SciELO - Scientific Electronic Library Online

 
vol.11 suppl.2Estimación del factor de transporte del índice de fósforo con climatologías y escenarios de cambio climático en tierras de Jalisco, MéxicoMitigación y adaptación al cambio climático mediante la implementación de modelos integrados para el manejo y aprovechamiento de los residuos pecuarios. Revisión 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


Revista mexicana de ciencias pecuarias

versión On-line ISSN 2448-6698versión impresa ISSN 2007-1124

Rev. mex. de cienc. pecuarias vol.11  supl.2 Mérida mar. 2020  Epub 30-Jun-2020

https://doi.org/10.22319/rmcp.v11s2.4705 

Articles

Impact of climate change on the potential distribution of Tithonia diversifolia (Hemsl.) A. Gray in Mexico

Noé Durán Pugaa 

José Lenin Loya Olguínb 

José Ariel Ruiz Corrala  * 

Diego Raymundo González Eguiartea 

Juan Diego García Paredesb 

Sergio Martínez Gonzálezb 

Marcos Rafael Crespo Gonzáleza 

a Universidad de Guadalajara. Centro Universitario de Ciencias Biológicas y Agropecuarias. Camino Ramón Padilla Sánchez No. 2100 Nextipac, 44600, Zapopan, Jalisco, México.

b Universidad Autónoma de Nayarit. Unidad Académica de Agricultura. Carretera Tepic-Compostela, km 9. Xalisco, Nayarit, México.


Abstract

The aim of this study was to estimate the possible impact of future climate changes on the potential distribution of T. diversifolia in Mexico. Distribution niches were modelled with MaxEnt for the 1951-2000, 2041-2060 and 2061-2080 climatologies, considering 20 bioclimatic and two topographical variables. For future climates, a HadGEM2-ES general circulation model (GCM) was considered, with two representative concentration pathways of greenhouse gases (RCP4.5 and RCP8.5). This information was obtained from the Global Climate Data Website WorldClim and processed with the Idrisi Selva system as raster images with 2.5 arc min resolution. The environmental variables that contributed the most to explain the geographical distribution of T. diversifolia were the May-October mean accumulated precipitation (pa5-10) and the mean maximum temperature of the warmest month (MMAX). The 10th percentile training presence logistic threshold reported a predicted suitable area (for the reference climatology) accounting for 30.71 % of the extent of Mexico. Niche modeling under climate change scenarios reported expansion as well as retraction areas for environmental suitability; however, after computing them, suitable areas are expected to present a small increment with respect to the climate reference period, 1950-2000: 31.62 %, 31.83 %, 32.45 % and 32.45 % of Mexican territory in scenarios for 2041-2060 in RCP 8.5, 2041-2060 in RCP 4.5, 2061-2080 in RCP 4.5 and 2061-2080 in RCP 8.5, respectively. Thus, climate change would bring more benefits than constrains for T. diversifolia dispersion.

Key words Climate change; Tithonia diversifolia; Ecological descriptors; Niche distribution

Resumen

El objetivo fue estimar el posible impacto del cambio climático en la distribución potencial de T. diversifolia en México. Nichos de distribución fueron modelados con MaxEnt para la climatología de referencia 1951-2000, y dos escenarios de cambio climático 2041-2060 y 2061-2080, considerando 20 variables bioclimáticas y dos topográficas. Para las climatologías futuras se consideró el modelo de circulación global (MCG) HadGEM2-ES, con dos vías representativas de concentración de gases efecto invernadero (rcp4.5 y rcp8.5). Esta información se obtuvo de la página web Global Climate Data de WorldClim y se trabajó con el sistema Idrisi Selva como imágenes raster con una resolución de 2.5 min de arco. Las variables que más contribuyeron a explicar la distribución geográfica de T. diversifolia fueron la precipitación acumulada mayo-octubre (pa5-10) y la temperatura máxima media del mes más cálido (txmc). El umbral del percentil 10th de la presencia logística predice una superficie (para la climatología de referencia 1951-2000), que representa el 30.71 % de la extensión de México. La modelación de nichos en los diversos escenarios climáticos registra zonas de expansión, así como zonas de retracción para la distribución potencial de la especie, pero después de hacer un balance, se obtiene que las áreas potenciales para T. diversifolia se incrementarán ligeramente con respecto a la climatología de referencia 1950-2000, con valores estimados de 31.62 %, 31.83 %, 32.45 % y 32.45 % del territorio mexicano en los escenarios 2041-2060 rcp8.5, 2041-2060 rcp4.5, 2061-2080 y 2061-2080 rcp4.5 y rcp8.5, respectivamente. Se concluye que el cambio climático traería más beneficios que restricciones para la dispersión de T. diversifolia.

Palabras clave Cambio climático; Tithonia diversifolia; Descriptores ecológicos; Nichos de distribución

Introduction

Food production is becoming a real challenge in a climate that is changing. Under this context, the diversification of animal feed is a key aspect to better adapt to climate change. Hence, the need to assess the impact of climate change on the presence and potential distribution of forage species. This is the specific case of T. diversifolia, a native species with current distribution in the lowlands of southeastern Mexico, Central and South America, and which is considered an important genetic resource and an exceptional plant as an alternative in animal feed(1,2 ). The future risk for many plant species due to climate change has not been well established yet.

The global temperature has increased 0.85 ° C in the period from 1901 to 20113. Based on climate prediction models, it has been established that the saturation pressure of the vapor has been very sensitive to temperature changes; therefore, ruptures in the global water cycle are expected to manifest in the future4. On the other hand, the availability of water will reportedly exhibit a marked annual reduction in the southeastern United States, the Caribbean and various parts of Mexico5. In addition, agriculture expands and is intensified mainly in the tropics6; it is estimated that approximately 80% of the new farmland will be based on forest substitution7. Therefore, delaying or restraining the expansion of agriculture in the tropics will reduce carbon emissions and loss of biodiversity8.

Climate change can modify the diversity of climates, as well as the composition of ecosystems9; these modifications would include alterations in the distribution, phenology and an increased risk of endangered species10. Several components of climate change are considered to affect all levels of biodiversity, from organisms and populations to biotic areas11. At basic levels, climate change can reduce genetic diversity in populations due to directional selection and dynamic migration, which may interfere with the resilience and functionality of ecosystems12; therefore, it may cause the modification of the network of interactions at the community level13. In addition, an effect of climate change may induce the invasion of potentially dangerous species to an ecosystem14; affecting the physiology, phenology and behavior of species15.

In the face of the above, Mexico is interested in building an inventory related to the expected effects of climate change on desirable wild species. Thus, the aim of this study was to estimate the impact of climate change on the potential distribution of T. diversifolia.

As reported by several authors, the use of Tithonia include forage production, soil erosion control, building material and bird shelter16. As a counterpart, T. diversifolia is considered an allelopathic herb with water-soluble allelochemicals in parts of the plant and with such phytotoxic potency, that it could suppress the growth and accumulation of nutrients in associated crops17.

Material and methods

Database

Data from 52 population sites of T. diversifolia were considered, with current distribution in the lowlands of southeastern Mexico1,2. The databases were obtained from two sources: the National Forest and Soil Inventory (INFyS) of the National Forestry Commission of Mexico (CONAFOR) and the website http://www.tropicos.org/.

Potential distribution areas

In this study, the MaxEnt model (maximum entropy)18 was utilized to model the ecological niche and predict the most likely geographic distribution of T. diversifolia. MaxEnt has been widely used to estimate potential plant and animal distributions with high precision; certain terrestrial groups are excellent examples19.

In the MaxEnt model, the distribution of a species is represented by a probability function P on a set of sites X in the study area. A P model is constructed using a set of restrictions derived from empirical data on the presence of the species20. Restrictions are expressed as simple functions of known environmental variables. The MaxEnt algorithm forces the average of each function of each variable to approach the actual average in those areas where the species is present. Of all the possible combinations of functions, the one that minimizes the entropy function, measured with the Shannon index, is selected. The general expression of the probability function for i environmental variables is21:

Px=eλf(x)/Zλ

Where:

P(x)= probability function;

λ= weighting coefficient vector;

f= corresponding vector of environmental variable functions;

The P (x) values thus obtained represent values of relative suitability for the presence of the species, thus constituting the basis for a potential distribution model. A high value of the distribution function in each pixel indicates that it has favorable conditions for the species22.

In this study, a model was generated using layers of environmental parameters and data of the occurrence of the species. 75 % of the occurrence records were used as training points, and 25 %, as validation points; in addition the AUC (area under the curve) index was used to evaluate the statistical model, since this index is one of the most commonly utilized to measure the quality of the models20. The settings that can be determined in the MaxEnt model affect its accuracy, since the model can work with simple or complex environmental variables.

Databases and environmental parameters

The data of monthly, seasonal and annual precipitation, maximum temperature, minimum temperature, average temperature and thermal oscillation of the 1950-2000 (reference climatology) and 2041-2060 and 2061-2080 (future climatology) periods were utilized for an environmental characterization of the sites where the species occurs. These climatic data were obtained from the WorldClim Global Climate Data website, and the images were processed in ASCII and raster formats, with a resolution of 2.5 min arc. For the 2041-2060 and 2061-2080 periods, the HadGEM2-ES GCM was considered with two representative concentration pathways of greenhouse gases: RCP 4.5 and RCP 8.5. This model has been used in the new climate change scenarios for Mexico; study presented in the Fifth Climate Change Assessment Report by the IPCC (Intergovernmental Panel on Climate Change)23. Other variables that were included in the modeling of the distribution of T. diversifolia, were the altitude and the slope of the land, images that were obtained from the National Environmental Information System (SIAN) of the National Institute for Research on Forestry, Agriculture and Livestock (INIFAP)24. The list of variables used for the space niche model were: slope of the land (%), altitude (m), mean annual maximum temperature (MAMT), May-October mean maximum temperature (MMT5-10), November -April mean maximum temperature (MAX11-4), maximum temperature of the warmest month (MMAX), mean annual temperature (MAT), mean temperature from May to October (MT5-10), mean temperature November-April (MT11-4), minimum temperature of the coldest month (MTCM), average temperature of the coldest month (ATCM), mean minimum annual temperature (MATmin), minimum average temperature from May to October (Tmin5-10), minimum average temperature November-April (Tmin11-4) , accumulated annual precipitation (Pann), accumulated precipitation May to October (P5-10) accumulated precipitation November-April (P11-4), accumulated precipitation of the driest month (Pmin), precipitation of the wettest month (Pmax), mean annual thermal oscillation (MATO), thermal oscillation from May to October (TO5-10), and thermal oscillation from November to April (TO11-4)

Areas with probability of environmental suitability

The model developed from the MaxEnt prediction for the potential distribution of T. diversifolia was examined with the Idrisi Selva system25, and a map was made with the threshold values of the pixels corresponding to the 10th percentile26. For the calculation of the potential distribution area of the species, the areas occupied by water bodies and urban centers were not considered. These thematic layers were obtained by manipulating the use of the soil and the vegetation layer27.

Ecological descriptors

The ecological descriptors for T. diversifolia were determined on the basis of environmental intervals derived from the characterization of the parameters of sites of species presence. This was done using the IDRISI system and raster images of each environmental variable and the geographic coordinates of each site where the species is present.

Results and discussion

In the model of the potential distribution niche of T. diversifolia, the operational curve showed satisfactory results (Figure 1). When considering the training data (75 %), the area under the curve (AUC) of T. diversifolia was 0.979; while when the test data is used (25 %), the validation of the model reported an AUC of 0.966, which indicates that the ability of the model to represent the presence of species was satisfactory28.

Figure 1 Operational curve for T. diversifolia 

Figure 2 shows the current geographical distribution of T. diversifolia. Its presence is mainly in the Central-South region, the Gulf of Mexico, the Pacific coastal areas and the Yucatan Peninsula, which is the tropical part of the country (<23 ° 26 '14' 'North). Most of the populations of T. diversifolia are concentrated in the southeastern part of Mexico, which corresponds to the asseveration about their geographical origin (Mexico and Central and South America)2, as well as in tropical and subtropical areas around the world, which has allowed it to grow under a wide range of environmental conditions and, therefore, develop a wide range of adaptation29.

Figure 2 Current distribution of T. diversifolia, and suitable, unsuitable, retraction and expansion areas for this species under four climate change scenarios in Mexico 

The ecological descriptors of T. diversifolia and the environmental variables that contribute to the construction of the model can be observed in Table 1. According to these results, the distribution of T. diversifolia is mainly determined by variable P5-10 (78.2 %), i.e., the amount of precipitation during the May-October period, which is a key element for the occurrence of this species, as well as for other species, where precipitation in small but frequent events during the summer is key for its good development30. The descriptor of T. diversifolia for this variable indicates a range of 297 to 3,404 mm (for the annual rainfall of 356 to 3,828 mm), which allows for a wide variation of the spatial precipitation. These results are consistent in the sense that this species grows properly under a wide range of precipitation. However, annual rainfall limits of 600 and 5,000 mm30 have been established. Therefore, it can be concluded that T. diversifolia can grow even with less seasonal and annual rainfall than that cited by the current literature.

Table 1 Ecological descriptors of environmental variables that condition the geographic distribution of T. diversifolia to the largest extent 

Variable Minimum value Maximum value Contribution (%)
P5-10, mm 297 3404 78.2
MMAX, °C 21.9 34.4 5.9
Slope, % 0 14 3
Tmin, °C 11 26.15 2.8
Pmax, mm 67 713 2.1

The second most contributing ecological descriptor was MMAX (mean maximum temperature of the warmest month), and T. diversifolia also exhibited a broad thermal range, of 21.9 to 34.4 ° C. Given that T. diversifolia also has a wide range for the MTmin descriptor (mean minimum temperature of the coldest month, 1.8 to 19.8 ° C, Table 1), it can be concluded that this species is capable of exploring and colonizing thermally extreme environments. Other authors2 had previously indicated that it adapts to different climates and altitudes.

The results of potential niches for T. diversifolia under the reference climatology (1951-2000) as well as future climatologies, are also shown in Figure 2. The maps in this figure show the presence of adequate, inadequate, retraction areas and expansion for Tithonia diversifies compared to the scenarios of the reference climatology and future climatologies. In all future scenarios, both retraction areas and expansion areas will appear, showing the potential dynamics of the areas with environmental suitability according to the future climatic changes.

The training presence logistic threshold of the 10th percentile estimates an area with environmental suitability (for the reference climatology) that represents 30.71 % of the territorial extension of Mexico. After a balance between the areas of contraction and expansion of climate change climatologies, the optimal areas will increase to 31.62 %, 31.83 %, 32.45 % and 32.45 % of the Mexican territory, in the 2041-2060 RCP8.5, 2041 -2060 RCP 4.5 and 2061-2080 RCP 4.5, 2061-2080 RCP 8.5 scenarios, respectively. These results show that future climate changes will be apparently beneficial to T. diversifolia. Considering that the most significant variables in the models of the distribution niches are the accumulated precipitation from May to October and the average maximum temperature of the warmest month, it may be inferred that the combinations of these parameters under climate change scenarios will have a positive effect on areas with environmental suitability for T. diversifolia.

The effect of the impact of climate change on the distribution of species may result in ‘winner’ or ‘loser’ species31. The final result will depend on their evolutionary adaptation. For example, expansions of suitable areas have been reported due to climate change for other species like Leucospermum hypophyllocarpodendron subsp. hypophyllocarpodendron11).

Most studies on the effects of climate change on species distribution have reported negative impacts, with the retraction areas being greater than the expansion areas32. In the present study the opposite was obtained, i.e., larger areas of expansion than the retraction areas. As can be seen in Figure 2, these two types of areas have a geographical pattern; contraction areas are mainly located in the western region of the country and near the Pacific coast, while expansion areas are located mainly on the eastern side and near the coast of the Gulf of Mexico. This fact points to two regions of the country where future weather patterns differ in environmental conditions and, therefore, in their level of suitability for T. diversifolia. This particular situation has already become manifest, since the region corresponding to the retraction areas has been previously reported with changes in crop patterns as a result of regional climate changes33,34.

According to the maps exemplified in Figure 2, no current population of T. diversifolia appears to be affected by the retraction of environmental fitness. However, in the event that some populations of T. diversifolia are located in retraction areas in the coming years, it should be considered that these populations will have only two options to survive -a) migrate to more favorable environments, or b) adapt to the new climate conditions35-, which will depend mainly on the genetic diversity of the species36,37. In the case of the present study, it can be hypothesized that environmental suitability will gain a little more surface area due to climate change, and apparently this may promote the future dispersion of T. diversifolia to new areas. However, it is important to consider the potential capabilities of this species to compete against others in the ecosystem or against invasive species that constitute a threat to the stability of the ecosystem38,39,40. In addition, it must be considered that, in the future, T. diversifolia will also depend on its ability to adapt to climate change41, which is a function of its ability to colonize new areas, or (when necessary) its ability to implement physiological modifications in order to adapt to the new environment42. The ecological plasticity of T. diversifolia1,30 is a key aspect for adapting to climate change and new climates43.

According to the results obtained, the expected climate change for both periods will have a more positive than negative effect on the environmental suitability of T. diversifolia (Figure 3), thus allowing its potential territorial expansion in the future. However, it is clear that this environmental advantage to be caused by future climate change will be favorable as well for other species that may compete with T. diversifolia. Under this type of scenario, the wide range for most of the ecological descriptors of T. diversifolia (Table 1) could be an advantageous feature, all the more if one considers the genetic diversity of this species in other parts of the world, which may increase the adaptation and colonization capacities of T. diversifolia36,44.

Figure 3 Dynamics of the surface area with environmental suitability for T. diversifolia in the current climate and four future climatologies 

Although the current and future climate scenarios appear to be favorable for T. diversifolia, measures should be taken to preserve the current populations of the species in order to ensure its presence in its natural environment. On the other hand, since it has been found that T. diversifolia is an allelopathic weed with water-soluble allelochemicals in various parts of the plant17, attention should be paid to supervise its possible territorial expansion due to possible effects on associated crops. Also, the fight against the presence of T. diversifolia as a weed in crops should also be monitored, as it may threaten the natural populations of this species.

Conclusions and implications

At present, the areas with environmental suitability for T. diversifolia represent about one third of the Mexican territory, adapting to a broad range of temperature and humidity conditions. The presence and spatial distribution of this species is determined by the amount of accumulated rainfall during the May-October period, ranging from 297 to 3,404 mm. Future climate changes will cause both retraction and expansion of areas with environmental suitability in different parts of the country; an analysis between these two effects leads to the conclusion that in the present century these areas will increase slightly, between 0.91 and 1.74 %, according to the models utilized. Therefore, it may be considered that this species can be a good alternative for forage production for the future. On the other hand, given that T. diversifolia has been found to be an allelopathic herb, attention should be paid to the supervision of its possible territorial expansion, due to the possible effects on associated crops. The fight against the presence of T. diversifolia in crop fields should also be monitored in order to prevent it from constituting a threat to the natural populations of this species.

Acknowledgements

To the National Forest and Soil Inventory (INFYS) of the National Forest Commission (CONAFOR), for providing valuable information on the presence sites of T. diversifolia in Mexico. Also, the authors are also grateful to the National Institute for Research on Forestry, Agriculture and Livestock (INIFAP) for having allowed the use of raster images of the altitude and slope of the land of the National Environmental Information System.

REFERENCES

1. Pérez A, Montejo I, Iglesias JM, López O, Martín GJ, García DE, Milián I, Hernández A. Tithonia diversifolia (Hemsl.) A. Gray. Pastos y Forrajes 2009;32(1):1-15. [ Links ]

2. Loya OJL, Martínez GS, Prado ROF, Valdés GYS, Gómez, DAA, Escalera V F, Macedo BR, Durán PN. El Sistema Silvopastoril. Sistema Superior Editorial. Nayarit, México; 2014:102. [ Links ]

3. Intergovernmental Panel on Climate Change. Climate Change. The Physical Science Basis. Working group I. Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Summary for Policymakers (Thomas FS, Dahe Q, Gian-Kasper P, Melinda MBT, Simon KA, Judith B, Alexander N, Yu X, Vincent B, Pauline MM. Cambridge University Press; 2013:29. [ Links ]

4. Milly PCD, Dunne KA, Vecchia AV. Global pattern of trends in stream low and water availability in a changing climate. Nature 2005;438(17):347-350. [ Links ]

5. William GJ. Causes of observed changes in extremes and projections of future changes. In: Thomas R. Karl, et al editors. Weather and climate extremes in changing climate regions of focus: North America, Hawaii, Caribbean and Pacific Islands, USA. A Report by the US Climate Change Science Program. Department of Commerce, NOAA’s National Climatic Data Center, Washington, D.C., USA; 2008:81-116. [ Links ]

6. DeFries R, Rosenzweig C. Toward a whole-landscape approach for sustainable land use in the tropics. Proc Natl Acad Sci 2010;107:19627-19632. [ Links ]

7. Gibbs HK, Ruesch AS, Achard F, Clayton MK, Holmgren P, Ramankuttly N, Foley JA. Tropical forests were the primary sources of new agricultural land in the 1980s and 1990s. Proc Natl AcadSci 2010;107(38):16732-16737. [ Links ]

8. Foley JA, Asner GP, Heil C M, Coe MT, DeFries R, Gibbs HK, Howard EA, Olson S, Patz J, Ramankutty N, Snyder P. Amazonia revealed: forest degradation and loss of ecosystem goods and services in the Amazon Basin. Front Ecol Environ 2007;5(1):25-32. [ Links ]

9. Schneider RR, Hamann A, Farr D, Wang, X, Boutin S. Potential effects of climate change on ecosystem distribution in Alberta. Can J For Res 2009;39:1001-1010. [ Links ]

10. Vale MM, Alves MAS, Lorini ML. Mudanças climáticas: desafios e oportunidades para a conservação da biodiversidade brasileira. Oecol Bras 2009;13:518-535. [ Links ]

11. Parmesan C. Ecological and evolutionary responses to recent climate change. Annu Rev Ecol Syst 2006;37:637-669. [ Links ]

12. Schneider RR, Hamann A, Farr D, Wang X, Boutin S. Potential effects of climate change on ecosystem distribution in Alberta. Can J For Res 2009;39:1001-1010. [ Links ]

13. Bellard C, Bertelsmeier C, Leadley P, Thuiller W, Courchamp F. Impacts of climate change on the future of biodiversity. Ecology Letters 2012;15:365-377. [ Links ]

14. Kirilenko AP, Belotelov NV, Bogatyrev BG. Global model of vegetation migration: incorporation of climatic variability. Ecological Modelling 2000;132:125-133. [ Links ]

15. Morueta HN, Fløjgaard C, Svenning JC. Climate change risks and conservation implications for a threatened small-range mammal species. PLoS ONE 2010;5(4):10360-101371. [ Links ]

16. Fasuyi AO, Dairo FAS, Ibitayo FJ. Ensiling wild sunflower (Tithonia diversifolia) leaves with sugar cane molasses. 2010. Livest Res Rural Develop 2015;42(22) from http://www.lrrd.org/lrrd22/3/fasu22042.htm. [ Links ]

17. Otusanya OO, Ilori OJ, Adelusi A.A. Allelopathic effects of Tithonia diversifolia (Hemsl) A. Gray on germination and growth of Amaranthus cruentus. Res J Environ Sci 2010;1(6):285-293. [ Links ]

18. Phillips SJ, Dudík M, Elith J, Graham CH, Lehmann A, Leathwick J, Ferrier S. Sample selection bias and presence-only distribution models: implications for background and pseudo-absence data. Ecological Applications 2009;19:181-197. [ Links ]

19. Sobek S, Kluza S, Cuddington DAK, Lyons DB. Potential distribution of emerald ash borer: What can we learns from ecological niche models using Maxent and GARP? Forest Ecol Management 2012;281(1):23-31. [ Links ]

20. Moreno R, Zamora R, Molina JR, Vásquez A, Herrera M. Predictive modeling of microhabitats for endemic birds in South Chilean temperate forest using Maximum entropy (Maxent). Ecol Informat 2011;6:364-370. [ Links ]

21. Phillips SJ, Dudík M. Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography 2008;31:161-175. [ Links ]

22. Morales SN. Modelos de distribución de especies: Software Maxent y sus aplicaciones en conservación. Rev Conserv Ambi 2012;2(1):1-5. [ Links ]

23. Magaña RVO. Guía para generar y aplicar escenarios probabilísticos regionales de cambio climático en la toma de decisiones. Centro de Ciencias de la Atmósfera. Universidad Nacional Autónoma de México. México; 2010. [ Links ]

24. Díaz PG, Guajardo PRA, Medina GG, Sánchez CI, Soria RJ, Vázquez AMP, et al. Potencial productivo de especies agrícolas de importancia socioeconómica en México. 1a. ed. INIFAP. Xalapa, Ver., México; 2012. [ Links ]

25. Eastman JR. Idrisi Selva Manual, Version 17. Clark Labs, Clark University. Worcester, Mass, USA; 2012. [ Links ]

26. Escalante T, Rodríguez TG, Linaje M, Illoldi RP, González LR. Identification of areas of endemism from species distribution models: threshold selection and nearctic mammals. Rev Especializada Cienc Químico-Biol 2013;16(1):5-17. [ Links ]

27. Instituto Nacional de Estadística Geografía e Informática (INEGI). Guía para interpretación cartográfica: Uso de suelo-vegetación Serie III. DF., México. 2009. [ Links ]

28. Parolo G, Rossi G, Ferrarini A. Toward improved species niche modelling: Arnica montana in the Alps as a case study. J Appl Ecol 2008;45(5):1410-1418. [ Links ]

29. Chapin III FS, Bloom AJ, Field CB, Waring RH. Plant responses to multiple environmental factors. BioScience 1987;37(1):49-57. [ Links ]

30. Murgueitio E, Rosales M, Gómez ME. Agroforestería para la producción animal sostenible. Centro para la Investigación en Sistemas Sostenibles de Producción Agropecuaria. Cali, Colombia; 2001. [ Links ]

31. Hoffman AA, Sgró CM. Climate change and evolutionary adaptation. Nature 2011;470:479-485. [ Links ]

32. Sork V, Davis F, Westfall R, Flints A, Ikegami M, Wang H, Grivet DD. Gene movements and genetic association with regional gradients in California valley oak (Quercus lobata Née) in the face of climate change. Mol Ecol 2010;19:3806-3823. [ Links ]

33. Ramírez LMR, Ruiz CJA, Medina GG, Jacobo CJL, Parra QRA, Ávila MMR, Pilar AJ. Perspectivas del sistema de producción de manzano en Chihuahua, ante el cambio climático. Rev Mex Cienc Agr 2011;(Pub Esp 2):265-279. [ Links ]

34. Santillán ELE, Blanco MF, Magallanes QR, García HJL, Cerano PJ, Delgadillo RO, Valdez CRD. Tendencias de temperatura extremas en Zacatecas, México. Rev Mex Cienc Agr 2011;( Pub Esp 2):207-219. [ Links ]

35. Jump AS, Peñuelas J. Running to stand still: Adaptation and the response of plants to rapid climate change. Ecology Letters 2005;8:1010-1020. [ Links ]

36. Burke MB, Lobell DB, Guarino L. Shifts in African crop climates by 2050, and the implications for crop improvement and genetic resources conservation. Global Environmental Change 2009;19(3):317-325. [ Links ]

37. Mercer KL, Perales HR. Evolutionary response of landraces to climate change in centers of crop diversity. Evol Appl 2010;3(5-6):480-493. [ Links ]

38. Hellmann JJ, Byers JE, Bierwagen BG, Dukes JS. Five potential consequences of climate change for invasive species. Conservation Biol 2008;22(3):534-543. [ Links ]

39. Howard G, Ziller SR. Alien alert: Plants for biofuel may be invasive. Bioenergy Business 2008:14-16. [ Links ]

40. Mainka SA, Howard GW. Climate change and invasive species: Double jeopardy. Integrative Zoology 2010;5:102-111. [ Links ]

41. Alsos IG, Alm T, Normand S, Brochmann C. Past and future range shifts and loss of diversity in dwarf willow (Salix herbacea L.) inferred from genetics, fossils and modelling. Global Ecol Biogeogr 2009;18:223-239. [ Links ]

42. Chown SL, Hoffman AA, Kristensen TN, Angilletta Jr, Stenseth MJ, Pertoldi NCh, C. Adapting to climate change: a perspective from evolutionary physiology. Climate Res 2010;43:3-15. [ Links ]

43. Peters K, Breitsameter L, Gerowitt B. Impact of climate change on weeds in agriculture: a review Agron Sustain Dev 2014;34:707-721. [ Links ]

44. Ruiz CJA, Hernández CJM, Sánchez GJJ, Ortega CA, Ramírez OG, Guerreo HMJ, et al. Ecología, adaptación y distribución actual y potencial de las razas mexicanas de maíz. Libro Técnico Núm. 5. INIFAP-CIRPAC-Campo Exp Centro Altos de Jalisco. 2013. [ Links ]

Received: November 20, 2017; Accepted: July 04, 2018

Creative Commons License Este es un artículo publicado en acceso abierto bajo una licencia Creative Commons