SciELO - Scientific Electronic Library Online

 
vol.7 número especial 13Índices de cambio climático en el estado de Chiapas, México, en el periodo 1960-2009Adaptación del modelo de vulnerabilidad costera en el litoral tabasqueño ante el cambio climático 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 agrícolas

versión impresa ISSN 2007-0934

Rev. Mex. Cienc. Agríc vol.7 spe 13 Texcoco ene./feb. 2016

 

Articles

Monthly trend of maximum and minimum temperatures in Aguascalientes, Mexico

Osías Ruíz Álvarez1  § 

Daniel Espejel Trujano2 

Ronald Ernesto Ontiveros Capurata3 

Juan M. Enciso4 

Manuel Antonio Galindo Reyes5 

Mario Leonel Quesada Parga6 

José Grageda Grageda7 

Rodimiro Ramos Reyes8 

José Ariel Ruíz Corral9 

1 Campo Experimental Pabellón-INIFAP. Carretera Aguascalientes-Zacatecas, km. 32.5. 20660. Pabellón de Arteaga, Aguascalientes.

2 Innovación agrícola. Avenida Tecnológico 510, Colonia Linda Vista. 38010. Celaya, Guanajuato. (despejel@nansa.com.mx).

3 Instituto Mexicano de Tecnología del Agua. Paseo Cuauhnáhuac Núm. 8532. 62550. Jiutepec, Morelos. (reontiverosca@conacyt.mx).

4 Department of Biological and Agricultural Engineering, Texas A&M AgriLife Research, 2415 E. Highway 83 Weslaco, TX 78596. (j-enciso@tamu.edu).

5 Campo Experimental Pabellón-INIFAP. (galindo.manuel@inifap.gob.mx).

6 Instituto Tecnológico El Llano Aguascalientes. Carretera Aguascalientes-San Luis Potosí km. 70, 20330. El Llano, Aguascalientes. (fpaguascalientes@gmail.com).

7 Campo Experimental Costa de Hermosillo-INIFAP. Boulevard del Bosque 7, Colonia Valle Verde. 83200. Hermosillo, Sonora. (grageda.jose@inifap.gob.mx).

8 El Colegio de la Frontera Sur (ECOSUR)- Unidad Villahermosa. Carretera Villahermosa-Reforma, km 15.5. 86280. Villahermosa, Tabasco. (rramos@ecosur.mx).

9 Campo Experimental Centro-Altos de Jalisco-INIFAP. Carretera libre Tepatitlán-Lagos de Moreno, km. 8, 47600. Tepatitlán, Jalisco. (ruiz.ariel@inifap.gob.mx).


Abstract

The objective of this study was to analyze trends in monthly average maximum (Tmax) and minimum (Tmin) temperature in Aguascalientes and time series of at least 30 years. Statistical analysis was performed using the Mann-Kendall methodology and the exchange rate was calculated with Theil slope. From 612 Tmin series, 365 were not significant, whereas 247 were significant, of these series with significant trend (p≤ 0.05), 129 had positive trend and 118 had negative trend. From 672 Tmax series, 448 were not significant, whereas 224 were significant, of these series with significant trend (p≤ 0.05), 167 had positive trend and 57 had negative trend. The highest positive trend (1.33 °C decade-1) from Tmin was in April and the lowest (0.09 °C decade-1) in July and September, the average annual positive trends was 0.50 °C decade-1; December had the highest negative trend (-1.42 °C decade-1) and August had the lowest (-0.14 °C decade-1), the average annual negative trends was -0.59 °C decade-1. Regarding Tmax, April had the highest positive trend (1.59 °C decade-1) and December had the lowest (0.18 °C decade-1), the average annual positive trend was 0.61 °C decade-1; October had the highest negative trend (-1.51 °C decade-1) and September had the lowest (-0.19 °C decade-1), the average annual negative trends was -0.68 °C decade-1. In Aguascalientes, a small number of localities are cooling and a significant portion are warming, the region that is showing temperature changes is having a more extreme weather and changes in the diurnal temperature range (RTD).

Keywords: climate change in Aguascalientes; diurnal temperature range; Mann-Kendall regression; time series; weather stations

Resumen

El objetivo de esta investigación fue analizar tendencias en los promedios de temperatura máxima (Tmax) y mínima (Tmin) mensual en Aguascalientes y series de tiempo de por lo menos 30 años. El análisis estadístico se realizó con la metodología de Mann-Kendall y la tasa de cambio se calculó con la pendiente de Theil. De 612 series de Tmin, 365 no fueron significativas, mientras que 247 si lo fueron, de estas series con tendencia significativa (p≤ 0.05), 129 tuvieron tendencia positiva y 118 tuvieron tendencia negativa. De 672 series de Tmax, 448 no fueron significativas, mientras que 224 si lo fueron, de estas series con tendencia significativa (p≤ 0.05), 167 tuvieron tendencia positiva y 57 tuvieron tendencia negativa. La mayor tendencia positiva (1.33 °C década-1) de Tmin fue en abril y la menor (0.09 °C década-1) en julio y septiembre, el promedio anual de las tendencias positivas fue 0.50 °C década-1; diciembre tuvo la mayor tendencia negativa (-1.42 °C década-1) y agosto tuvo la menor (-0.14 °C década-1), el promedio anual de tendencias negativas fue -0.59 °C década-1. En relación con Tmax, abril tuvo la mayor tendencia positiva (1.59 °C década-1) y diciembre tuvo la menor (0.18°C década-1), el promedio anual de tendencias positivas fue 0.61 °C década-1; octubre tuvo la mayor tendencia negativa (-1.51 °C década-1) y septiembre tuvo la menor (-0.19 °C década-1), el promedio anual de tendencias negativas fue -0.68 °C década-1. En Aguascalientes, un número pequeño de localidades presenta enfriamiento y una parte importante se está calentando, la región que manifiesta cambios en la temperatura está presentando un clima más extremo y cambios en el rango térmico diurno (RTD).

Palabras clave: cambio climático en Aguascalientes; estaciones meteorológicas; rango térmico diurno; regresión de Mann-Kendall; series de tiempo

Introduction

The Intergovernmental Panel on Climate Change (IPCC) indicates that global warming is caused by human activities; it has led to increasing air and ocean temperatures, melting snow and ice caps and sea level rise (IPCC, 2007). Different studies show that at the end of the second half of the twentieth century, the average temperature of the atmosphere increased about 0.6 °C ± 0.2 °C (Folland et al., 2001; Nicholls and Collins, 2006); and brought collateral disasters among them severe heat waves, agricultural losses, damage to health, increased energy demand and loss of humans life (Knapp et al., 1993).

Changes in temperature affect construction activities, power generation and consumption, and crop growth and development (Rehman and Al-Hadhrami, 2012); so the study of its trend over time is vital to estimate their future impact on society, take measures to ensure food production, prevent damage to health and natural resources, and provide information to facilitate the decision-making to political leaders (Skaggs and Irmak, 2012).

The trend analysis of meteorological variable is a statistical procedure, which through testing hypotheses explains the change in the variable, shows the change rate over time, it helps identify vulnerable regions and times to these changes; and answers the question: ¿the climate is changing or remains constant?(Lee et al., 2013). The current trend of global average temperature (IPCC, 2007) is known; but it requires to quantify the change rate at local and regional level to define strategies for prevention, adaptation / mitigation in specific places as spatial and local temporary trend does not follow the global trend (McGuire et al., 2012). Often temperature trend studies are performed on a monthly or annual average value, although this approach helps to detect the pattern of the variable, some details are missed on thermal extremes that are important in agriculture and natural resources; for this reason it is suggested that trend estimation is made for both maximum temperature (Tmax) and minimum temperature (Tmin) (Braganza et al., 2004).

Globally there are a variety of studies on Tmax and Tmin trends. Lee et al. (2013) analyzed changes in extreme temperatures in South Korea, mentioned that trends showed significant discrepancies in space and time, thus in winter the increase of Tmin was 0.80 °C decade-1 in urban areas, and 0.54 °C decade-1 in rural areas; this authors attribute these trends to changes in land use and growth of urban area. Qu et al. (2014) studied the seasonal diurnal temperature range (RTD) on the United States during 1911-2011, trends found varied widely in four regions and for every season; RTD decreases because Tmin increases at a higher rate than Tmax, also summer and autumn showed higher rates of reduction in RTD than spring and winter. In a study of RTD trend, Braganza et al. (2004) found that globally, the average trend of Tmin was higher (0.9 °C) than the trend of Tmax (0.6 °C), these explains the reduction of RTD. Price et al. (1999) report that in the island of Cyprus, Tmin increases at a higher rate than Tmax causing a reduction in RTD from -0.5 to -3.5 °C century-1; changes are attributed to the increased concentration of greenhouse gases, changes in land use and increased urbanization. Other researchers have studied the temperature trend and its relation with agro-meteorological variables and extreme frosts of different thresholds (De Gaetano, 1996; Skaggs and Irmak, 2012).

For Mexico there are few studies on temperature trends. In Zacatecas, Santillan et al. (2011) studied the trends of Tmax and Tmin in 24 meteorological stations; the average positive trends in Tmax was 0.794 °C decade -1 and the only positive trend of Tmin had a rate of 0.496 °C decade-1. The average negative trend of Tmax was -0648 °C decade-1, while for Tmin was -0705 °C decade-1. Zarazúa et al., (2011) point out that by the end of the twenty first century, in Cienega de Chapala, Jalisco; Tmax could increase 6.4 °C; this rise may contribute to increased evapotranspiration, rapid accumulation of heat units and decreased chilling hours which could jeopardize the climatic ability of the region to produce wheat. Lemus and Gay (1988), in a study with seven meteorological stations in Aguascalientes, detected an increase of 0.4 °C in average temperature for 1978 to 1985 regarding the period 1921-1985; and point out that this heat exchange contributed to the reduction of net primary productivity of the ecosystem.

The objective of this research was to determine and quantify the average monthly trend in maximum temperature (Tmax) and minimum temperature (Tmin) in weather stations (56 stations for Tmax and 51 stations for Tmin) from the state of Aguascalientes with information in a continuous period ≥ 30 years.

Materials and methods

Study area

The state of Aguascalientes is 22° 27' and 21° 38' north latitude, 101° 53' and 102° 52' west longitude; and it has a land area of 5 680.33 km2; neighboring to the north, northeast and west with Zacatecas and south and east with Jalisco (INEGI, 1995). The state has three types of climate: temperate semi-dry (BS1k), semi-dry warm (BS1h) and temperate sub-humid with summer rains C (w). In such climates the average annual temperature is 17.1, 20.1 and 14.5 °C, respectively; and average precipitation of 488, 579.1 and 688.3 mm respectively (Garcia, 1973; INEGI, 2013).

Weather information

Daily data of Tmax and Tmin from traditional weather stations in the state of Aguascalientes were used, these stations belong to the National Weather Service (SMN) (Table 1). For Tmax the information available from 56 stations was used, while for Tmin the information available in 51 stations.

Table 1 Weather Stations used in the study. 

With daily data of Tmax and Tmin of each month, the monthly average for each station was obtained, the monthly average was the one used in the trends analysis. For Tmax it counted with 672 time series; i.e. 56 stations by 12 months; for Tmin 612 time series were obtained; i.e. stations 51 by 12 months.

Mann-Kendall test

The Mann-Kendall test is a nonparametric statistical technique that evaluates the significance of a trend. The test uses the Mann-Kendall statistic which is obtained with (Hamed, 2008):

Where: xj and xk represent the variable in years j and k, and also j> k.

Then S variance is given by:

Where: g is the number of consolidated groups and tp is the number of data in the pth group.

Then with S and VAR (S) the statistic test is estimated:

A positive value for Z indicates upward trend, while a negative value indicates a downward trend. H0 is rejected in favor of HA if the absolute value of Z is greater than Z1-α/2 (Hamed, 2008); the value of α used in this study was 0.05.

After the hypothesis test, the change rate of the variable is calculated with (Theil, 1950):

Where: xi' and xi are the data values in the times i' and i, also i'> i. Then it makes N' the number of data where i'> i. Median of N' values of Q is Sen slope or the change rate of the variable (Sen, 1968).

Results and discussion

In this research the monthly trend of Tmax and Tmin was studied with data from meteorological stations (56 weather stations for maximum temperature and 51 weather stations for minimum temperature) in the state of Aguascalientes, the statistical analysis was performed using the Mann-Kendall regression. The results were similar to those reported for other parts of the world; the trends vary depending on the type of temperature (Tmax or Tmin), location (meteorological station) and season (Qu et al., 2014).

Minimum temperature

Table 2 shows the most outstanding parameters from the trend analysis of Tmin. Of the total (612) series of Tmin studied, 365 (59. 64%) series showed no statistically significant trend, the remaining 247 (40.36%) series had a statistically significant trend (p≤ 0.05). Of Series with statistical significance, 129 (21.08%) had positive trend and 118 (19.28%) had negative trend.

Table 2 Summary of time series and trends (°C decade-1) for minimum temperature in Aguascalientes. 

T= total, (%) = porcentaje; NO SIG= series con tendencia no significativa; SIG = series con tendencia significativa, SIG+= series con tendencia positiva significativa; SIG- = series con tendencia negativa significativa, MAX + = tendencia positiva significativa máxima; MIN += tendencia positiva significativa mínima; MEDIA+= tendencia positiva significativa media; MAX-= tendencia negativa significativa máxima; MIN-= tendencia negativa significativa mínima y MEDIA- = tendencia negativa significativa media.

The positive trends are present in the twelve months of the year, but most of them concentrated in February and from April to October (Table 2). The month with the highest positive trend was April with 1.33 °C decade-1 in the station Los Negritos, meanwhile the lowest positive trend occurred in July and September with 0.09 °C decade-1 in the station Presa Plutarco Elias Calles. The average of significant positive trends for the 12 months was 0.50 °C decade-1; this value is similar to the average (0.496 °C decade-1) that Santillan et al. (2011) reported for the neighboring state of Zacatecas. There is also similarity with the average trend reported by Al Buhairi (2010) (0.50 °C decade-1) for the semi-arid City of Taiz, Republic of Yemen, and Al-Hadhrami and Rehman (2012) (0.508 °C decade-1) for Saudi Arabia; besides the average positive trend of each month varied between 0.43 and 0.57 °C decade-1 for January and October, respectively.

Negative trends are also present in the twelve months of the year, but most of them concentrated in January, from April to September and December (Table 2). The month with the highest negative trend was December with -1.42 °C decade-1 in the station El Ocote I, meanwhile the lowest negative trend was in August with -0.14 °C decade-1 in the station Calvillo. The average of all significant negative trends of the 12 months was -0.59 °C decade-1; this value is similar to the average (-0.705 °C decade-1) that Santillan et al. (2011) reported for the neighboring state of Zacatecas. Meanwhile the average of the negative trends of each month ranged from -0.49 in June and September to -0.81 °C decade-1 in December.

Maximum temperature

Table 3 shows the most outstanding parameter from the trend analysis in Tmax. Of the total (672) series of Tmax studied, 448 (66.67%) series had no statistically significant trend, the remaining 224 (33.33%) series had statistically significant trend (p≤ 0.05). Of the total series with significant trend, 167 (24.85%) had positive trend and 57 (8.48%) had negative trend.

Table 3 Summary of time series and trends (°C decade-1) for maximum temperature in Aguascalientes. 

T= total, (%) = porcentaje; NO SIG= series con tendencia no significativa; SIG = series con tendencia significativa, SIG+= series con tendencia positiva significativa; SIG- = series con tendencia negativa significativa, MAX + = tendencia positiva significativa máxima; MIN += tendencia positiva significativa mínima; MEDIA+= tendencia positiva significativa media; MAX-= tendencia negativa significativa máxima; MIN-= tendencia negativa significativa mínima y MEDIA- = tendencia negativa significativa media.

The months with fewer numbers of positive trends in Tmax were in September and November with seven and six series, respectively. The month with the highest positive trend was April with 1.59 °C decade-1 in the station La Tinaja II, while the lowest positive trend was in December with 0.18 °C decade-1 in the station Presa Plutarco Elias Calles. The average positive trend during the 12 months was 0.61 °C decade-1; this value is similar to the average (0.79 °C decade-1) that Santillan et al. (2011) report for the neighboring state of Zacatecas. Our average positive trend is slightly below to average (0.86 °C decade-1) of positive trends cited by Al-Hadhrami and Rehman (2012) for Saudi Arabia. The monthly average of the positive trend in maximum temperature was between 0.37 and 0.82 °C decade-1 in November and April, respectively.

Arelatively small proportion (57) of series with negative trends in Tmax was obtained. The month with the highest negative trend was October with -1.51 °C decade-1 in the station in San Francisco de los Romo, meanwhile the less negative trend was in September with -0.19 °C decade-1 in the station Campo Experimental Pavbellón. The average signif icant negative trend during the 12 months was -0.68 °C decade-1. The average value of the negative trend of Tmax was similar to that reported (0.648 °C decade-1) by Santillan et al. (2011) for the state of Zacatecas. Moreover, the average monthly negative trend ranged between -0.38 and -0.86 °C decade-1 in May and October respectively.

The sum of the series with positive trends in Tmax and Tmin resulted in a total of 296, while the sum of series with negative trends in both temperatures produced a total of 175; this gives an idea that in Aguascalientes there is an upward condition to warming than to cooling which agrees with the situation prevailing in Zacatecas (Santillán et al., 2011) and in other regions of Mexico (Pavia et al., 2009).

Origin of temperature changes

There are different arguments associated to change in temperature patterns. Some reasons are the increased cloud cover, changes in evaporation and precipitation patterns, direct absorption of the infrared portion of the incoming solar radiation, the presence of aerosols, chemical reactions of atmospheric water vapor and increased soil moisture (Henderson, 1992; Hansen, 1995). McNider et al. (1995) comment that other cause could be the change in surface roughness due to the onset of trees or establishment of buildings, this alteration of roughness leads to warmer temperatures due to reduced wind speed. In the 1980's there was already evidence that some regions of Aguascalientes had significant reduction in forest area and the appearance of shrubs, these changes in vegetation coincided with the period of increase in the average temperature of 0.4 °C (Lemus and Gay, 1988).

Environmental effects of temperature changes

According to Santillán et al. (2011) when Tmax and Tmin trends are analyzed there are nine possibilities: 1) Tmax and Tmin stable; 2) Tmax stable and negative trend of Tmin; 3) Tmax stable and positive trend of Tmin; 4) negative trend of Tmax and Tmin stable; 5) Positive trend of Tmax and Tmin stable; 6) negative trend of Tmax and Tmin; 7) positive trend in Tmax and Tmin; 8) negative trend in Tmax and positive trend in Tmin; and 9) positive trend in Tmax and negative trend in Tmin. When the possibilities two, five and nine are present; there is an increase of RTD by increasing RTD, it increases the evaporation capacity of the atmosphere, the relative humidity decreases, increases crop reference evapotranspiration (ETO), increases the actual crop evapotranspiration (ETC) and increases the irrigation water volume required (Tabari et al., 2011; Xiaomang et al., 2011). When the possibilities three, four and eight are present; there is a reduction in RTD by reducing RTD, decreases the evaporation capacity of the atmosphere, the relative humidity increases, decreases crop reference evapotranspiration (ETO), decreases the actual crop evapotranspiration (ETC) and decreases irrigation water volume required. When possibilities six and seven are present could be an increase or decrease in RTD, this will depend on the change rate of each of the temperatures.

Table 4 shows the nine possible cases of Tmax and Tmin obtained for Aguascalientes; the month column appears the number of stations present for each case. It is appreciated that case one had the greatest number of stations (237). Note that adding stations involved in cases 5, 2 and 9, 180 stations are obtained, all of them show the condition of increased RTD. Moreover, if the stations involved in cases 3, 4 and 8 are added, 110 stations are obtained and all of them show the condition of reduced RTD. It is important to add that cases 7 and 6 could also have modifications of RTD as trend lines are not quite parallel, i.e. differ in the change rate, for these cases 67 stations are involved.

Table 4 Trend cases of maximum and minimum temperature and number of stations per month. 

*Se refiere al número de estaciones en cada mes que no contaron con temperatura mínima.

Finally, Figure 1 shows examples of time series for Tmax, Tmin and RTD in the station Sandovales. For Tmax case there was a significant trend for six months (February to June and December), while the trend in Tmin was significant for eight months (April to November). Consequently, RTD showed significant trend in 11 months (February to December), note that in those months the trend line is positive which gives an idea that the climate of the locality is more extreme and has more deficit of environmental humidity.

Figure 1 Time series for Tmax, Tmin and RTD from January to December in the station Sandovales 

Conclusions

The monthly trends of Tmax and Tmin in weather stations from the state of Aguascalientes that count with time series ≥ 30 years were studied. In this state the trend of these temperatures is not uniform; there is temperature variation by temperature type, location and month.

There are increments of Tmin in the 12 months and in most stations; however the highest rate of increase occurs in spring (April), while the lowest in summer months (July to September). Negative trends of Tmin are also present the 12 months; but highest occurs in winter (December) and the lowest in summer (August).

The increments of Tmax occurs year round, despite the month with highest positive trend is April; and December is the month with the lowest positive trend. The highest negative trend is in October, and the lowest in September.

An important part of Aguascalientes does not present changes in weather, which is demonstrated by the 237 series presenting case one of maximum and minimum temperature. Most of the state shows changes, and the evidence is that 290 series have one of the cases 2, 3, 4, 5, 8 and 9 which affects RTD.

It is recognized that the region of the state with changes in atmospheric pattern, a small number of locations show cooling while an important part shows warming.

Literatura citada

Al Buhairi, M. H. 2010. Analysis of monthly, seasonal and annual air temperature variability and trends in Taiz City-Republic of Yemen. J. Environ. Protec. 1:401-409. [ Links ]

Braganza, K.; Karoly, D. J. and Arblaster, J. M. 2004. Diurnal temperature range as an index of global climate change during the twentieth century. Geophysical Research Letters. 31: doi: 10.1029/2004GL019998. [ Links ]

De Gaetano, A. T. 1996. Recent trends in maximum and minimum temperatures threshold exceedences in the Northeastern United States. Journal of Climate. 9: 1646-1660. [ Links ]

Folland, C.K.; Karl, T.R.; Christy, J.R.; Clarke, R.A.; Gruza, G.V.; Jouzel, J.; Mann, M. E.; Oerlemans, J.; Salinger, M. J. and Wang, S. W. 2001. Observed climate variability and changes. In: Houghton, J. T.; Ding, Y.; Griggs, D. J.; Noguer, M.; van den Linden, P. J.; Dai, X.; Maskell, K. and Johnson, C. A. (Eds.). Climate change 2001: the scientific basis. Contribution of working group I to the third assessment report of the intergovernmental panel on climate change. Cambridge University Press. Cambridge, UK. 99-181 pp. [ Links ]

García, E. 1973. Modificaciones al sistema de clasificación climática de Köppen (para adaptarlo a las condiciones de la República Mexicana). Universidad Nacional Autónoma de México (UNAM). México, D. F., 246 p. [ Links ]

Hamed, K. H. 2008. Trend detection in hydrologic data: the Mann-Kendall trend test under the scaling hypothesis. J. Hydrol. 349:350-363. [ Links ]

Hansen, J.; Sato, M. and Ruedy, R. 1995. Long-Term changes of the diurnal temperature cycle: implications about mechanisms of global climate change. Atmosphere Res. 37:175-209. [ Links ]

Henderson, S. A. 1992. Continental cloudiness changes this century. Geo Journal. 27: 255-262. [ Links ]

INEGI (Instituto Nacional de Estadística, Geografía e Informática). 1995. Anuario estadístico del estado de Aguascalientes. Gobierno del estado de Aguascalientes. Aguascalientes, México. 317 p. [ Links ]

Instituto Nacional de Estadística, Geografía e Informática (INEGI). 2013. Anuario estadístico de los Estados Unidos Mexicanos. Gobierno del estado de Aguascalientes. Aguascalientes, México. 785 p. [ Links ]

IPCC (Intergovernmental Panel on Climate Change). 2007. In Climate Change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change. In: Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K. B.; Tignor, M. and Miller, H. L. (Eds.). Cambridge University Press: Cambridge, New York; 996. [ Links ]

Knapp, W. W.; Eggleston, K. L.; De Gaetano, A. T.; Vreeland, K. and Schultz, J. D. 1993. Northeast climate impacts. Northeast Regional Climate Center. 93(7):1-7. [ Links ]

Lee, K.; Baek, H. J. and Cho, Ch. 2013. Analysis of changes in extreme temperatures using quantile regression. Asia-Pacific J. Atmospheric Sci. 49:313-323. [ Links ]

Lemus, L. and Gay, C. 1988. Temperature, precipitation variations and local effects Aguascalientes 1921-1985. Atmósfera. 1:39-44. [ Links ]

McGuire, Ch. R.; Nufio, C. R.; Bowers, M. D. and Guralnick, R. P. 2012. Elevation- dependent temperature trends in the rocky mountain front range: changes over a 56- and 20- year record. PLos one. 7(9):e44370. doi:10.1371/journal.pone.0044370. [ Links ]

Nicholls, N. and Collins, D. 2006. Observed climate change in Australia over the past century. Energy and Environment. 17:1-12. [ Links ]

McNider, R. T.; England, D. E.; Friedman, M. J. and Shi, X. 1995. Predictability of the stable atmospheric boundary layer. J. Atmospheric Sci. 52:1602-1614. [ Links ]

Pavia, E. G.; Graef, F. and Reyes, J. 2009. Annual and seasonal surface air temperature trends in Mexico. Int. J. Climatol. 29:1324-1329. [ Links ]

Price, C.; Michaelides, S.; Pashiardis, S. and Alpert, P. 1999. Long term changes in diurnal temperature range in Cyprus. Atmospheric Research. 51:85-98. [ Links ]

Qu, M.; Wan, X. and Hao, X. 2014. Analysis of diurnal air temperature range change in the continental United States. Weather and Climate Extremes. 4:86-95. [ Links ]

Rehman, S. and Al-Hadhrami, L. M. 2012. Extreme temperatures trends on the West Coast of Saudi Arabia. Atmospheric and Climate Sciences. 2:351-361. [ Links ]

Santillán, E. L. E.; Blanco-Macías, F.; Magallanes-Quintanar, R.; García- Hernández, J. L.; Cerano-Paredes, J.; Delgadillo-Ruiz, O. y Valdez-Cepeda, R. 2011. Tendencias de temperaturas extremas en Zacatecas, México. Rev. Mex. Cienc. Agríc. 2:207-219. [ Links ]

Sen, P. K. 1968. Estimates of the regression coefficient based on Kendall’s tau. J. Am. Statistical Assoc. 63:1379-1389. [ Links ]

Skaggs, K. E. and Irmak, S. 2012. Long-term trends in air temperature distribution and extremes, growing degree-days, and spring and fall frosts for climate impact assessments on agricultural practices in Nebraska. J. Appl. Meteorol. Climatol. 51:2060-2073. [ Links ]

Tabari, H.; Marofi, S.; Aeini, A.; Talaee, P. H. and Mohammadi, K. 2011. Trend analysis of reference evapotranspiration in the western half of Iran. Agric. Forest. Meteorol. 151:128-136. [ Links ]

Theil, H. 1950. A rank-invariant method of linear and polynomial regression analysis. Part 3. Proceedings of Koninalijke Nederlandse Akademie van Weinenschatpen Amsterdam. 53:1397-1412. [ Links ]

Xiaomang, L.; Hongxing, Z.; Minghua, Z. and Changming, L. 2011. Identification of dominant climate factor for pan evaporation trend in the Tibetan Plateau. J. Geographical Sci. 21:594-608. [ Links ]

Zarazúa, V.P.; Ruíz-Corral, J.A.; González-Eguiarte, D.R.; Flores-López, H. E. y Ron-Parra, J. 2011. Cambio climático y agroclimático para el ciclo otoño-invierno en la región Ciénega de Chapala. Rev. Mex. Cienc. Agríc. 2:295-308. [ Links ]

Received: November 2015; Accepted: February 2016

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