SciELO - Scientific Electronic Library Online

 
vol.28 número3Nueva ecuación dinámica de altura dominante e índice de sitio para Pinus chiapensis (Martínez) Andresen en Puebla y Veracruz, MéxicoValor económico del almacén de carbono en biomasa aérea de manglares de Sonora, México 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 Chapingo serie ciencias forestales y del ambiente

versión On-line ISSN 2007-4018versión impresa ISSN 2007-3828

Rev. Chapingo ser. cienc. for. ambient vol.28 no.3 Chapingo sep./dic. 2022  Epub 08-Mar-2024

https://doi.org/10.5154/r.rchscfa.2021.08.052 

Scientific articles

Precipitation variability (1660-2018) for the western part of Chihuahua induced with tree rings

Ricardo Miranda-Briones1 

Julián Cerano-Paredes2  * 

Gerardo Esquivel-Arriaga2 

Jordi Morató-Farreras3 

Rosalinda Cervantes-Martínez2 

Ignacio Sánchez-Cohen2 

Alejandro Gómez-Nísino4 

1Universitat Politècnica de Catalunya (UPC - Barcelona Tech), Instituto en Ciencia y Tecnología de la Sostenibilidad (IS.UPC). Carrer Jordi Girona 29. C. P. 08034. Barcelona, España.

2Instituto Nacional de Investigaciones Forestales, Agrícolas y Pecuarias (INIFAP), Centro Nacional de Investigación Disciplinaria en Relación Agua, Suelo, Planta, Atmósfera (CENID-RASPA). km 6.5 Margen Derecha del Canal Sacramento. C. P. 35140. Gómez Palacio, Durango, México.

3Universitat Politècnica de Catalunya (UPC-Barcelona Tech), Cátedra UNESCO de Sostenibilidad. Carrer Colom 1, edificio TR1, planta 0 puerta 108. C. P. 08222. Terrassa, España.

4Comisión Nacional de Áreas Naturales Protegidas (CONANP), Área de Protección de Flora y Fauna (APFF) Campo Verde. Calle 8a s/n entre Allende y Ahumada, col. Centro. C. P. 31940. Madera, Chihuahua, México.


Abstract

Introduction:

Tree rings play a significant role for the reconstruction of climate variables on century scales.

Objective:

To reconstruct the variability of precipitation for the western part of the state of Chihuahua, based on tree rings of Pseudotsuga menziesii (Mirb.) Franco, Pinus arizonica Engelm. and Pinus strobiformis Engelm.

Materials and methods:

Samples were extracted with a Pressler drill, dated and growth indexes were generated based on dendrochronological techniques. A correlation analysis was run between regional chronology and observed precipitation data. A regression model was used to reconstruct precipitation. A spectral analysis determined the recurrence of droughts and the influence of El Niño Southern Oscillation (ENSO) on precipitation variability.

Results and discussion:

The relationship between the chronologies was statistically significant and a regional series of 359 years was developed. October-June precipitation had the highest impact on growth (r = 0.73, P < 0.01), so the variability of precipitation for that period was reconstructed for the last 359 years. Severe droughts were detected in the periods 1751-1768, 1807-1810 and 1945-1956 and a recurrence of droughts every 50 years was identified. The reconstructed precipitation variability is significantly modulated by ENSO at different time scales and frequencies.

Conclusion:

Paleoclimatic series help us to spread the knowledge of climate in specific geographic regions and to analyze the recurrence of extreme events (droughts).

Keywords: Pseudotsuga menziesii; Pinus arizonica; Pinus strobiformis; dendrochronology, drought severity.

Resumen

Introducción:

Los anillos de los árboles juegan un papel relevante para la reconstrucción de variables climáticas en escalas de siglos.

Objetivo:

Reconstruir la variabilidad de la precipitación para el oeste de Chihuahua, con base en anillos de árboles Pseudotsuga menziesii (Mirb.) Franco, Pinus arizonica Engelm. y Pinus strobiformis Engelm.

Materiales y métodos:

Se extrajeron muestras con taladro de Pressler, se fecharon y generaron índices de crecimiento con base en técnicas dendrocronológicas. Se corrió un análisis de correlación entre la cronología regional y los datos de precipitación observados. Se generó un modelo de regresión para reconstruir la precipitación. Un análisis espectral determinó la recurrencia de sequías y la influencia de El Niño Oscilación del Sur (ENSO) en la variabilidad de la precipitación.

Resultados y discusión:

La relación fue significativa entre las cronologías y se desarrolló una serie regional de 359 años. La precipitación octubre-junio es la que más influye en el crecimiento (r = 0.73, P< 0.01), por lo que se reconstruyó la variabilidad de la precipitación de dicho periodo para los últimos 359 años. Se detectaron sequías severas en los periodos 1751-1768, 1807-1810 y 1945-1956 y se determinó una recurrencia de sequías cada 50 años. La variabilidad de la precipitación reconstruida es modulada significativamente por ENSO a diferentes escalas de tiempo y frecuencias.

Conclusión:

Las series paleoclimáticas permiten extender el conocimiento del clima en regiones geográficas específicas y analizar la recurrencia de eventos extremos (sequías).

Palabras clave: Pseudotsuga menziesii; Pinus arizonica; Pinus strobiformis; dendrocronología, severidad de sequía

Highlights:

  • Precipitation reconstructed with tree rings of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis.

  • A 359-year regional series (1660-2018) was developed.

  • October-June precipitation had the highest impact on annual ring´s growth.

  • A drought recurrence every 50 years was identified.

  • The reconstructed precipitation variability is significantly modulated by ENSO.

Introduction

Droughts are one of the most important processes of natural climate variability, where precipitation levels are significantly lower than normal; due to current climate change, these events are likely to be magnified (Pokhrel et al., 2021). Even if we meet with the Paris Agreement to limit the global average temperature increase to 1.5 °C compared to pre-industrial levels, an increase in the intensity or frequency of droughts and precipitation deficits is expected in some regions of the world (Grupo Intergubernamental de Expertos sobre el Cambio Climático [IPCC], 2019). While there is still uncertainty, there is enough knowledge to affirm that the north of Mexico will be impacted with more prolonged and intense droughts (Magaña-Rueda & Neri, 2012).

The border region of the states of Chihuahua and Sonora, located in an arid to semi-arid zone in the northwest part of Mexico, is frequently affected by severe droughts with considerable impact on the agricultural and livestock economy (Reyes-Gómez, Díaz, Brito-Castillo, & Núñez-López, 2013). Thanks to the diversity of these ecosystems, the area is a refuge for a significant variety of flora and fauna species that provide important environmental services for the population. The region is also rich in water resources due to the headwaters of the Fuerte, Yaqui and Humaya rivers that flow on both sides of the Sierra Madre Occidental, which recharge the aquifers (González-Elizondo, González-Elizondo, Tena-Flores, Ruacho-González, & López-Enríquez, 2012).

Therefore, it is essential to analyze trends in hydroclimatic variability to understand the impact of atmospheric circulation phenomena, to help developing plans to address future water scarcity events. However, the availability of climate records is limited because stations located in the area lack extensive and quality data (Servicio Meteorológico Nacional [SMN], 2019).

Annual tree growth rings are one option to generate information on past climate (Guibal & Guiot, 2021). It is possible to obtain information on the historical variability of the climate of a region (Lüttge, 2017). In the bordering areas of Chihuahua and Sonora, paleoclimatic reconstructions are scarce (Irby, Fulé, Yocom, & Villanueva-Díaz, 2013; Villanueva-Díaz et al., 2015), thus, the generation of a greater number of series could help in the improvement of knowledge and understanding of the climate variability experienced and possible repercussions.

Therefore, the present study has the following objectives: 1) to reconstruct the variability of precipitation for the western region of the state of Chihuahua based on annual growth rings of Pseudotsuga menziesii (Mirb.) Franco, Pinus arizonica Engelm. and Pinus strobiformis Engelm., 2) to determine the frequency of extreme events (droughts) and 3) to analyze the influence of El Niño Southern Oscillation (ENSO) on the historical variability of climate.

Materials and Methods

Study area

The Santuario de la Cotorra Serrana conservation area (29° 19’ 25” N, 108° 12’ 49° W) and the La Norteña site (29° 39' 08" N, 108° 28' 20" W), the latter corresponding to the Campo Verde Flora and Fauna Protection Area, are found in the municipality of Madera, Chihuahua, at an elevation gradient of 2 500 to 3 000 m (Figure 1). Both sites are located in the Yaqui river basin, which in turn belongs to the Región Hidrológica 9 Sonora Sur, with an area of 72 859.05 km2 and a perimeter of 2 333.03 km, within the Sierra Madre Occidental (Instituto Nacional de Estadística y Geografía [INEGI], 2019). Also, both sites belong to Regiones Terrestres Prioritarias (RTP) of Mexico (Comisión Nacional para el Conocimiento y Uso de la Biodiversidad [CONABIO], 2019).

The predominant climate in the area is semi-cold-subhumid with long cool summer (Cb’(w2)x’) (García, 1970). Mean annual precipitation ranges between 650 and 700 mm. Mean annual temperature of 11 °C with maximum temperatures of 28 °C in June and July, and minimum temperatures of -9 °C in December and January (SMN, 2019). The dominant vegetation type is conifers and oaks, while the soil type is Feozem (CONABIO, 2019).

Figure 1 Geographical location of the study sites in the Yaqui river basin in western part of Chihuahua. 

Dendrochronological sampling and sample preparation

Tree ring samples were obtained by selective sampling. To maximize the climate signal, long-lived and climate-sensitive trees located in dry, rocky terrain were selected, seeking sites with minimal human disturbance (Fritts, 2001; Speer, 2010). Two to three samples were extracted per individual using a Pressler drill of 5 and 12 mm diameter and variable lengths (Hägloff). A total of 40 trees were sampled (18 P. menziesii, 15 P. arizonica and seven P. strobiformis) and a total of 91 samples were obtained at El Santuario. Sixty P. arizonica trees were sampled and a total of 132 samples were extracted at La Norteña. These were processed and predated following conventional techniques (Stokes & Smiley, 1996).

Dating and development of chronologies

For the predated samples, each of the annual growths (total ring, earlywood and latewood) were measured using a VELMEX sliding-phase system with 0.001 mm accuracy. Dating and measurement accuracy were verified with the COFECHA program (Grissino-Mayer, 2001; Holmes, 1983). Only series with intercorrelation values > 0.3281 (P < 0.01) were included in the construction of the final chronology (Holmes, 1983). Chronologies or growth indices (standard, residual and arstan), with mean 1.0 and homogeneous variance (Fritts, 2001), were generated with the ARSTAN program (Cook, 1987). This study used the standard and residual versions of the chronology because they provide the best association with precipitation.

Data analysis

A regional chronology was developed, and a Pearson correlation analysis was run between regional indices and observed regional precipitation data. Precipitation comprises quadrangles 51934 and 53212 (Terán-Cuevas, Gutiérrez-Castillo, & Davydova-Belitskaya, 2010) and the climatological stations Madera and Mesa del Huracán (SMN, 2019) (Figure 1; Table 1).

Table 1 Weather stations used in the dendrochronological analysis for the western region of Chihuahua. 

Weather stations Latitude (N) Longitude (W) Elevation (m) Period
Madera 8097 29° 11’ 24” 108° 08’ 29” 2 102 1957-2007
Mesa del Huracán 8103 29° 38’ 00” 108° 14’ 01” 2 107 1957-1990
Cuadrante 51934 29° 19’ 22” 108° 13’ 01” --- 1950-2015
Cuadrante 53212 29° 39’ 08” 108° 28’ 22” --- 1950-2015

The relationship between monthly precipitation and regional indices of total ring width, earlywood and latewood in their standard and residual versions was evaluated by correlation analysis. Eighteen months of data were considered: six months before the growing year and 12 months during the growing season. According to Cleaveland, Stahle, Therrell, Villanueva-Díaz, and Burns (2003), climate conditions recorded before the beginning of the growing season influence the development of subsequent annual growth.

For the highest significant correlation (P < 0.05) between cumulative precipitation and regional total ring index, a transfer equation was developed, using the program R version 3.6.1 (R Core Team, 2019). Precipitation was reconstructed with this model for the time span covered by the observed records. Using the Verify subroutine of the University of Arizona Dendrochronological Programs package (Holmes, 2001), a calibration test was applied to half of the observed and reconstructed climate data (1998-2015, n = 18) and the remaining half (1980-1997, n = 17) was assigned a verification test (Fritts, 2001). The equation obtained was applied to reconstruct precipitation over the total period of the regional series. A 10-year spline curve was fitted to the precipitation reconstruction to highlight dry and rainy events (Cook, 1987). A spectral analysis was applied to the reconstructed series to identify recurrence of dominant anomalies (droughts). For the most severe droughts, a spatial correlation analysis was performed between the reconstructed precipitation and the Palmer Drought Severity Index (PDSI) (Stahle et al., 2016). Finally, to determine the impact of ENSO on hydroclimatic variability, the reconstructed precipitation and the Niño 3.4 SST index (Rayner et al., 2003) were related using MATLAB version 8.5 (The MathWorks Inc., 2015). For this purpose, annual resolution plots of wavelet spectra and wavelet coherence plots were used.

Results

Dating of samples

From a total of 223 growth cores, 159 samples (71 %) were dated to the exact year of formation. The COFECHA results indicated that the intercorrelation of the series was statistically significant (Table 2). Two chronologies were identified: El Santuario of 156 years (1863-2018) and La Norteña of 359 years (1660-2018). A significant correlation (r = 0.580; P < 0.01) was determined between the two chronologies and a regional chronology was developed (Figure 2). This showed a expressed population signal (EPS) > 0.85 with more than 10 cores starting in 1739, a statistically robust sample size (Figure 2).

Figure 2 Composite total ring chronology of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis. The time series shows the climate variability over the last 359 years for the western region of Chihuahua. The bottom gray line represents the annual variability; the dotted horizontal line, the mean; the black highlighted line is a 10-year spline that allows the observation of low frequency events; and the gray area indicates the sample size for each segment of the regional chronology. The red line represents an expressed population signal value (EPS) > 0.85. 

Table 2 Statistical parameters of total ring chronologies of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis for the western region of Chihuahua. 

Site Total ring correlation Length of chronology Sensitivity
El Santuario 0.610 (P < 0.01) 1863-2018 (156 years) 0.265
La Norteña 0.605 (P < 0.01) 1660-2018 (359 years) 0.313
Regional 0.580 (P < 0.01) 1660-2018 (359 years) 0.3

Response function

Figure 3 shows that the highest correlations between precipitation and chronologies are found with the standard version. Regarding monthly precipitation, the period from September of the previous year to June of the current year showed a positive relationship with the regional chronology of total ring, earlywood and latewood in its standard and residual version. The months of October and November of the previous year and February of the current year showed significant correlation (P < 0.05) with both standard and residual chronology for total ring, earlywood and latewood (Figure 3A and 3C).

In the case of cumulative precipitation, correlations were significant (P < 0.01) in all three indices (total ring, earlywood and latewood) throughout the year for standard and residual chronology (Figure 3B and 3D). However, the seasonal October-June cumulative precipitation period of the standard chronology recorded the highest correlation for total ring (r = 0.73, P < 0.01), earlywood (r = 0.71, P < 0.01), and latewood (r = 0.71, P < 0.01) (Figure 3B). The regional standard total ring chronology has the greatest potential to be used as a proxy for seasonal precipitation variability in the October-June period (r = 0.73, P < 0.01).

Figure 3 Response function analysis between precipitation (monthly and cumulative) and regional standard (A and B) and residual (C and D) chronology of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis for the western region of Chihuahua. *P< 0.05 and **P< 0.01 indicate the degree of significance of correlation between both variables. 

Precipitation reconstruction

Figure 4A shows highly significant association (r = 0.73, P < 0.01) between the seasonal precipitation period October-June and the total ring width index for the common period 1980-2015. A linear regression model was developed that allowed reconstructing the variability of observed data precipitation in that period (Figure 4B). The model was as follows:

Yt = -170.41 + 454.15 * Xt

where,

Yt = precipitation value (mm) October-June, reconstructed for a specific t year

Xt = total ring width index for a specific t year.

Figure 4 Association between the ring width index (RWI) of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis and regional October-June precipitation for the period 1980-2015 in the western region of Chihuahua (A); linear regression model between both variables (B); and comparison of reconstructed October-June precipitation (solid line) and measured precipitation (dotted line) in the period 1980-2015, verification and calibration of the regression model (C). 

Model calibration and verification

Figure 4C indicates the variability of reconstructed precipitation for the common period of measured data (1980-2015). The model calibration process indicated highly significant correlation (r = 0.76, r2 = 0.58, P < 0.01) between measured and reconstructed precipitation, and explained 58 % of the precipitation variability. On the other hand, the verification showed highly significant correlation r = 0.66 (r2 = 0.44, P< 0.01), which explained 44 % of precipitation variability (Tables 3 and 4). Since both subperiods show significant correlations, the model for the total period of measured data (1980-2015) was considered statistically valid with r = 0.75 (r2 = 0.56, P < 0.01) and was applied to the total chronology to reconstruct precipitation variability (Table 3).

Table 3 Calibration and statistics of the regression between October-June precipitation and total ring standard chronology index of Pseudotsuga menziesii, Pinus arizonica and Pinus strobiformis for the western part of Chihuahua. 

Period R2 adj Coefficients Standard error T test Probability
β0 β1 β0 β1 β0 β1 β0 β1
1980-1997 0.44 -136.28 463.61 155.5 148.6 -0.87 3.11 0.39 0.007
1998-2015 0.58 -110.69 342.51 65.33 71.48 -1.69 4.79 0.1 0
1980-2015 0.56 -170.41 454.15 74.93 76.59 -2.27 5.93 0.029 0

Table 4 Verification statistics for seasonal precipitation October-June for the western part of Chihuahua. 

Period Pearson correlation (r) Error reduction Sign test T test First significant difference
1980-1997 0.66** 0.69* 2* 4.2* 3*
1998-2015 0.76** 0.82* 4* 4.3* 3*
1980-2015 0.75** 0.55* 6* 4.7* 6*

*P < 0.05, ** P < 0.01

Variability of reconstructed precipitation

Figure 5 shows the October-June reconstructed precipitation variability covering the period 1660-2018. The series indicates high interannual and multiannual variability with dry and wet periods that varied in extent and intensity. For the last three and a half centuries, significant frequencies (P < 0.05) of extreme events were determined every 50 years (Figure 6), at approximately mid-century and between the end and beginning of the century. Spatial correlation analysis between PDSI and four of the most severe droughts (1695-1707, 1751-1768, 1807-1810, 1945-1956) indicated areas of significant correlation (P < 0.05), showing the magnitude and geographic extent of these episodes (Figure 7).

Figure 5 Reconstruction of winter-spring precipitation (359 years) for the western region of Chihuahua. The gray line at the bottom indicates annual variability and the smoothed 10-year curve, highlighted in black, indicates low-frequency events (in blue dry periods and in red wet periods). Average precipitation is 277.4 mm. 

Figure 6 Spectral analysis of reconstructed regional precipitation for the western part of Chihuahua. Numbers indicate the frequencies reported in the series; in bold, the frequencies of 2 and 5 years indicate relationship with the ENSO phenomenon and the frequencies of 51 and 91 years correspond to extreme events with cycles of about 50 years. 

Figure 7 Spatial correlation between reconstructed precipitation and Palmer Drought Severity Index (PDSI) for the months of June, July and August (Stahle et al., 2016) for four extreme drought events. The star indicates the study area. 

Impact of ENSO on precipitation variability

Figure 8A shows the wavelet spectrum for the reconstructed precipitation, significant areas (P < 0.05) corresponding to periods of one (1902-1904, 1988-1989), three (1995-2002) and six years (2006-2018). On the other hand, Figure 8B indicates the wavelet spectrum of the ENSO phenomenon and its irregular cycle from one to seven years for the intervals 1912-1920, 1942-1958, 1968-1973, 1978-2006 and 2008-2015. Spectral wavelet coherence analysis determined significant (P < 0.05) areas of in-phase relationship between the reconstructed precipitation and the Niño 3.4 index for the period 1900-2018 at frequencies of one to seven years (Figure 8C). Significant (P < 0.05) periods in phase were determined for the intervals 1900-1910, 1932-1940, 1934-1963, 1920-1950 and 1995-2018 at frequencies of 1-7, 1-2, 4-7, 10-16 and 5-7 years, respectively.

Figure 8 Wavelet spectra for the reconstructed precipitation (A) and the ENSO 3.4 index (B). Wavelet coherence analysis between the reconstructed precipitation series and the ENSO 3.4 index (C). Areas in red delimited by a black line indicate a significant relationship (P < 0.05) between the two variables. Horizontal arrows to the right indicate that both phenomena are in phase (positive relationship). On the other hand, arrows to the left indicate an anti-phase or negative relationship. 

Discussion

Series sensitivity

The regional chronology generated exceeds the correlation parameter determined by the COFECHA program (Holmes, 1983), which establishes r = 0.3281 as the minimum significant correlation (P < 0.01) between growth series. This indicates that the species used have good sensitivity to environmental changes, showing potential for the reconstruction of regional climate variability.

Common variability of the series

P. menziesii, P. arizonica and P. strobiformis all show common climate variability through time, suggesting that annual growth is influenced by meteorological conditions modulating the regional climate. Multiple studies have found a common response between species from the Sierra Madre Oriental (Cerano-Paredes et al., 2011a; Cerano-Paredes, Villanueva-Díaz, Valdez-Cepeda, Méndez-González, & Constante-García, 2011) and the Sierra Madre Occidental (Chávez-Gándara et al., 2017; Irby et al., 2013; Villanueva-Díaz et al., 2015), even when they are separated; this implies that species capture the variability of larger scale climatic phenomena. A similar fluctuation in growth among species is due to the common limiting factor that is precipitation (Fritts, 2001).

Response of chronology to precipitation variability

Precipitation significantly influences the growth of tree species in the western part of Chihuahua, especially the seasonal October-June precipitation. Villanueva-Díaz et al. (2015) report similar results in the northwestern region of Chihuahua. The response of conifer growth to winter-spring precipitation in the northwestern part of Mexico and southwestern United States has been corroborated by several paleoclimatic studies (Cerano-Paredes et al., 2011a, 2012; Chávez-Gándara et al., 2017; Cleaveland et al., 2003; D’Arrigo & Jacoby, 1991). An explanation for the above is that rainfall during winter is usually of low intensity, favoring infiltration and soil saturation; meanwhile, low evapotranspiration in those months makes water saturate the soil and be available for the following tree growth period (Zywiec et al., 2017). Several paleoclimatic studies have documented and discussed the response of conifer growth to winter-spring precipitation in Chihuahua (Cerano-Paredes et al., 2009; Irby et al., 2013; Villanueva-Díaz et al., 2015). This common species response has made it possible to reconstruct droughts and analyze their impact at national level (Stahle et al., 2011).

Reconstructed droughts

There is little historical information available on the occurrence of droughts and their ravages in Mexico; with respect to the northern part, records begin in the second half of the eighteenth century when the area was colonized and populated (García-Acosta, 1993). In Chihuahua, the records report drought affectations during several periods: in the mid-1720s, the late 1730s and early 1740s, between 1748 and 1766, in the first three years of the 1770s, in the mid-1780s, the first two decades of the 19th century, between 1890 and 1897, and during the 20th century in the 1930s, 1950s and 1990s (Endfield & Fernández-Tejedo, 2006). The present reconstruction recorded all reported events except for the period from the late 1730s to the early 1740s, which validates the annual resolution quality of the reconstruction.

Drought in the late 17th and early 18th centuries (1695-1707) has been reported in other paleoclimatic reconstructions from Chihuahua (Villanueva-Díaz et al., 2015) and Durango (Cerano-Paredes et al., 2012); moreover, PDSI spatial correlation analysis shows that the greatest intensity of this drought was concentrated in those states (Figure 7A). Drought affectation in the mid-18th century, with a duration of almost 20 years (1751-1768), extended over a large part of northern Mexico and southern United States (Figure 7B), corroborating that reported by Cook et al. (2009) and Cerano-Paredes et al. (2011a). On the other hand, the effects of one of the widely documented droughts known as "El Año del Hambre (1785-1786)", considered the most severe event in the history of agriculture in New Spain (Therrell, Stahle, Villanueva-Díaz, Cornejo-Oviedo, & Cleaveland 2006), were less severe in the northern part compared to central Mexico, a region that suffered from famine and epidemics, with more than 300 000 people dying (Florescano, 1986).

The beginning of the armed struggle for Mexican independence was preceded by droughts and bad harvests in 1807, 1808 and 1809, which, in 1810, caused excessively high maize prices, food shortages and crises throughout the viceroyalty (O'Hara & Metcalfe, 1997). This created unemployment in the countryside and cities, social tension in the populations, epidemics and popular riots (Florescano, 1980). The reconstruction of precipitation variability for the western part of Chihuahua, as well as the spatial correlation analysis with the PDSI (Figure 7C) confirm this information, as it shows that the entire territory of New Spain was severely affected.

For the last years of the period known as “El Porfiriato” (1908-1911), below-average precipitation was again reconstructed. This drought caused massive food shortages, which subsequently led to social unrest (Irby et al., 2013). Florescano (1980) points out a close relationship between the drought conditions that affected mainly Querétaro and Hidalgo in the center of the country and Chihuahua, Coahuila and Nuevo León in the northern part, with the outbreak of the Mexican Revolution, key locations where insurrection took place.

The droughts of the 1930s, 1950s and 1990s have been recorded in other paleoclimatic reconstructions of Chihuahua (Cerano-Paredes et al., 2009; Irby et al., 2013; Villanueva-Díaz et al., 2015) recording water shortages, crop and livestock loss. Spatial correlation analysis (Figure 7D) shows that the mid-20th century drought impacted much of the southwestern United States and northern Mexico (Cleaveland et al., 2003; O'Hara & Metcalfe, 1997). This episode caused a great economic crisis in the rural areas of northern Mexico, causing tens of thousands of farmers to migrate to large Mexican cities and to the United States in search of opportunities (Florescano, 1980).

Severe droughts have been observed every mid-century and between the end or beginning of each century, a recurrence of 50 years (Figure 6), which synchronizes with results from other studies for both central (Therrell et al., 2006) and northern Mexico (Cerano-Paredes et al., 2011b, 2012; Chávez-Gándara et al., 2017). The most severe drought episodes reconstructed for the western part of Chihuahua synchronize with events recorded in most of northern and central Mexico and southern United States, with variations in intensity between regions. This can be attributed to large-scale climate phenomena affecting geographic regions with different intensity.

ENSO impact on precipitation

ENSO strongly influences the interannual variability of the winter-spring climate of northern Mexico (Magaña-Rueda, 1999; Stahle et al., 1998). The prolonged dry and wet regimes in this region reflect the high relationship between decadal variability and annual variance in both cool and warm season precipitation (Cleaveland et al., 2003). Cool-season precipitation in northern Mexico and the southwestern United States correlates significantly with ENSO (Magaña-Rueda, Vázquez, Pérez, & Pérez, 2003). This teleconnection of ENSO with the climate of northern Mexico is clearly recorded in the growth of trees in the Sierra Madre Occidental (Cerano-Paredes et al., 2009; Chávez-Gándara et al., 2017; Cleaveland et al., 2003).

The October-June precipitation reconstruction contain has a strong ENSO signal in the time and frequency domains. Spectral ondelta coherence analysis shows that the precipitation variability is significantly modulated by ENSO at different time and frequency scales over the last 100 years (Figure 8). Significant coherence was determined between the reconstructed precipitation and the Niño 3.4 index. According to Gergis and Fowler (2009), the last three most important drought events in this northern region of the country at the beginning, middle and end of the 20th century and the beginning of the 21st century are significantly related to ENSO. Trenberth (1997) notes that two of the most intense ENSO events in its cold phase occurred in the mid-20th century (1950-1951 and 1954-1956) and caused significant reductions in precipitation and severe droughts in a large part of northern Mexico, forcing tens of thousands of farmers to migrate (Florescano, 1980).

Conclusions

The reconstruction of 359 years of October-June precipitation variability for the western part of Chihuahua has provide an analysis of the recurrence of extreme events every 50 years and the influence of ENSO on the hydroclimatology of the region. Currently, thanks to technological advances in ENSO monitoring, its behavior can be known in advance and, consequently, the positive or negative impact it will have on the region, which is related to a greater or lesser recharge of aquifers. The generation of knowledge on climate variability for this region provides greater elements on an annual and multidecadal scale to contribute to the sustainable management of natural resources, especially the water that is captured in these forests and drains to the Yaqui river basin, a necessary resource for the agricultural and livestock development of the region which thousands of people need for their survival.

Acknowledgments

This research was funded under the project “Procesos de mitigación del impacto por cambio climático en la agricultura del sur de Sonora” with code 1450034669. Thanks to Área de Protección de Flora y Fauna “Campo Verde” for granting permission to access and for helping during the development of the research. Thanks to Ing. José Luis González Castañeda, Ing. Edgar Guzmán Arreola and Jorge Adrián Ochoa Chacón for their valuable collaboration during the field study.

References

Cerano-Paredes, J., Villanueva-Díaz, J., Fulé, P. Z., Arreola-Ávila, J. G., Sánchez-Cohen, I., & Valdez-Cepeda, R. D. (2009). Reconstrucción de 350 años de precipitación para el suroeste de Chihuahua, México. Madera y Bosques, 15(2), 27-44. Retrieved from http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S1405-04712009000200002Links ]

Cerano-Paredes, J., Villanueva-Díaz, J., Valdez-Cepeda, R. D., Constante-García, V., González-Barrios J. L., & Estrada-Ávalos J. (2012). Precipitación reconstruida para la parte alta de la cuenca del Río Nazas, Durango. Revista Mexicana de Ciencias Forestales, 3(10), 7-23. Retrieved from http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S2007-11322012000200002Links ]

Cerano-Paredes, J., Villanueva-Díaz, J., Valdez-Cepeda, R. D., Cornejo-Oviedo, H. E., Sánchez-Cohen, I., & Constante-García, V. (2011a). Variabilidad histórica de la precipitación reconstruida con anillos de árboles para el sureste de Coahuila. Revista Mexicana de Ciencias Forestales, 2(4), 33-47. Retrieved from http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S2007-11322011000200004Links ]

Cerano-Paredes, J., Villanueva-Díaz, J., Valdez-Cepeda, R. D., Méndez-González, J., & Constante-García, V. (2011b). Sequías reconstruidas en los últimos 600 años para el noreste de México. Revista Mexicana de Ciencias Agrícolas, 2(1), 235-249. Retrieved from http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S2007-09342011000800006Links ]

Chávez-Gándara, M. P., Cerano-Paredes, J., Nájera-Luna, J. A., Pereda-Breceda, V., Esquivel-Arriaga, G., Cervantes-Martínez, R., & Corral-Rivas, S. (2017). Reconstrucción de la precipitación invierno-primavera con base en anillos de crecimiento de árboles para la región de San Dimas, Durango, México. Bosque, 38(2), 387-399. doi: 10.4067/S0717-92002017000200016 [ Links ]

Cleaveland, M. K., Stahle, D. W., Therrell, M. D., Villanueva-Díaz, J., & Burns, B. T. (2003). Tree-ring reconstructed winter precipitation and tropical teleconnections in Durango, Mexico. Climate Change, 59, 369-388. doi: 10.1023/A:1024835630188 [ Links ]

Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO). (2019). Portal de geoinformación. Sistema Nacional de Información sobre Biodiversidad. Retrieved from http://www.conabio.gob.mx/informacion/gis/Links ]

Cook, E. R. (1987). The decomposition of tree-ring series for environmental studies. Tree-Ring Bulletin, 47, 37-59. Retrieved from http://hdl.handle.net/10150/261788Links ]

Cook, E. R., Seager, R., Heim Jr, R. R., Vose, R. S., Herweijer, C., & Woodhouse, C. (2009). Megadroughts in North America: placing IPCC projections of hydroclimatic change in a long-term palaeoclimate context. Journal of Quaternary Science, 25(1), 48-61. doi: 10.1002/jqs.1303 [ Links ]

D’Arrigo, R. D., & Jacoby, G. C. (1991). A 1000-year record of winter precipitation from northwestern New Mexico, USA: A reconstruction from tree-rings and its relation to El Niño and the Southern Oscillation. The Holocene, 1(2), 95-101. doi: 10.1177/095968369100100201 [ Links ]

Endfield, G. H., & Fernández-Tejedo, I. (2006). Decades of drought, years of hunger: archival investigations of multiple year droughts in late colonial Chihuahua. Climate Change, 75, 391-419. doi: 10.1007/s10584-006-3492-7 [ Links ]

Florescano, E. M. (1980). Análisis histórico de las sequías en México. México. Secretaría de Agricultura y Recursos Hidráulicos. [ Links ]

Florescano, E. M. (1986). Precios del maíz y crisis agrícolas en México: 1708-1810. México: Editorial Era. [ Links ]

Fritts, H.C. (2001). Tree rings and climate. Tucson, AZ, USA: The University of Arizona Press. [ Links ]

García, E. (1970). Modificaciones al sistema de clasificación climática de Köppen para adaptarlo a las condiciones de la República Mexicana. México: Instituto de Geografía, Universidad Nacional Autónoma de México (UNAM). [ Links ]

García-Acosta, V. (1993). Las sequías históricas de México. Desastres y sociedad. Desastres y Sociedad, 1, 1-18. Retrieved from https://www.desenredando.org/public/revistas/dys/rdys01/dys1-LSHM-oct-24-2001.pdfLinks ]

Gergis, J. L., & Fowler, A. M. (2009). A history of ENSO events since A.D. 1525: implications for future climate change. Climate Change, 92, 343-387. doi: 10.1007/s10584-008-9476-z [ Links ]

González-Elizondo, M. S., González-Elizondo, M., Tena-Flores, J. A., Ruacho-González, L., & López-Enríquez, I. L. (2012). Vegetación de la Sierra Madre Occidental, México: una síntesis. Acta Botánica Mexicana, 100, 351-403. doi: 10.21829/ABM100.2012.40 [ Links ]

Grissino-Mayer, H. D. (2001). Evaluating crossdating accuracy: A manual and tutorial for the computer program COFECHA. Tree-Ring Research, 57(2), 205-221. Retrieved from http://hdl.handle.net/10150/251654Links ]

Grupo Intergubernamental de Expertos sobre el Cambio Climático (IPCC). (2019). Resumen para responsables de políticas. Calentamiento global de 1,5 °C, Informe especial del IPCC sobre los impactos del calentamiento global de 1,5 °C con respecto a los niveles preindustriales y las trayectorias correspondientes que deberían seguir las emisiones mundiales de gases de efecto invernadero, en el contexto del reforzamiento de la respuesta mundial a la amenaza del cambio climático, el desarrollo sostenible y los esfuerzos por erradicar la pobreza. Suiza: Author. Retrieved from https://www.ipcc.ch/site/assets/uploads/sites/2/2019/09/SR15_Summary_Volume_spanish.pdfLinks ]

Guibal, F., & Guiot, J. (2021). Dendrochronology. In G. Ramstein, A. Landais, N. Bouttes, P. Sepulchre, & A. Govin (Eds.), Paleoclimatology (pp. 117-122). Cham: Springer. doi: 10.1007/978-3-030-24982-3_8 [ Links ]

Holmes, R. L. (1983). Computer-assisted quality control in tree-ring dating and measurement. Tree-Ring Bulletin, 43(1), 69-78. doi: 10.1016/j.ecoleng.2008.01.004 [ Links ]

Holmes, R. L. (2001). Dendrochronology program library. Retrieved from https://www.ltrr.arizona.edu/software.htmlLinks ]

Instituto Nacional de Estadística y Geografía (INEGI). (2019). Simulador de flujos de agua de cuencas hidrográficas. Retrieved from http://antares.inegi.org.mx/analisis/red_hidro/siatl/#Links ]

Irby, C. M., Fulé, P. Z., Yocom, L. L., & Villanueva-Díaz, J. (2013). Dendrochronological reconstruction of long-term precipitation patterns in Basaseachi National Park, Chihuahua, Mexico. Madera y Bosques, 19(1), 93-105. Retrieved from https://www.redalyc.org/articulo.oa?id=61727444007Links ]

Lüttge, U. (2017). From dendrochronology and dendroclimatology to dendrobiochemistry. Trees, 31, 1743-1744. doi: 10.1007/s00468-017-1608-5 [ Links ]

Magaña-Rueda, V. O., Vázquez, J. L., Pérez, J. L., & Pérez, J. B. (2003). Impact of El Niño on precipitation in Mexico. Geofísica Internacional, 42(3), 313-330. Retrieved from http://areas.geofisica.unam.mx/geofisica_internacional/anteriores/2003/03/magana.pdfLinks ]

Magaña-Rueda, V. O. (1999). Los impactos de El Niño en México. México. SEP-CONACYT. Retrieved from http://centro.paot.org.mx/index.php/porinstituciones/otros-organismos/73-varios/745-los-impactos-de-el-nino-en-mexicoLinks ]

Magaña-Rueda, V. O., & Neri, C. (2012). Cambio climático y sequías en México. Ciencia, 63(4), 26-35. Retrieved from http://revistaciencia.amc.edu.mx/images/revista/63_4/PDF/sequiasMexico.pdfLinks ]

O’Hara, S. L., & Metcalfe, S. E. (1997). The climate of Mexico since the Aztec period. Quaternary International, 43/44, 25-31. doi: 10.1016/S1040-6182(97)00017-7 [ Links ]

Pokhrel, Y., Felfelani, F., Satoh, Y., Boulange, J., Burek, P., Gädeke, A., …Wada, Y. (2021). Global terrestrial water storage and drought severity under climate change. Nature Climate Change, 11, 226-233. doi: 10.1038/s41558-020-00972-w [ Links ]

R Core Team. (2019). R: A language and environment for statistical computing. Austria: R Foundation for Statistical Computing. [ Links ]

Rayner, N. A., Parker, D. E., Horton, E. B., Folland, C. K., Alexander, L. V., & Rowell, D. P. (2003). Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. Journal of Geophysical Research, 108(D14). doi: 10.1029/2002JD002670 [ Links ]

Reyes-Gómez, V. M., Díaz, S. C., Brito-Castillo, L., & Núñez-López, D. (2013). ENSO drought effects and their impact in the ecology and economy of the state of Chihuahua, Mexico. In A. Yáñez-Arancibia, & R. Dávalos-Sotelo (Eds.), Ecological dimensions for sustainable socio economic development (pp. 102-117). USA: WIT Press. Retrieved from https://www.witpress.com/Secure/elibrary/papers/9781845647568/9781845647568007FU1.pdfLinks ]

Servicio Meteorológico Nacional (SMN). (2019). Información climatológica. Retrieved from https://smn.conagua.gob.mx/es/climatologia/informacion-climatologica/normales-climatologicas-por-estadoLinks ]

Speer, J. H. (2010). Fundamentals of tree-ring research. Tucson, AZ, USA: The University of Arizona Press. doi: 10.1002/gea.20357 [ Links ]

Stahle, D. W., Cook, E. R., Burnette, D. J., Villanueva-Díaz, J., Cerano-Paredes, J., Burns, J. N., …Howard, I. M. (2016). The Mexican Drought Atlas: Tree-ring reconstructions of the soil moisture balance during the late pre-Hispanic, colonial, and modern eras. Quaternary Science Reviews, 149, 34-60. doi: 10.1016/j.quascirev.2016.06.018 [ Links ]

Stahle, D. W., D’Arrigo, R. D., Krusic, P. J., Cleaveland, M. K., Cook, E. R., Allan, R. J., …Thompson, L. G. (1998). Experimental dendroclimatic reconstruction of the Southern oscillation. Bulletin of the American Meteorological Society, 79(10), 2137-2152. doi: 10.1175/1520-0477(1998)079<2137:EDROTS>2.0.CO;2 [ Links ]

Stahle, D. W., Villanueva-Diaz, J., Burnette, D. J., Cerano-Paredes, J., Heim Jr., R. R., Fye, F. K., …Stahle, D. K. (2011). Major Mesoamerican droughts of the past millennium. Geophysical Research Letters, 38(5), L05703. doi: 10.1029/2010GL046472 [ Links ]

Stokes, M. A., & Smiley, T. L. (1996). An introduction to tree-ring dating. Tucson, AZ, USA: The University of Arizona Press. [ Links ]

Terán-Cuevas, A. R., Gutiérrez-Castillo, M. E., & Davydova-Belitskaya, V. (2010). Escenarios de lluvia en México. México: Instituto Politécnico Nacional. Retrieved from https://tesis.ipn.mx/jspui/handle/123456789/23294Links ]

The MathWorks Inc. (2015). MATLAB. Natick, MA, USA: The MathWorks Inc. [ Links ]

Therrell, M. D., Stahle, D. W., Villanueva-Díaz, J., Cornejo-Oviedo, H. E., & Cleaveland, M. K. (2006). Tree-ring reconstructed maize yield in central Mexico: 1474-2001. Climate Change, 74, 493-504. doi: 10.1007/s10584-006-6865-z [ Links ]

Trenberth, K. E. (1997). The definition of El Niño. Bulletin of the American Meteorological Society, 78(12), 2771-2777. doi: 10.1175/1520-0477(1997)078<2771:TDOENO>2.0.CO;2 [ Links ]

Villanueva-Díaz, J., Cerano-Paredes, J., Fulé, P. Z., Cortés-Montaño, C., Vázquez-Selem, L., Yocom, L. L., & Ruiz-Corral, J. A. (2015). Cuatro siglos de variabilidad hidroclimática en el noroeste de Chihuahua, México, reconstruida con anillos de árboles. Investigaciones Geográficas, 87, 141-153. doi: 10.14350/rig.44485 [ Links ]

Zywiec, M., Muter, E., Zielonka, T., Delibes, M., Calvo, G., & Fedriani, J. M. (2017). Long-term effect of temperature and precipitation on radial growth in a threatened thermo-Mediterranean tree population. Trees, 31, 491-501. doi: 10.1007/s00468-016-1472-8 [ Links ]

Received: August 20, 2021; Accepted: May 26, 2022

*Corresponding author: cerano.julian@gmail.com; tel.: +52 871 159 0104.

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