and climate change mitigation. Integrated Crop Management. FAO. ISBN   978-92-5-106494-8. OCLC   890677450.
  179. J Turpie, P Botha, K Coldrey, K Forsythe, T Knowles, GLetley, J Allen and R de Wet (2019). "Towards a Policy on Indigenous Bush Encroachment in South Africa" (PDF). Department of Environmental Affairs.CS1 maint: multiple names: authors list (link)
  180. Sankey, Temuulen; Shrestha, Rupesh; Sankey, Joel B.; Hardegree, Stuart; Strand, Eva (2013). "Lidar-derived estimate and uncertainty of carbon sink in successional phases of woody encroachment: LIDAR AND WOODY CARBON SINK". Journal of Geophysical Research: Biogeosciences. 118 (3): 1144–1155. doi:10.1002/jgrg.20088.
  181. Nuñez, Martin A; Davis, Kimberley T; Dimarco, Romina D; Peltzer, Duane A; Paritsis, Juan; Maxwell, Bruce D; Pauchard, Aníbal (3 May 2021). "Should tree invasions be used in treeless ecosystems to mitigate climate change?". Frontiers in Ecology and the Environment. 19 (6): 334–341. doi:10.1002/fee.2346. ISSN   1540-9295. S2CID   235564362.
  182. "When it comes to carbon capture, tree invasions can do more harm than good". Mongabay Environmental News. 21 June 2021. Retrieved 10 July 2021.
  183. Kumar, Dushyant; Pfeiffer, Mirjam; Gaillard, Camille; Langan, Liam; Martens, Carola; Scheiter, Simon (2020). "Misinterpretation of Asian savannas as degraded forest can mislead management and conservation policy under climate change". Biological Conservation. 241: 108–293. doi:10.1016/j.biocon.2019.108293.
  184. Veldman, Joseph W.; Overbeck, Gerhard E.; Negreiros, Daniel; Mahy, Gregory; Le Stradic, Soizig; Fernandes, G. Wilson; Durigan, Giselda; Buisson, Elise; Putz, Francis E.; Bond, William J. (1 October 2015). "Where Tree Planting and Forest Expansion are Bad for Biodiversity and Ecosystem Services". BioScience. 65 (10): 1011–1018. doi:10.1093/biosci/biv118. ISSN   1525-3244.
  185. Bond, William J.; Stevens, Nicola; Midgley, Guy F.; Lehmann, Caroline E.R. (2019). "The Trouble with Trees: Afforestation Plans for Africa". Trends in Ecology & Evolution. 34 (11): 963–965. doi:10.1016/j.tree.2019.08.003. PMID   31515117.
  186. Nackley, Lloyd L.; West, Adam G.; Skowno, Andrew L.; Bond, William J. (2017). "The Nebulous Ecology of Native Invasions". Trends in Ecology & Evolution. 32 (11): 814–824. doi:10.1016/j.tree.2017.08.003. PMID   28890126.
  187. Stevens, N.; Lehmann, C.E.R.; Murphy, B.P.; Durigan, G. (2017). "Savanna woody encroachment is widespread across three continents". Glob. Change Biol. 23 (1): 235–244. Bibcode:2017GCBio..23..235S. doi:10.1111/gcb.13409. PMID   27371937.
  188. Venter, Z. S.; Cramer, M. D.; Hawkins, H.-J. (2018). "Drivers of woody plant encroachment over Africa". Nature Communications. 9 (1): 2272. Bibcode:2018NatCo...9.2272V. doi:10.1038/s41467-018-04616-8. ISSN   2041-1723. PMC   5995890 . PMID   29891933.
  189. Saha, M. V.; Scanlon, T. M.; D'Odorico, P. (2015). "Examining the linkage between shrub encroachment and recent greening in water-limited southern Africa". Ecosphere. 6 (9): art156. doi:10.1890/ES15-00098.1. ISSN   2150-8925.
  190. Mitchard, Edward T. A.; Flintrop, Clara M. (5 September 2013). "Woody encroachment and forest degradation in sub-Saharan Africa's woodlands and savannas 1982–2006". Philosophical Transactions of the Royal Society B: Biological Sciences. 368 (1625): 20120406. doi:10.1098/rstb.2012.0406. ISSN   0962-8436. PMC   3720033 . PMID   23878342.
  191. Fuchs, R.; Herold, M.; Verburg, P. H.; Clevers, J. G. P. W. (7 March 2013). "A high-resolution and harmonized model approach for reconstructing and analysing historic land changes in Europe". Biogeosciences. 10 (3): 1543–1559. Bibcode:2013BGeo...10.1543F. doi:10.5194/bg-10-1543-2013. ISSN   1726-4189.
  192. Liu, Xu; Feng, Siwen; Liu, Hongyan; Ji, Jue (2021). "Patterns and determinants of woody encroachment in the eastern Eurasian steppe". Land Degradation & Development. 32 (13): 3536–3549. doi:10.1002/ldr.3938. ISSN   1099-145X. S2CID   233663989.
  193. Marsman, F.; Nystuen, K.O.; Opedal, Ø.H.; Foest, J.J.; Sørensen, M.V.; De Frenne, P.; Graae, B.J.; Limpens, J. (2020). "Determinants of tree seedling establishment in alpine tundra". J Veg Sci. 32. doi: 10.1111/jvs.12948 .
  194. Rosén, Ejvind; Eddy Van der Maarel (2000). "Restoration of Alvar Vegetation on Öland, Sweden". Applied Vegetation Science. 3 (1): 65–72. doi:10.2307/1478919. JSTOR   1478919.
  195. Kose, Marika (2021). Coastal meadows : maintenance, restoration and recovery. Kauer, Karin, Tali, Kadri. Eesti Maaülikool. pp. 6.999Mb. doi:10.15159/EMU.69. ISBN   9789949698837. ISSN   2382-7076.
  196. Pedro., Silva, João (2008). LIFE and Europe's grasslands : restoring a forgotten habitat. Office for Official Publications of the European Communities. ISBN   978-92-79-10159-5. OCLC   780915358.
  197. Zehnder, T.; Lüscher, A.; Ritzmann, C. (2020). "Dominant shrub species are a strong predictor of plant species diversity along subalpine pasture-shrub transects". Alp Botany. 130 (2): 141–156. doi: 10.1007/s00035-020-00241-8 .
  198. Urbina, Ifigenia; Grau, Oriol; Sardans, Jordi; Ninot, Josep M.; Peñuelas, Josep (2020). "Encroachment of shrubs into subalpine grasslands in the Pyrenees changes the plant-soil stoichiometry spectrum". Plant and Soil. 448 (1–2): 37–53. doi:10.1007/s11104-019-04420-3. ISSN   0032-079X. S2CID   209897210.
  199. Tasser, Erich; Walde, Janette; Tappeiner, Ulrike; Teutsch, Alexandra; Noggler, Werner (January 2007). "Land-use changes and natural reforestation in the Eastern Central Alps". Agriculture, Ecosystems & Environment. 118 (1–4): 115–129. doi:10.1016/j.agee.2006.05.004.
  200. Teleki, Balázs; Sonkoly, Judit; Erdős, László; Tóthmérész, Béla; Prommer, Mátyás; Török, Péter (April 2020). Hölzel, Norbert (ed.). "High resistance of plant biodiversity to moderate native woody encroachment in loess steppe grassland fragments". Applied Vegetation Science. 23 (2): 175–184. doi:10.1111/avsc.12474. ISSN   1402-2001.
  201. Chauchard, Sandrine; Beilhe, Fabien; Denis, Nicole; Carcaillet, Christopher (March 2010). "An increase in the upper tree-limit of silver fir (Abies alba Mill.) in the Alps since the mid-20th century: A land-use change phenomenon". Forest Ecology and Management. 259 (8): 1406–1415. doi:10.1016/j.foreco.2010.01.009.
  202. Henkin, Zalmen (9 August 2021). "The role of brush encroachment in Mediterranean ecosystems: a review". Israel Journal of Plant Sciences: 1–12. doi:10.1163/22238980-bja10039. ISSN   0792-9978.
  203. Ortiz, Carlos; Fernández-Alonso, María José; Kitzler, Barbara; Díaz-Pinés, Eugenio; Saiz, Gustavo; Rubio, Agustín; Benito, Marta (January 2022). "Variations in soil aggregation, microbial community structure and soil organic matter cycling associated to long-term afforestation and woody encroachment in a Mediterranean alpine ecotone". Geoderma. 405: 115450. doi:10.1016/j.geoderma.2021.115450.
  204. Hongwei Zeng; Bingfang Wu; Miao Zhang; Ning Zhang; Abdelrazek Elnashar; Liang Zhu; Weiwei Zhu; Fangming Wu; Nana Yan; Wenjun Liu (2021). "Dryland ecosystem dynamic change and its drivers in Mediterranean region". Current Opinion in Environmental Sustainability. 48: 59–67. doi:10.1016/j.cosust.2020.10.013. S2CID   229411318.
  205. Rolo, Victor; Moreno, Gerardo (April 2019). "Shrub encroachment and climate change increase the exposure to drought of Mediterranean wood-pastures". Science of the Total Environment. 660: 550–558. Bibcode:2019ScTEn.660..550R. doi:10.1016/j.scitotenv.2019.01.029. PMID   30641382.
  206. Nadal-Romero, Estela; Rubio, Pablo; Kremyda, Vasiliki; Absalah, Samira; Cammeraat, Erik; Jansen, Boris; Lasanta, Teodoro (October 2021). "Effects of agricultural land abandonment on soil organic carbon stocks and composition of soil organic matter in the Central Spanish Pyrenees". CATENA. 205: 105441. doi:10.1016/j.catena.2021.105441.
  207. Lasanta, Teodoro; Nadal-Romero, Estela; Errea, Paz; Arnáez, José (February 2016). "The Effect of Landscape Conservation Measures in Changing Landscape Patterns: A Case Study in Mediterranean Mountains: The Effect of Landscape Conservation in Chaning Landscape Patterns". Land Degradation & Development. 27 (2): 373–386. doi:10.1002/ldr.2359.
  208. Gelabert, P.J.; Rodrigues, M.; de la Riva, J.; Ameztegui, A.; Sebastià, M.T.; Vega-Garcia, C. (September 2021). "LandTrendr smoothed spectral profiles enhance woody encroachment monitoring". Remote Sensing of Environment. 262: 112521. Bibcode:2021RSEnv.262k2521G. doi:10.1016/j.rse.2021.112521.
  209. McBride, Joe; Heady, Harold F. (1968). "Invasion of Grassland by Baccharis pilularis DC". Journal of Range Management. 21 (2): 106. doi:10.2307/3896366. ISSN   0022-409X. JSTOR   3896366.
  210. Hamilton, W. T.; Ueckert, D. N. (2004). Brush management: past, present, future. Texas A & M University Press. pp. 3–13. ISBN   978-1-60344-628-0. OCLC   605342910.
  211. Fuhlendorf, Samuel D.; Archer, Steven A.; Smeins, Fred; Engle, David M.; Taylor, Charles A. (2008), "The Combined Influence of Grazing, Fire, and Herbaceous Productivity on Tree–Grass Interactions", Western North American Juniperus Communities, New York, NY: Springer New York, pp. 219–238, doi:10.1007/978-0-387-34003-6_12, ISBN   978-0-387-34002-9 , retrieved 12 March 2021
  212. Jones, Matthew O.; Naugle, David E.; Twidwell, Dirac; Uden, Daniel R.; Maestas, Jeremy D.; Allred, Brady W. (2020). "Beyond Inventories: Emergence of a New Era in Rangeland Monitoring". Rangeland Ecology & Management. 73 (5): 577–583. doi:10.1016/j.rama.2020.06.009. ISSN   1550-7424. S2CID   221472237.
  213. Scharnagl, Klara; Johnson, David; Ebert-May, Diane (3 September 2019). "Shrub expansion and alpine plant community change: 40-year record from Niwot Ridge, Colorado". Plant Ecology & Diversity. 12 (5): 407–416. doi:10.1080/17550874.2019.1641757. ISSN   1755-0874. S2CID   199635665.
  214. Brodie, Jedediah F.; Roland, Carl A.; Stehn, Sarah E.; Smirnova, Ekaterina (May 2019). "Variability in the expansion of trees and shrubs in boreal Alaska". Ecology. 100 (5): e02660. doi:10.1002/ecy.2660. ISSN   0012-9658. PMID   30770560.
  215. Leite PA, Wilcox BP, McInnes KJ (2020). "Woody plant encroachment enhances soil infiltrability of a semiarid karst savanna". Environmental Research Communications. 2 (11): 115005. Bibcode:2020ERCom...2k5005L. doi: 10.1088/2515-7620/abc92f .
  216. Asner, Gregory P.; Archer, Steve; Hughes, R. Flint; Ansley, R. James; Wessman, Carol A. (March 2003). "Net changes in regional woody vegetation cover and carbon storage in Texas Drylands, 1937–1999: Net Changes in Woody Plants and Carbon Storage". Global Change Biology. 9 (3): 316–335. doi:10.1046/j.1365-2486.2003.00594.x.
  217. Kennedy, Tony (12 March 2018). "Conservation groups make point: Sometimes trees have to go". Star Tribune. Retrieved 23 February 2021.
  218. Fogarty, Dillon T.; de Vries, Caitlin; Bielski, Christine; Twidwell, Dirac (September 2021). "Rapid Re-encroachment by Juniperus virginiana After a Single Restoration Treatment". Rangeland Ecology & Management. 78: 112–116. doi:10.1016/j.rama.2021.06.002.
  219. Wang, X.; Jiang, L.; Yang, X.; Shi, Z.; Yu, P. (2020). "Does Shrub Encroachment Indicate Ecosystem Degradation? A Perspective Based on the Spatial Patterns of Woody Plants in a Temperate Savanna-Like Ecosystem of Inner Mongolia, China". Forests. 11 (12): 1248. doi: 10.3390/f11121248 .
  220. Hao, Guang; Dong, Ke; Yang, Nan; Xu, Yujuan; Ding, Xinfeng; Chen, Lei; Wang, Jinlong; Zhao, Nianxi; Gao, Yubao (December 2021). "Both fencing duration and shrub cover facilitate the restoration of shrub-encroached grasslands". CATENA. 207: 105587. doi:10.1016/j.catena.2021.105587.
  221. A. Jayadevan; S. Mukherjee; A. T. Vanak (2018). "Bush encroachment influences nocturnal rodent community and behaviour in a semi-arid grassland in Gujarat, India". Journal of Arid Environments. 153: 32–38. Bibcode:2018JArEn.153...32J. doi:10.1016/j.jaridenv.2017.12.009.
  222. Lunt, I.; Winsemius, L.; McDonald, S.; Morgan, J.; Dehaan, R.; Bowman, D. (2010). "How widespread is woody plant encroachment in temperate Australia? Changes in woody vegetation cover in lowland woodland and coastal ecosystems in Victoria from 1989 to 2005". Journal of Biogeography. 37 (4): 722–732. doi: 10.1111/j.1365-2699.2009.02255.x .
  223. Saintilan, N.; Bowen, S.; Maguire, O. (2021). "Resilience of trees and the vulnerability of grasslands to climate change in temperate Australian wetlands". Landscape Ecol. 36 (3): 803–814. doi:10.1007/s10980-020-01176-5. S2CID   231590107.
  224. Cabral, A.C.; Miguel, J.M.; Rescia, A.J.; Schmitz, M.F.; Pineda, F.D. (April 2003). "Shrub encroachment in Argentinean savannas". Journal of Vegetation Science. 14 (2): 145–152. doi:10.1111/j.1654-1103.2003.tb02139.x. ISSN   1100-9233.
  225. Rosan, T. M.; Aragão, L. E. O. C.; Oliveras, I.; Phillips, O. L.; Malhi, Y.; Gloor, E.; Wagner, F. H. (2019). "Extensive 21st‐century woody encroachment in South America's savanna". Geophysical Research Letters. 46 (12): 6594–6603. Bibcode:2019GeoRL..46.6594R. doi: 10.1029/2019GL082327 .
  226. "Can tree campaigns curb climate change without harming grasslands?". Scienceline. 28 May 2021. Retrieved 1 June 2021.
  227. Honda E. A.; Durigan G. (2016). "Woody encroachment and its consequences on hydrological processes in the savannah". Phil. Trans. R. Soc. B37120150313 (1703). doi: 10.1098/rstb.2015.0313 . PMC   4978871 . PMID   27502378.
  228. Gonçalves, Rogério Victor S.; Cardoso, João Custódio F.; Oliveira, Paulo Eugênio; Oliveira, Denis Coelho (30 March 2021). "Changes in the Cerrado vegetation structure: insights from more than three decades of ecological succession". Web Ecology. 21 (1): 55–64. doi:10.5194/we-21-55-2021. ISSN   1399-1183. S2CID   233417795.
  229. Haddad, Thaís Mazzafera; Pilon, Natashi Aparecida Lima; Durigan, Giselda; Viani, Ricardo Augusto Gorne (2021). "Restoration of the Brazilian savanna after pine silviculture: Pine clearcutting is effective but not enough". Forest Ecology and Management. 491: 119158. doi:10.1016/j.foreco.2021.119158.
  230. Overbeck, G; Muller, S; Fidelis, A; Pfadenhauer, J; Pillar, V; Blanco, C; Boldrini, I; Both, R; Forneck, E (11 December 2007). "Brazil's neglected biome: The South Brazilian Campos". Perspectives in Plant Ecology, Evolution and Systematics. 9 (2): 101–116. doi:10.1016/j.ppees.2007.07.005.
  231. Ravera, F.; D. Tarrasón; E. Simelton (2011). "Envisioning adaptive strategies to change: participatory scenarios for agropastoral semiarid systems in Nicaragua". Ecology and Society. 16 (1): 20. doi: 10.5751/ES-03764-160120 .
  232. Prins, Herbert H. T.; van der Jeugd, Henk P. (1993). "Herbivore Population Crashes and Woodland Structure in East Africa". The Journal of Ecology. 81 (2): 305. doi:10.2307/2261500. ISSN   0022-0477. JSTOR   2261500.
  233. J., Gwynne, M. D. Pratt, D. (1978). Rangeland management and ecology in East Africa. Hodder and Stoughton. ISBN   0-340-19766-8. OCLC   270888555.
  234. Fenetahun, Y.; Yong-dong, W.; You, Y. (2020). "Dynamics of forage and land cover changes in Teltele district of Borana rangelands, southern Ethiopia: using geospatial and field survey data". BMC Ecol. 20, 55 (1): 55. doi: 10.1186/s12898-020-00320-8 . PMC   7539436 . PMID   33028276.
  235. Coppock, D. L.; Gebru, G.; Mesele, S.; Desta, S. (2008). "Are Drought-Related Crashes in Pastoral Cattle Herds Predictable? More Evidence of Equilibrium Dynamics from Southern Ethiopian Rangelands". undefined. S2CID   129766182 . Retrieved 24 May 2021.
  236. Dalle, Gemedo; Maass, Brigitte L.; Isselstein, Johannes (June 2006). "Encroachment of woody plants and its impact on pastoral livestock production in the Borana lowlands, southern Oromia, Ethiopia". African Journal of Ecology. 44 (2): 237–246. doi:10.1111/j.1365-2028.2006.00638.x. ISSN   0141-6707.
  237. Haile, Mebrahtu; Birhane, Emiru; Rannestad, Meley Mekonen; Adaramola, Muyiwa S. (1 June 2021). "Expansive shrubs: Expansion factors and ecological impacts in northern Ethiopia". Journal for Nature Conservation. 61: 125996. doi:10.1016/j.jnc.2021.125996. ISSN   1617-1381.
  238. Legese, Behailu; Balew, Abel (July 2021). "Land-use and land-cover change in the lowlands of Bale Zone, Ethiopia: its driving factors and impacts of rangeland dynamics in livestock mobility". Environmental Monitoring and Assessment. 193 (7): 453. doi:10.1007/s10661-021-09222-8. ISSN   0167-6369. PMID   34181091. S2CID   235656444.
  239. Angassa, Ayana (2005). "The ecological impact of bush encroachment on the yield of grasses in Borana rangeland ecosystem". African Journal of Ecology. 43 (1): 14–20. doi:10.1111/j.1365-2028.2005.00429.x. ISSN   0141-6707.
  240. Megersa, Bekele; Markemann, André; Angassa, Ayana; Valle Zárate, Anne (February 2014). "The role of livestock diversification in ensuring household food security under a changing climate in Borana, Ethiopia". Food Security. 6 (1): 15–28. doi:10.1007/s12571-013-0314-4. ISSN   1876-4517. S2CID   2064196.
  241. Forrest, Brigham W.; Coppock, D. Layne; Bailey, DeeVon; Ward, Ruby A. (March 2016). "Economic Analysis of Land and Livestock Management Interventions to Improve Resilience of a Pastoral Community in Southern Ethiopia". Journal of African Economies. 25 (2): 233–266. doi:10.1093/jae/ejv021. ISSN   0963-8024.
  242. Schwartzstein, Peter (9 April 2019). "An invasive, thorny mesquite tree is taking over Africa—can it be stopped?". National Geographic. Retrieved 23 February 2021.
  243. Ilukor, John; Rettberg, Simone; Treydte, Anna; Birner, Regina (2016). "To eradicate or not to eradicate? Recommendations on Prosopis juliflora management in Afar, Ethiopia, from an interdisciplinary perspective". Pastoralism. 6 (1): 14. doi:10.1186/s13570-016-0061-1. ISSN   2041-7136. S2CID   56094169.
  244. Mehari, Zeraye H. (2015). "The invasion of Prosopis juliflora and Afar pastoral livelihoods in the Middle Awash area of Ethiopia". Ecological Processes. 4 (1): 13. doi:10.1186/s13717-015-0039-8. ISSN   2192-1709. S2CID   53124626.
  245. Tilahun, Minyahel; Angassa, Ayana; Abebe, Aster; Mengistu, Alemayehu (2016). "Perception and attitude of pastoralists on the use and conservation of rangeland resources in Afar Region, Ethiopia". Ecological Processes. 5 (1): 18. doi:10.1186/s13717-016-0062-4. ISSN   2192-1709. S2CID   54845709.
  246. Bekele, Ketema; Haji, Jema; Legesse, Belaineh; Shiferaw, Hailu; Schaffner, Urs (2018). "Impacts of woody invasive alien plant species on rural livelihood: Generalized propensity score evidence from Prosopis spp. invasion in Afar Region in Ethiopia". Pastoralism. 8 (1): 28. doi:10.1186/s13570-018-0124-6. ISSN   2041-7136. S2CID   53600422.
  247. Haile, Mebrahtu; Birhane, Emiru; Mekonen Rannestad, Meley; S. Adaramola, Muyiwa (29 May 2021). Hasanagas, Nikolaos D. (ed.). "Carbon Stock and Soil Characteristics under Expansive Shrubs in the Dry Afromontane Forest in Northern Ethiopia". International Journal of Forestry Research. 2021: 1–10. doi: 10.1155/2021/6647443 . ISSN   1687-9376.
  248. Becker, Mathias; Alvarez, Miguel; Heller, Gereon; Leparmarai, Paul; Maina, Damaris; Malombe, Itambo; Bollig, Michael; Vehrs, Hauke (2 January 2016). "Land-use changes and the invasion dynamics of shrubs in Baringo". Journal of Eastern African Studies. 10 (1): 111–129. doi:10.1080/17531055.2016.1138664. ISSN   1753-1055. S2CID   147128387.
  249. Petersen, Maike; Bergmann, Christoph; Roden, Paul; Nüsser, Marcus (6 April 2021). "Contextualizing land‐use and land‐cover change with local knowledge: a case study from Pokot Central, Kenya". Land Degradation & Development. 32 (10): 2992–3007. doi:10.1002/ldr.3961. ISSN   1085-3278. S2CID   233558265.
  250. Kibet, Staline; Nyangito, Moses; MacOpiyo, Laban; Kenfack, David (July 2021). "Savanna woody plants responses to mammalian herbivory and implications for management of livestock–wildlife landscape". Ecological Solutions and Evidence. 2 (3). doi:10.1002/2688-8319.12083. ISSN   2688-8319.
  251. Wahungu, Geoffrey M.; Gichohi, Nathan W.; Onyango, Irene A.; Mureu, Lucy K.; Kamaru, Douglas; Mutisya, Samuel; Mulama, Martin; Makau, Joseph K.; Kimuyu, Duncan M. (March 2013). "Encroachment of open grasslands and Acacia drepanolobium Harms ex B.Y.Sjöstedt habitats by Euclea divinorum Hiern in Ol Pejeta Conservancy, Kenya". African Journal of Ecology. 51 (1): 130–138. doi:10.1111/aje.12017.
  252. Kagunyu, Anastasia W; Wanjohi, Joseph (December 2014). "Camel rearing replacing cattle production among the Borana community in Isiolo County of Northern Kenya, as climate variability bites". Pastoralism. 4 (1): 13. doi:10.1186/s13570-014-0013-6. ISSN   2041-7136. S2CID   54071088.
  253. Vehrs, Hauke-Peter (2 January 2016). "Changes in landscape vegetation, forage plant composition and herding structure in the pastoralist livelihoods of East Pokot, Kenya". Journal of Eastern African Studies. 10 (1): 88–110. doi:10.1080/17531055.2015.1134401. ISSN   1753-1055. S2CID   147067911.
  254. Kimiti, David W.; Ganguli, Amy C.; Herrick, Jeffrey E.; Bailey, Derek W. (June 2020). "Evaluation of Restoration Success to Inform Future Restoration Efforts in Acacia reficiens Invaded Rangelands in Northern Kenya". Ecological Restoration. 38 (2): 105–113. doi:10.3368/er.38.2.105. ISSN   1543-4060. S2CID   219223577.
  255. Mbaabu, Purity Rima; Ng, Wai-Tim; Schaffner, Urs; Gichaba, Maina; Olago, Daniel; Choge, Simon; Oriaso, Silas; Eckert, Sandra (22 May 2019). "Spatial Evolution of Prosopis Invasion and its Effects on LULC and Livelihoods in Baringo, Kenya". Remote Sensing. 11 (10): 1217. Bibcode:2019RemS...11.1217M. doi: 10.3390/rs11101217 . ISSN   2072-4292.
  256. Eschen, René; Bekele, Ketema; Mbaabu, Purity Rima; Kilawe, Charles Joseph; Eckert, Sandra (28 March 2021). Zenni, Rafael (ed.). "Prosopis juliflora management and grassland restoration in Baringo County, Kenya: Opportunities for soil carbon sequestration and local livelihoods". Journal of Applied Ecology. 58 (6): 1365–2664.13854. doi:10.1111/1365-2664.13854. ISSN   0021-8901. S2CID   233683243.
  257. Kimaro, Houssein Samwel; Treydte, Anna C. (13 May 2021). "Rainfall, fire and large‐mammal‐induced drivers of Vachellia drepanolobium establishment: Implications for woody plant encroachment in Maswa, Tanzania". African Journal of Ecology. n/a (n/a): aje.12881. doi:10.1111/aje.12881. ISSN   0141-6707. S2CID   236570118.
  258. Mugasi, S. K.; Sabiiti, E. N.; Tayebwa, B. M. (1 March 2000). "The economic implications of bush encroachment on livestock farming in rangelands of Uganda". African Journal of Range & Forage Science. 17 (1–3): 64–69. doi:10.2989/10220110009485741. ISSN   1022-0119. S2CID   85259906.
  259. Egeru, Anthony; Wasonga, Oliver; Kyagulanyi, Joseph; Majaliwa, GJ; MacOpiyo, Laban; Mburu, John (2014). "Spatio-temporal dynamics of forage and land cover changes in Karamoja sub-region, Uganda". Pastoralism: Research, Policy and Practice. 4 (1): 6. doi:10.1186/2041-7136-4-6. ISSN   2041-7136. S2CID   52203524.
  260. N’Dri, Aya Brigitte; Kpré, Aka Jean-Noel; Kpangba, Koffi Prosper; Soro, Tionhonkélé Drissa; Kouassi, Koffi Vincent; Koffi, Kouamé Fulgence; Kouamé, Yao Anicet Gervais; Koffi, Ahou Blandine; Konan, Louis N’Guessan (2021), "Experimental Study of Fire Behavior in Annually Burned Humid Savanna of West Africa in the Context of Bush Encroachment", Sustainable Development in Africa, Cham: Springer International Publishing, pp. 491–505, doi:10.1007/978-3-030-74693-3_27, ISBN   978-3-030-74692-6 , retrieved 14 July 2021
  261. 1 2 Bassett, TJ & Boutrais, J 2017, Cattle and trees in the West African savanna. in Contesting Forestry in West Africa. Taylor and Francis Inc., pp. 242-263.
  262. "Bush encroachment must be curbed". Namibia Economist. Retrieved 23 October 2015.
  263. Garrard, Svenja (2017). Environmental awareness for sustainable development : a resource book for Namibia. P. Heyns, Michelle Pfaffenthaler, Gabi Schneider (1 ed.). Windhoek. ISBN   978-99945-79-88-4. OCLC   1031052763.
  264. Mapani, Benjamin; Shikangalah, Rosemary; Mapaure, Isaac; Musimba, Aansbert (2021), Leal Filho, Walter; Ogugu, Nicholas; Adelake, Lydia; Ayal, Desalegn (eds.), "Dichrostachys cinerea Growth Rings as Natural Archives for Climatic Variation in Namibia", African Handbook of Climate Change Adaptation, Cham: Springer International Publishing, pp. 1–14, doi:10.1007/978-3-030-42091-8_257-1, ISBN   978-3-030-42091-8 , retrieved 1 February 2021
  265. Government of Namibia (2017). Fifth National Development Plan (NDP5) 2017/18 – 2021/2022. National Planning Commission of Namibia.
  266. "UNCCD – Namibia | Knowledge Hub – Overview of LDN Targets". UNCCD. 2018. Retrieved 1 April 2021.
  267. Charis, Gratitude; Danha, Gwiranai; Muzenda, Edison (1 January 2019). "Waste valorisation opportunities for bush encroacher biomass in savannah ecosystems: A comparative case analysis of Botswana and Namibia". Procedia Manufacturing. 35: 974–979. doi: 10.1016/j.promfg.2019.06.044 .
  268. Birch C., Harper-Simmonds L., Lindeque P., and Middleton A. (2017). Benefits of bush control in Namibia. A national economic study for Namibia and a case for the Otjozondjupa Region. Economics of Land Degradation (ELD) Initiative.CS1 maint: multiple names: authors list (link)
  269. FSC Africa (17 March 2020). "From Bush to Charcoal: the Greenest Charcoal Comes from Namibia". FSC Africa. Retrieved 14 May 2020.
  270. Hoffmann, Jürgen. "De-bushing initiatives are coordinated". South African Institute of International Affairs. Retrieved 24 February 2015.
  271. Shigwedha, Absalom. "De-bushing advisory service set up". The Namibian. Retrieved 13 November 2016.
  272. C. Nott; J. Boys; J. Nzehengwa (2019). NRMP Best Practice Strategy Document, Reviving Namibia’s Livestock Industry (PDF). Windhoek, Namibia: Government of Namibia - Ministry of Agriculture Water and Forestry.CS1 maint: multiple names: authors list (link)
  273. "The FSC National Forest Stewardship Standard for the Republic of Namibia". FSC International. Retrieved 17 February 2020.
  274. FSC Africa (10 April 2020). "1,6 million hectares: Namibia reaches new heights in FSC certification". FSC Africa. Retrieved 14 May 2020.
  275. 1 2 Charis, Gratitude; Danha, Gwiranai; Muzenda, Edison (1 January 2019). "Waste valorisation opportunities for bush encroacher biomass in savannah ecosystems: A comparative case analysis of Botswana and Namibia". Procedia Manufacturing. 35: 974–979. doi: 10.1016/j.promfg.2019.06.044 .
  276. Vegten, J. A. (April 1984). "Thornbush invasion in a savanna ecosystem in eastern Botswana". Vegetatio. 56 (1): 3–7. doi:10.1007/bf00036129. ISSN   0042-3106. S2CID   29215146.
  277. Moleele, N.M.; Ringrose, S.; Matheson, W.; Vanderpost, C. (January 2002). "More woody plants? the status of bush encroachment in Botswana's grazing areas". Journal of Environmental Management. 64 (1): 3–11. doi:10.1006/jema.2001.0486. PMID   11876072.
  278. Dougill, Andrew J.; Akanyang, Lawrence; Perkins, Jeremy S.; Eckardt, Frank D.; Stringer, Lindsay C.; Favretto, Nicola; Atlhopheng, Julius; Mulale, Kutlwano (March 2016). "Land use, rangeland degradation and ecological changes in the southern Kalahari, Botswana". African Journal of Ecology. 54 (1): 59–67. doi: 10.1111/aje.12265 .
  279. Ringrose, Susan; Vanderpost, Cornelis; Matheson, Wilma (July 1996). "The use of integrated remotely sensed and GIS data to determine causes of vegetation cover change in southern Botswana". Applied Geography. 16 (3): 225–242. doi:10.1016/0143-6228(96)00005-7.
  280. Mmolai, Esther (23 January 2018). "Botswana: Savannah Degradation Threatens Country". AllAfrica. Retrieved 8 June 2020.
  281. Keakabetse, Boniface (5 December 2017). "North-West communities pilot climate smart projects". Mmegi Online. Retrieved 8 June 2020.
  282. Motsamai, Mmoniemang (23 March 2021). "Botswana: Kereng Outlines Projects". allAfrica.com. Retrieved 29 March 2021.
  283. Jane Turpie; Pieter Botha; Kevin Coldrey; Katherine Forsythe; Tony Knowles; Gwyneth Letley; Jessica Allen; Ruan de Wet (2019). "Towards a Policy on Indigenous Bush Encroachment in South Africa" (PDF). Department of Environmental Affairs.
  284. T.K.J. Sebitloane; H. Coetzee; K. Kellner; P. Malan (2020). "The socio-economic impacts of bush encroachment in Manthestad, Taung, South Africa". Environmental & Socio-economic Studies. 8: Issue 3 (3): 1–11. doi: 10.2478/environ-2020-0013 .
  285. 1 2 Grasslands of the world. Suttie, J. M., Reynolds, Stephen G., Batello, Caterina., Food and Agriculture Organization of the United Nations. Rome: Food and Agricultural Organization of the United Nations. 2005. ISBN   92-5-105337-5. OCLC   61697614.CS1 maint: others (link)
  286. van Wilgen, Brian W; Wannenburgh, Andrew (2016). "Co-facilitating invasive species control, water conservation and poverty relief: achievements and challenges in South Africa's Working for Water programme". Current Opinion in Environmental Sustainability. 19: 7–17. doi:10.1016/j.cosust.2015.08.012. hdl:10019.1/112990.
  287. Kellner, Klaus, Mangani, Reletile T., Sebitloane, Tshegofatso J.K., Chirima, Johannes G., Meyer, Nadine, Coetzee, Hendri C., Malan, Pieter W., & Koch, Jaco (2021). "Restoration after bush control in selected range-land areas of semi-arid savannas in South Africa". African Biodiversity & Conservation. 51 (1): 1–13. doi:10.38201/btha.abc.v51.i1.7. S2CID   232410555.CS1 maint: multiple names: authors list (link)
  288. Zhou, Yong; Tingley, Morgan W.; Case, Madelon F.; Coetsee, Corli; Kiker, Gregory A.; Scholtz, Rheinhardt; Venter, Freek J.; Staver, A. Carla (9 August 2021). "Woody encroachment happens via intensification, not extensification, of species ranges in an African savanna". Ecological Applications. n/a (n/a): e02437. doi:10.1002/eap.2437. ISSN   1051-0761. PMID   34374155.
  289. Marake M, Mokuku C, Majoro M, Mokitimi N. (1998). "Global change and subsistence rangelands in southern Africa: Resource variability, access and use in relation to rural livelihoods and welfare. A preliminary report and literature review for Lesotho". National University of Lesotho.CS1 maint: multiple names: authors list (link)
  290. Turpie, Jane; Benn, Grant; Thompson, Mark; Barker, Nigel (12 March 2021). "Accounting for land cover changes and degradation in the Katse and Mohale Dam catchments of the Lesotho highlands". African Journal of Range & Forage Science. 38 (1): 53–66. doi:10.2989/10220119.2020.1846214. ISSN   1022-0119. S2CID   232105650.
  291. Beyene, S. T. (2015). "Rangeland Degradation in a Semi‐Arid Communal Savannah of Swaziland: Long–Term DIP‐Tank Use Effects on Woody Plant Structure, Cover and their Indigenous Use in Three Soil Types". Land Degrad. Develop. 26 (4): 311–323. doi:10.1002/ldr.2203.
  292. Blaser, Wilma J. (2013). Impact of woody encroachment on soil-plant-herbivore interactions in the Kafue Flats floodplain ecosystem (Thesis). ETH Zürich. doi:10.3929/ethz-a-009933926. hdl:20.500.11850/70796.
  293. Mzezewa, J. and J. Gotosa. (2009). "Bush encroachment in Zimbabwe: a preliminary observation on soil properties". Journal of Sustainable Development in Africa. 11: 298–318. S2CID   127911174.
  294. Mohammed, Elmugheira M. I.; H., Elhag A. M.; Ndakidemi, Patrick A.; Treydte, Anna C. (4 March 2021). "Anthropogenic Pressure on Tree Species Diversity, Composition, and Growth of Balanites aegyptiaca in Dinder Biosphere Reserve, Sudan". Plants. 10 (3): 483. doi: 10.3390/plants10030483 . ISSN   2223-7747. PMC   8000727 . PMID   33806457. S2CID   232384911.

Sources

Videos

Websites

Articles