Next Article in Journal
The Effects of Localized Plant–Soil–Microbe Interactions on Soil Nitrogen Cycle in Maize Rhizosphere Soil under Long-Term Fertilizers
Next Article in Special Issue
Interactive Effects of Inorganic–Organic Compounds on Passivation of Cadmium in Weakly Alkaline Soil
Previous Article in Journal
Plant–Nanoparticle Interactions: Transcriptomic and Proteomic Insights
Previous Article in Special Issue
Reducing Deep Percolation Losses Using a Geotextile Layer at Different Soil Depths and Irrigation Levels for Lettuce Crop (Lactuca sativa L. var. capitata) (Limor)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of Smart Irrigation Management for Improving Water Productivity under Climate Change in Drylands

1
Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi 830011, China
2
Cele National Station of Observation and Research for Desert-Grassland Ecosystem, Cele 848300, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
4
Faculty of Environmental Science and Engineering, Kunming University of Science and Technology, Kunming 650500, China
*
Author to whom correspondence should be addressed.
Agronomy 2023, 13(8), 2113; https://doi.org/10.3390/agronomy13082113
Submission received: 30 June 2023 / Revised: 7 August 2023 / Accepted: 9 August 2023 / Published: 11 August 2023

Abstract

:
Global drylands, covering about 41% of Earth’s surface and inhabited by 38% of the world’s population, are facing the stark challenges of water scarcity, low water productivity, and food insecurity. This paper highlights the major constraints to agricultural productivity, traditional irrigation scheduling methods, and associated challenges, efforts, and progress to enhance water use efficiency (WUE), conserve water, and guarantee food security by overviewing different smart irrigation approaches. Widely used traditional irrigation scheduling methods (based on weather, plant, and soil moisture conditions) usually lack important information needed for precise irrigation, which leads to over- or under-irrigation of fields. On the other hand, by using several factors, including soil and climate variation, soil properties, plant responses to water deficits, and changes in weather factors, smart irrigation can drive better irrigation decisions that can help save water and increase yields. Various smart irrigation approaches, such as artificial intelligence and deep learning (artificial neural network, fuzzy logic, expert system, hybrid intelligent system, and deep learning), model predictive irrigation systems, variable rate irrigation (VRI) technology, and unmanned aerial vehicles (UAVs) could ensure high water use efficiency in water-scarce regions. These smart irrigation technologies can improve water management and accelerate the progress in achieving multiple Sustainable Development Goals (SDGs), where no one gets left behind.

1. Introduction

Drylands (hyper-arid, arid, semiarid, and dry sub-humid parts) occupy 41% of Earth’s surface, supporting 38% of global population [1,2]. Agriculture and pastoralism are the major livelihood sources for most of the population, largely dependent upon natural resources [3]. About 70% of the world’s drylands exist in developing countries where people are confronting the stark challenge of poverty, food insecurity, malnourishment, poor economic conditions, and marginalization [4,5]. Water availability and agricultural productivity are the most pressing issues associated with drylands and land degradation [6]. Globally, water scarcity is already affecting 1–2 billion people, and a majority of them are concentrated in drylands, where the supply of water is insufficient to meet the user demands [7]. Future climate projections also suggest that in coming decades more people will be facing huge shortages of water. Consequently, climate change and water management decisions will adversely affect drylands and their inhabitants [8]. As global population is increasing rapidly, agricultural productivity in drylands needs improvement to meet food security demands. Therefore, adopting smart irrigation approaches is a viable option to better utilize the available water resources and improve water productivity in drylands. Water scarcity has become one of the critical issues and threatens the sustainable development in drylands [9]. Water scarcity occurs when water demand becomes equal or even exceeds the total available fresh water resources [10]. Water scarcity should be considered from both physical and economic perspectives [11]. Physical water scarcity has two aspects: green water scarcity (soil moisture in root zone is insufficient to meet crop water demands), and blue water scarcity (both surface and ground water availability is unable to meet human water needs) [12]. The economic water scarcity occurs when water resources are physically available, but lack of institutional capacity and socioeconomic conditions limit the use of that water [13]. Water scarcity negatively impacts social integrity and sustainable economic development, especially in drylands. The primary sector, which is seriously affected, is agriculture, utilizing more than 80% of total fresh water [14]. Intensification of agricultural water scarcity could affect food production and threaten food security in drylands in the future [15]. Further, it may seriously impact the associated Sustainable Development Goals of SDG-2 (Zero hunger), SDG-6 (Clean water), SDG-7 (Clean and affordable energy), SDG 15.3 (Desertification control) and SDG-14 (Life below water), which are directly or indirectly dependent on water availability [16,17].
Since the available water resources are limited and to obtain more yields with less water use, efficient management of available water with improved water productivity is direly needed to meet future food demands [18]. Managing irrigation efficiently is challenging in drylands because there are so many factors to take into account, such as crop type, climate, soil type, and irrigation methods [19]. Drylands are characterized by high potential evapotranspiration, low and erratic rainfall and high temperature [20]. Additionally, predicted extreme weather events due to climate change will further worsen the situation. Besides hostile environmental conditions, increasing water scarcity in these regions is posing serious threats to irrigated agriculture and sustained food production [16]. Although agriculture consumes about 80% of the total water utilized in the agriculture sector globally [14], this irrigation generates a lower return per unit of water used than other economic sectors [21]. The use of traditional irrigation methods and low water use efficiency (35–40%) caused by poor management are major constraints to sustainable crop production in drylands [22,23]. Moreover, farmers in these regions still rely on traditional irrigation systems that manifest the lowest WUE. This situation puts enormous pressure on the agriculture sector to become more efficient in irrigation water use and evoked the call for a “Blue Revolution” in water-limited agricultural regions “to produce more crop per drop of water.” Efficient water-saving irrigation approaches, especially the application of smart irrigation systems, have the potential to meet this critical challenge in dryland productivity.
A smart irrigation system applies water in the right amount, at the right time and place, in a field [24]. Smart irrigation offers better irrigation decision-making by using several factors, including soil and climate variation, soil hydraulic properties, plant responses to water deficits, and changes in weather factors, that can help save water and increase yields [25]. By using smart irrigation systems, farmers can save precious resources without exposing plants to moisture deficiencies [26]. Smart irrigation has been argued as a way to manage soil variability and gain economic benefits by fulfilling the specific irrigation demands of individual crops [27]. It is also implied that the smart irrigation system will be managed in such a way that will enable nutrients and water to be delivered directly to the plant roots [28]. To the best of our knowledge, studies addressing the issue of low water productivity in dryland agriculture and its improvement through adoption of smart irrigation approaches is limited. Therefore, the specific objectives of this article were (i) to present an overview of the constraints of low water use efficiency in dryland agriculture, (ii) to assess the conventional irrigation scheduling methods, and (iii) to examine the feasibility and benefits of smart irrigation systems for better irrigation management to enhance water productivity in water-scarce regions.

2. Major Constraints of Agricultural Productivity in Drylands

2.1. Land Degradation

Natural processes such as vegetation loss, wildfires, overgrazing, climate change, wind and water erosion and other adverse/destructive anthropogenic activities cause land degradation (Figure 1) [3,29]. This causes a substantial decline in the functional capabilities of those specific areas, negatively influences agricultural activities and productivity and natural resources management, creates economic loss, and loss in biological activity [30]. Generally, land degradation is more critical or serious in dryland, semiarid and arid areas [31]. These include some parts of central Asia, China, Africa and the Mediterranean basin [32]. Land degradation in Africa highly impacts Somalia, Eritrea, and Ethiopia (horn of Africa) [33]. A prominent sign of land degradation is the occurrence of unexpected climatic conditions and vegetative stress [34]. Moreover, low levels of soil nitrogen and organic carbon (C) show a state of land degradation that leads to very low soil fertility [35]. Stavi and Lal [36] reported that land degradation increases by 5–10 million hectares every year globally. Ibrahim et al. [37] found that the increase in land degradation observed in the Africa Saharan region was mainly due to the increase in drought frequency from 1968–1990.
To gain insight into land degradation intensity, statistical methods and models have been utilized using data collected over the years. These models allow precise predictions about land degradation intensity [38]. Normalized difference vegetation index (NDVI) is one of the methods used to measure the vegetation mass over a specific area [39]. NDVI represents the vegetation state of an area in terms of numerical values. If numerical values are negative, it means there is a reduction in vegetation. A study carried out by Ibrahim et al. [37] using residual trend analysis of NDVI showed substantial evidence of soil degradation in sub-Saharan West Africa. The study also highlighted the drought occurrence caused by vegetation decline in Africa. Pravalie et al. [40] reported that the impending increase in land degradation is posing serious threats to the people of developing countries. Disease proliferation, decrease in crop yields and increasing armed conflicts are the major outcomes of land deterioration. A possible solution to land degradation is to achieve a state of land degradation neutrality (LDN) [41]. LDN is a state where land is retained in a stable condition capable of enduring biological functions including facilitating appropriate food security [42]. Sustainable land management practices such as soil amendments (addition of biochar, manuring, composting) to restore degraded lands could be used to decrease the effects of land degradation [43].

2.2. Water Scarcity Issues and Sustainable Development Goals

Water availability is one of the main indicators of land degradation, and upsurge in land degradation is potentially aggravating water scarcity [44,45]. Availability of water is jeopardized by lessening of ground and surface water due to reduction in biomass [46]. This results in less water available for agricultural (Figure 2) [16] and domestic use. Scarcity of water is an indicator of safe water; hence, water scarcity is a deficiency in fresh water resources to meet the standard water demand [11]. Successful achievement of SDGs is dependent upon the water security of both human and environmental systems (Figure 3) because water is directly linked to all SDGs [7]. The leading reasons for water scarcity are droughts, climate change, and inaccessibility and inadequate management of resources [47]. Rosa et al. [16] reported that in 2012, almost 2.3 billion people did not have access to safe water. Access to safe water is an important factor as it decreases disease frequency and social problems, including unemployment, malnutrition, and poverty [48]. Agriculture and industrial sectors use the largest proportion of water, and this water is usually drawn from below ground, lakes and rivers [49]. Falkenmark [50] reported that water scarcity poses substantial threats to the agriculture sector, which requires sufficient quantity of water for irrigation. Crop production and food security are directly dependent on sufficient water availability [51]. Water scarcity leads to plant stress, resulting in various environmental problems, such as intensified soil erosion and salt concentration [52]. Increasing climate change, land degradation, and population necessitate the development of effective management systems to wisely mitigate water scarcity [53,54,55].
Liu et al. [12] described the common indicators employed to evaluate water scarcity, such as IWMI, (a system that evaluates the economic and physical changes influencing the availability of water within a country), the criticality ratio (a ratio of water consumption to the available water resources) and Falkenmark indicator (which compares the quantity of available water against the number of people who consume that water). Water stress and crowding indices may also be employed to measure water scarcity in a country. These indices also estimate the decrease or increase in water scarcity. Efforts are required to curtail the water scarcity issues. In this regard, implementing efficient water use practices and managing water in high-risk regions are important aspects to address [45]. These efforts could decrease water scarcity and increase access to adequate agricultural and drinking water. Moreover, by taking care of structural and social customs, governments can effectively resolve major conflicts. Hence, joint efforts are needed from governments and research institutes to resolve the water scarcity problem successfully [56].

2.3. Climate Variability

Variations/changes in climatic conditions often impact human, biological and agricultural systems through decreasing water resources, rising global temperatures, heavy precipitation, elevation in permafrost thawing, worsening water and air quality, rise in sea level, health risk, food supply and availability, intense drought, disturbing rainfall periods, and devastating effects on coastal infrastructure [57,58]. East Africa is an example of climatic anomalies, and the progression of increased rainy seasons, droughts and temperatures is detrimental to the development of this area [59]. In Tanzania, crop data techniques predicted a yield reduction of 7.6%, 8.8%, and 13% in rice, sorghum and maize by 2050, respectively [3]. The impacts of variation in climatic events are significantly negative [60], for example, the Mediterranean Basin has shown a prominent rise in average temperatures beyond global changes, with significant impacts on plant processes and water resources [61]. Climatic variation may induce a significant reduction in crop growth and productivity and increase vector-borne diseases, thus threatening food security [62]. Heavy rain and early frost may have negative effects on flowering periods and induce frost damage, while dry seasons or droughts largely decrease the decomposition of soil organic matter [63]. Such alterations induce negative impacts on biological systems. Fitness of the population is one of the main factors negatively affecting community structures and population dynamics [64]. Thus, it becomes vital to develop mitigation practices to decrease climate change impacts on agriculture. Several mitigation protocols for climate change impacts suggested by Haussmann et al. [65] and Gomez-Zavaglia et al. [66] are: (i) developing skills for water management, (ii) improving methods of breeding selection, (iii) capacity building of farmers, (iv) genetically improved seeds with better adaptation to increased temperatures, (v) reducing livestock and crop emissions, (vi) changes in net irrigation, and (vii) sequestering carbon in soils.

2.4. Overexploitation of Groundwater

Increasing economic growth has resulted in continuous demand for water, leading to groundwater overexploitation, especially in big cities [67]. Both humans and the environment rely heavily upon groundwater; therefore, understanding its environmental implications is vital [68]. Over-pumping of water is a global issue, primarily caused via agricultural water use. Among its effects are the drying of wetlands and streams, phreatophytic vegetation elimination, soil subsidence, storage loss, decline in groundwater level, increased pumping cost, and soil salinization [69,70]. Recently, the exploitation of groundwater resources has become increasingly critical, especially in semiarid and arid coastal areas. In these regions, the coastal aquifers are at risk due to the intrusion of salty marine water. For example, Tripoli in northwestern Libya has been experiencing progressive seawater intrusion into its coastal aquifers since the 1930s due to its ever-growing demand for water [71]. Groundwater overexploitation not only causes water-quality degradation and aquifer depletion but also influences the ecological stability of wetlands and streams, resulting in substantial losses of biodiversity and habitat. Therefore, societies should realize that water resources are vulnerable and finite and discover methods to resolve the stresses of human development on nature’s tolerance. Educating the public about human influences on the environment is the first step toward sustainable water use.

2.5. Socioeconomic Drivers

The adverse impacts on water availability, agricultural productivity, and economic feasibility have been correlated with the displacement of residents [49]. Countries dependent on other countries’ agricultural productive capabilities are also seriously affected. Food insecurity and hunger are common outcomes of poor agricultural production phases for millions who are dependent on agriculture [72]. The 1980s famine in Africa resulted from a decline in agricultural productivity [73,74]. Generally, droughts indicate decreased food security [75]. Between 1992 and 1995, drought periods aggravated already problematic conditions in Africa. There was income loss for farmers, a rise in unemployment, and a decline in maize export to neighboring states. Moreover, there was an emergence in incurred service debts through farmers [76]. Gebremskel et al. [33] stated that drought periods in East Africa have caused more than 0.5 million deaths and financial losses up to US $1,500,000. Several rural residents heavily rely on land productivity. Thus, rising water scarcity, land degradation and drought frequency cause adverse effects on people’s lives [77]. Couttenier et al. [78] reported that social conflict could erupt as a result of decreased availability of farmland and water (e.g., civil war in Darfur). Rifts for agricultural resources and farmland often generate deadly circumstances. Ongoing conflicts among Fulani herdsmen farmers in Nigeria are another example of such conflicts. These conflicts resulted in 68% of total deaths in north-central Nigeria [79].

2.6. Droughts

Drought is a condition of abnormally dry weather sufficiently prolonged, lacking water to induce hydrologic imbalance and constraining agricultural activities in the affected region [80]. Drought types such as inter-annual droughts stay longer and decrease crop productivity, whereas inter-seasonal drought may be short and controlled via efficient water management [81]. Hermans and McLeman [82] reported that repeated occurrence of drought contributed towards degradation of land. Henchiri [83] reported that sub-Saharan Africa has faced long periods of drought with greater intensity. Nijbroek et al. [84] demonstrated that Namibia is an arid country and faced drought for many years, as well as neighboring counties of South Africa being liable to droughts repeatedly triggered by El Niño—Southern Oscillation (ENSO), a system of warm seawater that passes over the Pacific about every 10 years. Climate change could trigger desertification, induce natural calamities and environmental susceptibility. An increase in drought frequency is also expected because of climate change [85]. Droughts cause extreme distress to food security, agricultural productivity, and the economy [81]. Losses of up to US $120,000,000,000 were noted over a period of 30 years in Europe. In southern Africa, more than 10,000,000 tons of food was needed in terms of aid to drought-affected areas [3]. Africa has an unpleasant history of droughts with devastating effects on food security [86]. The most disturbing drought was during 1991–1992. It caused a huge loss in agricultural productivity, massive unemployment, and economic despair. Given the detrimental effects of drought, it is important to improve agricultural practices in response to droughts. One such practice is environmental restoration. Land restoration helps to reduce land degradation impacts [87]. Rigorous land restoration can be attained through tree plantations in degraded lands. Alternatively, in situ planting of seeds that could lead to natural regeneration may also be used [87]. Predicting droughts is also useful for proper mitigation. Recently, scientists in Kenya introduced a technology by satellite analysis, which can forecast droughts with 90% accuracy [88]. Moreover, better water management techniques and planting drought-resistant trees and crops may remove the distressing damage droughts have on the agriculture. Thus, most African nations should focus on irrigation as a method to reduce drought impacts (Figure 4) [89].

2.7. Conventional Technology

In dryland regions, most farmers employ old farming techniques that result in failure to manage food for increasing populations [90]. The traditional farming techniques generate little food [3]. Agriculture conservation (crop rotation, soil cover and minimum tillage) could help to enhance crop yields with increasing profitability and decreasing soil degradation [91]. Some techniques are not used by subsistence farmers mainly because of unfamiliarity [92]. The implication of microbial-resistant varieties or seeds is less common in Africa [93]. These advancements have the capability to improve yield and increase stress tolerance. Nonetheless, most farmers in rural areas are lacking access to services and information to be effectively used in their favor [94]. Sarkar et al. [95] reported that the use of modern techniques is helpful and provides opportunities to farmers for increasing the crop yields. Digitizing farming methods permits farmers to guess the yield and weather forecast, select suitable crops according to the area, and improve irrigation systems. Nuclear technology is also used as a tool to increase yield via radioactive isotope utilization. These are used as early detectors and tracers of the existence of diseases. Moreover, applying nuclear technology in agricultural practices can increase crop productivity by 40%, improve soil structure and texture, and decrease labor and input costs [96]. Nonetheless, the main hindrance is that modern farming techniques and technologies are not available to the majority of farmers in dryland areas.

3. Traditional Approaches Used for Irrigation Scheduling

The amount of water and its application timing is crucial in irrigation scheduling (IS), either in agriculture or landscapes [97,98]. Irrigation water requirement is measured following a criterion that determines irrigation needs and methods to apply a calculated amount of water [18]. In order to use irrigation water efficiently, we have to understand the dynamics of plant water use, together with weather, plant physiology, and soil properties. Among the various irrigation scheduling approaches developed and suggested, three types are most important: weather-based, soil moisture-based, and plant water status-based [28,99].

3.1. Weather-Based Irrigation Scheduling

In weather-based irrigation planning, reference evapotranspiration (ET0) is calculated by measuring the weather elements that reflect the amount of water lost via plants and soil [28]. Solar radiation, humidity, air temperature and wind speed influence the quantity of water lost through evapotranspiration. In the absence of soil and plant measurements, weather attributes are used to determine irrigation schedules based on evapotranspiration [100]. Reference evapotranspiration can be calculated following the FAO Penman–Monteith equation by measuring the solar radiation, wind speed, air temperature and humidity [100,101]. Daily crop water use can be calculated by:
ETc = Kc × ET0
where ETc = crop evapotranspiration (mm day−1), Kc = crop coefficient, and ET0 = reference evapotranspiration (mm day−1).
The method is strongly dependent on (1) the accurate calculation of ET0, (2) better Kc curve development over the entire crop-growing season, (3) determination of soil water-holding capacity by analyzing soil properties, and (4) quantifying site-specific rainfall [102]. Mostly, real-time weather monitoring systems are equipped with an automatic weather station containing sensors for temperature, rainfall, wind speed, humidity, atmospheric pressure, and solar radiation [103]. These data loggers are designed to obtain data automatically at periodic intervals, and these data are transferred to an online data access portal. Data loggers communicate with remote servers using a wireless sensor network (WSN) or Internet of Things (IoT) framework [18]. WSN is one of the most popular technological methods that is used to precisely monitor the weather and environmental parameters [104,105,106,107,108]. These data finally reach smart irrigation controllers, which in combination with site-specific variables (e.g., soil type), set up the irrigation schedule. The selection and performance of a weather monitoring system depends upon different accuracy, installation, robustness, data acquisition, maintenance, and power requirements. An IoT-based weather monitoring system demonstrated by Wasson et al. [109] monitors and analyzes the crop environment in terms of wind speed, temperature, solar radiation, soil moisture, and humidity using various weather-based sensors connected through a wireless network for data transfer and web-based services. Likewise, Khoa et al. [110] implemented an IoT platform for smart irrigation management. They suggested an innovative topology of sensor nodes with low cost. The authors were satisfied with the performance of the LoRa LPWAN (long-range low-power wide area network) technology transmission module system. A multiagent-based monitoring approach consisting of an open-source platform (PANGEA) was used to collect data on weather elements and soil moisture via different sensors. This platform is equipped with several master and slave nodes connected through sensors for data transfer [111]. Many researchers [112,113,114] have also employed WSN and IoT based platforms for weather-based monitoring and reported satisfactory performance of the systems. Although weather-based irrigation scheduling is widely practiced, the heterogeneity of soil properties used to estimate soil water volume affects the amount of available soil water. In addition to that, spatiotemporal, variability in large-scale evapotranspiration is another challenge confronted by this approach.

3.2. Plant-Based Irrigation Scheduling

Plant-based irrigation scheduling mainly relies on several indices indicating plant water status [115]. The relationship between soil moisture deficit and crop water stress helps to determine irrigation scheduling. Plant-based irrigation scheduling is sensitive to measurements conducted at a specific crop stage to determine water deficit in plants [18]. Since varying plant species, plant tissues, and crop growth stages have variable sensitivity to moisture deficit, several plant-based stress measurements have been suggested for irrigation scheduling [99]. There are two principal categories based on plant variable measurements used for irrigation scheduling: firstly, plant water status-based direct measurements including leaf, stem, and xylem water potential status and indirect measurements pertinent to leaf thickness, turgor pressure, and trunk diameter [116,117]; and secondly, plant physiology-based estimates including sap flow, stomatal conductance, xylem cavitation, and thermal sensing [118]. A leaf turgor pressure sensor estimates the relative change in leaf turgor pressure to determine leaf water stress [119]. In addition to transpiration water loss, root water uptake and cellular osmotic pressure determine the magnitude of turgor pressure. For example, a ZIM probe (leaf turgor pressure sensor) is a noninvasive leaf patch clamp pressure probe that can detect leaf turgor pressure. ZIM probes are capable of measuring even minute shifts of turgor pressure within leaves in real time [119,120].
Due to advanced electronic technologies, researchers have developed small leaf sensors and tested them against cowpea (Vigna unguiculata L.) and tomato (Solanum lycopersicum L.) plants. Leaf thickness-based irrigation timing improved WUE by 25–45% compared to preset irrigation plans [121]. In another study, Afzal et al. [122] reported that leaf thickness and leaf electrical capacitance (CAP) could be employed for leaf water status monitoring. Based on energy balance and heat pulse, thermal sensors have been developed to determine sap flow from plant stems, assisting irrigation scheduling. Sap-flow methods are able to provide in situ measurements of plant water use and transpiration dynamics. The Dynagage sap-flow sensors are the latest ones used to estimate sap flow and thus the water consumption by plant. The amount of heat utilized by the sap is measured by the energy balance sensors and gives the real-time sap flow in grams or kilograms per hour. These sensors require no calibration and offer an efficient and affordable method to determine the water use of plants [28]. During transpiration, the water in the xylem subjected to tension is directly proportional to the deficit in water to the point where the water columns can rupture or cavitate [123]. This cavitation leads to the eruptive formation of a bubble that contains water vapor [124]. Audio or ultrasonic frequency signals can detect these cavitation events, and the associated embolisms can hinder water flow [116]. Detection of such ultrasonic acoustic emissions (AEs) indicates plant stress and cavitation events. Thus, the AE rate can be used as a sensor to detect plant stress.

Stem Diameter Fluctuations

Stem and fruit diameter experience diurnal fluctuations due to changing water content [125]. Various water stress indicators can be determined by analyzing the daily patterns of stem diameter variation (SDV) [126]. The maximum daily shrinkage (MDS) and stem growth rate (SGR) are commonly used indicators for scheduling irrigation [115]. Currently, different optical sensors are used to detect plant water status, nutrient level and health condition. Two types of optical sensors (contact and noncontact) are primarily used [18]. Contact sensors are physically connected to plants, whereas noncontact sensors are vehicle-mounted, fixed, handled, or remotely controlled (aerial vehicles or satellite data) [28]. In many studies, monitoring through unmanned aerial vehicles (UAVs) with high-resolution cameras helped to generate irrigation maps over large cropped areas [127,128]. In another study, Lozoya et al. [129] used a sensor network to control green pepper growth under four different irrigation designs.
Bauer and Aschenbruck [130] employed IoT and sensor network integration to monitor leaf area for optimizing irrigation. Several factors other than water content can affect the SDV-derived index, such as plant age, crop load, and field management practices. Furthermore, SDV estimates are usually affected by small raindrops and animals [131]. Canopy temperature measurement is another key method commonly employed for irrigation scheduling. Major canopy temperature-based methods include the crop water stress index (CWSI), temperature–time threshold (TTT), and temperature stress day (TSD) [132,133,134]. Infrared thermometers and thermal cameras are used to detect temperature. However, diurnal dynamics of temperature remain a major challenge for all the canopy temperature-based methods [99]. A major drawback of using plant-based sensors for irrigation management is that they lack direct estimates of irrigation water amount to be applied. Soil textural variability is another challenge affecting the accurate estimation of irrigation amounts needed.

3.3. Irrigation Scheduling Based on Soil Moisture

Soil moisture monitoring is one of the fundamental approaches used for irrigation scheduling, and it is conducted by determining the soil water content or the soil water potential [135]. Monitoring soil moisture at high spatial and temporal resolution is critical for optimal irrigation scheduling [28]. Different types of sensors, such as time-domain transmission, neutron probes, granular matrix, and capacitance, are commonly implemented for soil moisture determination [136]. Gravimetric sampling to estimate soil moisture fluxes and a tensiometer is also used to measure soil matric potential, reflecting the amount of soil water available for plant use [99]. With the advancement of technology, satellite and groundwater sensors are becoming popular as irrigation tools. Soil moisture sensors can be installed at multiple depths in the field and capture soil moisture dynamics. They enhance accuracy and improve understanding of changes appearing in soil water content pertinent to crop water use and irrigation [137]. Soil sensors also provide information about soil chemical, physical and mechanical properties obtained in the form of optical, electrical, mechanical, electromagnetic, acoustic, radiometric, and pneumatic measurements [138]. Measurement of these attributes assists in the estimation of maximum allowable depletion [139]. Soil moisture sensors estimate the volumetric moisture content (VMC) by detecting changes in soil electrical and thermal properties [140].
Frequency-domain reflectometry sensors (FDR) can estimate field soil moisture content [129]. The sensors are put near the crop roots and show a moisture content range of 0–50% with 0.1% resolution, thus optimizing water use for vegetables. Shigeta et al. [141] found that real-time soil moisture sensing can be used in practical measurements of soil moisture fluxes by correlating the VWC of the soil with the capacitance of sensors inserted in the soil. In TDR sensors, two parallel rods are inserted at the desired depth to measure the soil moisture content. The rate of the electromagnetic pulse, which radiates from the sensor into the soil and returns to the soil surface, is directly proportional to soil water content. However, this is an expensive method for farmers. Other studies [142,143,144] interconnected IoT-based field monitoring with cloud-based monitoring and data analysis using an Ardunio controller. They found that the collected data were used to make predictions that helped reduce water consumption and improve crop yields. In another study, six capacitance-based sensors were used at three locations with a data logger [145]. This method improved the WUE compared to traditional approaches. Soil moisture-based irrigation scheduling has the disadvantage that plant water uptake and stress are affected by soil moisture content and also influenced by environmental conditions, pests and diseases, root zone salinity, and nutrient availability. Variation in soil properties also affects irrigation scheduling, which necessitates soil testing at multiple points for accurate estimation of soil moisture content.

4. Innovative Smart Irrigation Approaches

A smart irrigation system consists of firmware, software, and hardware interconnected via various computational techniques, including artificial intelligence (AI) and deep learning (DL) etc., which ensures the right amount of water at the appropriate time in crops to improve WUE, increase yield, reduce fertilizer use, reduce labor cost, and save energy [146]. Various control methods are employed to improve irrigation system efficiency by monitoring variables such as canopy and air temperature, evapotranspiration, rainfall, and solar radiation. By integrating information from multiple sources, smart irrigation systems can significantly improve crop production and resource management [147]. The following section presents various recent techniques associated with smart irrigation systems in agriculture.

4.1. State-of-the-Art Smart Irrigation Technologies

4.1.1. Artificial Intelligence (AI) and Deep Learning

AI is a machine’s ability to learn and implement tasks similar to those of a human brain, and it is powered by computers [148]. When applied to a certain problem domain, AI algorithms can mimic human decision-making. Irrigation systems have been integrated with AI for adaptive decision-making through fuzzy logic, expert systems, and ANNs [149].
An artificial neural network (ANN) is an algorithm for processing information that is inspired by the working of the human brain [150]. Like human brain neurons, an ANN also contains a neural network, but synapses are substituted with biased connections and weights [151]. This facilitates the mapping of input and output relationships [152]. ANN-based control systems can learn and adapt to the variable dynamics, making them ideal for irrigation systems. Additionally, ANNs have been used as smart strategies in dealing with the issue of formulating mathematical models based on first principles. Recently, many researchers have employed ANN methods for irrigation scheduling. Using the AQUACROP model integrated with a dynamic neural network, Adeyemi et al. (2018) [149] simulated soil moisture for a potato crop. Karasekreter et al. [153] demonstrated energy and water savings up to 23.9% and 20.5%, respectively, by implementing an ANN integrated with soil physical properties and moisture content in a strawberry orchard. Umair and Muhammad [154] designed an ANN-based controller model in MATLAB using climate variables as input.
A fuzzy logic system is an extension of Boolean logic that expresses logical values in the form of true or false and demonstrates the nonlinearity and uncertainty in real-world problems [155]. The fuzzy system uses different sets of input data to categorize data in membership classes, and then applies a decision rule to every set to produce human-like decision outputs [103]. Many researchers have recommended the use of fuzzy logic in irrigation control systems. Mendes et al. [156] designed a fuzzy inference system that can control the speed of the central pivot according to the spatial field variability. A fuzzy irrigation system developed by Mousa et al. [157] was used to compute evapotranspiration (ETo) via a fuzzy inference system using weather variables as input. They found that the fuzzy model was accurate and quick in obtaining the required evapotranspiration and net irrigation to recover the water loss. An expert system is another type of intelligent system used for irrigation control. Basically, an expert system is a computer program that simulates the verdict and behavior of an individual or organization with expertise in a certain area through the use of artificial intelligence (AI) technologies [158].
Expert systems can be used for problem-solving activities such as monitoring, control, planning, forecasting, prescribing, fusion, and decision-making [159]. An expert-controlled irrigation system enables farmers to quantify the water amount needed by crops at the appropriate time by considering the weather and soil conditions. Many researchers [160,161] have implemented expert systems for irrigation management. The expert system uses various knowledge-based inputs for accurate decision-making about irrigation scheduling. However, errors in knowledge-based input can seriously affect the performance and reliability of expert systems [162].
A hybrid intelligence system is another type of intelligent control system in which at least two artificial intelligence algorithms such as fuzzy logic and neural network are combined, known as “neuro-fuzzy” [149]. Other examples of such hybrid intelligence systems include fuzzy PID and GAPSO. Tsang et al. [152] employed seven different machine-learning algorithms to assess soil moisture conditions using aerial images of agricultural fields to control irrigation. The results demonstrated a 52% reduction in water consumption by reducing timing, irrigation level, and location errors. Similarly, a combination of ANN, genetic algorithm (GA) and the Bayesian framework was implemented to forecast daily irrigation demand under limited data conditions [163]. The results exhibited an improvement in forecast precision by 3% and 11%. Many other studies also demonstrated the use of intelligent hybrid systems and reported precise forecasting and improved control of irrigation systems [164,165].
The deep learning method is now applied to deal with millions of weights among neurons for a better understanding of behaviors owing to recent developments in computing technology in parallel processing, software and hardware. Deep learning has developed a revolutionary epoch, since it can solve the problems confronted by artificial intelligence for a long period [166]. Deep learning has been applied in the agriculture and hydrology fields due to difficulty in software data availability, budget, and complexity, such as crop evapotranspiration modeling and approximation [167]. Wang and Ma [168] reported that the traditional machine learning and deep learning models work similarly as a data-driven artificial intelligence technique and could be applied to model the convoluted correlation between input and output (Table 1). Nonetheless, deep learning has an advantage over traditional machine learning techniques because of its great hierarchical structure model [169].

4.1.2. Model Predictive Irrigation Systems

Development in smart agriculture through internet usage and increasing computational power facilitated large data collection from agricultural systems [180]. The model predictive system has been employed in irrigation scheduling, irrigation canal control, soil moisture, and stem water potential regulation [55]. Model predictive control (MPC) has manifested applicability to gate operation and control the canal flow. The management goal of model predictive control for canals is to maintain the level of water as close to the set-points as possible [181]. Thus, an appropriate model regulating the dynamics of canal-water levels is required. A model predictive control system has been employed to model water movement in the canals, keeping a specific level of water at different locations and the flow of water that affects these water levels [182,183,184]. The controlling instruments maintain the flow of water, by which the regulator can attain the management goals [180]. Nonetheless, attaining this goal is not straightforward, as variations in inflows and outflows interrupt the whole water system. To estimate future water flows and levels in response to control actions and disturbances, the water system (controller, canal reaches, disturbances and structures) needs to be modeled. Several authors have applied MPC in driving irrigation flows of canals. For instance, Puig et al. [185] applied MPC to create flow control approaches from the source of water to the user and Guadiana River’s irrigation territory. The results exhibited the usefulness of the MPC application. Zhang et al. [186] developed a non-cooperative distributed MPC algorithm based on Nash optimality for the regulation of water levels in canals. The simulation of system results indicated the efficacy of the advocated algorithm. To efficiently deliver the flow of a canal without oscillations, MPC was combined with online water storage to allow for a delay and evade wave distraction. The results indicated the significant development of canal setups using automation [180].

4.1.3. Variable-Rate Irrigation (VRI)

VRI is a method of applying irrigation at variable rates in different irrigation management zones over the entire field in an optimized way [187]. Normally, the application of irrigation water is uniform in the entire field. However, owing to soil spatial variability in soil topography, hydraulic properties and vegetation condition, the soil moisture content remains nonuniform [188]. When such soil spatial variability becomes significant, the field is split into different management zones consisting of those field areas with the same soil properties and crop conditions [150]. Then, irrigation is applied at differential rates in different management zones [189]. Such variable irrigation management may enhance the economic value of irrigation by improving WUE, increasing productivity and reduction in nutrient leaching [190]. This enables an accurate and timely water application based on soil spatiotemporal properties and plant demand [191].
In other words, VRI technology ensures the application of the right amount of water at the right time in the right field zone, resulting in significant water savings. The main components of VRI technology include sensors, prescription maps, spatial information, and a unit system to apply VRI prescription (lateral irrigation) in the crop field [192]. Optimization of VRI prescriptions is usually determined by using remote sensing, yield maps, topography, soil apparent electrical conductivity and soil maps [193,194]. There are different types of irrigation systems used for VRI applications. The variable-rate lateral irrigation system contains a global navigation satellite system (GNSS) or global positioning system (GPS) receiver, custom software-operated relays, and valves, thus supplying water at variable rates using the nozzle-pulsing method with a speed controller [26]. This system has high accuracy in controlling the irrigation rate and forward speed [195]. Likewise, the center pivot VRI system consists of a VRI and pivot control panels, control nodes, solenoid valves, a GNSS, a remote sensing control system, and a variable-frequency drive (VFD) [191].
The speed and operation of the pivot are regulated by pivot control. The VRI controller panel governs the irrigation application based on pivot location and the prescription map. The flow of sprinkler heads is controlled by solenoid valves [26]. The pivot positioning is regulated by the GNSS system, and the control nodes attached to the pivot govern the valve opening and closing. VFD regulates the pressure by altering the irrigation rate at different points in the field [196]. The rotation speed of the pump impeller is also controlled by the VFD in response to the input communicated by the pressure switch mounted on the pump. It helps to maintain the pressure within the predefined threshold limits [197]. The use of VRI technology offers several advantages over conventional irrigation methods. VRI can substantially improve overall yields by avoiding under-irrigation and/or over-irrigation. The growers usually set up soil moisture sensors in those field areas with low soil water-holding capacity (WHC) to prevent under-irrigation in the field [198]. This practice may increase irrigation frequency, leading to over- irrigation with high soil WHC. Over-irrigation may result in yield loss due to nutrient leaching and depletion of oxygen in the root zone [39]. The prevention of under-irrigation in those field areas with higher yield potential could help optimize water input.
Another advantage of using VRI is that irrigation can be withheld over those field areas that are not arable [199]. VRI also supports fertilizer application at variable rates that would benefit in matching the variability in crop nutrient requirements [200]. A two-year study led by Sui and Yan [192] demonstrated that crop water productivity for corn and soybean was much better under VRI than uniform rate irrigation (URI) in Mississippi. In another investigation, application of VRI with delineated management zones based on the difference in WHC showed better crop water productivity for maize and winter wheat, which was higher than the overall average of the field [201]. Besides its several advantages, VRI also has some disadvantages, including higher cost, complexity in developing soil maps, and maintenance of the system [202]. Overall, VRI technology is a good option to precisely utilize precious water resources, but considerable efforts are needed to make this technology affordable and more user-friendly.

4.1.4. Unmanned Aerial Vehicles (UAVs) for Irrigation Management

UAVs, also called drones [203], are frequently linked with military operations, as they are used as weapons for targeting aircraft and involved in intelligence services. Recently, drones have been used in a wide range of applications, including delivery services, weather monitoring, traffic monitoring, surveillance, and rescue [204]. Several studies emphasized UAV utilization for forecasting and monitoring in agriculture to maintain crop health [205]. Drones are also useful for irrigation monitoring, as they use infrared or thermal imaging cameras in the IOT network [206]. Manual spraying of pesticides induces lethal diseases to workers globally, as described by the World Health Organization and Food and Agriculture Organization [207]. Thus, UAVs could be a potential alternative to manual pesticide spraying, reducing the potential ecological/environmental risks and health problems [205].
Recently, UAVs with IoT-based sensor networks have been used for smart irrigation purposes, thus significantly improving crop productivity [208]. Chebrolu et al. [209] suggested a technique of UAV images to rebuild a three-dimensional crop model that mediates crop growth monitoring based on a plant level. Likewise, the plant height of sorghum and maize plants was measured based on UAV images and a three-dimensional model [210]. Roth et al. [211] reported that the RMSE (root-mean-square error) was 0.33 m for a single sorghum’s height. The soybean leaf area index was extracted using 3D and UAV plant models. In another study, carried out by Deng et al. [212], different cameras were fixed on UAVs for smart farming. The results indicated that the UAV-based multiband images are useful and showed substantial ability for precise irrigation and agriculture management. RGB (red, green and blue) cameras can be used with a drone to determine crop biomass using visible reflectance for assessing vegetation indices [213]. According to Rokhmana [214], using UAVs for remote sensing can support precision farming. They can be used to obtain periodic information from the field, i.e., stock evaluation, plant health and vegetation monitoring.
Several researchers examined the chances of applying IoT systems to govern crop health and irrigation monitoring. Automated water irrigation was innovated and employed by mobile applications [215]. The designed smartphone application can process and develop the soil images near the root surface of plants to ascertain sensor-less water quality. A smart drip irrigation method was established using an ARM9 processor, involving environmental conditions including CO2 amount, low moisture, and high temperature [216]. Zaier et al. [217] suggested an irrigation system controlled wirelessly to promote groundwater usage in large-scale fields of Oman. Oksanen et al. [218] stated that big data and IoT could be used to establish a real-time crop growth monitoring system, facilitating precise irrigation estimates. Based on the collection of plant data, they suggested establishing a central unit to develop a crop growth model, Oksanen et al. [218] proposed a method to forecast and diagnose wheat diseases, weeds, and pests using a computerized IoT-based method. A fungicide and insecticide management system based on predictive models for pests and crop diseases has also been developed using the IoT [219]. A new irrigation method based on the IoT was proposed by [220] where humidity sensors were applied, and high humidity values of 90–95% were recorded (Table 2).

4.2. Forecasting Smart Irrigation Technology with DSSIS

A decision support system (DSS) is an interactive software-based system used to identify, analyze, and improve decisions based on raw data, documents, and personal knowledge [99]. Various decision support systems (DSSs) have been designed for managing irrigation water to improve WUE [230,231]. A smart and efficient DSS has to consider several factors, such as soil water status, crop type, irrigation method, weather information, and application, to develop irrigation scheduling [19]. To facilitate precise irrigation scheduling by minimizing errors in field soil moisture estimates, DSSs provide irrigation schedules not only for the current day but also to forecast irrigation events for future days.
Based on the idea of forecasting irrigation, recently a prototype of an irrigation scheduling DSS called decision support system for irrigation scheduling (DSSIS) has been developed for arid regions [232]. This DSSIS has the ability to predict irrigation events for the current day as well as forecast irrigation for the future by using the weather information of the next 4 days. The DSSIS prototype consists of irrigation pipelines, software and hardware to control irrigation and peripheral equipment The irrigation pipelines consist of a drip irrigation system, valves and polyvinyl chloride pipes. The software controlling the irrigation system includes RZWQM2 (Root Zone Water Quality Model) integrated with an irrigation scheduling software (RZ Irrsch and an online weather data acquisition system [233]. The irrigation-controlling hardware contains automatic control equipment. A peripheral equipment consists of a water reservoir, circulating pumps, and strainers. In DSSIS, the RZWQM2 model works as an engine and facilitates decision-making about irrigation scheduling. The RZWQM2 is first calibrated and validated according to site-specific experimental data (crop, weather, and soil data). The IrrSch software generates daily weather data from the nearby weather stations and also forecasts upcoming 4-day weather using a weather application program interface (API) then transfer this information to RZWQM2 [99].
Based on the information given by IrrSch, RZWQM2 predicts crop evapotranspiration, soil water stress factor (SWFAC), and soil moisture content for the current and next 4 days. When the current day’s water stress level falls below the preset threshold, an irrigation event is initiated. The amount of water to be supplied is computed by RZWQM2 using field capacity and the predicted soil water content and rooting depth. This system (DSSIS) has been tested for cotton irrigation scheduling under full, deficit, experience, and sensor-based irrigation treatments in an arid region. Under deficit irrigation, DSSIS saved 50% of irrigation water with a 4% increase in yield and up to 80% increase in water productivity over experience-based irrigation [232]. In another study, Chen et al. [234] evaluated the effect of irrigation scheduling by DSSIS on water productivity, seed cotton yield, and economic profitability under an arid desert climate. Under DSSIS, water productivity, seed cotton yield, and economic benefit were higher than soil moisture sensor-based irrigation scheduling. Full irrigation DSS also maintained crop yield under deficit irrigation treatment. In a three-year field study, Chen et al. [233] evaluated irrigation water use efficiency (IWUE) in cotton using the RZWQM2 model with DSSIS in an arid oasis. After validation, the RZWQM2 model was run with seven irrigation scenarios (from 850 to 350 mm water), and the long-term weather data (1990–2019) were used to estimate the best IWUE. The results manifested that the irrigation with 660 mm water produced the highest seed cotton yield (4.09 Mg ha−1), whereas irrigation with 550 mm water exhibited the highest IWUE (6.53 kg ha−1mm−1). These findings provided important guidelines for farmers to use deficit irrigation strategies. This will also help the farmers to develop and improve irrigation scheduling strategies with respect to their specific crop production settings.
This irrigation forecasting DSSIS has been tested on a small scale in an arid oasis and provided satisfactory results regarding water productivity. It could be promoted over a large scale. To further improve the site-specific irrigation management, the soil textural variability could be analyzed and a local-level soil database could be developed through extensive soil sampling and analysis. This soil database could be integrated with DSSIS for scheduling irrigation for a specific field. This concept of irrigation is termed soil test-based irrigation prescription (STIP). Availability of site-specific soil information may result in potential gains in improved WUE and higher profitability in arid regions where existing irrigation strategies are poorly connected with local agronomic and biophysical settings. Hence, development of the STIP concept could be a way forward to improve WUE and further strengthen efforts to conserve and efficiently utilize the limited water resources in arid and hyperarid regions.

5. Future Prospects

This paper presented an overview of advanced and smart irrigation practices for improving WUE in water-limited regions, but still there are some challenges that are important to consider for designing smart, sustainable and user-friendly irrigation systems.
I.
Variability in soil texture is a vital source of uncertainty because it influences the current and potential soil water storage estimates both vertically and latterly in a field. Therefore, site-specific soil analysis is one way to rectify this problem and obtain the exact soil parameter information needed for accurate irrigation scheduling. Site-specific soil test-based information integrated with smart irrigation systems can help to improve WUE in arid and semiarid regions. This method is called soil test-based irrigation prescription (STIP). The proper execution of STIP needs specific field soil sampling, analysis of soil properties and development of a soil database. This soil information with crop and weather data can be integrated with a model or decision support system to forecast an irrigation event.
II.
Most of the experiments related to smart irrigation systems were conducted on a small scale in research fields or under controlled environmental conditions, which cannot represent commercial farming practices. Therefore, more on-farm studies in large fields are needed for a clear understanding about the implementation of smart irrigation technology.
III.
Most of the commercial smart irrigation systems offered by different irrigation companies help to improve water use efficiency, but the high cost of these state-of-the-art devices is a serious challenge for farmers. Moreover, these commercial smart irrigation systems are custom-built, meaning difficulty in control and adaptability. Therefore, affordable and user-friendly equipment should be manufactured at a local level.
IV.
Most of the farmers in dryland regions are not well educated and should be trained through practical demonstration of smart irrigation systems by expert extension workers. Furthermore, governments should provide subsidies to farmers for dissemination of such technologies on a large scale.

6. Conclusions

This article provided an overview of the major constraints to agricultural productivity, traditional irrigation scheduling methods, and efforts and advancements that have been achieved to enhance WUE, conserve water, and most importantly guarantee food security through the adoption of different smart irrigation approaches in dryland regions. Dryland agriculture is largely affected by low WUE because farmers are relying upon traditional irrigation scheduling methods, resulting in over- and/or under-irrigation of fields and yield reduction. In this situation, adoption of smart irrigation approaches or technologies including artificial intelligence and deep learning (ANN, fuzzy logic, expert system, hybrid intelligent system, and deep learning), model predictive irrigation systems, VRI technology, and UAVs could ensure high water use efficiency and productivity in water-scarce regions. These technologies consider several factors, including soil and climate variation, soil structure and hydraulic properties, plant responses to water deficits, and changes in weather factors to apply the right amount of water at the right time and place. However, all these methods face some challenges regarding accurate execution and performance under field conditions, which could be rectified by incorporating indigenous knowledge and through practical demonstrations to the farmers. Smart irrigation technologies are revolutionizing global agriculture. Such technologies are highly desirable to achieve the SDGs and improve the living standards of poor farmers in drylands.

Author Contributions

Conceptualization, Z.A. and D.G.; methodology, D.G. and G.M.; software, L.Y. and S.A.; investigation, Z.A. and D.G.; resources, D.G.; data curation, S.A., G.M. and L.Y.; writing—original draft preparation, Z.A., D.G. and G.M.; writing—review and editing, S.A., L.Y., G.M.; supervision, D.G.; project administration, Z.A.; funding acquisition, D.G. and Z.A. All authors have read and agreed to the published version of the manuscript.

Funding

The funding was provided by the Natural Science Foundation of Xinjiang Uygur Autonomous Region (2022D01E099).

Acknowledgments

The authors highly acknowledge the Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences for supporting this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maestre, F.T.; Quero, J.L. Plant Species Richness and Ecosystem Multi-functionality in Global Drylands. Science 2012, 335, 214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Tariq, A.; Ullah, A.; Sardans, J.; Zeng, F.; Graciano, C.; Li, X.; Peñuelas, J. Alhagi sparsifolia: An ideal phreatophyte for combating desertification and land degradation. Sci. Total Environ. 2022, 844, 157228. [Google Scholar] [PubMed]
  3. Chimwamurombe, P.M.; Mataranyika, P.N. Factors influencing dryland agricultural productivity. J. Arid Environ. 2021, 189, 104489. [Google Scholar] [CrossRef]
  4. Stringer, L.C.; Reed, M.S.; Fleskens, L.; Thomas, R.J.; Le, Q.B.; Lala-Pritchard, T. A new dryland development paradigm grounded in empirical analysis of dryland systems science. Land Degrad. Dev. 2017, 28, 1952–1961. [Google Scholar] [CrossRef] [Green Version]
  5. Plaza, C.; Zaccone, C. Soil resources and element stocks in drylands to face global issues. Sci. Rep. 2018, 8, 13788. [Google Scholar] [CrossRef] [Green Version]
  6. Pravalie, R. Drylands extent and environmental issues. A global approach. Earth-Sci. Rev. 2016, 161, 259–278. [Google Scholar] [CrossRef]
  7. Stringer, L.C.; Mirzabaev, A. Climate change impacts on water security in global drylands. One Earth 2021, 4, 851–864. [Google Scholar]
  8. Byers, E.; Gidden, M. Global exposure and vulnerability to multi-sector development and climate change hotspots. Environ. Res. Lett. 2018, 13, 055012. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, Z.; Yuan, X.; Liu, X. The key drivers for the changes in global water scarcity: Water withdrawal versus water availability. J. Hydrol. 2021, 601, 126658. [Google Scholar] [CrossRef]
  10. Kummu, M.; Guillaume, J.H.A. The world’s road to water scarcity: Shortage and stress in the 20th century and pathways towards sustainability. Sci. Rep. 2016, 6, 38495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Rosa, L.; Chiarelli, D.D.; Rulli, M.C.; Dell’Angelo, J.; D’Odorico, P. Global agricultural economic water scarcity. Sci. Adv. 2020, 6, eaaz6031. [Google Scholar] [CrossRef]
  12. Liu, J.; Yang, H. Water scarcity assessments in the past, present, and future. Earth’s Future 2017, 5, 545–559. [Google Scholar] [CrossRef] [PubMed]
  13. Stroosnijder, L.; Moore, D.; Alharbi, A.; Argaman, E.; Biazin, B.; van den Elsen, E. Improving water use efficiency in drylands. Curr. Opin. Environ. Sustain. 2012, 4, 497–506. [Google Scholar] [CrossRef]
  14. Dalezios, N.R.; Angelakis, A.N.; Eslamian, S. Water scarcity management: Part 1: Methodological framework. Int. J. Glob. Environ. Issues 2018, 17, 1–40. [Google Scholar] [CrossRef]
  15. Pastor, A.V.; Palazzo, A.; Havlik, P.; Biemans, H.; Wada, Y.; Obersteiner, M.; Ludwig, F. The global nexus of food–trade–water sustaining environmental flows by 2050. Nat. Sustain. 2019, 2, 499–507. [Google Scholar] [CrossRef] [Green Version]
  16. Rosa, L.; Chiarelli, D.D.; Sangiorgio, M.; Beltran-Peña, A.A.; Rulli, M.C.; D’Odorico, P.; Fung, I. Potential for sustainable irrigation expansion in a 3 C warmer climate. Proc. Nat. Acad. Sci. USA 2020, 117, 29526–29534. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, X.; Liu, W. Global agricultural water scarcity assessment incorporating blue and green water availability under future climate change. Earth’s Future 2022, 10, e2021EF002567. [Google Scholar] [CrossRef]
  18. Bwambale, E.; Abagale, F.K.; Anornu, G.K. Smart irrigation monitoring and control strategies for improving water use efficiency in precision agriculture: A review. Agric. Water Manag. 2022, 260, 107324. [Google Scholar]
  19. Dabach, S.; Lazarovitch, N.; Simunek, J.; Shani, U. Numerical investigation of irrigation scheduling based on soil water status. Irrig. Sci. 2013, 31, 27–36. [Google Scholar] [CrossRef]
  20. Modarres, R.; da Silva, V.R. Rainfall Trends in Arid and Semi-Arid Regions of Iran. J. Arid Environ. 2007, 70, 344–355. [Google Scholar] [CrossRef]
  21. Zhang, J.; Guan, K.; Peng, B.; Jiang, C.; Zhou, W.; Yang, Y.; Cai, Y. Challenges and opportunities in precision irrigation decision-support systems for center pivots. Environ. Res. Lett. 2021, 16, 053003. [Google Scholar]
  22. Huang, Y.; Li, Y.P.; Chen, X.; Ma, Y.G. Optimization of the irrigation water resources for agricultural sustainability in Tarim River Basin, China. Agric. Water Manag. 2012, 107, 74–85. [Google Scholar] [CrossRef]
  23. Yu, L.; Zhao, X.; Gao, X.; Jia, R.; Yang, M.; Yang, X.; Siddique, K.H. Effect of natural factors and management practices on agricultural water use efficiency under drought: A meta-analysis of global drylands. J. Hydrol. 2021, 594, 125977. [Google Scholar] [CrossRef]
  24. Singh, U.; Praharaj, C.S. Precision irrigation management: Concepts and applications for higher use efficiency in field crops. In Scaling Water Productivity and Resource Conservation in Upland Field Crops Ensuring More Crop Per Drop; ICAR-Indian Institute of Pulses Research: Kampur, India, 2019. [Google Scholar]
  25. Bitella, G.; Rossi, R.; Bochicchio, R.; Perniola, M.; Amato, M. A novel low-cost open-hardware platform for monitoring soil water content and multiple soil-air-vegetation parameters. Sensors 2014, 14, 19639–19659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Neupane, J.; Guo, W. Agronomic basis and strategies for precision water management: A review. Agronomy 2019, 9, 87. [Google Scholar] [CrossRef] [Green Version]
  27. Cambra, C.; Sendra, S. Smart system for bicarbonate control in irrigation for hydroponic precision farming. Sensors 2018, 18, 1333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Abioye, E.A.; Abidin, M.S.Z.; Mahmud, M.S.A.; Buyamin, S.; Ishak, M.I.; Ramli, M.K.I. A review on monitoring and advanced control strategies for precision irrigation. Comput. Electron. Agric. 2020, 173, 105441. [Google Scholar]
  29. Food and Agriculture Organization of the United Nations (FAO). The State of the World’s Land and Water Resources for Food and Agriculture-Systems at Breaking Points; Main Report; FAO: Rome, Italy, 2022. [Google Scholar] [CrossRef]
  30. Zamfir, R.; Smiraglia, D.; Quaranta, G.; Salvia, R.; Salvati, L.; Giménez-Morera, A. Land degradation and mitigation policies in the Mediterranean region: A brief commentary. Sustainability 2020, 12, 8313. [Google Scholar] [CrossRef]
  31. Burrell, A.L.; Evans, J.P.; De Kauwe, M.G. Anthropogenic climate change has driven over 5 million km2 of drylands towards desertification. Nat. Comm. 2020, 1, 3853. [Google Scholar] [CrossRef] [PubMed]
  32. Dameneh Eskandari, H. Desertification of Iran in the early twenty-first century: Assessment using climate and vegetation indices. Sci. Rep. 2021, 11, 20548. [Google Scholar]
  33. Gebremeskel, G.; Tang, Q. Droughts in East Africa: Causes, impacts and resilience. Earth-Sci. Rev. 2019, 193, 146–161. [Google Scholar]
  34. Imbrenda, V.; Quaranta, G.; Salvia, R.; Egidi, G.; Salvati, L.; Prokopova, M.; Lanfredi, M. Land degradation and metropolitan expansion in a peri-urban environment. Geomat. Nat. Hazards Risk 2021, 12, 1797–1818. [Google Scholar] [CrossRef]
  35. Jin, J.; Wang, L.; Muller, K. A 10-year monitoring of soil properties dynamics and soil fertility evaluation in Chinese hickory plantation regions of southeastern China. Sci. Rep. 2021, 11, 23531. [Google Scholar] [PubMed]
  36. Stavi, I.; Lal, R. Achieving zero net land degradation: Challenges and opportunities. J. Arid Environ. 2015, 112, 44–51. [Google Scholar] [CrossRef]
  37. Ibrahim, Y.Z.; Balzter, H.; Kaduk, J.; Tucker, C.J. Land degradation assessment using residual trend analysis of GIMMS NDVI3g, soil moisture and rainfall in Sub- Saharan West Africa from 1982 to 2012. Remote Sens. 2015, 7, 5471–5494. [Google Scholar] [CrossRef] [Green Version]
  38. Minea, G.; Ciobotaru, N.; Ioana-Toroimac, G. Designing grazing susceptibility to land degradation index (GSLDI) in hilly areas. Sci. Rep. 2022, 12, 9393. [Google Scholar] [PubMed]
  39. Li, H.; Cohen, A.; Li, Z.; Zhang, M. The impacts of socioeconomic development on rural drinking water safety in China: A provincial-level comparative analysis. Sustainability 2018, 11, 85. [Google Scholar] [CrossRef] [Green Version]
  40. Prăvălie, R.; Bandoc, G.; Patriche, C.; Sternberg, T. Recent changes in global drylands: Evidences from two major aridity databases. Catena 2019, 178, 209–231. [Google Scholar] [CrossRef]
  41. Allen, C.; Metternicht, G.; Verburg, P.; Akhtar-Schuster, M.; da Cunha, M.I.; Santivañez, M.S. Delivering an enabling environment and multiple benefits for land degradation neutrality: Stakeholder perceptions and progress. Environ. Sci. Pol. 2020, 114, 109–118. [Google Scholar] [CrossRef]
  42. Grainger, A. Is land degradation neutrality feasible in dry areas? J. Arid Environ. 2015, 112, 14–24. [Google Scholar] [CrossRef]
  43. Pandit, R.; Parrotta, J.A. A framework to evaluate land degradation and restoration responses for improved planning and decision-making. Ecosyst. People 2019, 16, 1–18. [Google Scholar] [CrossRef] [Green Version]
  44. Boretti, A.; Rosa, L. Reassessing the projections of the World Water Development Report. NPJ Clean Water 2019, 2, 15. [Google Scholar] [CrossRef] [Green Version]
  45. Wijitkosum, S. Factor influencing land degradation sensitivity and desertification in a drought prone watershed in Thailand. Int. Soil Water Conserv. Res. 2020, 9, 217–228. [Google Scholar]
  46. Vanham, D.; Alfieri, L.; Feyen, L. National water shortage for low to high environmental flow protection. Sci. Rep. 2022, 12, 3037. [Google Scholar]
  47. Vallino, E.; Ridolfi, L.; Laio, F. Measuring economic water scarcity in agriculture: A cross-country empirical investigation. Environ. Sci. Res. Pollut. 2020, 114, 73–85. [Google Scholar] [CrossRef]
  48. Vilar-Compte, M.; Burrola-Mendez, S. Urban poverty and nutrition challenges associated with accessibility to a healthy diet: A global systematic literature review. Int. J. Eq. Health 2021, 20, 40. [Google Scholar] [CrossRef] [PubMed]
  49. Borsato, E.; Rosa, L. Weak and Strong Sustainability of Irrigation: A Framework for Irrigation Practices Under Limited Water Availability. Front. Sustain. Food Syst. 2020, 4, 17. [Google Scholar] [CrossRef]
  50. Falkenmark, M. Growing water scarcity in agriculture: Future challenge to global water security. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2013, 37, 20120410. [Google Scholar] [CrossRef] [Green Version]
  51. Assouline, S.; Russo, D.; Silber, A.; Or, D. Balancing water scarcity and quality for sustainable irrigated agriculture. Water Resour. Res. 2015, 51, 3419–3436. [Google Scholar] [CrossRef]
  52. Seleiman, M.F.; Al-Suhaibani, N.; Ali, N.; Akmal, M.; Alotaibi, M.; Refay, Y.; Battaglia, M.L. Drought stress impacts on plants and different approaches to alleviate its adverse effects. Plants 2021, 10, 259. [Google Scholar] [CrossRef]
  53. Ali, S.; Chen, Y.; Azmat, M.; Kayumba, P.M.; Ahmed, Z.; Mind’je, R.; Ghaffar, A.; Qin, J.; Tariq, A. Long-Term Performance Evaluation of the Latest Multi-Source Weighted-Ensemble Precipitation (MSWEP) over the Highlands of Indo-Pak (1981–2009). Remote Sens. 2022, 14, 4773. [Google Scholar] [CrossRef]
  54. Okello, C.; Tomasello, B.; Greggio, N.; Wambiji, N.; Antonellini, M. Impact of population growth and climate change on the freshwater resources of Lamu Island, Kenya. Water 2015, 7, 1264–1290. [Google Scholar] [CrossRef]
  55. Cianconi, P.; Betro, S.; Janiri, L. The impact of climate change on mental health: A systematic descriptive review. Front. Psychiat. 2020, 11, 74. [Google Scholar]
  56. Garcia, E.N.; Ulibarri, N. Plan writing as a policy tool: Instrumental, conceptual, and tactical uses of water management plans in California. J. Environ. Stud. Sci. 2022, 12, 475–489. [Google Scholar] [CrossRef]
  57. Filho, L.W.; Balasubramanian, M. Handling the health impacts of extreme climate events. Environ. Sci. Euro. 2022, 34, 45. [Google Scholar] [CrossRef]
  58. Hegerl, G.C.; Brönnimann, S.; Cowan, T.; Friedman, A.R.; Hawkins, E.; Iles, C.; Undorf, S. Causes of climate change over the historical record. Environ. Res. Lett. 2019, 14, 123006. [Google Scholar] [CrossRef]
  59. Chersich, M.F.; Wright, C.Y. Climate change adaptation in South Africa: A case study on the role of the health sector. Glob. Health 2019, 15, 22. [Google Scholar]
  60. Van der Wiel, K.; Bintanja, R. Contribution of climatic changes in mean and variability to monthly temperature and precipitation extremes. Commun. Earth Environ. 2021, 2, 1. [Google Scholar] [CrossRef]
  61. Tzanakakis, V.A.; Paranychianakis, N.V.; Angelakis, A.N. Water supply and water scarcity. Water 2020, 12, 2347. [Google Scholar] [CrossRef]
  62. Dasgupta, S.; Robinson, E.J.Z. Attributing changes in food insecurity to a changing climate. Sci. Rep. 2022, 12, 4709. [Google Scholar] [CrossRef] [PubMed]
  63. Boczon, A.; Hilszczanska, D.; Wrzosek, M. Drought in the forest breaks plant–fungi interactions. Euro. J. For. Res. 2021, 140, 1301–1321. [Google Scholar]
  64. Garcia-Cervigon, A.I.; Quintana-Ascencio, P.F.; Escudero, A. Demographic effects of interacting species: Exploring stable coexistence under increased climatic variability in a semiarid shrub community. Sci. Rep. 2021, 11, 3099. [Google Scholar] [CrossRef] [PubMed]
  65. Haussmann, B.I.; Fred Rattunde, H.; Weltzien-Rattunde, E.; Traoré, P.S.; Vom Brocke, K.; Parzies, H.K. Breeding strategies for adaptation of pearl millet and sorghum to climate variability and change in West Africa. J. Agron. Crop Sci. 2012, 198, 327–339. [Google Scholar] [CrossRef] [Green Version]
  66. Gomez-Zavaglia, A.; Mejuto, J.C.; Simal-Gandara, J. Mitigation of emerging implications of climate change on food production systems. Food Res. Int. 2020, 134, 109256. [Google Scholar]
  67. Zhang, H.; Yu, J.; Du, C.; Xia, J.; Wang, X. Assessing risks from groundwater exploitation and utilization: Case study of the Shanghai megacity, China. Water 2019, 11, 1775. [Google Scholar] [CrossRef] [Green Version]
  68. Esteller, M.V.; Diaz-Delgado, C. Environmental effects of aquifer overexploitation: A case study in the highlands of Mexico. Environ. Manag. 2002, 29, 266–278. [Google Scholar] [CrossRef] [Green Version]
  69. Cohen, D.; Teodorescu, K. On the Effect of Practice on Exploration and Exploitation of Options and Strategies. Front. Psychol. 2021, 12, 725690. [Google Scholar] [CrossRef] [PubMed]
  70. Golian, M.; Saffarzadeh, A.; Katibeh, H.; Mahdad, M.; Saadat, H.; Khazaei, M.; Dashti Barmaki, M. Consequences of groundwater overexploitation on land subsidence in Fars Province of Iran and its mitigation management programme. Water Environ. J. 2021, 35, 975–985. [Google Scholar] [CrossRef]
  71. Alfarrah, N.; Walraevens, K. Groundwater overexploitation and seawater intrusion in coastal areas of arid and semi-arid regions. Water 2018, 10, 143. [Google Scholar] [CrossRef] [Green Version]
  72. Beyer, R.M.; Hua, F.; Martin, P.A. Relocating croplands could drastically reduce the environmental impacts of global food production. Commun. Earth Environ. 2022, 3, 74–85. [Google Scholar] [CrossRef]
  73. Bjornlund, V.; Bjornlund, H.; van Rooyen, A. Why food insecurity persists in sub-Saharan Africa: A review of existing evidence. Food Secur. 2022, 14, 845–864. [Google Scholar] [PubMed]
  74. Xia, L.; Robock, A. Global food insecurity and famine from reduced crop, marine fishery and livestock production due to climate disruption from nuclear war soot injection. Nat. Food 2022, 3, 586–596. [Google Scholar] [CrossRef] [PubMed]
  75. Van Ginkel, M.; Biradar, C. Drought Early Warning in Agri-Food Systems. Climate 2021, 9, 134. [Google Scholar] [CrossRef]
  76. Ayugi, B.; Eresanya, E.O.; Onyango, A.O.; Ogou, F.K.; Okoro, E.C.; Okoye, C.O.; Ongoma, V. Review of meteorological drought in Africa: Historical trends, impacts, mitigation measures, and prospects. Pure Appl. Geo. 2022, 179, 1365–1386. [Google Scholar]
  77. Hunegnaw, Y.; Alemayehu, G. Plant Density and Time of White Lupine (Lupinus albus L.) Relay Cropping with Tef Eragrostis tef Zucc. Trotter. in Additive Design in the Highlands of Northwest Ethiopia. Int. J. Agron. 2022, 2022, 8730191. [Google Scholar] [CrossRef]
  78. Couttenier, M.; Soubeyran, R. Drought and civil war in sub-saharan Africa. Econ. J. 2014, 124, 201–244. [Google Scholar] [CrossRef] [Green Version]
  79. Li, X.; Zhao, W.; Li, J.; Li, Y. Crop yield and water use efficiency as affected by different soil-based management methods for variable-rate irrigation in a semi-humid climate. Trans. ASABE 2018, 61, 1915–1922. [Google Scholar] [CrossRef]
  80. Williams, A.P.; Cook, B.I.; Smerdon, J.E. Rapid intensification of the emerging southwestern North American mega drought in 2020–2021. Nat. Clim. Chang. 2022, 12, 232–234. [Google Scholar] [CrossRef]
  81. Nickayin, S.S.; Coluzzi, R.; Marucci, A. Desertification risk fuels spatial polarization in ‘affected’ and ‘unaffected’ landscapes in Italy. Sci. Rep. 2022, 12, 747. [Google Scholar] [CrossRef] [PubMed]
  82. Hermans, K.; McLeman, R. Climate change, drought, land degradation and migration: Exploring the linkages. Curr. Opin. Environ. Sustain. 2021, 50, 236–244. [Google Scholar] [CrossRef]
  83. Henchiri, M.; Igbawua, T.; Javed, T.; Bai, Y.; Zhang, S.; Essifi, B.; Zhang, J. Meteorological drought analysis and return periods over north and west africa and linkage with el niño–southern oscillation (Enso). Remote Sens. 2021, 13, 4730. [Google Scholar] [CrossRef]
  84. Nijbroek, R.; Piikki, K.; Söderström, M.; Kempen, B.; Turner, K.G.; Hengari, S.; Mutua, J. Soil organic carbon baselines for land degradation neutrality: Map accuracy and cost tradeoffs with respect to complexity in Otjozondjupa, Namibia. Sustainability 2018, 10, 1610. [Google Scholar] [CrossRef] [Green Version]
  85. Sloggy, M.R.; Suter, J.F.; Rad, M.R.; Manning, D.T.; Goemans, C. Changing climate, changing minds? The effects of natural disasters on public perceptions of climate change. Clim. Chang. 2021, 168, 25. [Google Scholar] [CrossRef] [PubMed]
  86. Baudoin, M.A.; Vogel, C. Living with drought in South Africa: Lessons learnt from the recent El Niño drought period. Int. J. Disast. Risk Reduct. 2017, 23, 128–137. [Google Scholar] [CrossRef]
  87. Pacheco, F.A.L.; Sanches Fernandes, L.F. Land degradation: Multiple environmental consequences and routes to neutrality. Curr. Opin. Environ. Sci. Health 2018, 5, 79–86. [Google Scholar] [CrossRef]
  88. Barrett, A.B.; Duivenvoorden, S.; Salakpi, E.E.; Muthoka, J.M.; Mwangi, J.; Oliver, S.; Rowhani, P. Forecasting vegetation condition for drought early warning systems in pastoral communities in Kenya. Rem. Sens. Environ. 2020, 248, 111886. [Google Scholar]
  89. United Nations Convention to Combat Desertification. Drought in Numbers-Restoration for Readiness and Resilience. 2022. Available online: https//www.catalogue.unccd.int/1872_Drought_in_Numbers_(English).pdf (accessed on 25 June 2023).
  90. Giller, K.E.; Delaune, T.; Silva, J.V. Small farms and development in sub-Saharan Africa: Farming for food, for income or for lack of better options? Food Secur. 2021, 13, 1431–1454. [Google Scholar] [CrossRef]
  91. Choden, T.; Ghaley, B.B. A Portfolio of Effective Water and Soil Conservation Practices for Arable Production Systems in Europe and North Africa. Sustainability 2021, 13, 2726. [Google Scholar] [CrossRef]
  92. Treich, N. Cultured Meat: Promises and Challenges. Environ. Resour. Eco. 2021, 79, 33–61. [Google Scholar] [CrossRef]
  93. Ikhimiukor, O.O.; Odih, E.E.; Donado-Godoy, P. A bottom-up view of antimicrobial resistance transmission in developing countries. Nat. Microbiol. 2022, 7, 757–765. [Google Scholar] [CrossRef] [PubMed]
  94. Altieri, M.A.; Nicholls, C.I.; Henao, A.; Lana, M.A. Agroecology and the design of climate change-resilient farming systems. Agron. Sustain. Develop. 2015, 35, 869–890. [Google Scholar] [CrossRef] [Green Version]
  95. Sarker, N.I.; Islam Shahidul Ali, A.; Islam Saiful Salam, A.; Mahmud, H.S.M. Promoting digital agriculture through big data for sustainable farm management. Int. J. Innovat. Appl. Stud. 2019, 25, 1235–1240. [Google Scholar]
  96. Kassie, M.; Shiferaw, B.; Muricho, G. Agricultural technology, crop income, and poverty alleviation in Uganda. World Dev. 2011, 39, 1784–1795. [Google Scholar] [CrossRef]
  97. McCready, M.; Dukes, M. Landscape irrigation scheduling efficiency and adequacy by various control technologies. Agric. Water Manag. 2011, 98, 697–704. [Google Scholar] [CrossRef]
  98. Vellidis, G.; Liakos, V. Irrigation scheduling for cotton using soil moisture sensors, smartphone apps, and traditional methods. In Proceedings of the Beltwide Cotton Conference, New Orleans, LA, USA, 5–7 January 2016; National Cotton Council Memphis: Cordova, TN, USA, 2016. [Google Scholar]
  99. Gu, Z.; Qi, Z.; Burghate, R.; Yuan, S.; Jiao, X.; Xu, J. Irrigation scheduling approaches and applications: A review. J. Irrig. Drain. Eng. 2020, 146, 04020007. [Google Scholar] [CrossRef]
  100. White, S.C.; Raine, S.R. A Grower Guide to Plant Based Sensing for Irrigation Scheduling; National Centre for Engineering in Agriculture: Toowoomba, QC, Australia, 2008; NCEA Publication 1001574/6. [Google Scholar]
  101. Allen, R.G.; Pereira, L.S.; Raes, D.; Smith, M. Crop Evapotranspiration-Guidelines for computing crop water requirements-FAO Irrigation and drainage paper 56. FAO Rome 1998, 300, D05109. [Google Scholar]
  102. Davis, S.; Dukes, M. Irrigation scheduling performance by evapotranspiration-based controllers. Agric. Water Manag. 2010, 98, 19–28. [Google Scholar] [CrossRef]
  103. Adeyemi, O.; Grove, I.; Peets, S.; Norton, T. Advanced monitoring and management systems for improving sustainability in precision irrigation. Sustainability 2017, 9, 353. [Google Scholar] [CrossRef] [Green Version]
  104. Patil, P.; Desai, L.B. Intelligent irrigation control system by employing wireless sensor networks. Int. J. Comput. Appl. 2013, 79, 33–40. [Google Scholar] [CrossRef]
  105. Hamouda, Y.E.M. Smart irrigation decision support based on fuzzy logic using wireless sensor network. In Proceedings of the International Conference on Promising Electronic Technologies, Deir El-Balah, Palestine, 16–17 October 2017; pp. 109–113. [Google Scholar]
  106. Viani, F.; Bertolli, M. Low-cost wireless monitoring and decision support for water saving in agriculture. IEEE Sens. J. 2017, 17, 4299–4309. [Google Scholar] [CrossRef]
  107. Keswani, B.; Mohapatra, A.G.; Mohanty, A.; Khanna, A.; Rodrigues, J.J.; Gupta, D.; De Albuquerque, V.H.C. Adapting weather conditions based IoT enabled smart irrigation technique in precision agriculture mechanisms. Neural Comput. Appl. 2019, 31, 277–292. [Google Scholar] [CrossRef]
  108. Difallah, W.; Benahmed, K.; Bounnama, F.; Draoui, B.; Saaidi, A. Intelligent irrigation management system. Int. J. Adv. Comput. Sci. Appl. 2018, 9, 429–433. [Google Scholar] [CrossRef] [Green Version]
  109. Wasson, T.; Choudhury, T. Integration of Rfid and sensor in agriculture using IoT. In Proceedings of the International Conference on Smart Technology for Smart Nation, Bengaluru, India, 17–19 August 2017; pp. 217–222. [Google Scholar]
  110. Khoa, T.A.; Man, M.M.; Nguyen, T.Y.; Nguyen, V.; Nam, N.H. Smart agriculture using IoT multi-sensors: A novel watering management system. J. Sens. Actuator Netw. 2019, 8, 45. [Google Scholar] [CrossRef] [Green Version]
  111. Villarrubia, G.; De Paz, J.F. Combining multi-agent systems and wireless sensor networks for monitoring crop irrigation. Sensors 2017, 17, 8. [Google Scholar] [CrossRef] [Green Version]
  112. Shahzadi, R.; Ferzund, J.; Tausif, M.; Suryani, M.A. Internet of things based expert system for smart agriculture. Int. J. Adv. Comput. Sci. Appl. 2016, 7. [Google Scholar] [CrossRef] [Green Version]
  113. Rahman, M.K.I.A.; Abidin, M.S.Z.; Azimi, M.S.; Mahmud, S.B.; Ishak, M.H.; Emmanuel, A.A. Advancement of a smart fibrous capillary irrigation management system with an internet of things intgration. Bull. Electr. Eng. Inform. 2019, 8, 1402–1410. [Google Scholar]
  114. Coelho, A.D.; Dias, B.G.; de Oliveira Assis, W.; de Almeida Martins, F.; Pires, R.C. Monitoring of soil moisture and atmospheric sensors with internet of things (IoT) applied in precision agriculture. In Proceedings of the 2020 XIV Technologies Applied to Electronics Teaching Conference (TAEE), Porto, Portugal, 8–10 July 2020; pp. 1–8. [Google Scholar]
  115. Fernandez, J.E. Plant-Based Methods for Irrigation Scheduling of Woody Crops. Horticulturae 2017, 3, 35. [Google Scholar] [CrossRef] [Green Version]
  116. Jones, H.G. Monitoring plant and soil water status: Established and novel methods revisited and their relevance to studies of drought tolerance. J. Exp. Bot. 2017, 58, 119–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Padilla-Diaz, C.M.; Rodriguez-Dominguez, C.M. Scheduling regulated deficit irrigation in a hedgerow olive orchard from leaf turgor pressure related measurements. Agric. Water Manag. 2016, 164, 28–37. [Google Scholar] [CrossRef] [Green Version]
  118. Fernandez, J.E. Plant-based sensing to monitor water stress: Applicability to commercial orchards. Agric. Water Manag. 2014, 142, 99–109. [Google Scholar] [CrossRef] [Green Version]
  119. Zimmermann, U.; Bitter, R.; Marchiori, P.E.R.; Rüger, S.; Ehrenberger, W.; Sukhorukov, V.L.; Ribeiro, R.V. A non-invasive plant-based probe for continuous monitoring of water stress in real time: A new tool for irrigation scheduling and deeper insight into drought and salinity stress physiology. Theo. Exp. Plant Physiol. 2013, 25, 2–11. [Google Scholar] [CrossRef]
  120. Zimmermann, U.; Ruger, S. Effects of environmental parameters and irrigation on the turgor pressure of banana plants measured using the non-invasive, online monitoring leaf patch clamp pressure probe. Plant Biol. 2010, 12, 424–436. [Google Scholar] [CrossRef] [PubMed]
  121. Seelig, H.D.; Stoner, R.J.; Linden, J.C. Irrigation control of cowpea plants using the measurement of leaf thickness under greenhouse conditions. Irrig. Sci. 2012, 30, 247–257. [Google Scholar] [CrossRef]
  122. Afzal, A.; Duiker, S.W.; Watson, J.E. Leaf thickness to predict plant water status. Biosyst. Eng. 2017, 156, 148–156. [Google Scholar] [CrossRef]
  123. Steudle, E. The cohesion-tension mechanism and the acquisition of water by plant roots. Ann. Rev. Plant Physiol. Plant Mol. Biol. 2001, 52, 847–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Jones, H.G. Irrigation scheduling: Advantages and pitfalls of plant based methods. J. Exp. Bot. 2004, 55, 2427–2436. [Google Scholar] [CrossRef] [Green Version]
  125. Guo, D.; Chen, Z.; Huang, D.; Zhang, J. Evapotranspiration model-based scheduling strategy for baby pakchoi irrigation in greenhouse. Hort. Sci. 2021, 56, 204–209. [Google Scholar] [CrossRef]
  126. Goldhamer, D.A.; Fereres, E. Irrigation scheduling protocols using continuously recorded trunk diameter measurements. Irrig. Sci. 2011, 20, 115–125. [Google Scholar] [CrossRef]
  127. Uddin, M.A.; Mansour, A. Agriculture internet of things: AG-IoT. In Proceedings of the 2017 27th International Telecommunication Networks and Applications Conference, ITNAC, Melbourne, VIC, Australia, 22–24 November 2017; pp. 1–6. [Google Scholar]
  128. Jia, X.; Huang, Y.; Wang, Y.; Sun, D. Research on water and fertilizer irrigation system of tea plantation. Int. J. Distrib. Sens. Net. 2019, 15, 1550147719840182. [Google Scholar] [CrossRef]
  129. Lozoya, C.; Mendoza, C. Sensor-based model driven control strategy for precision irrigation. J. Sens. 2016, 2016, 9784071. [Google Scholar] [CrossRef] [Green Version]
  130. Bauer, J.; Aschenbruck, N. Design and implementation of an agricultural monitoring system for smart farming. In Proceedings of the 2018 IoT Vertical and Topical Summit on Agriculture—Tuscany, IOT Tuscany, Tuscany, Italy, 8–9 May 2018; pp. 1–6. [Google Scholar]
  131. Fernandez, J.; Cuevas, M. Irrigation scheduling from stem diameter variations: A review. Agric. For. Meteorol. 2010, 150, 135–151. [Google Scholar] [CrossRef]
  132. Idso, S.B.; Jackson, R.D.; Pinter, P.J., Jr.; Reginato, R.J.; Hatfield, J.L. Normalizing the stress-degree-day parameter for environmental variability. Agric. Meteorol. 1981, 24, 45–55. [Google Scholar] [CrossRef]
  133. Wanjura, D.; Upchurch, D.; Mahan, J. Behavior of temperature-based water stress indicators in biotic-controlled irrigation. Irrig. Sci. 2006, 24, 223–232. [Google Scholar] [CrossRef]
  134. Clawson, K.L.; Blad, B.L. Infrared thermometry for scheduling irrigation of corn. Agron. J. 1982, 74, 311–316. [Google Scholar] [CrossRef]
  135. Delgoda, D.; Saleem, S.K.; Malano, H.; Halgamuge, M.N. Root zone soil moisture prediction models based on system identification: Formulation of the theory and validation using field and AQUACROP data. Agric. Water Manag. 2016, 163, 344–353. [Google Scholar] [CrossRef]
  136. Hedley, C.; Yule, I. A Method for Spatial Prediction of Daily Soil Water Status for Precise Irrigation Scheduling. Agric. Water Manag. 2009, 96, 1737–1745. [Google Scholar] [CrossRef]
  137. Soulis, K.X.; Elmaloglou, S.; Dercas, N. Investigating the effects of soil moisture sensors positioning and accuracy on soil moisture based drip irrigation scheduling systems. Agric. Water Manag. 2015, 148, 258–268. [Google Scholar] [CrossRef]
  138. Zinkernagel, J.; Maestre-Valero, J.F. New technologies and practical approaches to improve irrigation management of open field vegetable crops. Agric. Water Manag. 2020, 242, 106404. [Google Scholar] [CrossRef]
  139. Li, W.; Awais, M. Review of sensor network-based irrigation systems using IoT and remote sensing. Adv. Meteorol. 2020, 2020, 8396164. [Google Scholar] [CrossRef]
  140. Peddinti, S.R.; Hopmans, J.W.; Abou Najm, M.; Kisekka, I. Assessing effects of salinity on the performance of a low-cost wireless soil water sensor. Sensors 2020, 20, 7041. [Google Scholar] [CrossRef] [PubMed]
  141. Shigeta, R.; Kawahara, Y. Capacitive-touch-based soil monitoring device with exchangeable sensor probe. In Proceedings of the 2018 IEEE Sensors, New Delhi, India, 28–31 October 2018; pp. 1–4. [Google Scholar]
  142. Jha, R.K.; Kumar, S. Field monitoring using IoT in agriculture. In Proceedings of the 2017 International Conference on Intelligent Computing, Instrumentation and Control Technologies, Kerala, India, 6–7 July 2017; pp. 1417–1420. [Google Scholar]
  143. Salvi, S.; Sanjay, H.A. Cloud based data analysis and monitoring of smart multi-level irrigation system using IoT. In Proceedings of the International conference on I-SMAC (IoT in Social, Mobile, Analytics and Cloud) (I-SMAC), Palladam, India, 10–11 February 2017; pp. 752–757. [Google Scholar]
  144. Yashaswini, L.S.; Vani, H.U. Smart automated irrigation system with disease prediction. In Proceedings of the 2017 IEEE International Conference on Power, Control, Signals and Instrumentation Engineering (ICPCSI), Chennai, India, 21–22 September 2017; pp. 422–427. [Google Scholar]
  145. Yadav, P.K.; Sharma, F.C. Soil Moisture Sensor-Based Irrigation Scheduling to Optimize Water Use Efficiency in Vegetables. Irrig. Assoc. 2020, 1–7, [WWW Document]. Available online: http://www.irrigation.org/IA/FileUploads/IA/Resources/TechnicalPapers/2018/Soil_Moisture_Sensor_based_Irrigation_YADAV.pdf (accessed on 25 June 2023).
  146. Boman, B.; Smith, S.; Tullos, B. Control and Automation in Citrus Micro Irrigation Systems; Agricultural and Biological Engineering Department, UF/IFAS Extension; University of Florida: Gainesville, FL, USA, 2015; pp. 1–15. [Google Scholar]
  147. Al-Ghobari, H.M.; Mohammad, F.S.; El Marazky, M.S.A. Assessment of smart irrigation controllers under subsurface and drip-irrigation systems for tomato yield in arid regions. Crop Pasture Sci. 2015, 66, 1086–1095. [Google Scholar] [CrossRef]
  148. Fuentes, B.S.; Tongson, E. Advances and requirements for machine learning and artificial intelligence applications in viticulture. Wine Vitic. J. 2018, 33, 47–51. [Google Scholar]
  149. Adeyemi, O.; Grove, I.; Peets, S.; Domun, Y.; Norton, T. Dynamic neural network modelling of soil moisture content for predictive irrigation scheduling. Sensors 2018, 18, 3408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Liakos, K.G.; Busato, P.; Moshou, D.; Pearson, S.; Bochtis, D. Machine learning in agriculture: A review. Sensors 2018, 18, 2674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Campos, J.; Gallart, M.; Llop, J.; Ortega, P.; Salcedo, R.; Gil, E. On-farm evaluation of prescription map-based variable rate application of pesticides in vineyards. Agronomy 2020, 10, 102. [Google Scholar] [CrossRef] [Green Version]
  152. Tsang, S.W.; Jim, C.Y. Applying artificial intelligence modeling to optimize green roof irrigation. Energy Build. 2016, 127, 360–369. [Google Scholar] [CrossRef]
  153. Karasekreter, N.; Basciftci, F.; Fidan, U. A new suggestion for an irrigation schedule with an artificial neural network. J. Exp. Theor. Artif. Intell. 2013, 25, 93–104. [Google Scholar] [CrossRef]
  154. Umair, S.; Muhammad, R.U. Automation of irrigation system using ANN based controller. Int. J. Elec. Comp. Sci. 2010, 10, 41–47. [Google Scholar]
  155. Hasan, M.F.; Haque, M.M.; Khan, M.R.; Ruhi, R.I.; Charkabarty, A. Implementation of fuzzy logic in autonomous irrigation system for efficient use of water. In Proceedings of the 2018 Joint 7th International Conference on Informatics, Electronics & Vision (ICIEV) and 2018 2nd International Conference on Imaging, Vision & Pattern Recognition, Kitakyushu, Japan, 25–29 June 2018; pp. 234–238. [Google Scholar]
  156. Mendes, W.R.; Araujo, F.M.U.; Dutta, R.; Heeren, D.M. Fuzzy control system for variable rate irrigation using remote sensing. Expert Syst. Appl. 2019, 124, 13–24. [Google Scholar] [CrossRef]
  157. Mousa, A.K.; Crook, M.S.; Abdullah, M.N. Fuzzy based decision support model for irrigation system management. Int. J. Comput. Appl. 2014, 104, 14–20. [Google Scholar]
  158. Janjanam, D.; Ganesh, B.; Manjunatha, L. Design of an expert system architecture: An overview. J. Phys. Conf. Ser. 2021, 1767, 012036. [Google Scholar] [CrossRef]
  159. Khamkar, M.N.U. Design and Implementation of Expert System in Irrigation of Sugarcane: Conceptual Study; Sinh Gad Institute of Management and Computer Application (SIM CA): Maharashtra, India, 2014; pp. 55–58. [Google Scholar]
  160. Hazman, M. Crop irrigation schedule expert system. In Proceedings of the International conference ICT Knowledge Engineering, Bangkok, Thailand, 18–20 November 2015; pp. 78–83. [Google Scholar]
  161. Eid, S.; Abdrabbo, M. Developments of an expert system for on-farm irrigation water management under arid conditions. J. Soil Sci. Agric. Eng. 2018, 9, 69–76. [Google Scholar] [CrossRef]
  162. Ragab, S.A.; El-Gindy, A.M.; Arafa, Y.E.; Gaballah, M.S. An expert system for selecting the technical specifications of drip irrigation control unit. Arab. Univ. J. Agric. Sci. 2018, 26, 601–609. [Google Scholar] [CrossRef]
  163. Perea, R.G.; Poyato, E.C.; Montesinos, P.; Díaz, J.A.R. Optimisation of water demand forecasting by artificial intelligence with short data sets. Biosyst. Eng. 2019, 177, 59–66. [Google Scholar] [CrossRef]
  164. Cam, Z.G.; Çimen, S.; Yildirim, T. Learning parameter optimization of multi-layer perceptron using artificial bee colony, genetic algorithm and particle swarm optimization. In Proceedings of the SAMI 2015—IEEE 13th International Symposium on Applied Machine Intelligence and Informatics, Herl’any, Slovakia, 22–24 January 2015; Volume 1, pp. 329–332. [Google Scholar]
  165. Mehra, M.; Saxena, S.; Sankaranarayanan, S.; Tom, R.J.; Veeramanikandan, M. IoT based hydroponics system using Deep Neural Networks. Comput. Electron. Agric. 2018, 155, 473–486. [Google Scholar] [CrossRef]
  166. Saggi, M.K.; Jain, S.A. Survey Towards Decision Support System on Smart Irrigation Scheduling Using Machine Learning approaches. Arch. Comput. Method. Eng. 2022, 29, 4455–4478. [Google Scholar]
  167. Ferreira, L.B.; da Cunha, F.F. New approach to estimate daily reference evapotranspiration based on hourly temperature and relative humidity using machine learning and deep learning. Agric. Water Manag. 2020, 234, 106113. [Google Scholar] [CrossRef]
  168. Wang, J.; Ma, Y. Deep learning for smart manufacturing: Methods and applications. J. Manufact. Syst. 2018, 48, 144–156. [Google Scholar] [CrossRef]
  169. Kia, P.J.; Far, A.T. Intelligent control based fuzzy logic for automation of greenhouse irrigation system and evaluation in relation to conventional systems. World Appl. Sci. J. 2009, 6, 16–23. [Google Scholar]
  170. Obaideen, K.; Yousef, B.A.; AlMallahi, M.N.; Tan, Y.C.; Mahmoud, M.; Jaber, H.; Ramadan, M. An overview of smart irrigation systems using IoT. Energy Nexus 2022. [Google Scholar] [CrossRef]
  171. Al-Ali, A.R.; Qasaimeh, M.; Al-Mardini, M.; Radder, S.; Zualkernan, I.A. ZigBee-based irrigation system for home gardens. In Proceedings of the 2015 International Conference on Communications, Signal Processing, and their Applications, Sharjah, United Arab Emirates, 17–19 February 2015; pp. 1–5. [Google Scholar]
  172. Anand, K.; Jayakumar, C.; Muthu, M.; Amirneni, S. Automatic drip irrigation system using fuzzy logic and mobile technology. In Proceedings of the 2015 IEEE technological innovation in ict for agriculture and rural development, Chennai, India, 10–12 July 2015; pp. 54–58. [Google Scholar]
  173. Giusti, E.; Marsili-Libelli, S.A. Fuzzy Decision Support System for irrigation and water conservation in agriculture. Environ. Modell. Soft. 2015, 63, 73–86. [Google Scholar] [CrossRef]
  174. Dela Cruz, J.R.; Baldovino, R.G.; Bandala, A.A.; Dadios, E.P. Water usage optimization of Smart Farm Automated Irrigation System using artificial neural network. In Proceedings of the 2017 5th International Conference on Information and Communication Technology, Melaka, Malaysia, 17–19 May 2017; pp. 1–5. [Google Scholar]
  175. Arvind, G.; Athira, V.G.; Haripriya, H.; Rani, R.A.; Aravind, S. Automated irrigation with advanced seed germination and pest control. In Proceedings of the 2017 IEEE Technological Innovations in ICT for Agriculture and Rural Development, Chennai, India, 7–8 April 2017; pp. 64–67. [Google Scholar]
  176. Krishnan, R.S.; Julie, E.G.; Robinson, Y.H.; Raja, S.; Kumar, R.; Thong, P.H. Fuzzy logic based smart irrigation system using internet of things. J. Clean. Prod. 2020, 252, 119902. [Google Scholar] [CrossRef]
  177. King, B.A.; Shellie, K.C.; Tarkalson, D.D.; Levin, A.D.; Sharma, V.; Bjorneberg, D.L. Data-driven models for canopy temperature-based irrigation scheduling. Trans. ASABE 2020, 63, 1579–1592. [Google Scholar] [CrossRef]
  178. Sidhu, R.K.; Kumar, R.; Rana, P.S. Long short-term memory neural network-based multi-level model for smart irrigation. Mod. Phy. Lett. B 2020, 34, 2050418. [Google Scholar] [CrossRef]
  179. Lakshmi, G.P.; Asha, P.N.; Sandhya, G.; Sharma, S.V.; Shilpashree, S.; Subramanya, S.G. An intelligent IOT sensor coupled precision irrigation model for agriculture. Measurement. Sensors 2023, 25, 100608. [Google Scholar]
  180. Bwambal, E.; Abagale, F.K.; Anornu, G.K. Data-driven model predictive control for precision irrigation management. Smart Agric. Technol. 2023, 3, 100074. [Google Scholar] [CrossRef]
  181. Kassing, R. Model Predictive Control of Open Water Systems with Mobile Operators Delft University of Technology. Master’s Thesis, Delft University of Technology, Delft, The Netherlands, 2018. [Google Scholar]
  182. Menon, J.; Mudgal, B.; Guruprasath, M.; Sivalingam, S. Control of an irrigation branch canal using model predictive controller. Curr. Sci. 2020, 118, 1255. [Google Scholar] [CrossRef]
  183. Wang, Y.; Riaz, S. Accelerated iterative learning control for linear discrete time invariant switched systems. Mathematic. Prob. Eng. 2022, 2022, 5738826. [Google Scholar] [CrossRef]
  184. Zheng, Z.; Wang, Z.; Zhao, J.; Zheng, H. Constrained model predictive control algorithm for cascaded irrigation canals. J. Irrig. Drain. Eng. 2019, 145, 04019009. [Google Scholar] [CrossRef]
  185. Puig, V.; Ocampo-Martínez, C.; Romera, J.; Quevedo, J.; Negenborn, R.; Rodríguez, P.; de Campos, S. Model predictive control of combined irrigation and water supply systems: Application to the Guadiana river. Proceedings of 2012 9th IEEE International Conference on Networking, Sensing and Control, Beijing, China, 11–14 April 2012; pp. 85–90. [Google Scholar]
  186. Zhang, R.; Liu, A.; Yu, L.; Zhang, W. Distributed model predictive control based on Nash optimality for large scale irrigation systems. IFAC-Pap. 2015, 48, 551–555. [Google Scholar]
  187. Bhakta, I.; Phadikar, S.; Majumder, K. State-of-the-art technologies in precision agriculture: A systematic review. J. Sci. Food Agric. 2019, 99, 4878–4888. [Google Scholar] [CrossRef] [PubMed]
  188. Fontanet, M.; Scudiero, E.; Skaggs, T.H.; Fernàndez-Garcia, D.; Ferrer, F.; Rodrigo, G.; Bellvert, J. Dynamic management zones for irrigation scheduling. Agric. Water Manag. 2020, 238, 106207. [Google Scholar] [CrossRef]
  189. Thorp, K.R. Long-term simulations of site-specific irrigation management for Arizona cotton production. Irrig. Sci. 2020, 38, 49–64. [Google Scholar] [CrossRef]
  190. Liang, X.; Liakos, V.; Wendroth, O.; Vellidis, G. Scheduling irrigation using an approach based on the van Genuchten model. Agric. Water Manag. 2016, 176, 170–179. [Google Scholar] [CrossRef] [Green Version]
  191. Serrano, J.; Shahidian, S.; Marques da Silva, J.; Paixão, L.; Moral, F.; Carmona-Cabezas, R.; Noéme, J. Mapping management zones based on soil apparent electrical conductivity and remote sensing for implementation of variable rate irrigation—Case study of corn under a center pivot. Water 2020, 12, 3427. [Google Scholar] [CrossRef]
  192. Sui, R.; Yan, H. Field study of variable rate irrigation management in humid Climates. Irrig. Drain. 2017, 66, 327–339. [Google Scholar] [CrossRef]
  193. Scudiero, E.; Teatini, P. Workflow to Establish Time-Specific Zones in Precision Agriculture by Spatiotemporal Integration of Plant and Soil Sensing Data. Agronomy 2018, 8, 253. [Google Scholar] [CrossRef] [Green Version]
  194. Ohana-Levi, N.; Bahat, I.; Peeters, A.; Shtein, A.; Netzer, Y.; Cohen, Y.; Ben-Gal, A. A weighted multivariate spatial clustering model to determine irrigation management zones. Comput. Electron. Agric. 2019, 162, 719–731. [Google Scholar] [CrossRef]
  195. Han, Y.J.; Khalilian, A.; Owino, T.O.; Farahani, H.J.; Moore, S. Development of Clemson variable-rate lateral irrigation system. Comput. Electron. Agric. 2009, 68, 108–113. [Google Scholar] [CrossRef]
  196. Yari, A.; Madramootoo, C.A. Performance evaluation of constant versus variable rate irrigation. Irrig. Drain. 2017, 66, 501–509. [Google Scholar] [CrossRef]
  197. Vories, E.; Stevens, W.G. Investigating irrigation scheduling for rice using variable rate irrigation. Agric. Water Manag. 2017, 179, 314–323. [Google Scholar] [CrossRef]
  198. Daccache, A.; Knox, J.W. Implementing precision irrigation in a humid climate: Recent experiences and on-going challenges. Agric. Water Manag. 2015, 147, 135–143. [Google Scholar] [CrossRef] [Green Version]
  199. Evans, R.G.; LaRue, J. Adoption of site-specific variable rate irrigation systems. Irrig. Sci. 2013, 31, 871–887. [Google Scholar] [CrossRef] [Green Version]
  200. Lo, T.H.; Heeren, D.M. Field characterization of field capacity and root zone available water capacity for variable rate irrigation. Appl. Eng. Agric. 2017, 33, 559–572. [Google Scholar] [CrossRef] [Green Version]
  201. Zhao, W.; Li, J.; Yang, R.; Li, Y. Crop yield and water productivity responses in management zones for variable-rate irrigation based on available soil water holding capacity. Trans. ASABE 2017, 60, 1659–1667. [Google Scholar] [CrossRef]
  202. O’Shaughnessy, S.A.; Evett, S.R. Identifying advantages and disadvantages of variable rate irrigation: An updated review. Appl. Eng. Agric. 2019, 35, 837–852. [Google Scholar] [CrossRef]
  203. Shi, W.; Zhou, H.; Li, J.; Xu, W.; Zhang, N.; Shen, X. Drone assisted vehicular networks: Architecture, challenges and opportunities. IEEE Netw. 2018, 32, 130–137. [Google Scholar] [CrossRef]
  204. Yazdinejad, A.; Parizi, R.M. Enabling drones in the internet of things with decentralized blockchain-based security. IEEE Internet Things J. 2020, 8, 6406–6415. [Google Scholar] [CrossRef]
  205. Vihari, M.M.; Nelakuditi, U.R.; Teja, M.P. IoT based Unmanned Aerial Vehicle system for Agriculture applications. In Proceedings of the 2018 International Conference on Smart Systems and Inventive Technology (ICSSIT), Tirunelveli, India, 13–14 December 2018; pp. 26–28. [Google Scholar]
  206. Roopaei, M.; Rad, P.; Choo, K.K.R. Cloud of things in smart agriculture: Intelligent irrigation monitoring by thermal imaging. IEEE Cloud Comput. 2017, 4, 10–15. [Google Scholar] [CrossRef]
  207. Elijah, O.; Rahman, T.A.; Orikumhi, I.; Leow, C.Y.; Hindia, M.N. An overview of Internet of Things (IoT) and data analytics in agriculture: Benefits and challenges. IEEE Internet Things J. 2018, 5, 3758–3773. [Google Scholar]
  208. Boursianis, A.D.; Papadopoulou, M.S. Internet of Things (IoT) and Agricultural Unmanned Aerial Vehicles (UAVs) in smart farming: A comprehensive review. Internet Things 2020, 18, 100187. [Google Scholar]
  209. Chebrolu, N.; Labe, T.; Stachniss, C. Robust Long-Term Registration of UAV Images of Crop Fields for Precision Agriculture. IEEE Robot. Autom. Lett. 2018, 3, 3097–3104. [Google Scholar] [CrossRef]
  210. Chang, A.; Jung, J.; Maeda, M.M.; Landivar, J. Crop height monitoring with digital imagery from Unmanned Aerial System (UAS). Comput. Electron. Agric. 2017, 141, 232–237. [Google Scholar] [CrossRef]
  211. Roth, L.; Aasen, H. Extracting leaf area index using viewing geometry effects-A new perspective on high-resolution unmanned aerial system photography. ISPRS 2018, 141, 161–175. [Google Scholar] [CrossRef]
  212. Deng, L.; Mao, Z.; Li, X.; Hu, Z.; Duan, F.; Yan, Y. UAV-based multispectral remote sensing for precision agriculture: A comparison between different cameras. ISPRS 2018, 146, 124–136. [Google Scholar] [CrossRef]
  213. Jannoura, R.; Brinkmann, K.; Uteau, D.; Bruns, C.; Joergensen, R.G. Monitoring of crop biomass using true colour aerial photographs taken from a remote controlled hexacopter. Biosyst. Eng. 2015, 129, 341–351. [Google Scholar] [CrossRef]
  214. Rokhmana, C.A. The Potential of UAV-based Remote Sensing for Supporting Precision Agriculture in Indonesia. Proc. Environ. Sci. 2015, 24, 245–253. [Google Scholar] [CrossRef] [Green Version]
  215. Jagüey, J.G.; Villa-Medina, J.F.; López-Guzmán, A.; Porta-Gándara, M.Á. Smartphone irrigation sensor. IEEE Sens. J. 2015, 15, 5122–5127. [Google Scholar] [CrossRef]
  216. Kavianand, G.; Nivas, V.M.; Kiruthika, R.; Lalitha, S. Smart drip irrigation system for sustainable agriculture. In Proceedings of the 2016 IEEE Technological Innovations in ICT for Agriculture and Rural Development (TIAR), Chennai, India, 15–16 July 2016; pp. 19–22. [Google Scholar]
  217. Zaier, R.; Zekri, S.; Jayasuriya, H.; Teirab, A.; Hamza, N.; Al-Busaidi, H. Design and implementation of smart irrigation system for groundwater use at farm scale. In Proceedings of the 2015 7th International Conference on Modelling, Identification and Control (ICMIC), Sousse, Tunisia, 18–20 December 2015; pp. 1–6. [Google Scholar]
  218. Oksanen, T.; Linkolehto, R.; Seilonen, I. Adapting an industrial automation protocol to remote monitoring of mobile agricultural machinery: A combine harvester with IoT. IFAC-Pap. 2016, 49, 127–131. [Google Scholar] [CrossRef]
  219. Lee, H.; Moon, A.; Moon, K.; Lee, Y. Disease and pest prediction IoT system in orchard: A preliminary study. In Proceedings of the 2017 Ninth International Conference on Ubiquitous and Future Networks (ICUFN), Milan, Italy, 4–7 July 2017; pp. 525–527. [Google Scholar]
  220. Chieochan, O.; Saokaew, A.; Boonchieng, E. IOT for smart farm: A case study of the Lingzhi mushroom farm at Maejo University. In Proceedings of the 2017 14th International Joint Conference on Computer Science and Software Engineering, Nakhon Si Thammarat, Thailand, 12–14 July 2017; pp. 1–6. [Google Scholar]
  221. Berni, J.A.; Zarco-Tejada, P.J.; Suárez, L.; Fereres, E. Thermal and narrowband multispectral remote sensing for vegetation monitoring from an unmanned aerial vehicle. IEEE Trans. Geosci. Remote Sens. 2009, 47, 722–738. [Google Scholar]
  222. Baluja, J.; Diago, M.P.; Balda, P.; Zorer, R.; Meggio, F.; Morales, F.; Tardaguila, J. Assessment of vineyard water status variability by thermal and multispectral imagery using an unmanned aerial vehicle (UAV). Irrig. Sci. 2012, 30, 511–522. [Google Scholar] [CrossRef]
  223. Gonzalez-Dugo, V.; Zarco-Tejada, P.; Berni, J.A.; Suarez, L.; Goldhamer, D.; Fereres, E. Almond tree canopy temperature reveals intra-crown variability that is water stress-dependent. Agric. For. Meteorol. 2012, 154, 156–165. [Google Scholar]
  224. Zarco-Tejada, P.J.; González-Dugo, V.; Berni, J.A. Fluorescence, temperature and narrow-band indices acquired from a UAV platform for water stress detection using a micro-hyperspectral imager and a thermal camera. Rem. Sens. Environ. 2012, 117, 322–337. [Google Scholar] [CrossRef]
  225. Gonzalez-Dugo, V.; Zarco-Tejada, P.; Nicolás, E.; Nortes, P.A.; Alarcón, J.J.; Intrigliolo, D.S.; Fereres, E.J.P.A. Using high resolution UAV thermal imagery to assess the variability in the water status of five fruit tree species within a commercial orchard. Precis. Agric. 2013, 14, 660–678. [Google Scholar] [CrossRef]
  226. Aboutalebi, M.; Allen, L.N.; Torres-Rua, A.F.; McKee, M.; Coopmans, C. Estimation of soil moisture at different soil levels using machine learning techniques and unmanned aerial vehicle (UAV) multispectral imagery. Auton. Air Ground Sens. Syst. Agric. Optim. Phenotyping 2019, 11008, 216–226. [Google Scholar]
  227. Chen, A.; Orlov-Levin, V.; Meron, M. Applying high-resolution visible-channel aerial imaging of crop canopy to precision irrigation management. Agric. Water Manag. 2019, 216, 196–205. [Google Scholar] [CrossRef]
  228. Jorge, J.; Vallb, M.; Soler, J.A. Detection of irrigation in homogeneities in an olive grove using the NDRE vegetation index obtained from UAV images. Euro. J. Remote Sens. 2019, 52, 169–177. [Google Scholar] [CrossRef] [Green Version]
  229. Zhang, Y.; Han, W.; Zhang, H.; Niu, X.; Shao, G. Evaluating soil moisture content under maize coverage using UAV multimodal data by machine learning algorithms. J. Hydrol. 2023, 617, 129086. [Google Scholar]
  230. Kropp, I.; Nejadhashemi, A.P.; Deb, K.; Abouali, M.; Roy, P.C.; Adhikari, U.; Hoogenboom, G. A multi-objective approach to water and nutrient efficiency for sustainable agricultural intensification. Agric. Syst. 2019, 173, 289–302. [Google Scholar] [CrossRef]
  231. Nawandar, N.; Satpute, V. IoT based low cost and intelligent module for smart irrigation system. Comput. Electron. Agric. 2019, 162, 979–990. [Google Scholar] [CrossRef]
  232. Chen, X.; Qi, Z.; Gui, D.; Gu, Z.; Ma, L.; Zeng, F.; Sima, M.W. A model-based real-time decision support system for irrigation scheduling to improve water productivity. Agronomy 2019, 9, 686. [Google Scholar] [CrossRef] [Green Version]
  233. Chen, X.; Feng, S.; Qi, Z.; Sima, M.W.; Zeng, F.; Li, L.; Wu, H. Optimizing Irrigation Strategies to Improve Water Use Efficiency of Cotton in Northwest China Using RZWQM2. Agriculture 2022, 12, 383. [Google Scholar] [CrossRef]
  234. Chen, X.; Qi, Z.; Gui, D.; Sima, M.W.; Zeng, F.; Li, L.; Gu, Z. Evaluation of a new irrigation decision support system in improving cotton yield and water productivity in an arid climate. Agric. Water Manag. 2020, 234, 106139. [Google Scholar] [CrossRef]
Figure 1. Global distribution of land degradation based on severity of human-induced pressure [29].
Figure 1. Global distribution of land degradation based on severity of human-induced pressure [29].
Agronomy 13 02113 g001
Figure 2. The global distribution of agricultural blue water scarcity (BWS), green water scarcity (GWS), and economic water scarcity (EWS) [16].
Figure 2. The global distribution of agricultural blue water scarcity (BWS), green water scarcity (GWS), and economic water scarcity (EWS) [16].
Agronomy 13 02113 g002
Figure 3. Interlinkage of SDGs and water highlighting the importance of water security for humans and environment.
Figure 3. Interlinkage of SDGs and water highlighting the importance of water security for humans and environment.
Agronomy 13 02113 g003
Figure 4. Countries affected by drought in 2020–2022 [89].
Figure 4. Countries affected by drought in 2020–2022 [89].
Agronomy 13 02113 g004
Table 1. Application of various artificial intelligence technologies for irrigation management.
Table 1. Application of various artificial intelligence technologies for irrigation management.
StrategyOutcomesReferences
Fuzzy logicOptimization[169]
ANNDecrease in evaporation due to schedule and savings observed in water and electrical energy[170]
Fuzzy logicThe fuzzy controller system can be effectively applied to PA applications such as water-saving agriculture areas, for example, the croplands, the nursery gardens and the greenhouses.[171]
Fuzzy logic controllerDrip irrigation prevents wastage of water and evaporation[172]
Fuzzy decision support
system
The system provided improved irrigation suggestions in terms of timing and water saving.[173]
ANN feedforwardOptimization of water resources in a smart farm[174]
Machine learning algorithmPrediction and tackles drought conditions[175]
Fuzzy logicObtained a higher level of accuracy to expertly use water for irrigation[176]
ANNNeural network models with one hidden layer with four neurons for sugar beet and five neurons for wine grape provided excellent predictions of well-watered canopy temperature[177]
ANNThe proposed model was able to predict the timing and quantity of irrigation water[178]
LoRa-based machine
learning
This system led to a 46% reduction in water usage, and the plants looked better than they would have with conventional watering[179]
Table 2. Application of different types of UAVs for irrigation management.
Table 2. Application of different types of UAVs for irrigation management.
Type of UAVs UsedPurposeReferences
Unmanned helicopterMapping of crop water stress, index for irrigation scheduling[221]
Unmanned helicopterAssess water stress variability in a commercial vineyard[222]
Unmanned helicopterThe fuzzy controller system can be effectively applied to water stress detection in an almond orchard[223]
Multi-copter enginesWater stress detection in grapevine[224]
Fixed wingDetection of water stress in citrus cultivars[224]
Fixed wingWater stress detection in fruit tress[225]
Fixed wingSoil moisture estimation at different soil levels[226]
QuadcopterEstimation of canopy cover maps for irrigation management of peanut and cotton[227]
QuadcopterIdentification of nonuniformly irrigated areas in olive groves and vineyard crops[228]
HexacopterSoil moisture content prediction under different irrigation treatments in maize crop[229]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahmed, Z.; Gui, D.; Murtaza, G.; Yunfei, L.; Ali, S. An Overview of Smart Irrigation Management for Improving Water Productivity under Climate Change in Drylands. Agronomy 2023, 13, 2113. https://doi.org/10.3390/agronomy13082113

AMA Style

Ahmed Z, Gui D, Murtaza G, Yunfei L, Ali S. An Overview of Smart Irrigation Management for Improving Water Productivity under Climate Change in Drylands. Agronomy. 2023; 13(8):2113. https://doi.org/10.3390/agronomy13082113

Chicago/Turabian Style

Ahmed, Zeeshan, Dongwei Gui, Ghulam Murtaza, Liu Yunfei, and Sikandar Ali. 2023. "An Overview of Smart Irrigation Management for Improving Water Productivity under Climate Change in Drylands" Agronomy 13, no. 8: 2113. https://doi.org/10.3390/agronomy13082113

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop