SciELO - Scientific Electronic Library Online

 
vol.100 número especialConservación de plantas silvestres en México en el siglo XXIEl reto de la restauración participativa en áreas rurales índice de autoresíndice de materiabúsqueda de artículos
Home Pagelista alfabética de revistas  

Servicios Personalizados

Revista

Articulo

Indicadores

Links relacionados

  • No hay artículos similaresSimilares en SciELO

Compartir


Botanical Sciences

versión On-line ISSN 2007-4476versión impresa ISSN 2007-4298

Bot. sci vol.100 spe México  2022  Epub 17-Oct-2022

https://doi.org/10.17129/botsci.3074 

Ecology

Ecosystem science: a new approach in the analysis of functional processes in natural and human transformed terrestrial ecosystems

La ciencia de los ecosistemas: un nuevo enfoque en el análisis de los procesos funcionales en ecosistemas terrestres naturales y transformados

Bruno Chávez-Vergara, Writing – original draft, Conceptualization1  2 
http://orcid.org/0000-0003-1061-5685

Enrico A. Yepez, Writing – original draft, Conceptualization3  2 
http://orcid.org/0000-0003-4746-573X

Felipe García-Oliva, Writing – original draft, Conceptualization4  * 
http://orcid.org/0000-0003-4138-1850

1 Instituto de Geología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico.

2 Laboratorio Nacional de Geoquímica y Mineralogía, Ciudad de México, Mexico.

3 Departamento de Ciencias del Agua y Medio Ambiente, Instituto Tecnológico de Sonora, Ciudad Obregón, Sonora, Mexico.

4 Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico.


Abstract

Although Tansley originally proposed the ecosystem concept in 1935, ecosystem science underwent significant development in the last 20 years, as in this period it has been consolidated with concepts and methods arisen in convergent disciplines, such as ecosystem genetics, ecological stoichiometry, global ecology, and ecosystem services. The objective of this paper is to review new concepts and methods of water, energy, and nutrient dynamics research in terrestrial ecosystems to contribute to generate a new theoretical framework in the field of ecosystem science. From this review, a new conceptual definition of ecosystem is required based on three key issues: (a) the integration of functional processes at different spatial and temporal scales to understand the ecosystem dynamics in its environmental context; (b) the concept of resource (i.e., water or nutrients) use efficiency as a key metric for ecosystem function; and (c) the role of biological species in ecosystem functioning, using the genetic framework. These new concepts and methods are necessary to advance in the research on ecosystem functioning and resilience in the context of the current environmental crisis that includes processes such as ecosystem degradation, biodiversity loss, and global climate change. Finally, this new conceptual definition must be linked to evolutionary theory and global ecology research.

Keywords: ecosystems; energy; functional biodiversity nutrients; resilience; resource use efficiency; water

Resumen

A pesar de que el concepto de ecosistema fue propuesto por Tansley desde 1935, la ciencia de los ecosistemas ha tenido un desarrollo significativo en los últimos 20 años, ya que en este periodo se fortaleció con conceptos y métodos surgidos en disciplinas convergentes como la genética de ecosistemas, la ecología estequiométrica, la ecología global y los servicios ecosistémicos. El objetivo del presente trabajo es revisar los nuevos marcos conceptuales y metodológicos en los estudios de la dinámica del agua, energía y nutrientes en ecosistemas terrestres. De esta revisión, destacan tres aspectos relevantes: (a) la integración de los procesos funcionales a diferentes escalas espaciales y temporales para entender la dinámica de los ecosistemas en su contexto ambiental, (b) el concepto de eficiencia del uso de recursos (agua o nutrientes) como una medición clave en el funcionamiento de los ecosistemas, y (c) el papel de las especies biológicas en el funcionamiento de los ecosistemas en el contexto de la genética de los ecosistemas. Una nueva definición del concepto del ecosistema es necesaria considerando los avances de los marcos teóricos, y es necesario aplicar nuevas metodologías a los estudios de ecosistemas. Estos nuevos conceptos y herramientas son necesarias para mejorar la investigación sobre la resiliencia de los ecosistemas en la actual crisis ambiental que incluye la degradación de ecosistemas, la pérdida de la biodiversidad y el cambio climático global. Finalmente, esta nueva definición de ecosistema debe estar ligada con la teoría de la evolución y la investigación en ecología global.

Palabras clave: agua; biodiversidad funcional; ecosistemas; eficiencia de uso de recursos; energía; nutrientes; resiliencia

Among the disciplines of ecology, ecosystem science is the most dynamic and has undergone the greatest development over the last 20 years. The term “ecosystem” was originally proposed by Tansley (1935), then Lindeman (1942) gave a framework to the ecosystem concept using the concept of trophic-dynamics of energy flow in lakes and, Odum (1969) integrated biodiversity, energy flow and nutrient dynamics into a conceptual model for ecosystem development. This conceptual model defined the ecosystem as a cybernetic system with a holistic approach (Margalef 1968, Odum 1969). However, this model had two main problems: the definition of ecosystem boundaries (Levin 1992) and the heterogeneity and scales, all of which were unclear within the black-box vision (Turner 1989, Pickett & Cadenasso 2002, Currie 2011, Carpenter & Turner 2017). Indeed, several contemporary ecosystem-level studies face these problems. In terrestrial ecosystems, the problem of boundaries was assumed to be a methodological issue encountered in the context of watersheds. The studies of Bormann & Likens (1979) and Swank & Crossley (1988) used small experimental watersheds for measuring water and nutrient budgets at the ecosystem level. Moreover, Bormann et al. (1974) proposed that forest ecosystems have the resilience capacity to cope with severe disturbances, such as complete deforestation; this study represented an important experiment to analyze the response of a forest ecosystem to disturbances. Although these studies do not solve the problem of boundaries, they address several conceptual and methodological problems for ecosystem science. O’Neill et al. (1986) suggested that the ecosystem is the highest hierarchical level of the components of a functional organization, and that the processes at the ecosystem level define the boundaries. However, this conceptual proposal is not fully accepted because of several limitations of the ecosystem concept as a closed biotic system (Currie 2011). The spatial heterogeneity and the integration of spatial and temporal scales have been addressed through landscape ecology (Turner 1989). All of these limitations of the ecosystem concept complicate the definition of ecosystem as an operational unit (Currie 2011). Ecosystem science thus requires convergent disciplines to overcome its limitations. Finally, in 1998, a scientific forum on ecosystem science was published in the Ecosystems journal (Carpenter & Turner 1998), that served to define a contemporary approach for ecosystem ecology research (Carpenter & Turner 2017).

Over the last two and a half decades, several concepts and methods have arisen in different convergent disciplines, contributing to the consolidation of ecosystem science. These disciplines include ecosystem genetics (Whitham et al. 2006, Sardans et al. 2011, Peñuelas et al. 2013), ecological stoichiometry (Sterner & Elser 2002, Peñuelas et al. 2019), global ecology (Schimel et al. 2019, Schlesinger & Bernhardt 2020), and ecosystem services (Millennium Ecosystem Assessment 2003).

The role of biological species in ecosystem functioning has been a relevant subject for ecosystem science and several authors have proposed different models to understand the effect of species on ecosystem resilience (Peterson et al. 1998). However, the role of species in different ecosystem processes has been determined with the development of omic molecular methods, such as genomics, proteomics, and transcriptomics (Whitham et al. 2006, Sardans et al. 2011, Peñuelas et al. 2013). These omic methods were applied successfully to the understanding of the biological richness and their interactions in microbial and plant communities, allowing the development of a conceptual framework with which to understand the genomic basis of ecosystem processes (Whitham et al. 2006).

In the two last decades of the 20th century, several authors considered that specific ecosystems are limited mainly by a single nutrient; for example, temperate forests and deserts are mainly limited by nitrogen (N), while tropical forests are limited by phosphorus (P) (Vitousek 1984, Noy-Meir 1985, Vitousek & Sanford 1986, Aber & Melillo 1991). At present, the net primary productivity of ecosystems can be co-limited by carbon (C), N, and P; expressed as C:N, C:P and N:P ratios (Sterner & Elser 2002, Sinsabaugh et al. 2009, Tapia-Torres et al. 2015a). In order for biological species to maintain their C:N:P ratios, they must invest energy in nutrient acquisition, rather than in population growth (Sterner & Elser 2002). The conceptual framework of these processes is known as ecological stoichiometry (Sterner & Elser 2002, Peñuelas et al. 2019), and it serves to analyze nutrient uptake in relation to the metabolic efficiency of specific species or group of species (Sterner & Elser 2002, Peñuelas et al. 2019). This theoretical framework allows the study of nutrient dynamics in an evolutionary context. Wider aspects of these interactions across trophic levels and functional assemblages at different scales are expressed as eco-evolutionary dynamics (Brunner et al. 2019).

Global ecology is focused on the integrated study of ecosystems from regional to global scales including carbon and nutrient dynamics, water dynamics and global energy budgets (Schimel et al. 2019; Schlesinger & Bernhardt 2020, Yu et al. 2021). This discipline is based on the development of methods that allow measurements of processes at different scales, such as plant physiological (i.e., photosynthesis and transpiration), soil (i.e., respiration, C accumulation) and atmospheric (i.e., turbulence, emissions) processes and their interactions. In this context, flux towers, remote sensing methods and isotopic ratios are fundamental tools in global ecology (Fry 2006, Schimel et al. 2019).

The human dimension as a key element for ecosystem functioning has been recognized since Odum (1969), but adoption of the concept of ecosystem services as a framework for research increased dramatically at the end of the 1990s (Costanza et al. 1997; Daily et al. 2000; Carpenter et al. 2009, Carpenter & Turner 2017). The advent of the Millennium Ecosystem Assessment synthesis (2005) clearly established the conceptual framework of ecosystem services and made it critical to current ecosystem management research. In addition, Currie (2011) and Yu et al. (2021) proposed ecosystem science as a tool for studying the effects of climate change and anthropic management on ecosystems, in terms of achieving sustainable development.

The objective of the present paper is to review novel concepts and methods of convergent disciplines that contributed to the theoretical framework of ecosystem science. In the two following sections, we review examples of concepts and methods relating to water, energy, and nutrient dynamics in terrestrial ecosystems. We then analyze how these new concepts and methods have influenced in the understanding of the ecosystem resilience processes.

Ecosystem water and energy dynamics

Water fluxes through soils, rocks, organisms, and the atmosphere essentially control all functional processes in terrestrial ecosystems (Biederman et al. 2016, Rodríguez-Robles et al. 2020). This circulation of water can be conceptualized and measured as energy and matter exchanges across ecosystem components that are central to the understanding of ecosystem processes. Given the close relationship between water fluxes and biological productivity, contemporary interdisciplinary fields such as ecohydrology have recently presented a research agenda to study ecosystem functions and applications (Newman et al. 2006). Ecohydrology embraces theory and methods from ecology and hydrology to identify couplings and feedbacks between organisms, populations and communities, as well as water-related processes across a broad range of spatial and temporal scales. This perspective has gained attention due to the significant impacts on ecosystems for the growing demand of ecosystem services (Sun et al. 2017) and climate change (Tague et al. 2020).

Early conceptual tools of ecohydrology, such as rain use efficiency (RUE), defined as the annual net primary production/mean annual precipitation ratio (Leith 1975, Le Houérou 1984), were widely used as basic process-integrators and to investigate environmental constraints on biological productivity (Lauenroth & Sala 1992, Paruelo et al. 1999). Since a fundamental relationship emerges when rainfall and biological production parameters are related, the role of water as a key driver for ecosystem function is clear. Indeed, when temperate or tropical-wet ecosystems experience water limitations, the slopes of rainfall and biological production converge to a common RUE (Huxman et al. 2004). This variation on the RUE slope indicates that the relationship itself is of interest, since RUE deviations in particular ecosystems, at specific times and under varying environmental conditions, could indicate ecophysiological adaptations (Plaut et al. 2012), organismic interactions (Rodríguez-Robles et al. 2020), constraints to nutrient availability (Porporato et al. 2003) and levels of ecosystem degradation (Ehleringer 2001).

When scaled to ecosystem level (Chapin III et al. 2006), photosynthesis (A) and gross primary production (GPP) express the CO2 assimilation process in plants. This flux is therefore the main path for radiant energy and inorganic atmospheric compounds to be incorporated into ecosystems as organic matter. However, such a fundamental flux has an expensive water cost, since a water loss of two orders of magnitude occurs at leaf level for each CO2 molecule fixed (Nobel 1999). This process led to the establishment of H2O and CO2 molecules as central currencies of matter and energy flow in studies addressing ecosystem function (Mooney 1972, Pearcy et al. 1989, Chapin III et al. 2002, Shuttleworth 2007).

Theoretical approaches to assess the availability of water that controls biological productivity, and contemporary methods that rely on advanced technology capable of monitoring CO2 and water fluxes in temporal and spatial scales to determine ecosystem function, are now central to ecosystem science research (Keenan & Williams 2018, Schimel et al. 2019, Baldocchi 2008, Yu et al. 2021). Careful tracing of H2O and CO2 fluxes across ecological scales of organization within an ecosystem has emerged as a powerful practical and conceptual tool with which to advance theory concerning ecosystem functioning (Tang et al. 2014), carbon cycle analysis (Vargas et al. 2012), as well as to consider ecosystems as natural climate solutions (NCS) in this age of chronic environmental climate change (Hemes et al. 2021).

Carbon dioxide and water measurements taken at the ecosystem scale with the eddy covariance technique have provided key metrics (Chapin III et al. 2006), such as evapotranspiration (ET), gross primary production (GPP), net ecosystem exchange (NEE), net ecosystem production (NEP) and ecosystem-level water use efficiency (EWUE), with which to assess ecosystem functioning across continents (Baldocchi 2008, Tang et al. 2014, Baldocchi et al. 2018, Keenan & Williams 2018, Delgado-Balbuena et al. 2019, Tarin et al. 2020, Kang & Cho 2021, Heiskanen et al. 2022). One example of these applications in the semiarid region of western North America comes from an integration of ecosystem water and CO2 measurements in combination with remote sensing tools (Biederman et al. 2016, 2017), to use time by space substitutions to explore the relationships of rainfall and ET to the NEP and GPP, and to compare the carbon sink potential and controls of arid and forested ecosystems. Taking advantage of the RUE-like criteria (see above), the authors invoked “fast” (i.e., weather, soil moisture, leaf area, microbial communities) and “slow” (i.e., climate, atmospheric CO2 concentration, soil development, biogeochemical pools, plant community) factors as differential controls of ecosystem processes across sites. Since most of the unexplained NEP variability was related to site-specific functions, further research is necessary to obtain more conclusive evidence regarding how slow factors drive carbon and water relations in ecosystems (Biederman et al. 2016, Baldocchi et al. 2018). However, since common temporal and spatial sensitivity to interannual environmental changes can be extrapolated to predict fundamental water controls on CO2 exchanges and biological productivity over long-term temporal scales, an important tool emerges to improve our understanding of ecosystem resilience in the context of climate change and human pressures (Yu et al. 2021).

Moreover, when disaggregated ecosystem fluxes are related among themselves (i.e., ET and GPP), there is a clearer biological response to the environmental drivers. For instance, the RUE slope for widespread ecosystems is asymptotic, while an ET vs. GPP relationship is linear (Biederman et al. 2016). As part of understanding these differences, the issue of temporal and spatial scales is unavoidable, but this also presents a challenge to ecosystem science regarding how to implement means to separate ecosystem fluxes into their fundamental components; for example, ET into transpiration (T) and evaporation (E), or NEP into GPP and ecosystem respiration (Reco), such that more in-depth explorations of the biological responses to environmental drivers can be made at ecosystem scales, while obtaining mechanistic information pertaining to ecosystem processes (Yepez et al. 2007, Stoy et al. 2019).

To confront this challenge, a combination of tools, including leaf cuvettes, porometer techniques, sap flow probes, soil chambers, lysimeters, and in situ spectral tools, has traditionally been employed. However, the propagation of error when scaling these methods up to an entire ecosystem is substantial (Wilson et al. 2001, Kool et al. 2014). A partial solution is provided by the eddy covariance technique (EC), which provides a direct and continuous estimate of aggregated fluxes, such as ET and NEE (NEE = NEP), within an ecosystem. In combination with modeling, data assimilation techniques and ancillary measurements have provided evidence of the capacities of these approaches to partition fluxes. Recently, Zhou et al. (2016) and Scanlon & Kustas (2010) were able to employ a water-use efficiency criterion to partition ET as the ratio of GPP and ET, simply by using the eddy covariance method, providing promising results. Although this approach is appealing, considerable uncertainties remain regarding the estimation of GPP with these criteria, since GPP is not a directly measured variable in EC operations but rather one that is modeled (Reichstein et al. 2005). Nevertheless, there is a promising future for disentangling spatial and temporal NEP variation with the application of remote sensing techniques (i.e., drones, microsatellite clouds, satellites, and space stations), relying on the spectral properties of surfaces, combined with numerical assimilation schemes that can offer greatly enhanced spatial resolution to detect functional traits and temporal dynamic changes in ecosystems (Schimel et al. 2019).

Fundamentally, the aim of separating water and carbon flux components demands a combination of temporal and spatial congruent techniques (Yepez et al. 2007, Stoy et al. 2019). Combining EC and stable isotope techniques constitutes an alternative that allows flux partitioning at the ecosystem scale (Good et al. 2012). This is possible since the stable isotopes of water and carbon can be used as tracers of hydrological and carbon cycles, because of their different fractionation processes (Yakir & Sternberg 2000).

Implementation of a concise scheme of combined methods, for incorporation into long-term monitoring programs to generate baselines of ecosystem functions responding to environmental variation at different scales, will undoubtedly contribute to the identification of effects of climate change on ecosystems and to the provision of ecosystem management strategies as options for natural climate solutions in the context of chronic climate change.

Ecosystem nutrient dynamics

The ecosystem nutrient dynamics (END) is broadly defined as the manner in which nutrients are taken up, retained, transferred, and recycled over time and space in an ecosystem (Allan & Castillo 2007). END is an emerging property of ecosystems, which, like the other biogeochemical processes, is the result of a convergence of geological, pedological, and ecological processes, each of which depends on the spatial and temporal scales at which they are observed (Berhe & Kleber 2013, Berhe et al. 2018). In most terrestrial ecosystems, END not only depends on the rate of input from extreme boundaries (i.e., atmospheric deposition, weathering) but also, and more importantly, from internal cycles, where soil organic matter is the main source of inorganic nutrients through its decomposition by the soil microbial community (Schlesinger et al. 2011, Banwart et al. 2019). In general, three stages help to summarize END: (a) absorption and allocation of inorganic nutrients by plant roots, (b) reabsorption of nutrients from senescent tissues and return of nutrients by litterfall and abscission of fine roots, and (c) release of inorganic nutrients from organic matter to the soil solution through microbial metabolism (mineralization).

In the first stage of END, nutrient absorption by the roots depends on the characteristics of the nutrients themselves, such as solubility, diffusion speed, and concentration. For example, N, P, and K, in the forms of HPO4 2- and H2PO4-, K+ and NH4 +, diffuse slowly due to the strong affinity with other components of the soil, cations in the case of P and clays in the case of K+ and NH4 +, while NO3 - is highly mobile and diffuses rapidly in the soil solution (Sparks 2019). One of the most common approaches to estimate the consequence of nutrient absorption in biomass production is through nutrient use efficiency (NUE), which is defined as the quantity of biomass production per unit of nutrient acquired. However, this does not allow direct measurement of the flow of nutrients to the plant (Silla & Escudero 2003, Chapin III et al. 2011). Once the roots acquire the nutrients, these are distributed through the different tissues. These requirements determine the demand for acquired nutrients and ultimately define their limitation (Peñuelas et al. 2019). The distribution of nutrients in plants is determined by their destination, including structural molecules, tissues with high metabolic activity such as photosynthetic organs, chemical defenses, or other defense structures. In general, three factors govern the allocation of nutrients in plants: (1) the genetic structure of the plant species community, (2) the environmental conditions in which the plants grow, and (3) the seasonal program of resource allocation (Hartmann et al. 2020).

In the second stage of END, leaves and fine roots have higher concentrations of nutrients (mainly N and P) than other tissues since both of these organs are physiologically active. Likewise, these structures are continually replaced due to depletion of their physiological function but, prior to detachment from the plant, they go through a process of preparation or deciduous senescence in which some nutrients are recovered and reabsorbed before abscission, then transported and stored in branches, stem or structural roots where they remain available for the formation of new functional or reproductive structures in the following phenological cycle (Leopold 1961, Killingbeck 1996, Cartaxana & Catarino 2002). Translocation or reabsorption of nutrients from senescent leaves has been recognized as an efficient mechanism for conservation of nutrients in several plant species (Killingbeck 1996) since it allows them to be less dependent on the availability of N and P in the soil (Aksekili et al. 2007). It has been calculated that between 50 and 80 % of foliar N and P can be reabsorbed, and that there are differences between these and other nutrients, (e.g., the cations), for which translocation is less efficient (Killingbeck & Costigan 1988, Rentería et al. 2005, Hagen-Thorn et al. 2006, Covelo et al. 2008b). One of the more commonly accepted metrics to calculate nutrient reabsorption is reabsorption efficiency, defined as the percentage of the nutrient that has been reabsorbed, based on the nutrient content in living leaves and the nutrient content in senescent leaves (Killingbeck 1996, Vergutz et al. 2012).

Multiple factors can define nutrient reabsorption patterns, including soil nutrient availability, foliar nutrient concentration, senescence time, air temperature, water stress, and factors associated with the physiological characteristics of each species (Hagen-Thorn et al. 2006). For example, it has been reported that between 1,300 and 1,900 genes are activated in Arabidopsis thaliana when senescence begins (Zentgraf et al. 2004). However, it is not known how many of these genes are directly related to nutrient resorption (Jones 2004). Studies have been carried out for different tree species, aiming to define their reabsorption patterns compared to other taxa (Killingbeck & Costigan 1988, Hagen-Thorn et al. 2006), considering the effect of forest management (Covelo et al. 2008b), and with climatic conditions such as a water gradient (Del Arco et al. 1991, Zhao et al. 2017). Moreover, studies have analyzed the effect of spatial patterns (Covelo et al. 2008a) and the foliar phenological strategies of different species (Huang et al. 2007, Brant & Chen 2015) on foliar nutrient resorption. In general, it has been determined that deciduous species have higher reabsorption efficiencies than perennials and, even within the same phenology, species with a longer leaf life-span present lower reabsorption efficiencies (Huang et al. 2007, Vergutz et al. 2012, Chavez-Vergara et al. 2014). It is widely accepted that nutrient reabsorption efficiency has an important genetic component at population level (Allan et al. 2012, Whitham et al. 2012) or at different taxonomic levels (Chávez-Vergara et al. 2015). The effect of physiological plant traits with a genetic basis on the chemical composition of the plant-derived material that returns to the soil is considered a part of ecosystem genetics, since these physiological traits can be analyzed within an evolutionary framework (Whitham et al. 2006, Peñuelas et al. 2013).

In the third stage of the END, the return of nutrients to the soil via organic residues follows different routes, such as canopy leaching (throughfall), litterfall, root death and root exudates. Litterfall quantity is regulated by primary productivity and seasonality by phenology, which is influenced by the availability of water, temperatures during the growing season, and the canopy morphology (Brant & Chen 2015). In forest ecosystems, the contribution from litterfall is dominated by the tree stratum, which represents 95 % of the litter mass that falls to the soil surface (Kögel-Knabner 2002). Of this percentage, about 80 % corresponds to foliar biomass (Rapp et al. 1999; Santa Regina & Tarazona 2001). The concentration of nutrients in the leaves that fall to the ground is regulated by the processes of nutrient resorption, environmental conditions, and plant physiological responses, while the intrinsic characteristics of the species themselves determine the quantity and chemical composition of the organic residues (Chavez-Vergara et al. 2014). Another way to return nutrients to the soil is through the abscission of fine roots, but very little is known about this process due to the difficulty of collecting samples. Nevertheless, in some cases, it can be as important as litter fall in terms of the mass and concentration of nutrients returned to the soil (e.g., in tropical dry forests) (Martinez-Yrizar et al. 1996).

Before organic residues can be transformed, they must be suspended or dissolved in the soil solution, when they become known as dissolved organic matter (DOM) and must be smaller than 0.45 µm (von Lützow et al. 2007). Some chemical characteristics that can hinder their decomposition include hydrophobicity, since they will tend to form micelles that are difficult to break, a high C:N ratio, non-polarity of the molecules, the presence of carbon covalent bonds (C-quaternary), functional groups with poorly hydrolyzable bonds (ester, phenyl, cyano), nitrogen heterocyclic (pyrrole, pyridine) and long hydrocarbon chains (Kiikkilä et al. 2005, van Hees et al. 2005, Roberts et al. 2009; Kiikkilä et al. 2011, Glanville et al. 2012, Jones et al. 2012). The use of solid-state 13C NMR and 31P NMR allows the definition of the molecular structure of soil organic C and P molecules and thus indicates their susceptibility to degradation (Almendros et al. 2000, Turner et al. 2003, 2007, Noack et al. 2012, Bonanomi et al. 2013, Merino et al. 2015, García-Oliva et al. 2018, Merino et al. 2021).

The dominant molecular size in dissolved organic matter is greater than 10 kDa, but most microorganisms can only assimilate small molecules (< 1 kDa) and the soil microbial community therefore requires a mechanism with which to cleave them (Neff & Asner 2001, Jones et al. 2012). This mechanism of depolymerization is primarily extracellular and carried out by ecoenzymes (exoenzymes). These are protein catalysts secreted into the soil solution and are essential for breaking the chemical bonds of organic molecules (Sinsabaugh 1994, Baldrian 2009, Bai et al. 2021). The enzyme can be produced if the genes that code for it are present in microbial community (Starke et al. 2021). For example, it is known that ascomycetes fungi have a greater specialization for the degradation of polysaccharides, while basidiomycetes, in addition to their capacity for polysaccharide degradation, are specialized in the degradation of lignin, polyphenols, and other aromatic compounds (Baldrian et al. 2011b, Culleton et al. 2013). The composition of the soil microbial community, as well as its genetic diversity, is therefore critical for soil organic matter decomposition, providing other example of the importance of genetic diversity to ecosystem functioning (Whitham et al. 2006, Sardans et al. 2011, Peñuelas et al. 2013). Likewise, to produce exoenzymes, a stimulus that generates the gene expression that produces the enzyme must be present (Ekschmitt et al. 2005, Baldrian et al. 2011a, Voříšková et al. 2011). For example, a high availability of nitrogen in the soil can reduce the production of exoenzymes associated with the depolymerization of nitrogenous molecules, while a low availability of nitrogen can stimulate the production of ecoenzymes, such as proteases, chitinases, and laccases. This suggests that microorganisms do not invest in producing these enzymes if they have access to available nutrients in the soil (Šnajdr et al. 2011, Glanville et al. 2012). Consequently, the soil microbial community must invest more energy in different ecoenzymes to acquire the limited element rather than growing the population (Sterner & Elser 2002, Sinsabaugh et al. 2009). However, the investment of energy in nutrient acquisition must be related to the C:N:P ratios of the soil microbial community. For example, Cleveland & Liptzin (2007) reported that the soil microbial community presents a relatively constant C:N:P ratio (60:7:1) in several terrestrial ecosystems. Furthermore, Sinsabaugh et al. (2009) reported an ecoenzymatic stoichiometry of 1:1 for C-N and C-P ecoenzymes in a different ecosystem, suggesting a strong relationship among ecoenzymatic activities. All of these results suggest that the C:N:P ratios are key indicators for understanding the integrated dynamic of nutrients in ecosystems, the conceptual framework of which is known as ecological stoichiometry (Sterner & Elser 2002, Sinsabaugh et al. 2009, Tapia-Torres et al. 2015a, Peñuelas et al. 2019).

Ecosystem resilience

In the two previous sections, we reviewed new theoretical frameworks and methods by which to understand the water, energy, and nutrient dynamics in terrestrial ecosystems, and which have served to consolidate ecosystem science. From this review, we highlighted three key issues: (a) the requirement for integration of functional processes at different spatial and temporal scales (e.g., from genomic based processes to global fluxes) to understand the dynamics of an ecosystem in its environmental context; (b) the resource use efficiency (energy, water or nutrients) as a key metric for ecosystem function; and (c) the role of biological species in ecosystems functioning using the genetic framework. As stated above, the development of new methodologies (i.e., eddy covariance, remote sensing, stable isotope, solid-state 13C NMR and 31P NMR, and the omic molecular methods) has allowed the three issues mentioned above to be adequately addressed.

One of the main problems of human well-being is related to the current environmental crisis, mainly through the problems derived from ecosystem degradation and global climate change (Millennium Ecosystem Assessment 2005, IPCC 2013). In 2005, one quarter of the continental area worldwide was converted to cultivated systems and approximately 60 % of the ecosystem services were degraded (Millennium Ecosystem Assessment 2005). Moreover, the last decade was the warmest since 1850 as a result of greenhouse gas emissions by human activities increasing the frequency of extreme climatic events (IPCC 2021) and disturbing several components and processes of ecosystems. For example, genetic diversity decreased globally (Millennium Ecosystem Assessment 2005), along with the capacity of soils to sequester carbon (Hoffmann et al. 2012, Albaladejo et al. 2013). Under this current scenario of environmental crisis, it is very important to understand the factors that determine the resilience capacity of ecosystems and find means by which to implement natural climate solutions (Griscom et al. 2017). Resilience is defined as the capacity of an ecosystem to return to the functional and structural condition it presented prior to perturbation (Holling 1973, Pimm 1984, Dell et al. 1986). López et al. (2011) then defined ecosystem resilience as how and to what extent the ecosystem condition changes between the different states prior to disturbance, in a model of different structural-functional states that includes threshold values of degradation that define the resilience capacity of an ecosystem for changing functional states. This definition considered that the ecosystem function has several equilibrium states, some of which could be far removed from its original condition. This latter definition therefore incorporates a dynamic concept of ecosystem resilience. Most of the papers related to ecosystem resilience analyzed changes of biological diversity following disturbance (MacGillivray et al. 1995, Tilman 1996, Griffiths et al. 2000, Wardle et al. 2000), but a few studies analyzed the recovery of functional processes (Herbert et al. 1999, Orwin & Wardle 2004, Hernández-Becerra et al. 2016). In addition, some authors proposed a quantitative index of ecosystem resilience (Herbert et al. 1999, Griffiths et al. 2001, Orwin & Wardle 2004).

Most studies are focused on a few variables, limiting the conclusions that can be made regarding the resilience capacity of ecosystems. For illustrative purposes, we present a resilience study conducted in a desert ecosystem in northern Mexico published by Hernández-Becerra et al. (2016). These authors analyzed soil microbial composition and soil nutrient dynamics in three plots with different management conditions (natural desert grassland, plot cultivated with alfalfa and plot abandoned for over 30 years) in the Cuatro Ciénegas Basin in northern Mexico. Briefly, the management consisted of cultivation of alfalfa (Medicago sativa L.) by flooding the fields and introducing large quantities of fertilizer, but the agriculture plots were abandoned because of soil degradation, mainly salinization (Hernández-Becerra et al. 2016, Tapia-Torres et al. 2018). Figure 1 shows the variables with the greatest differences among the three plots. The number of OTUs of soil bacteria and the Shannon diversity index decreased from the grassland to the abandoned plot, suggesting a reduction in the number of bacterial species as a result of the agricultural management. Soil pH and electrical conductivity were both reduced in the alfalfa plot, but these variables both increased dramatically in the abandoned plots. Soil pH is considered as a strong filter of several soil bacterial species (Delgado-Baquerizo et al. 2018, Malik et al. 2018). The bacterial species were therefore affected by two strong soil pH changes; one during the cultivation of alfalfa and the other after the agriculture activities were abandoned. The strongest changes in the bacterial community composition were observed during cultivation, in which the proportion of protobacteria decreased from 50 to 35 % in both the native grassland and alfalfa plot, and the species of cyanobacteria and chloroflexi bacteria were not observed in the latter plot. Finally, the available soil NO3 and HPO4 concentrations increased and decreased, respectively, in the abandoned plot. These two variables also had the lowest resilience index values among all the variables measured. These results suggest that nitrification increased with cultivation and soil P occlusion increased along with pH in the abandoned plots. Soil vulnerability to N loss therefore increased (Tapia-Torres et al. 2015b), as well as the soil capacity for geochemical occlusion of P (Perroni et al. 2014), and the soil processes that reduce the availability of N and P are therefore enhanced. These results suggest that the loss of bacterial species strongly affects the mechanisms of soil nutrient availability. The effect of the management could be worse than that of global climate change in this site, since the frequency of extreme climatic events had increased over the last decade (Montiel-González et al. 2021). In conclusion, soil bacterial composition is the key factor that controls the resilience of soil nutrient dynamics in this desert ecosystem. However, it is necessary to study the changes of expression of functional genes of bacteria that are related to nutrient dynamics. With this information, it is possible to design soil restoration or degradation mitigation methods, for example, by incorporating the functional bacterial species which were lost during the agriculture activities.

Figure 1 Soil bacteria OTUs, Shannon biodiversity index, and some soil characteristics (pH, EC: electrical conductivity, NO3: nitrate concentration and HPO4: available phosphate concentration) under three management conditions (G: natural desert grassland, C: cultivated plot with alfalfa, and A: abandoned plot) in the Cuatro Ciénegas Valley, Mexico. (Hernández-Becerra et al. 2016). 

According to the global ecology perspective, one future challenge is the development of simple but integrative metrics with which to assess the resilience capacity of ecosystems at large scales. As an exploratory tool with which to assess ecosystem response to disturbances, space for time substitutions in resource use efficiencies (i.e., RUE) can be used to measure resilience (i.e., as disturbance increases, the slope of RUE decreases, and vice versa, as ecosystems recover the slope RUE increases) (Ehleringer 2001). To develop these powerful metrics for the assessment of ecosystem function responding to disturbance, we first require the baseline information that only long-term monitoring schemes combining techniques at the appropriate scales can provide.

Integration of resource use efficiency, species functions, and ecosystems processes is critical for understanding changes in the resilience capacity of an ecosystem. Unfortunately, no studies have integrated these three components to date. Figure 2 shows the main processes driving ecosystem changes in four evolutionary stages. Prior to soil development in the foundational ecosystem stage, the parent material has a high concentration of mineral-derived nutrients but with low incorporation of organic matter, resulting in a low capacity for accumulating energy and water. Primary succession then occurs, promoting ecological interactions among the species with their eco-evolutionary dynamics, while the soils develop at the same time. In the developed ecosystem stage, the ecosystem shows a high water and energy storage capacity in the soils and biomass compared to the previous stages. In this stage, the availability of resources is strongly regulated by internal recycling mediated by the metabolic capacity of the organisms. This capacity depends on the species composition of the communities, where the resource use efficiency of species, defined by the genetic composition of species, could be an important driver of community composition. Incorporation of anthropic management into the ecological processes modifies the ecosystem function (human-transformed ecosystem stage), disrupting the feedback among ecological components and leading to soil degradation. In this condition, the storage capacity of energy and water is reduced due to the decline in functional, metabolic, and biodiversity. However, if the effect of anthropic management does not irreversibly deplete the biodiversity, it is possible for the ecosystem to return to its condition prior to management, depending on its resilience capacity. However, if the biodiversity is strongly depleted, a new human-induced foundational ecosystem stage is reached. This new stage is strongly limited by nutrients, energy, and water, with highly adverse physical and chemical conditions, representing a strongly degraded ecosystem.

Figure 2 Functional and ecological processes involved in the evolutionary stages of foundational, developed, human-transformed and human-induced foundational ecosystems. Created with BioRender.com  

Final remarks

In the last two and half decades, several concepts and methodologies from different convergent disciplines have contributed to the consolidation of ecosystem science. From these, three key issues emerge: (a) the requirement for integration of functional processes at different spatial and temporal scales to understand the dynamics of the ecosystem in its environmental context, (b) the concept of resource use efficiency (i.e., water or nutrients) as a key metric for ecosystem function, and (c) the role of biological species in ecosystem functioning, using a genetic framework. However, a new conceptual definition of the ecosystem is required in the context of the advance in theoretical bases, and it is necessary to apply these new methods to ecosystem studies. These new concepts and tools are necessary for improving ecosystem resilience research in the context of the current environmental crisis (i.e., ecosystem degradation, biodiversity erosion and global climate change). Finally, the new conceptual definition must be linked to evolution theory and global ecology research.

Acknowledgements

FGO received support from PAPIIT-DGAPA, UNAM (grant number IN207721) and EAY received support from PROFAPI-ITSON (2021-0030). BCV received support from PAPIIT-DGAPA, UNAM (grant number IN211820). We thank the anonymous reviewers for comments on preliminary drafts of the manuscript.

Literature cited

Aber JD, Melillo JM. 1991. Terrestrial Ecosystems. Philadelphia: Saunders College Publishing. ISBN: 978-0030474439 [ Links ]

Aksekili E, Kilic DD, Kutbay HG. 2007. Foliar nutrient dynamics and foliar resorption in Quercus brantii Lindley along an elevational gradient. Pakistan Journal of Biological Sciences 10: 3778-3785. DOI: https://doi.org/10.3923/pjbs.2007.3778.3785 [ Links ]

Albaladejo J, Ortiz R, Garcia-Franco N, Navarro AR, Almagro M, Garcia Pintado J, Martínez-Mena M. 2013. Land use and climate change impacts on soil organic carbon stocks in semi-arid Spain. Journal of Soils Sediments 13: 265-277. DOI: http://doi.org/10.1007/s11368-012-0617-7 [ Links ]

Allan GJ, Shuster SM, Woolbright SA, Walker F, Meneses N, Keith A, Bailey JK, Whitham TG. 2012. Perspective: interspecific indirect genetic effects (IIGEs). Linking genetics and genomics to community ecology and ecosystem processes. In: Ohgushi T, Schmitz OJ, Holt RD, eds. Trait-Mediated Indirect Interactions: Ecological and Evolutionary Perspectives. Cambridge: Cambridge University Press, pp. 295-323. ISBN: 978-0521173131 [ Links ]

Allan JD, Castillo MM. 2007. Stream Ecology: Structure and Function of Running Waters, 2nd ed. Dordrecht: Springer. DOI: https://doi.org/10.1007/978-1-4020-5583-6 [ Links ]

Almendros G, Dorado J, González-Villa FJ, Blanco MJ, Lankes U. 2000. 13C NMR assessment of decomposition patterns during composting of forest and shrub biomass. Soil Biology and Biochemistry 32: 793-804. DOI: https://doi.org/10.1016/S0038-0717(99)00202-3 [ Links ]

Bai X, Dippold MA, An S, Wang B, Zhang H, Loeppmann S. 2021. Extracellular enzyme activity and stoichiometry: The effect of soil microbial element limitation during leaf litter decomposition. Ecological Indicators 121: 107200. DOI: https://doi.org/10.1016/j.ecolind.2020.107200 [ Links ]

Baldocchi D. 2008. ‘Breathing’ of the terrestrial biosphere: lessons learned from a global network of carbon dioxide flux measurement systems. Australian Journal of Botany 56: 1-26. DOI: https://doi.org/10.1071/BT07151 [ Links ]

Baldocchi D, Chu H, Reichstein M. 2018. Inter-annual variability of net and gross ecosystem carbon fluxes: A review. Agricultural and Forest Meteorology 249: 520-533. DOI: https://doi.org/10.1016/j.agrformet.2017.05.015 [ Links ]

Baldrian P. 2009. Microbial enzyme-catalyzed processes in soils and their analysis. Plant, Soil and Environment 55: 370-378. DOI: https://doi.org/10.17221/134/2009-PSE [ Links ]

Baldrian P, Head IM, Prosser JI, Schloter M, Smalla K, Tebbe CC. 2011a. Ecology and metagenomics of soil microorganisms. FEMS Microbiology Ecology 78: 1-2. DOI: https://doi.org/10.1111/j.1574-6941.2011.01184.x [ Links ]

Baldrian P, Voříšková J, Dobiášová P, Merhautová V, Lisá L, Valášková V. 2011b. Production of extracellular enzymes and degradation of biopolymers by saprotrophic microfungi from the upper layers of forest soil. Plant and Soil 388: 111-125. DOI: https://doi.org/10.1007/s11104-010-0324-3 [ Links ]

Banwart SA, Nikolaidis NP, Zhu Y-G, Peacock CL, Sparks, DL. 2019. Soil functions: Connecting Earth’s critical zone. Annual Review of Earth and Planetary Sciences 47: 333-359. DOI: https://doi.org/10.1146/annurev-earth-063016-020544 [ Links ]

Berhe AA, Barnes RT, Six J, Marín-Spiotta E. 2018. Role of soil erosion in biogeochemical cycling of essential elements: carbon, nitrogen, and phosphorus. Annual Review of Earth and Planetary Sciences 46: 521-548. DOI: https://doi.org/10.1146/annurev-earth-082517-010018 [ Links ]

Berhe AA, Kleber M. 2013. Erosion, deposition, and the persistence of soil organic matter: mechanistic considerations and problems with terminology. Earth Surface Processes and Landforms 38: 908-912. DOI: https://doi.org/10.1002/esp.3408 [ Links ]

Biederman JA, Scott RL, Bell TW, Bowling DR, Dore S, Garatuza‐Payan J, Kolb TE, Krishnan P, Krofcheck DJ, Litvak ME, Maurer GE, Meyers TP, Oechel WC, Papuga SA, Ponce-Campos GE, Rodriguez JC, Smith WK, Vargas R, Watts CJ, Yepez EA, Goulden ML. 2017. CO2 exchange and evapotranspiration across dryland ecosystems of southwestern North America. Global Change Biology 23: 4204-4221. DOI: https://doi.org/10.1111/gcb.13686 [ Links ]

Biederman JA, Scott RL, Goulden ML, Vargas R, Litvak ME, Kolb TE, Yepez EA, Oechel WC, Blanken PD, Bell TW, Garatuza-Payan J, Maurer GE, Dore S, Burns SP. 2016. Terrestrial carbon balance in a drier world: the effects of water availability in southwestern North America. Global Change Biology 22: 1867-1879. DOI: https://doi.org/10.1111/gcb.13222 [ Links ]

Bonanomi G, Incerti G, Giannino F, Mingo A, Lanzotti V, Mazzoleni S. 2013. Litter quality assessed by solid state 13C NMR spectroscopy predicts decay rate better than C/N and Lignin/N ratios. Soil Biology and Biochemistry 56: 40-48. DOI: https://doi.org/10.1016/j.soilbio.2012.03.003 [ Links ]

Bormann FH, Likens GE. 1979. Pattern and Process in a Forested Ecosystem. New York: Springer Verlag. DOI: https://doi.org/10.1007/978-1-4612-6232-9 [ Links ]

Bormann FH, Likens GE, Siccama TG, Pierce RS, Eaton JS. 1974. The export nutrients and recovery of stable conditions following deforestation at Hubbard Brook. Ecological Monographs 44: 255-277. DOI: https://doi.org/10.2307/2937031 [ Links ]

Brant AN, Chen HYH. 2015. Patterns and mechanisms of nutrient resorption in plants. Critical Reviews in Plant Sciences 34: 471-486. DOI: https://doi.org/10.1080/07352689.2015.1078611 [ Links ]

Brunner FS, Deere JA, Egas M, Eizaguirre C, Raeymaekers JAM. 2019. The diversity of eco-evolutionary dynamics: Comparing the feedbacks between ecology and evolution across scales. Functional Ecology 33: 7-12. DOI: https://doi.org/10.1111/1365-2435.13268 [ Links ]

Carpenter SR, Mooney HA, Agard J, Capistrano D, DeFries RS, Díaz S, Dietz T, Duraiappah AK, Oteng-Yeboah A, Pereira MH, Perrings C, Reid WV, Sarukhan J, Scholes RJ, Whyte A. 2009. Science for managing ecosystem services: Beyond the Millennium Ecosystem Assessment. Proceedings of the National Academy Sciences of the United States of America 106: 1305-1312. DOI: https://doi.org/10.1073/pnas.0808772106 [ Links ]

Carpenter SR, Turner MG. 1998. At last: A journal devoted to ecosystem science. Ecosystems 1: 1-5. DOI: https://www.jstor.org/stable/3658700Links ]

Carpenter SR, Turner MG. 2017. Twenty years of Ecosystems: emerging questions and challenges. Ecosystems 20: 1-3. DOI: https://doi.org/10.1007/s10021-016-0056-1 [ Links ]

Cartaxana P, Catarino F. 2002. Nitrogen resorption from senescing leaves of three salt marsh plant species. Plant Ecology 159: 95-102. DOI: https://doi.org/10.1023/A:1015595430010 [ Links ]

Chapin III FS, Matson PA, Mooney HA. 2002. Principles of Terrestrial Ecosystem Ecology. New York: Springer Verlag . DOI: https://doi.org/10.1007/b97397 [ Links ]

Chapin III FS, Matson PA, Vitousek PM. 2011. Principles of Terrestrial Ecosystem Ecology 2nd ed. New York: Springer. DOI: https://doi.org/10.1007/978-1-4419-9504-9 [ Links ]

Chapin III FS, Woodwell GM, Randerson JT, Rastetter EB, Lovett GM, Baldocchi DD, Clark DA, Harmon ME, Schimel DS, Valentini R, Wirth C, Aber JD, Cole JJ, Goulden ML, Harden JW, Heimann M, Howarth RW, Matson PA, McGuire AD, Melillo JM, Mooney HA, Neff JC, Houghton RA, Pace ML, Ryan MG, Running SW, Sala OE, Schlesinger WH, Schulze E-D. 2006. Reconciling carbon-cycle concepts, terminology, and methods. Ecosystems 9: 1041-1050. DOI: https://doi.org/10.1007/s10021-005-0105-7 [ Links ]

Chavez-Vergara B, Merino A, Vázquez-Marrufo G, García-Oliva F. 2014. Organic matter dynamics and microbial activity during decomposition of forest floor under two native neotropical oak species in a temperate deciduous forest in Mexico. Geoderma 235-236: 133-145. DOI: https://doi.org/10.1016/j.geoderma.2014.07.005 [ Links ]

Chávez-Vergara BM, González-Rodríguez A, Etchevers JD, Oyama K, García-Oliva F. 2015. Foliar nutrient resorption constrains soil nutrient transformations under two native oak species in a temperate deciduous forest in Mexico. European Journal of Forest Research 134: 803-817. DOI: https://doi.org/10.1007/s10342-015-0891-1 [ Links ]

Cleveland CC, Liptzin D. 2007. C:N:P stoichiometry in soil: is there a ‘‘Redfield ratio’’ for the microbial biomass? Biogeochemistry 85: 235-252. DOI: https://doi.org/10.1007/s10533-007-9132-0 [ Links ]

Costanza R, d’Arge R, de Groot R, Farber S, Grasso M, Hannon B, Limburg K, Naeem S, O’Neill RV, Paruelo J, Raskin RG, Sutton P, van den Belt M. 1997. The value of the world’s ecosystem services and natural capital. Nature 387: 253-260. DOI: https://doi.org/10.1038/387253a0 [ Links ]

Covelo F, Durán J, Gallardo A. 2008a. Leaf resorption efficiency and proficiency in a Quercus robur population following forest harvest. Forest Ecology and Management 255: 2264-2271. DOI: https://doi.org/10.1016/j.foreco.2008.01.001 [ Links ]

Covelo F, Rodríguez A, Gallardo A. 2008b. Spatial pattern and scale of leaf N and P resorption efficiency and proficiency in a Quercus robur population. Plant and Soil 311: 109-119. DOI: https://doi.org/10.1007/s11104-008-9662-9 [ Links ]

Culleton H, McKie V, de Vries RP. 2013. Physiological and molecular aspects of degradation of plant polysaccharides by fungi: What have we learned from Aspergillus? Biotechnology Journal 8: 884-894. DOI: https://doi.org/10.1002/biot.201200382 [ Links ]

Currie WS. 2011. Units of nature or process across scales? The ecosystem concept at age 75. New Phytologist 190: 21-34. DOI: https://doi.org/10.1111/j.1469-8137.2011.03646.x [ Links ]

Daily GC, Söderqvist T, Aniyar S, Arrow K, Dasgupta P, Ehrlich PR, Folke C, Jansson A, Jansson B-O, Kautsky N, Levin S, Lubchenco J, Mäler K-G, Simpson D, Starrett D, Tilman D, Walker B. 2000. The value of nature and the nature of value. Science 289: 395-396. DOI: https://doi.org/10.1126/science.289.5478.395 [ Links ]

Del Arco JM, Escudero A, Vega Garrido M. 1991. Effects of site characteristics on nitrogen retranslocation from senescing leaves. Ecology 72: 701-708. DOI: https://doi.org/10.2307/2937209 [ Links ]

Delgado-Balbuena J, Yépez EA, Paz-Pellat F. Ángeles-Pérez G, Alvarado-Barrientos MS, Bullock SH, Castellanos AE, Arredondo T, Figueroa-Espinoza B, Garatuza-Payán J, González-del Castillo E., González-Sosa E, Maya-Delgado Y, Rodríguez JC, Vargas R, Vivoni ER, Watts CJ (eds). 2019. Flujos verticales de carbono en ecosistemas terrestres. In: Paz-Pellat F, Hernández-Ayón JM, Sosa-Ávalos R, Velázquez-Rodríguez AS, eds. Estado del Ciclo del Carbono en México: Agenda Azul y Verde. Texcoco: Programa Mexicano del Carbono. [ Links ]

Delgado-Baquerizo M, Oliverio AM, Brewer TE, Benavent-González A, Eldridge DJ, Bardgett RD, Maestre FT, Singh BK, Fierer N. 2018. A global atlas of the dominant bacteria found in soil. Science 359: 320-325. DOI: https://doi.org/10.1126/science.aap9516 [ Links ]

Dell B, Hopkins AJM, Lamont BB. 1986. Resilience in Mediterranean-Type Ecosystems. Dordrecht: Springer. DOI: https://doi.org/10.1007/978-94-009-4822-8 [ Links ]

Ehleringer JR. 2001. Productivity of deserts. In: Roy J, Saugier B, Mooney HA, eds. Terrestrial Global Productivity. New York: Academic Press, pp. 345-362. DOI: https://doi.org/10.1016/B978-0-12-505290-0.X5000-5 [ Links ]

Ekschmitt K, Liu M, Vetter S, Fox O, Wolters V. 2005. Strategies used by soil biota to overcome soil organic matter stability - why is dead organic matter left over in the soil? Geoderma 128: 167-176. DOI: https://doi.org/10.1016/j.geoderma.2004.12.024 [ Links ]

Fry B. 2006. Stable Isotope Ecology . New York: Springer . DOI: https://doi.org/10.1007/0-387-33745-8 [ Links ]

García-Oliva F, Merino A, Fonturbel MT, Omil B, Fernández C, Vega JA. 2018. Severe wildfire hinders renewal of soil P pools by thermal mineralization of organic P in forest soil: Analysis by sequential extraction and 31P NMR spectroscopy. Geoderma 309: 32-40. DOI: https://doi.org/10.1016/j.geoderma.2017.09.002 [ Links ]

Glanville H, Rousk J, Golyshin P, Jones DL. 2012. Mineralization of low molecular weight carbon substrates in soil solution under laboratory and field conditions. Soil Biology and Biochemistry 48: 88-95. DOI: https://doi.org/10.1016/j.soilbio.2012.01.015 [ Links ]

Good SP, Soderberg K, Wang L, Caylor KK. 2012. Uncertainties in the assessment of the isotopic composition of surface fluxes: A direct comparison of techniques using laser-based water vapor isotope analyzers. Journal of Geophysical Research 117: D15301. https://doi.org/10.1029/2011JD017168 [ Links ]

Griffiths BS, Bonkowski M, Roy J, Ritz K. 2001. Functional stability, substrate utilisation and biological indicators of soils following environmental impacts. Applied Soil Ecology 16: 49-61. DOI: https://doi.org/10.1016/S0929-1393(00)00081-0 [ Links ]

Griffiths BS, Ritz K, Bardgett RD, Cook R, Christensen S, Ekelund F, Sørensen SJ, Bååth E, Bloem J, de Ruiter PC, Dolfing J, Nicolardot B. 2000. Ecosystem response of pasture soil communities to fumigation-induced microbial diversity reductions: an examination of the biodiversity-ecosystem function relationship. Oikos 90: 279-294. DOI: https://doi.org/10.1034/j.1600-0706.2000.900208.x [ Links ]

Griscom BW, Adams J, Ellis PW, Houghton RA, Lomax G, Miteva DA, Schlesinger WH, Shoch D, Siikamäki JV, Smith P, Woodbury P, Zganjar C, Blackman A, Campari J, Conant RT, Delgado C, Elias P, Gopalakrishna T, Hamsik MR, Herrero M, Kiesecker J, Landis E, Laestadius L, Leavitt SM, Minnemeyer S, Polasky S, Potapov P, Putz FE, Sanderman J, Silvius M, Wollenberg E, Fargione J. 2017. Natural climate solutions. Proceedings of the National Academy of Science s of the United States of America 114: 11645-11650. DOI: https://doi.org/10.1073/pnas.1710465114 [ Links ]

Hagen-Thorn A, Varnagiryte I, Nihlgård B, Armolaitis K. 2006. Autumn nutrient resorption and losses in four deciduous forest tree species. Forest Ecology and Management 228: 33-39. DOI: https://doi.org/10.1016/j.foreco.2006.02.021 [ Links ]

Hartmann H, Bahn M, Carbone M, Richardson AD. 2020. Plant carbon allocation in a changing world - challenges and progress: introduction to a Virtual Issue on carbon allocation. New Phytologist 227: 981-988. DOI: https://doi.org/10.1111/nph.16757 [ Links ]

Heiskanen J, Brümmer C, Buchmann N, Calfapietra C, Chen H, Gielen B, Gkritzalis T, Hammer S, Hartman S, Herbst M, Janssens IA, Jordan A, Juurola E, Karstens U, Kasurinen V, Kruijt B, Lankreijer H, Levin I, Linderson M-L, Loustau D, Merbold L, Myhre CL, Papale D, Pavelka M, Pilegaard K, Ramonet M, Rebmann C, Rinne J, Rivier L, Saltikoff E, Sanders R, Steinbacher M, Steinhoff T, Watson A, Vermeulen AT, Vesala T, Vítková G, Kutsch W. 2022. The integrated carbon observation system in Europe. Bulletin of the American Meteorological Society 103: E855-E872. DOI: https://doi.org/10.1175/BAMS-D-19-0364.1 [ Links ]

Hemes KS, Runkle BRK, Novick KA, Baldocchi DD, Field CB. 2021. An ecosystem-scale flux measurement strategy to assess natural climate solutions. Environmental Science & Technology 55: 3494-3504. DOI: https://doi.org/10.1021/acs.est.0c06421 [ Links ]

Herbert DA, Fownes JH, Vitousek PM. 1999. Hurricane damage to a Hawaiian forest: Nutrient supply rate affects resistance and resilience. Ecology 80: 908-920. DOI: https://doi.org/10.1890/0012-9658(1999)080[0908:HDTAHF]2.0.CO;2 [ Links ]

Hernández-Becerra N, Tapia-Torres Y, Beltrán-Paz O, Blaz J, Souza V, García-Oliva F. 2016. Agricultural land-use change in a Mexican oligotrophic desert depletes ecosystem stability. PeerJ 4: e2365. DOI: https://doi.org/10.7717/peerj.2365 [ Links ]

Hoffmann U, Yair A, Hikel H, Kuhn NJ. 2012. Soil organic carbon in the rocky desert of northern Negev (Israel). Journal of Soils and Sediments 12: 811-825, DOI: https://doi.org/10.1007/s11368-012-0499-8 [ Links ]

Holling CS. 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics 4: 1-23. DOI: https://doi.org/10.1146/annurev.es.04.110173.000245 [ Links ]

Huang J, Wang X, Yan E. 2007. Leaf nutrient concentration, nutrient resorption and litter decomposition in an evergreen broad-leaved forest in eastern China. Forest Ecology and Management 239: 150-158. https://doi.org/10.1016/j.foreco.2006.11.019 [ Links ]

Huxman TE, Smith MD, Fay PA, Knapp AK, Shaw MR, Loik ME, Smith SD, Tissue DT, Zak JC, Weltzin JF, Pockman WT, Sala OE, Haddad BM, Harte J, Koch GW, Schwinning S, Small EE, Williams DG. 2004. Convergence across biomes to a common rain-use efficiency. Nature 429: 651-654. DOI: https://doi.org/10.1038/nature02561 [ Links ]

IPCC [ Intergovernmental Panel on Climate Change] . 2013. The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. In: Stocker TF, Qin D, Plattner GK, Tignor M, Allen SK, Boschung J, Nauels A, Xia Y, Bex V, Midgley PM, eds. Cambridge and New York: Cambridge University Press. [ Links ]

IPCC [ Intergovernmental Panel on Climate Change] . 2021. Summary for Policymakers. In: Masson-Delmotte V, Zhai P, Pirani A, Connors SL, Péan C, Berger S, Caud N, Chen Y, Goldfarb L, Gomis MI, Huang M, Leitzell K, Lonnoy E, Matthews JBR, Maycock TK, Waterfield T, Yelekçi O, Yu R, Zhou B, eds. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press. [ Links ]

Jones DL, Willett VB, Stockdale EA, Macdonald AJ, Murphy DV. 2012. Molecular weight of dissolved organic carbon, nitrogen, and phenolics in grassland soils. Soil Science Society of America Journal 76: 142-150. DOI: https://doi.org/10.2136/sssaj2011.0252 [ Links ]

Jones ML. 2004. Changes in gene expression during senescence. In: Noodén L, ed. Plant Cell Death Processes. San Diego: Academic Press, pp 51-71. DOI: https://doi.org/10.1016/B978-012520915-1/50007-2 [ Links ]

Kang M, Cho S. 2021. Progress in water and energy flux studies in Asia: A review focused on eddy covariance measurements. Journal of Agricultural Meteorology 77: 2-23. DOI: http://dx.doi.org/10.2480/agrmet.D-20-00036 [ Links ]

Keenan TF, Williams CA. 2018. The terrestrial carbon sink. Annual Review of Environment and Resources 43: 219-243. DOI: https://doi.org/10.1146/annurev-environ-102017-030204 [ Links ]

Kiikkilä O, Kitunen V, Smolander A. 2005. Degradability of dissolved soil organic carbon and nitrogen in relation to tree species. FEMS Microbiology Ecology 53: 33-40. DOI: https://doi.org/10.1016/j.femsec.2004.08.011 [ Links ]

Kiikkilä O, Kitunen V, Smolander A. 2011. Properties of dissolved organic matter derived from silver birch and Norway spruce stands: Degradability combined with chemical characteristics. Soil Biology and Biochemistry 43: 421-430. DOI: https://doi.org/10.1016/j.soilbio.2010.11.011 [ Links ]

Killingbeck KT. 1996. Nutrients in senesced leaves: Keys to the search for potential resorption and resorption proficiency. Ecology 77: 1716-1727. DOI: https://doi.org/10.2307/2265777 [ Links ]

Killingbeck KT, Costigan SA. 1988. Element resorption in a guild of understory shrub species: niche differentiation and resorption thresholds. Oikos 53: 366-374. DOI: https://doi.org/10.2307/3565537 [ Links ]

Kögel-Knabner I. 2002. The macromolecular organic composition of plant and microbial residues as inputs to soil organic matter. Soil Biology and Biochemistry 34: 139-162. DOI: https://doi.org/10.1016/S0038-0717(01)00158-4 [ Links ]

Kool D, Agam N, Lazarovitch N, Heitman JL, Sauer TJ, Ben-Gal A. 2014. A review of approaches for evapotranspiration partitioning. Agricultural and Forest Meteorology 184: 56-70. DOI: https://doi.org/10.1016/j.agrformet.2013.09.003 [ Links ]

Lauenroth WK, Sala OE. 1992 Long-term forage production of North American short grass steppe. Ecological Applications 2: 397-403. DOI: http://dx.doi.org/10.2307/1941874 [ Links ]

Le Houérou HN. 1984. Rain use-efficiency: a unifying concept in arid-land ecology. Journal of Arid Environments 7: 213-247. https://doi.org/10.1016/S0140-1963(18)31362-4 [ Links ]

Leith H. 1975. Modeling the primary productivity of the world. In: Leith H, Whittaker RH, eds. Primary Productivity of the Biosphere. Berlin, Heidelberg: Springer-Verlag. DOI: https://doi.org/10.1007/978-3-642-80913-2_12 [ Links ]

Leopold AC. 1961. Senescence in plant development. Science 3492: 1727-1732. DOI: https://doi.org/10.1126/science.134.3492.1727 [ Links ]

Levin SA. 1992. The problem of pattern and scale in ecology: The Robert H. MacArthur award lecture. Ecology 73: 1943-1967. DOI: https://doi.org/10.2307/1941447 [ Links ]

Lindeman RL. 1942. The trophic-dynamic aspect of ecology. Ecology 23: 399-417. DOI: https://doi.org/10.2307/1930126 [ Links ]

López DR, Cavallero L, Brizuela MA, Aguiar MR. 2011. Ecosystemic structural-functional approach of the state and transition model. Applied Vegetation Science 14: 6-16. DOI: https://doi.org/10.1111/j.1654-109X.2010.01095.x [ Links ]

MacGillivray CW, Grime JP, Band SR, Booth RE, Campbell B, Hendry GAF, Hillier SH, Hodgson JG, Hunt R, Jalili A, Mackey JML, Mowforth MA, Neal AM, Reader R, Rorison, IH, Spencer RE, Thompson K, Thorpe PC. 1995. Testing predictions of the resistance and resilience of vegetation subjected to extreme events. Functional Ecology 9: 640-649. DOI: https://doi.org/10.2307/2390156 [ Links ]

Malik AA, Puissant J, Buckeridge KM, Goodall T, Jehmlich N, Chowdhury S, Gweon HS, Peyton JM, Mason KE, van Agtmaal M, Blaud A, Clark IM, Whitaker J, Pywell RF, Ostle N, Gleixner G, Griffiths RI. 2018. Land use driven change in soil pH affects microbial carbon cycling processes. Nature Communications 9: 3591. DOI: https://doi.org/10.1038/s41467-018-05980-1 [ Links ]

Margalef R. 1968. Perspectives in Ecological Theory. Chicago: University of Chicago Press. DOI: https://doi.org/10.4319/lo.1969.14.2.0313 [ Links ]

Martinez-Yrizar A, Maass JM., Perez-Jimenez LA, Sarukhan, J. 1996. Net primary productivity of a tropical deciduous forest ecosystem in western Mexico. Journal of Tropical Ecology 12: 169-175. DOI: https://doi.org/10.1017/S026646740000938X [ Links ]

Merino A, Chávez-Vergara B, Salgado J, Fonturbel MT, García-Oliva F, Vega JA. 2015. Variability in the composition of charred litter generated by wildfire in different ecosystems. Catena 133: 52-63. DOI: https://doi.org/10.1016/j.catena.2015.04.016 [ Links ]

Merino A, García-Oliva F, Fontúrbel MT, Vega JA. 2021. The high content of mineral-free organic matter in soils increases their vulnerability to wildfire in humid-temperate zones. Geoderma 395: 115043. DOI: https://doi.org/10.1016/j.geoderma.2021.115043 [ Links ]

Millennium Ecosystem Assessment. 2003. Ecosystem and Human Well-being: A Framework for Assessment. Washington DC: Island Press. ISBN: 1-55963-402-2 [ Links ]

Millennium Ecosystem Assessment. 2005. Ecosystems and Human Well-being: Synthesis. Washington DC: Island Press . ISBN: 1-59726-040-1 [ Links ]

Montiel-González C, García-Oliva F, Bautista F, Sánchez-Meneses O. 2021. The global climate change footprint in a Mexican desert ecosystem: The increasing frequency of extreme climatic events. Tecnología y Ciencias del Agua 12: 219-268. DOI: https://doi.org/10.24850/j-tyca-2021-04-05 [ Links ]

Mooney HA. 1972. The carbon balance of plants. Annual Review of Ecology and Systematics 3: 315-346. DOI: https://doi.org/10.1146/annurev.es.03.110172.001531 [ Links ]

Neff JC, Asner GP. 2001. Dissolved organic carbon in terrestrial ecosystems: synthesis and a model. Ecosystems 4: 29-48. DOI: https://doi.org/10.1007/s100210000058 [ Links ]

Newman BD, Wilcox BP, Archer SR, Breshears DD, Dahm CN, Duffy CJ, McDowell NG, Phillips FM, Scanlon BR, Vivoni ER. 2006. Ecohydrology of water-limited environments: A scientific vision. Water Resources Research 42: 1-15. DOI: https://doi.org/10.1029/2005WR004141 [ Links ]

Noack SR, McLaughlin MJ, Smernik RJ, McBeath TM, Armstrong RD. 2012. Crop residue phosphorus: speciation and potential bio-availability. Plant and Soil 359: 375-385. DOI: https://doi.org/10.1007/s11104-012-1216-5 [ Links ]

Nobel PS. 1999. Physicochemical and Environmental Plant Physiology. London: Academic Press. ISBN: 0-12-520025-0 [ Links ]

Noy-Meir I. 1985. Desert ecosystem structure and function. In: Evenari M, Noy-Meir I, Goodall DW, eds. Ecosystems of the World 12A: Hot Deserts and Arid Shrublands. Amsterdam: Elsevier, pp. 93-103. ISBN: 978-0444422972 [ Links ]

Odum EP. 1969. The strategy of ecosystem development. Science 164: 262-270. DOI: https://doi.org/10.1126/science.164.3877.262 [ Links ]

O’Neill RV, DL DeAngelis, JBWaide, TFH Allen. 1986. A Hierarchical Concept of Ecosystems. New Jersey: Princeton University Press. ISBN: 0-691-08437-8 [ Links ]

Orwin KH, Wardle DA. 2004. New indices for quantifying the resistance and resilience of soil biota to exogenous disturbances. Soil Biology and Biochemistry 36: 1907-1912. DOI: https://doi.org/10.1016/j.soilbio.2004.04.036 [ Links ]

Paruelo JM, Lauenroth WK, Burke IC, Sala OE. 1999. Grassland precipitation-use efficiency varies across a resource gradient. Ecosystems 2: 64-68. DOI: https://doi.org/10.1007/s100219900058 [ Links ]

Pearcy RW, Schulze E-D, Zimmerman R. 1989 Measurement of transpiration and leaf conductance. In: Pearcy RW, Ehleringer JR, Mooney HA, Rundel PW, eds. Plant PhysiologicalEcology : Field Methods and Instrumentation. London: Chapman and Hall. pp. 137-160. DOI: https://doi.org/10.1007/978-94-009-2221-1_8 [ Links ]

Peñuelas J, Sardans J, Estiarte M, Ogaya R, Carnicer J, Coll M, Barbeta A, Rivas-Ubach A, Llusià J, Garbulsky M, Filella I, Jump AS. 2013. Evidence of current impact of climate change on life: a walk from genes to the biosphere. Global Change Biology 19: 2303-2338. DOI: https://doi.org/10.1111/gcb.12143 [ Links ]

Peñuelas J, Fernández-Martínez M, Ciais P, Jou D, Piao S, Obersteiner M, Vicca S, Janssens IA, Sardans J. 2019. The bioelements, the elementome, and biogeochemical niche. Ecology 100: e02652. DOI: https://doi.org/10.1002/ecy.2652 [ Links ]

Perroni Y, García-Oliva F, Tapia-Torres Y, Souza V. 2014. Relationship between soil P fractions and microbial biomass in an oligotrophic grassland-desert scrub system. Ecological Research 29: 463-472. DOI: https://doi.org/10.1007/s11284-014-1138-1 [ Links ]

Peterson G, Allen CR, Holling CS. 1998. Ecological resilience, biodiversity, and scale. Ecosystems 1: 6-18. DOI: https://doi.org/10.1007/s100219900002 [ Links ]

Pickett STA, Cadenasso ML 2002. The ecosystem as a multidimensional concept: meaning, model and metaphor. Ecosystems 5: 1-10. DOI: https://doi.org/10.1007/s10021-001-0051-y [ Links ]

Pimm SL. 1984. The complexity and stability of ecosystems. Nature 307: 321-326. DOI: https://doi.org/10.1038/307321a0 [ Links ]

Plaut JA, Yepez EA, Hill J, Pangle R, Sperry JS, Pockman WT, McDowell NG. 2012. Hydraulic limits preceding mortality in a piñon-juniper woodland under experimental drought. Plant, Cell & Environment 35: 1601-1617. DOI: https://doi.org/10.1111/j.1365-3040.2012.02512.x [ Links ]

Porporato A, D’Odorico P, Laio F, Rodriguez-Iturbe I. 2003. Hydrologic controls on soil carbon and nitrogen cycles. I. Modeling scheme. Advances in Water Resources 26: 45-58. DOI: https://doi.org/10.1016/S0309-1708(02)00094-5 [ Links ]

Rapp M, Santa Regina I, Rico M, Gallego HA. 1999. Biomass, nutrient content, litterfall and nutrient return to the soil in Mediterranean oak forests. Forest Ecology and Management 119: 39-49. DOI: https://doi.org/10.1016/S0378-1127(98)00508-8 [ Links ]

Reichstein M, Falge E, Baldocchi D, Papale D, Aubinet M, Berbigier P, Bernhofer C, Buchmann N, Gilmanov T, Granier A, Grünwald T, Havránková K, Ilvesniemi H, Janous D, Knohl A, Laurila T, Lohila A, Loustau D, Matteucci G, Meyers T, Miglietta F, Ourcival JM, Pumpanen J, Rambal S, Rotenberg E, Sanz M, Tenhunen J, Seufert G, Vaccari F, Vesala T, Yakir D, Valenteni R. 2005. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: review and improved algorithm. Global Change Biology 11: 1424-1439. DOI: https://doi.org/10.1111/j.1365-2486.2005.001002.x [ Links ]

Rentería LY, Jaramillo VJ, Martínez-Yrízar A, Pérez-Jiménez A. 2005. Nitrogen and phosphorus resorption in trees of a Mexican tropical dry forest. Trees 19: 431-441. DOI: https://doi.org/10.1007/s00468-004-0402-3 [ Links ]

Roberts P, Stockdale R, Khalid M, Iqbal Z, Jones DL. 2009. Carbon-to-nitrogen ratio is a poor predictor of low molecular weight organic nitrogen mineralization in soil. Soil Biology and Biochemistry 41: 1750-1752. DOI: https://doi.org/10.1016/j.soilbio.2009.04.020 [ Links ]

Rodríguez‐Robles U, Arredondo JT, Huber‐Sannwald E, Yépez EA, Ramos‐Leal JA. 2020. Coupled plant traits adapted to wetting/drying cycles of substrates co‐define niche multidimensionality. Plant, Cell & Environment 43: 2394-2408. DOI: https://doi.org/10.1111/pce.13837 [ Links ]

Santa Regina I, Tarazona T. 2001. Nutrient cycling in a natural beech forest and adjacent planted pine in northern Spain. Forestry 74: 11-28. DOI: https://doi.org/10.1093/forestry/74.1.11 [ Links ]

Sardans J, Peñuelas J, Rivas-Ubach A. 2011. Ecological metabolomics: overview of current developments and future challenges. Chemoecology 21: 191-225. DOI: http://dx.doi.org/10.1007/s00049-011-0083-5 [ Links ]

Scanlon TM, Kustas WP. 2010. Partitioning carbon dioxide and water vapor fluxes using correlation analysis. Agricultural and Forest Meteorology 150: 89-99. DOI: https://doi.org/10.1016/j.agrformet.2009.09.005 [ Links ]

Schimel D, Schneider FD, JPL Carbon Ecosystem Participants 2019. Flux towers in the sky: global ecology from space. New Phytologist 224: 570-584. DOI: https://doi.org/10.1111/nph.15934 [ Links ]

Schlesinger WH, Bernhardt ES. 2020. Biogeochemistry : An Analysis of Global Change. 4th ed, San Diego: Academic Press . ISBN: 978-0128146088 [ Links ]

Schlesinger WH, Cole JJ, Finzi AC, Holland EA. 2011. Introduction to coupled biogeochemical cycles. Frontiers in Ecology and the Environment 9: 5-8. DOI: https://doi.org/10.1890/090235 [ Links ]

Shuttleworth WJ. 2007. Putting the “vap” into evaporation. Hydrology and Earth System Sciences 11: 210-244. DOI: https://doi.org/10.5194/hess-11-210-2007 [ Links ]

Silla F, Escudero A. 2003. Uptake, demand and internal cycling of nitrogen in saplings of Mediterranean Quercus species. Oecologia 136: 28-36. DOI: https://doi.org/10.1007/s00442-003-1232-5 [ Links ]

Sinsabaugh RS. 1994. Enzymic analysis of microbial pattern and process. Biology and Fertility of Soils 17: 69-74. DOI: https://doi.org/10.1007/BF00418675 [ Links ]

Sinsabaugh RL, Hill BH, Follstad Shah JJ. 2009. Ecoenzymatic stoichiometry of microbial organic nutrient acquisition in soil and sediment. Nature 462: 795-798. DOI: https://doi.org/10.1038/nature08632 [ Links ]

Šnajdr J, Cajthaml T, Valášková V, Merhautová V, Petránková M, Spetz P, Leppänen K, Baldrian P. 2011. Transformation of Quercus petraea litter: successive changes in litter chemistry are reflected in differential enzyme activity and changes in the microbial community composition. FEMS Microbiology Ecology 75: 291-303. DOI: https://doi.org/10.1111/j.1574-6941.2010.00999.x [ Links ]

Sparks D. 2019. Fundamentals of soil chemistry. In: Maurice P, ed. Encyclopedia of Water: Science, Technology, and Society. Hoboken: John Wiley & Sons DOI: https://doi.org/10.1002/9781119300762.wsts0025 [ Links ]

Starke R, López Mondéjar R, Human ZR, Navrátilová D, Štursová M, Větrovský T, Olson HM, Orton DJ, Callister SJ, Lipton MS, Howe A, McCue LA, Pennacchio C, Grigoriev I, Baldrian P. 2021. Niche differentiation of bacteria and fungi in carbon and nitrogen cycling of different habitats in a temperate coniferous forest: A metaproteomic approach. Soil Biology and Biochemistry 155: 108170. DOI: https://doi.org/10.1016/j.soilbio.2021.108170 [ Links ]

Sterner RW, Elser JJ. 2002. Ecological Stoichiometry: The Biology of Elements From Molecules to the Biosphere. Princeton: Princeton University Press. ISBN: 978-0691074917 [ Links ]

Stoy PC, El-Madany TS, Fisher JB, Gentine P, Gerken T, Good SP, Klosterhalfen A, Liu S, Miralles DG, Perez-Priego O, Rigden AJ, Skaggs TH, Wohlfahrt G, Anderson RG, Coenders-Gerrits AMJ, Jung M, Maes WH, Mammarella I, Mauder M, Migliavacca M, Nelson JA, Poyatos R, Reichstein M, Scott RL, Wolf S. 2019. Reviews and syntheses: Turning the challenges of partitioning ecosystem evaporation and transpiration into opportunities. Biogeosciences 16: 3747-3775. DOI: https://doi.org/10.5194/bg-16-3747-2019 [ Links ]

Sun G, Hallema D, Asbjornsen H. 2017. Ecohydrological processes and ecosystem services in the Anthropocene: a review. Ecological Processes 6: 35 (2017). DOI: https://doi.org/10.1186/s13717-017-0104-6 [ Links ]

Swank WT, Croosley DA. 1988. Forest Hydrology and Ecology at Coweeta. New York: Springer -Verlag. DOI: https://doi.org/10.1007/978-1-4612-3732-7 [ Links ]

Tague CL, Papuga SA, Gerlein-Safdi C, Dymond S, Morrison RR, Boyer EW, Riveros-Iregui D. Agee E, Arora B, Dialynas YG, Hansen A, Krause S, Kuppel S, Loheide II SP, Schymanski SJ, Zipper SC. 2020. Adding our leaves: A community-wide perspective on research directions in ecohydrology. Hydrological Processes 34: 1665-1673. DOI: https://doi.org/10.1002/hyp.13693 [ Links ]

Tang X, Li H, Desai AR, Nagy Z, Luo J, Kolb TE, Olioso A, Xu X, Yao L, Kutsch W, Pilegaard K, Köstner B, Ammann C. 2014. How is water-use efficiency of terrestrial ecosystems distributed and changing on Earth? Scientific Reports 4: 7483. DOI: https://doi.org/10.1038/srep07483 [ Links ]

Tansley AG. 1935. The use and abuse of vegetational concepts and terms. Ecology 16: 284-307. DOI: https://doi.org/10.2307/1930070 [ Links ]

Tapia-Torres Y, Chávez Ortiz P, Hernández-Becerra N, Morón Cruz A, Beltrán O, García-Oliva F. 2018. How do agricultural practices modify soil nutrients dynamics in CCB? In: García-Oliva F, Elser J, Souza V, eds. Ecosystem Ecology and Geochemistry of Cuatro Cienegas: How to Survive in an Extremely Oligotrophic Site. Cham: Springer, pp. 153-161. DOI: https://doi.org/10.1007/978-3-319-95855-2_12 [ Links ]

Tapia-Torres Y, Elser JJ, Souza V, García-Oliva F. 2015a. Ecoenzymatic stoichiometry at the extremes: How microbes cope in an ultra-oligotrophic desert soil. Soil Biology and Biochemistry 87: 34-42. DOI: https://doi.org/10.1016/j.soilbio.2015.04.007 [ Links ]

Tapia-Torres Y, López-Lozano NE, Souza V, García-Oliva F. 2015b. Vegetation-soil system controls soil mechanisms for nitrogen transformations in an oligotrophic Mexican desert. Journal of Arid Environments 114: 62-69. DOI: https://doi.org/10.1016/j.jaridenv.2014.11.007 [ Links ]

Tarin T, Nolan RH, Eamus D, Cleverly J. 2020. Carbon and water fluxes in two adjacent Australian semi-arid ecosystems. Agricultural and Forest Meteorology 281: 107853. DOI: https://doi.org/10.1016/j.agrformet.2019.107853 [ Links ]

Tilman D. 1996. Biodiversity: population versus ecosystem stability. Ecology 77: 350-363. DOI: https://doi.org/10.2307/2265614 [ Links ]

Turner BL, Condron LM, Richardson SJ, Peltzer DA, Allison VJ. 2007. Soil organic phosphorus transformations during pedogenesis. Ecosystems 10: 116-1181. DOI: https://doi.org/10.1007/s10021-007-9086-z [ Links ]

Turner BL, Mahieu N, Condron LM. 2003. Phosphorus-31 nuclear magnetic resonance spectral assignments of phosphorus compounds in soil NaOH-EDTA extracts. Soil Science Society of America Journal 67: 497-510. DOI: https://doi.org/10.2136/sssaj2003.4970 [ Links ]

Turner MG. 1989. Landscape ecology: the effect of pattern on process. Annual Review of Ecology and Systematics 20: 171-197. DOI: https://doi.org/10.1146/annurev.es.20.110189.001131 [ Links ]

van Hees PAW, Jones DL, Finlay R, Godbold DL, Lundström US. 2005. The carbon we do not see-the impact of low molecular weight compounds on carbon dynamics and respiration in forest soils: a review. Soil Biology and Biochemistry 37: 1-13. DOI: https://doi.org/10.1016/j.soilbio.2004.06.010 [ Links ]

Vargas R, Loescher HW, Arredondo T, Huber-Sannwald E, Lara-Lara R, Yépez EA. 2012. Opportunities for advancing carbon cycle science in Mexico: towards a continental scale understanding. Environmental Science & Policy 21: 84-93. DOI: https://doi.org/10.1016/j.envsci.2012.04.003 [ Links ]

Vergutz L, Manzoni S, Porporato A, Novais RF, Jackson RB. 2012. Global resorption efficiencies and concentrations of carbon and nutrients in leaves of terrestrial plants. Ecological Monographs 82: 205-220. DOI: https://doi.org/10.1890/11-0416.1 [ Links ]

Vitousek PM. 1984. Litterfall, nutrient cycling, and nutrient limitation in tropical forest. Ecology 65: 285-298. DOI: https://doi.org/10.2307/1939481 [ Links ]

Vitousek PM, Sanford RL. 1986. Nutrient cycling in moist tropical forest. Annual Review of Ecology and Systematics 17: 137-167. DOI: https://doi.org/10.1146/annurev.es.17.110186.001033 [ Links ]

von Lützow M, Kögel-Knabner I, Ekschmitt K, Flessa H, Guggenberger G, Matzner E, Marschner B. 2007. SOM fractionation methods: Relevance to functional pools and to stabilization mechanisms. Soil Biology and Biochemistry 39: 2183-2207. DOI: https://doi.org/10.1016/j.soilbio.2007.03.007 [ Links ]

Voříšková J, Dobiášová P, Šnajdr J, Vaněk D, Cajthaml T, Šantrůčková H, Baldrian P. 2011. Chemical composition of litter affects the growth and enzyme production by the saprotrophic basidiomycete Hypholoma fasciculare. Fungal Ecology 4: 417-426. DOI: https://doi.org/10.1016/j.funeco.2011.03.005 [ Links ]

Wardle DA, Bonner KI, Barker GM. 2000. Stability of ecosystem properties in response to above-ground functional group richness and composition. Oikos 89: 11-23. DOI: https://doi.org/10.1034/j.1600-0706.2000.890102.x [ Links ]

Whitham TG, Bailey JK, Schweitzer JA, Shuster SM, Bangert RK, LeRoy CJ, Lonsdorf EV, Allan GJ, DiFazio SP, Potts BM, Fischer DG, Gehring CA, Lindroth RL, Marks JC, Hart SC, Wimp GM, Wooley SC. 2006. A framework for community and ecosystem genetics: from genes to ecosystems. Nature Reviews Genetics 7: 510-523. DOI: https://doi.org/10.1038/nrg1877 [ Links ]

Whitham TG, Gehring CA, Lamit LJ, Wojtowicz T, Evans LM, Keith AR, Smith DS. 2012. Community specificity: life and afterlife effects of genes. Trends in Plant Science 17: 271-281. DOI: https://doi.org/10.1016/j.tplants.2012.01.005 [ Links ]

Wilson KB, Hanson PJ, Mulholland PJ, Baldocchi DD, Wullschleger SD. 2001. A comparison of methods for determining forest evapotranspiration and its components: sap-flow, soil water budget, eddy covariance and catchment water balance. Agricultural and Forest Meteorology 106: 153-158. DOI: https://doi.org/10.1016/S0168-1923(00)00199-4 [ Links ]

Yakir D, Sternberg LSL. 2000. The use of stable isotopes to study ecosystem gas exchange. Oecologia 123: 297-311. DOI: https://doi.org/10.1007/s004420051016 [ Links ]

Yepez EA, Scott RL, Cable WL, Williams DG. 2007. Intraseasonal variation in water and carbon dioxide flux components in a semiarid riparian woodland. Ecosystems 10: 1100-1115. DOI: https://doi.org/10.1007/s10021-007-9079-y [ Links ]

Yu G, Piao S, Zhang Y, Liu L, Peng J, Niu S. 202. Moving toward a new era of ecosystem science. Geography and Sustainability 2: 151-162. DOI: https://doi.org/10.1016/j.geosus.2021.06.004 [ Links ]

Zentgraf U, Jobst J, Kolb D, Rentsch D. 2004. Senescence-related gene expression profiles of rosette leaves of Arabidopsis thaliana: leaf age versus plant age. Plant Biology 6: 178-183. DOI: https://doi.org/10.1055/s-2004-815735 [ Links ]

Zhao G, Shi P, Wu J, Xiong D, Zong N, Zhang X. 2017. Foliar nutrient resorption patterns of four functional plants along a precipitation gradient on the Tibetan Changtang Plateau. Ecology and Evolution 7: 7201-7212. DOI: https://doi.org/10.1002/ece3.3283 [ Links ]

Zhou S, Yu B, Zhang Y, Huang Y, Wang G. 2016. Partitioning evapotranspiration based on the concept of underlying water use efficiency. Water Resources Research 52: 1160-1175. DOI: https://doi.org/10.1002/2015wr017766 [ Links ]

Received: November 25, 2021; Accepted: June 13, 2022; Published: August 22, 2022

Author for correspondence: fgarcia@cieco.unam.mx

Guest Editors: Ken Oyama, Jorge A. Meave, Victoria Sosa, Arturo de Nova.

Author contributions: BCV, conceptualization and writing; EAY conceptualization and writing; FGO, conceptualization and writing.

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License