You are on page 1of 21

Journal of Hydrology 612 (2022) 128211

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Research papers

Analysis of recent rainfall trends and links to teleconnection patterns in


California (U.S.)
A. González-Pérez a, *, R. Álvarez-Esteban b, A. Penas c, S. del Río c
a
Department of Biodiversity and Environmental Management (Botany Area), Faculty of Biological and Environmental Science, University of Leon, Campus de Vegazana
s/n, 24071 León, Spain
b
Department of Economics and Statistics (Statistics and Operations Research Area), Faculty of Economics and Business. University of Leon, Campus de Vegazana s/n,
24071 León, Spain
c
Department of Biodiversity and Environmental Management (Botany Area), Faculty of Biological and Environmental Science, University of Leon, Mountain Livestock
Institute CSIC-UNILEON, Campus de Vegazana s/n, 24071 León, Spain

A R T I C L E I N F O A B S T R A C T

This manuscript was handled by Emmanouil The aim of this research is to enlarge knowledge of space-time evolution of precipitation in California over the
Anagnostou, Editor-in-Chief. last decade on a monthly, seasonal and annual time-scale. This study also analyses the relationship between
Original content: https://wrcc.dri.edu/ precipitation and teleconnection patterns with most influence on Californian climate. The homogeneity of the
data from the selected stations was verified and finally 165 meteorological stations were used with an obser­
Keywords: vation period that ranged from 1980 to 2019. In order to evaluate trends and statistical significance, both the
Precipitation non-parametric Mann-Kendall test and modified Sen’s slope method were used. Correlation analysis using the
Trends partial non-parametric Spearman Test (95% confidence level) was performed to find out relationships between
Teleconnection Patterns precipitation and nine teleconnection patterns in the State of California. Spatial analysis was achieved using
Regionalisation Empirical Bayesian Kriging (EBK). Finally, this research, as a novelty, shows regionalisation of California State as
California a function of significance in the correlation of precipitation with teleconnection patterns. To achieve this,
Principal Component Analysis (PCA) and Agglomerative Hierarchical Cluster analysis (HCA) was performed.
Results show a positive precipitation trend in winter and negative precipitation trend in late summer and
autumn. The teleconnection patterns more correlated with precipitation are El Niño along Southern Oscillation
(ENSO), Antarctic Oscillation (AAO) and North Atlantic Oscillation (NAO). Four areas were discerned outlining
the Mojave Climate Region (Area 4) by having ENSO as the teleconnection pattern that best account
precipitation.

1. Introduction damage each year (Cheung et al., 2016; Davenport et al., 2021) partic­
ularly in California, where in 2017 extreme precipitation and flooding
Climate change has been drawing the scientific community’s atten­ led to damage worth around $1.5 billion (Mallakpour et al., 2019).
tion for more than three decades (Stoddard et al., 2021). In this regard, Incidentally, and including other issues, during flood events public
the increase in the frequency of extreme weather events has boosted water systems fail to provide fresh water due to the destruction of
economic losses over recent decades due to tropical cyclones, winter treatment plants located near/ in close proximity to rivers (Exum et al.,
storms, floods, droughts, heatwaves and wildfires (Beniston, 2007; 2018). Another important factor leading to major economic and social
Beniston et al., 2021; Ehsani et al., 2020; Roksvåg et al., 2021). In fact, losses are aviation accidents. Most of these accidents are caused by
the occurrence and intensity of rainstorms has increased since the 1950s weather conditions such as fog and precipitation that reduce visibility
around the world, where trend analysis data are obtainable, owing to the (Gultepe et al., 2019).
surge in human CO2 emissions (Dong et al., 2021; Fischer and Knutti, At this point, it is important to mention that precipitation is any
2015; Held and Soden, 2006; Masson-Delmotte et al., 2021). One of the creation of the condensation of atmospheric water vapour that falls
most expensive natural hazards is flooding, causing billions of dollars in rapidly from a cloud. The principal forms include; drizzle, rain, sleet and

* Corresponding author.
E-mail addresses: agonp@unileon.es (A. González-Pérez), ramon.alvarez@unileon.es (R. Álvarez-Esteban), apenm@unileon.es (A. Penas), sriog@unileon.es (S. del
Río).

https://doi.org/10.1016/j.jhydrol.2022.128211
Received 17 March 2022; Received in revised form 4 July 2022; Accepted 10 July 2022
Available online 19 July 2022
0022-1694/© 2022 Elsevier B.V. All rights reserved.
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

snow (NASA Goddard Space Flight Center, 2018). In this regard, it is


important to identify that ice fog (suspended ice crystals) can affect local
climates due to the alteration of surface albedo. In fact, ice fog modu­
lates the heat and the moisture fluxes in the troposphere thereby altering
the climate (Gultepe et al., 2017). It is widely known that precipitation is
an important part of the water cycle and at the same time probably-one
of the main variables associated with atmospheric circulation in climate
studies (Sun et al., 2018). Moreover, snow precipitation can be either in
a liquid or ice phase, and plays an important role in the water cycle of
the mountain’s ecosystem (Gultepe, 2015). Precipitation and clouds,
along with water vapour mass exchanges, play a major role in climate
fluctuations at both global and regional levels (Levizzani and Cattani,
2019). It is of note that, not only do they make an impact on the climate,
but also on the weather at every scale, and affect water obtainability
(Levizzani and Cattani, 2019). In addition, precipitation is the primary
source of fresh water for numerous systems such as agriculture and
ecosystems (Lausier and Jain, 2018). Therefore, changes in precipitation
could cause unprecedented consequences for society, agriculture and
the environment (Guo et al., 2019).
Foreseeing precipitation trends is important for a country’s future
economic development (Ahmad et al., 2015). Carvalho (2020) have
claimed that precipitation is expected to increase in wet regions and
decrease in dry regions. Nevertheless, there is a growing body of liter­
ature that identifies precipitation trends becoming less predictable at
finer spatial scales even in observational data (Gultepe et al., 2016),
because of the increasingly strong influence of climate variability in
different regions (de Luis et al., 2009; Jiang et al., 2016; Lausier and
Jain, 2018; Mohammad and Goswami, 2019; Torres-Batlló and Martí-
Cardona, 2020; Treppiedi et al., 2021) and specifically in the United
Fig. 1. California physical map with the meteorological stations used in this States (Davenport et al., 2021; Powell and Keim, 2015; Sayemuzzaman
research (●). Some cities of the State are named to ease localization. and Jha, 2014). Precipitation differs greatly due to both topographic
effects and mesoscale atmospheric dynamical processes (Huang et al.,
2020a). It is important to mention that some studies agree that a

Table 1
Results of rainfall trends (mm year− 1) in the whole State of California. *Statistical significance (p-value) at α = 0.05.
January February March April May June July August September

Slope − 0.15 − 0.04 − 0.71 +0.28 +0.13 − 0.04 − 0.02 − 0.04 − 0.08
p-value 0.82 0.98 0.29 0.31 0.47 0.37 0.37 0.18 0.37

October November December Winter Spring Summer Autumn Annual

Slope − 0.35 − 0.68 +0.24 +0.02 − 0.06 − 0.06 − 0.38 − 0.03


p-value 0.27 0.18 0.68 0.98 0.86 0.21 *0.05 0.82

Fig. 2. Percentage of meteorological stations with positive and negative trends (monthly, seasonal and annual scale) in the State of California over the
period 1980–2019.

2
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 3. Boxplot for monthly, seasonal and annual trend values in the State of California over the period 1980–2019. The black line in the boxes (Interquartile range)
shows the median value slope for the entire State.

Fig. 4. Positive Trend analysis result and example of modified Sen slope for Yosemite Park meteorological station in the State of California over the
period 1980–2019.

reduction in the frequency of warm-season precipitation combined with atmospheric pattern that occurs over months to years as a result of
anthropogenic impact could lead to an increase in forest fires in the State changes in surface fluxes and covers widespread geographic areas (Yuan
of California (Abatzoglou and Williams, 2016; Diffenbaugh et al., 2015; et al., 2020). Additionally, sometimes these can last for several
Williams et al., 2019). Understanding the variability and trends in pre­ consecutive years. It is worth noting that teleconnection patterns have
cipitation could help to improve water resource management in been associated with alterations in seasonal rainfall in California (Zhang
different parts of the world such as California, where frequent droughts et al., 2021). In fact, there are some patterns of teleconnection that may
occur (He and Gautam, 2016; Serra-Llobet et al., 2016). In addition, be related to precipitation and do not occur in isolation (Wise et al.,
several research papers have shown a huge loss of snowpack due to early 2015). Together these research papers provide important insights into
thaws and more rain instead of snow (Huang et al., 2020a). Some re­ the understanding of climate unpredictability on a global scale including
searchers have claimed that drought risk and the decrease in precipi­ North America (Ge and Gong, 2009; Lau and Weng, 2002; Trenberth
tation in California are due to anthropogenic forcing and the greenhouse et al., 2014) and California (Abatzoglou, 2011; Allen and Anderson,
effect (Dasgupta et al., 2020; Diffenbaugh et al., 2015; Swain et al., 2018; Guan et al., 2013; Sheppard et al., 2002; Zhang et al., 2021; Zhou
2016). et al., 2020). Those research projects establish the main large-scale
Moreover, physical and dynamical processes force changes in pre­ teleconnection patterns that could affect the United States and, in
cipitation. For this reason, several studies have focused on the re­ particular, the State of California. They are: Antarctic Oscillation (AAO)
lationships between precipitation and teleconnection patterns. The term (Jian-Qi, 2010), Arctic Oscillation (AO) (Dai and Tan, 2017), North
teleconnection pattern is often defined as a large-scale repetitive Atlantic Oscillation (NAO) (Hurrell and Deser, 2010), Pacific-North

3
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 5. Negative Trend analysis result and example of modified Sen slope for Yosemite Park meteorological station in the State of California over the
period 1980–2019.

Fig. 6. Results of the geographic coverage of rainfall trends (mm year− 1) and their statistical significance (monthly, seasonally and annually). Area that shows
statistical significance (95%).

American Pattern (PNA) (Wallace and Gutzler, 1981), Madden Julian Oscillation (ENSO) (Whan and Zwiers, 2016) and Pacific Decadal
Oscillation (MJO) (Zhou et al., 2020), West Pacific Oscillation (WPO) Oscillation (PDO) (McAfee, 2014; Newman et al., 2016). Additional
that is West Pacific (WP) (Choi and Moon, 2012), previously described observational work in combination with modelling studies is essential to
by Barnston et al. (1991) and J. Wallace and Gutzler, (1981), Eastern acquire knowledge of the relationship between precipitation in Cali­
Pacific Oscillation (EPO) (Dai and Tan, 2019a), El Niño along Southern fornia and teleconnection patterns (Luković et al., 2021). As has been

4
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

previously mentioned, examining precipitation is an enduring concern 2. Materials and methodology


both in the long term and within extreme weather events.
This research is original and it is important to highlight that it shows 2.1. Study area
the most recent results in rainfall trends in California on a monthly,
seasonal and annual time scale. The leading aim of this study is to This research is conducted in the State of California, one of largest
investigate the space–time progression of precipitation at a regional states in the U.S. that covers an area of 423,955 km2. This State exhibits
level, in order to enlighten, on the one hand, the trend and its statistical two main mountain ranges, the Sierra Nevada and the Coast Range
significance for the entire State of California for the period of which give California a varied landscape. The Coast Range goes from the
1980–2019 at a monthly, seasonal and annual rate. On the other hand, northwest right down to the Mexican border, spanning 1300 km. The
this study, ground-breaking to our knowledge, examines the relation­ longest and largest range in California is the Sierra Nevada, where
ship between precipitation and up to a total of nine teleconnection Mount Whitney at 4421 m, is California’s highest peak (Luteyn and
patterns with its most probable influence on the Californian climate Hickman, 1993). In addition, the State has two distinct desert zones,
during the period mentioned above. It provides the possible links be­ Mojave and Colorado. California boasts several major rivers and lakes
tween different teleconnection patterns on rainfall in California. such as the Sacramento River and Lake Tahoe (Fig. 1). The Californian
Another unique feature of this research is to achieve objective region­ climate is highly variable, although a mediterranean climate prevails all
alisation by clustering meteorological stations according to their links over the State, except for the highland area of Klamath where there is a
with rainfall recorded in California and the teleconnection patterns temperate climate (Rivas-Martínez et al., 2011). Moreover, in the
selected. southeast of this State, there is a somewhat tropical climate in the
Finally, it is considered that the conclusions of this might support Sonoran Desert (Killam et al., 2014). Its complex topography and great
managers of land planning. Not only will they be able to put into latitudinal extension favour a wide variety of climates from desert to
practice confidence-building measures to tackle global warming, but subalpine environments (Pathak et al., 2018). Thus, its proximity to the
also to take measures in order to overcome the adverse effects of this Pacific coast is one of the determining factors in the climate of the State.
phenomenon. Subsequent work is needed to evaluate how trends may
evolve in response to future climate variability. Moreover, it will
become possible to anticipate future climate change. The recognition of 2.2. Data
upcoming trends across a series of thresholds allows for risk manage­
ment with more suitable adaptation goals. Rainfall values from the climatic database of the meteorological
stations were used and were available on the WRCC website (“Western
Regional Climate Center,” 2020) from 1980 to 2019. Initially, monthly
rainfall (mm) data were selected from 350 stations. The meteorological
stations were chosen due to criteria based on completeness, length and
homogeneity in most of the State (He and Gautam, 2016). Only stations

5
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

with less than 10 % of missing values were chosen. The gaps were (He and Gautam, 2016; Kukal and Irmak, 2016; Peña-Angulo et al.,
completed with the corresponding monthly long-term mean value (Ríos 2021; Sarricolea et al., 2019) particularly in hydrological ones (Ahmad
Cornejo et al., 2015). After that, analysis of the homogeneity of the data et al., 2015).
was tested (Blöschl et al., 2019; Gocic and Trajkovic, 2013; Karmeshu, In many areas of the globe it is complicated to assess the climate
2015; Song et al., 2019). This study was determined by the Run test since there is a lack in the number of weather stations available (Díaz-
(Thom, 1966) at a 95 % confidence level. This test is recommended by Padilla et al., 2011). Therefore, statistical interpolation of the values is
the WMO, because it is not necessary for the analysed series to follow necessary for some specific regions. Empirical Bayesian Kriging (EBK)
normal criteria and it has also been used previously in other climatic was used as a method of interpolation of geographic values as it is su­
studies (del Río et al., 2013; Meseguer-Ruiz and Sarricolea, 2017). perior to other predictors, and progressively so with data complexity. In
Finally, 165 stations were chosen for the study, adding their altitude addition, it provides both a straightforward and robust method of data
values and geographic coordinates (Fig. 1) to enable the results be interpolation (Barber et al., 2017; Gribov and Krivoruchko, 2020; Kri­
represented on maps, as will be discussed below. Moreover, for each one voruchko, 2012). To conduct this interpolation, ArcGIS 10.8© (Envi­
of the meteorological stations, seasonal and annual rainfall values were ronmental Systems Research Institute (ESRI), 2019) software was used
calculated. It is widely known that there are four seasons in California: and, to be more precise, an EBK geoprocessing tool was applied.
Spring (March, April and May), Summer (June, July and August), Furthermore, 17 rainfall trend contour maps were created using the
Autumn (September, October and November) and Winter (December, same software and areas with statistical significance (95 % confidence
January and February). Henceforth those months will be named MAM, level) were overlaid onto the contour maps.
JJA, SON, DJF. The annual rainfall series in California were computed
using Voronoi polygons, considering the weather stations in accordance 2.4. Selecting the atmospheric teleconnection patterns
with the area of each polygon (Hijmans et al., 2021).
Similar to other research (Cordero et al., 2011; Guirguis et al., 2019;
2.3. Trend analysis Yu et al., 2019) values of the atmospheric circulation pattern indices
were selected from the Climate Prediction Center available on the NOAA
Firstly, to attain the slope results from the selected stations, we National Climatic Data Center (NCDC) website (https://www.ncdc.
applied a modified Sen’s slope method, and the Mann-Kendall test (Liu noaa.gov/teleconnections/). It must be pointed out that the Real-Time
et al., 2020) with R package version 4.1.0. The Sen slope estimator is a Multivariate Madden-Julian Oscillation (RMM) indices were obtained
non-parametric technique that estimates variations per unit of time in a from the Australian Bureau of Meteorology (https://www.bom.gov.
series where there is a linear trend. Trend analysis was carried out on a au/climate/mjo/graphics /rmm.74toRealtime.txt). It is important to
monthly, seasonal and annual basis. Secondly, to estimate trends and mention that Madden-Julian Oscillation (MJO) is both a global-scale
statistical significance, the non-parametric Mann-Kendall test was used, complex teleconnection pattern of tropical atmosphere and the domi­
this Kendall tau test is one of the most commonly applied non- nant mode of inter-seasonal tropical variability (Zheng et al., 2018). The
parametric tests for noticing trends in environmental time series data, real-time multivariate index (RMM) is a commonly used index for

6
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

defining MJO. The two principal components are RMM1 and RMM2 and meteorological stations whose precipitation is related to the same tele­
they have been shown to be suitable indices of MJO and related vari­ connection patterns studied. The PCA was done based on eigenvalues
ability (Wheeler and Hendon, 2004). Consequently, MJO is a two- >1, and the HCA was conducted by the Ward Method and squared
dimensional phase space (RMM1 and RMM2) pattern and their union Euclidean distance. Once the number of both clusters and stations
results in 8 equatorial phases. belonging to each cluster was obtained, a Voronoi interpolation was
The teleconnection patterns chosen first that can influence the carried out with the data in order to obtain the regionalised map. This
climate in the United States and in the State of California were: El Niño was then divided into areas according to the patterns that have shown
3.4 (ENSO) and Pacific Decadal Oscillation (PDO) (Jiménez-Quiroz, the most correlation with precipitation.
2014; Lee and Grotjahn, 2019). Due to the connections in atmospheric
rivers that may modify the Californian climate, indices for Antarctic 3. Results and discussion
Oscillation (AAO), Arctic Oscillation (AO), Madden-Julian Oscillation
(MJO), North Atlantic Oscillation (NAO), Pacific-North American 3.1. Rainfall trends
Pattern (PNA), West Pacific Oscillation (WPO) and East Pacific Oscil­
lation (EPO) were also considered. Rainfall results from the global analysis in California (Table 1) show
To discover relationships between rainfall and the teleconnection differences in trend directions in the entire State. In April (+0.28 mm
patterns selected the partial non-parametric Spearman Test was applied year− 1), May (+0.13 mm year− 1) and December (+0.24 mm year− 1) the
at confidence level of 95 % (Liu et al., 2020). Its simple interpretation, highest values of increases are found. It is worth noting that both
robustness, and ability to capture nonlinear correlations makes it a November and March (− 0.68 and − 0.71 mm year− 1) have shown a
perfect choice for detecting monotonic trends. This method gives less negative trend during the period studied. In addition, Autumn shared
importance to outliers and eliminates the effect that the time variable the same trend (− 0.38 mm year− 1) and is the only season in which the
can have on rainfall and teleconnection pattern variables, which thereby decrease is statistically significant.
avoids false relationships (Ríos Cornejo et al., 2015). Finally, using R Some of the results from examining rainfall trends (positive and
software, both the correlation results and their statistical significance negative ones) at monthly seasonal and annual levels from 1980 to 2019
were shown on monthly maps. in the State of California are shown in Fig. 2. In general, negative trends
are found in most of the months except for April, May and December
(Fig. 3) when positive trends outweigh the negative ones. As it can be
2.5. Regionalisation seen in Fig. 3 the median slope values are located below the zero line in
the majority of the months excluding May, June, July, August,
In order to obtain homogeneous areas both Agglomerative Hierar­ September and summer where the slope is zero or close to. It should be
chical Cluster (HCA) and Principal Component Analysis (PCA) were specified that in the summer period, precipitation trends do not exist due
utilized with all the correlations obtained using the non-parametric to the fact that precipitation in those months is 0 mm for most years in
Spearman test (Ríos Cornejo et al., 2015). These areas contain the

7
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

the period studied. This lack of precipitation in summer is similar to for any territories and, remarkably in our case, none of the study areas
Mediterranean climate territories around the world. have shown statistically significant trends annually, seasonally or
One particular example of how trends vary each month, season or monthly but for June and Autumn.
annually can be observed in the station at Yosemite and is represented in A decrease in the precipitation trend could be observed in March (92
both Figs. 4 and 5. In this meteorological station both positive (April) % of the stations). This is consistent with previous research conducted in
and negative (September) trends are shown with a red line. other parts of the world which showed a negative trend during this
It is also important to mention the difference between the results month (Boé and Terray, 2008; da Silva et al., 2015; Mosaffa et al., 2020;
presented in Table 1 and Fig. 6. Both of them show rainfall trends but the Río et al., 2011). In this study empirical proof has shown, over the period
former is for the whole State in a general approach and the latter is an studied, a negative trend of − 2 mm year− 1 in some points of Sierra
Empirical Bayesian Kriging interpolation of all the meteorological sta­ Nevada, curiously from Bridgeport to Mount Whitney (− 0.8 mm year− 1)
tions selected in this study, giving a detailed result for each part of the which constitute higher altitude areas. It is worth pointing out that the
State of California. snow accumulation period became shorter and for this reason the
As was mentioned above, Fig. 6, shows the geographical distribution snowpack accumulation in Sierra Nevada will decrease. (Guan et al.,
of rainfall trends on a monthly, seasonal and annual time-scale. The 2013). In addition, a downward trend was found in most of the territory
results of the spatial distribution trends in rainfall and their statistical except for in the vicinity of Eureka. In this area a positive trend was
significance on annual levels in California show two localized areas found of +0.8 mm year− 1. Turning now to results in April and May, they
discriminated by the trends they present. From Yreka to Sacramento and show increasing trends, which is the general pattern in most of the State
from the San Joaquin Valley over to San Francisco and the rest of (58 % and 63 % of the stations respectively, Fig. 2). During April, mainly
southern California, a negative rainfall trend is presented. In fact, 70 % in the northern territories from Yreka, the Klamath mountains (Asarian
of the meteorological stations studied have shown that trend. In the and Walker, 2016), Redding to Oroville, Placerville, Auburn and over to
aforementioned areas there has been a gradual decline in rainfall, up to Mount Whitney, there is a sharp increase (+1.4 mm year− 1) in rainfall. It
− 0.2 mm year− 1, which is consistent with other research (Polade et al., is remarkable that in May on the Coast Ranges, San Francisco, Sacra­
2017). However, northern areas of the State such as Coastal Ranges mento Valley and Northern California, a steady decline in rainfall
(+0.49 mm year− 1), the Klamath Mountains, Cascade Ranges (+0.20 (− 0.15 mm year− 1) can be seen.
mm year− 1), Susanville and Lake Tahoe (+0.40 mm year− 1) have shown If we focus on our results in spring (− 0.06 mm year− 1), they are
positive trends. As is mentioned above, several researchers have found consistent, up to a point, with a decreasing trend (− 0.1 cm decade− 1)
that daily precipitation events had shown a steady rise over the last five found in other research (He and Gautam, 2016). In our research, it has
decades, showing a difference between the north and the south in terms been found that in the north, apart from the Sacramento valley, a pos­
of precipitation (Berg and Hall, 2015; Killam et al., 2014; Swain et al., itive trend appears from +0.12 mm to +0.6 mm year− 1 (Fig. 6). This is
2016). According to previous research (He and Gautam, 2016) average entirely different to what happens in southern areas such as in San
precipitation has shown no statistically significant trends in any month Diego, the Laguna Mountains and Riverside where there is negative

8
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

rainfall up to − 0.6 mm year− 1. In this regard, 73 % of the meteorological − 0.62 mm/month/year for November (Luković et al., 2021). Our results
stations show negative trends. According to this data, water storage and (Fig. 6) are differentiated into regional isoclines where different gradi­
supply systems are essentially dependent on winter/spring snowpack ents are observed, unlike in the aforementioned research where the
(Diffenbaugh et al., 2015). In addition, rainfall decreases for the months values are for the entire State of California. The rainfall trend observed
of March through May which explains much of the variation in the area in autumn by other authors establishes that, based on observational
destroyed by fires each year. In this regard, years with low spring rainfall precipitation records from the last few decades, decreases in rainfall
meant a disproportionately large percentage of the total area was trends were considered during SON season since 1979 (Luković et al.,
devastated by fire (Dennison et al., 2008). 2021). That statement coincides with our results. The whole of the State
If we now turn to the results in June, July and August, no increase in of California displays a decrease in rainfall over the period of study (92
rainfall trends were found. For instance, 55 % of the stations in June % of the stations). The negative trend found in Sierra Nevada areas is
show negative trends. It is important to mention that during these meaningful, − 1.35 mm year− 1, and the fact that a statistically signifi­
months most of the State has values of 0 mm over the period studied. In cant negative trend appears in the surroundings of San Francisco (− 0.9
fact, in June there is a significant negative trend (− 0.1 mm year− 1) in mm year− 1) is of interest. This coincides with the research that claimed
the Mojave Desert. In that area the results showed a steady decline in the that there is a statically significant decrease in rainfall in autumn − 0.25
last month (− 0.35 mm year− 1), however, it is not statistically signifi­ mm/month/year and suggests lower precipitation conditions at the
cant. Related to summer, similar results (90 % of stations) have been beginning of the wet season in recent decades. (Luković et al., 2021). In
found although a negative trend (− 0.25 mm year− 1) appears in Sierra addition, a decrease in non-extreme precipitation dominates the drier
Nevada and this suggests an intensification of the probability of dry autumn and spring (Dong et al., 2019).
summers in the future. This will become more common throughout According to the monthly results in January a steady rise in rainfall
California (He and Gautam, 2016) and it will be prone to occasional was observed over Sierra Nevada (+0.3 mm year− 1) and it spreads to the
drought episodes. Droughts in Southwestern United States have been San Rafael Mountains, Santa Lucia Ranges and the San Gabriel Moun­
intensified by unusual warm summer temperatures and by the lack of tains up to +0.9 mm year− 1. Meanwhile in the southern part of the State
rainfall (Cayan et al., 2013). and inland, in the vicinity of the Mojave Desert, there is a negative trend
According to the results in rainfall trends during the months of in rainfall (− 0.6 mm year− 1). This trend is found in 60 % of the stations
September, October and November, negative values are found in each of and can be clearly observed in the San Francisco Bay (− 0.9 mm year− 1),
those months (75 %, 82 % and 95 % of the stations studied respectively). it also spreads north through the Sacramento Valley (− 0.6 mm year− 1)
Negative trends bestrew the whole of California but the southern areas to Yreka (− 0.3 mm year− 1). Moving to February, the results show that
show the lowest values. For instance, the vicinities of Mount Ritter there is a positive trend (+2.25 mm year− 1) in the northern half of the
showed the highest negative values in September, − 1.26 mm year− 1 State, particularly in Sierra Nevada areas close to Oroville (+1.85 mm
while in November up to 2.8 mm year− 1. This result coincides, up to a year− 1) and Auburn. In contrast, during this month the areas in the south
point, with other research that claimed a negative precipitation trend of have shown decreases (− 1.35 mm year− 1) in rainfall, mainly in the area

9
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

adjacent to Mount Witney (− 0.40 mm year− 1), Independence and will be life-threatening, in preparation for risk to water resources and
Mariposa. This result in February coincides with other research that ecosystems, risk of wildfire and other crucial matters (Cayan et al.,
northern-mid and high-latitude regions are expected to become wetter, 2013). Climate models show that extended warming and intensification
while southern-subtropical areas are likely to become drier (Berg and of seasonal drying trends in the future are likely to result in further rises
Hall, 2015). in extreme autumn fire weather situations all over California (Goss et al.,
Finally, some research explains that in response to global warming, 2020).
there is usually a dry winter throughout the Mediterranean climate re­
gions as a consequence of poleward expansion in subtropical dry zones
(Dong and Leung, 2021). As can be observed in Fig. 6, a positive trend 3.2. Teleconnection patterns
(+1.2 mm year− 1) is shown in December except for in some coastal areas
with no trends, such as San Francisco Bay, Monterey, the San Rafael This section will show the results of the spatial and statistical analysis
Mountains, Diablo Range and San Luis Obispo. Moreover, the vicinities between rainfall and the nine teleconnection patterns chosen. Together
of El Centro show negative trends (− 0.6 mm year− 1). The trends are these results provide important insights into understanding the link that
mostly small, during the wet season, but there is a slight tendency to­ teleconnection patterns might have had on the rainfall over California
ward moisture being positive but similar to zero (+0.35 cm decade− 1) from 1980 to 2019. The results are summarised in Fig. 7 except for the
(Berg and Hall, 2015). Some research explains that more than half of the EPO data in December, which was not available.
annual precipitation drops in winter (Chang et al., 2015; He and Gau­ The results are also shown on monthly maps where the meteoro­
tam, 2016). It is worth noting that, in this research, the analysis of trend logical stations in red are significantly correlated (positive or negative)
anomalies in precipitation (low) during two consecutive winters, 2015/ with each one of the patterns studied (Figs. 8 to 17).
2016 and 2016/2017, with the influence of El Niño pattern, has been It is well known that Pacific Decadal Oscillation (PDO) affects the
considered (Wang et al., 2017). For example, during winter, in Sierra rainfall relationship with El Niño-Southern Oscillation (ENSO) in this
Nevada (California) a great number of storms are associated with At­ area with positive interference (Pavia et al., 2016). If we focus on our
mospheric Rivers (AR) (Guan et al., 2013). Moreover, AR will probably results, the meteorological stations which significantly correlate with
increase rainfall and snowmelt in Sierra Nevada (Huang et al., 2020b). PDO are located in northern coastal areas in April, May and August
Atmospheric rivers are modulated by teleconnection patterns (Guirguis except for in December where the correlations also appear in southern
et al., 2019), the following section elaborates on the correlations of coastal areas (Fig. 8). These results are consistent with other research in
precipitation with teleconnection patterns. which PDO/ENSO brings precipitation and takes place along the
Taking everything into consideration, these results suggest that the northern Californian coast and over Sierra Nevada (Guirguis et al.,
situation could be leading to a rise in seasonal rainfall in California. In 2019). It is important to observe that previous research work has
this regard, creating models of hydrological procedures already provide pointed out that in summer, over the centre and southern areas of Cal­
adequate spatial detail in order to evaluate hydroclimatic causes that ifornia, despite the lack of correlation (small or insignificant), the con­
tributions of these two patterns to causing precipitation are significant

10
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

(Cheng et al., 2021). In the same manner, as in other local research, Los addition, PDO, ENSO and PNA have been proven to have significant
Angeles has years where there is more than beyond average rainfall correlation with precipitation and groundwater fluctuations (Velasco
during a positive phase of PDO. This variability in precipitation is et al., 2017). It is worth noting that predicting winter precipitation over
influenced by ENSO and PDO whose relationships becomes stronger in a 3–4-week period in the southern and eastern US, it is highly correlated
the southern areas of the State (Killam et al., 2014). The increase in to NAO, ENSO 3.4 and PNA indices (Wang and Robertson, 2019).
winter pluvial risks over the entire State of California are connected with If we focus on North Atlantic Oscillation (NAO) (Fig. 11) it is the
positive phases of PDO while positive phases of ENSO affect Northern major large-scale teleconnection pattern over the extra-tropical Atlantic
California (Kam and Sheffield, 2016). Our study found that ENSO is Ocean (Whan and Zwiers, 2016). In this study, the results show that
likely to explain the precipitation trend in January, February and March February and May are the months with the highest correlation with
(Fig. 9), and accordingly, ENSO is a more meaningful predictor than precipitation (61.8 and 50.3 %). In addition, there is a lower value in
PDO over briefer periods, such as from October to March (Cheng et al., April (13.9 %) where the correlated meteorological stations are found
2021). In our study 18.1 % of the stations seem to be affected by ENSO in both in the coastal areas of Los Angeles, San Diego and Sierra Nevada
January and 11.5 % in both February and March. In other research, Ranges (Fig. 11). Decadal wet phases of the precipitation pattern in
Californian precipitation showed a strong correlation with extratropical California are associated with November to March circulation modes
cyclone weather conditions over the Eastern Pacific to the west of Cal­ such as low South-eastern pressure from the NAO pattern. This favours
ifornia, with positive correlation (0.8) (Chang et al., 2015). However, recurrent North Pacific winter storms shifting towards California (Cayan
Chang et al. (2015) mentioned that ENSO can only shed light on part of et al., 1998; Hurrell, 1995). In this regard, the NAO exerts a substantial
the link between Californian rainfall and extratropical cyclone activity, influence on precipitation in other Mediterranean climate type regions
due to the fact that the correlation between winter rainfall and el Niño (López-Moreno et al., 2011). NAO is a regional manifestation of a
3.4 index was just 0.38 between 1979/1980 and 2013/2014. hemispheric-wide mode known as Arctic Oscillation (AO). AO has been
Moving to the results of correlations between Pacific-North Amer­ defined as the atmospheric pressure at middle latitudes that fluctuates
ican teleconnection patterns (PNA) and precipitation, it is interesting between negative and positive phases. Its high frequency oscillation
that 37.0 % of the stations correlate in June (Fig. 10). Positive PNA from those phases impacts on both daily and weekly patterns (Brolley,
circulation phases make Northern California dry because of the associ­ 2007; Sellars et al., 2015). A typically positive (negative) AO phase is
ation of PNA with drier forms that seep into Northern California. This in linked with wetter (drier) conditions over California, in fact, AO has
turn leads to fewer dry conditions in Southern California and West been claimed to correlate with the forecast of precipitation differences
Coast/Southwest California (Lin et al., 2017). During the month of June, across California at least from 2015 to 2017 (Wang and Robertson,
most of the stations correlated with PNA are located in the central areas 2019). Moreover, this pattern correlates highly with the North Atlantic
of the State; this pattern might therefore be related to the distribution of Oscillation pattern (Walker and Bliss, 1932), and has been an issue of
rainfall. For example, as we saw in the previous section, there is a slight much interest over the last few years (e.g., Wallace and Gutzler, 1981;
decrease in the precipitation trend in June during the years of study. In Hurrell, 1995). Previous research has claimed that those two patterns

11
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 6. (continued).

Fig. 7. Heatmap with the percentage of meteorological stations that have shown positive (+) or negative (− ) statistical significance correlations between tele­
connection patterns and rainfall at a confidence level of 95 %l. Blank spaces are zero percentage values that have been removed for clarity. The top histogram (blue)
shows the percentage contribution of the teleconnection patterns to rainfall in the meteorological stations per month. The histogram in the right (green) shows the
percentage contribution of each teleconnection pattern. PDO (Pacific Decadal Oscillation), ENSO (El Niño-Southern Oscillation), PNA (Pacific-North American), NAO
(North Atlantic Oscillation), RMM1 and RMM2 (Real Multivariate MJO), EPO (East Pacific Oscillation), WPO (Western Pacific Oscillation), AO (Artic Oscillation) and
AAO (Antarctic Oscillation). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

12
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 8. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between PDO and rainfall. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 9. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between ENSO and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 10. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between PNA and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

correlate well together, there is a clear difference that could illustrate northern areas of the State, mainly in June.
how we attempt to unravel physical mechanisms in northern hemi­ According to our Madden–Julian Oscillation (MJO) results, RMM1
sphere atmospheric irregularity (Ambaum et al., 2001). According to seems to have a stronger relationship with precipitation than RMM2
this, our AO correlation results (Fig. 12) showed 29.1 % in the central- over the period studied (Figs. 13 and 14). It is important to note that

13
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 11. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between NAO and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 12. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between AO and rainfall. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 13. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between RMM1 and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

mainly meteorological stations located in San Francisco Bay and its vi­ et al., 2020).
cinities had shown any correlation. These results agree with other The Western Pacific Oscillation (WPO) pattern is one of the foremost
research that investigated the inter-annual variability in the range of teleconnection patterns over winter months in the Northern Hemisphere
MJO found, and apparently there are no direct and systematic re­ (Dai and Tan, 2019b). In addition, WPO/EPO is connected with more
lationships between extreme precipitation and this index (Jones, 2000). precipitation in California, which has been attributed to circulation
However, recent research has claimed that MJO would be suitable as a features that disturb the path and orientation of atmospheric rivers at
predictor of pentad precipitation at weeks 3–6 (15–42 days) (Nardi landfall (Guirguis et al., 2019). According to our results, EPO (Fig. 15)

14
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 14. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between RMM2 and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 15. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between EPO and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 16. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between WPO and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

seems to be connected with precipitation in October (40.6 %) and Despite these results, this teleconnection pattern mainly takes place in
related to trend results, the effect might mean a decrease in precipita­ the Southern Hemisphere. It was stated that positive phases could alter
tion. WPO (Fig. 16), on the other hand, could have a greater effect on the the circulation of MJO and ENSO in extratropical zones of the Pacific,
increase in precipitation over April and May (41.2 % and 21.2 % specifically in wintertime (Carvalho et al., 2005).
respectively) (Fig. 12). To sum up, it can be seen that the relationship between tele­
Finally, according to Antarctic Oscillation (AAO), high correlations connection patterns and precipitation is widespread all over the State of
have been found in February (54.5 %) and April (70.9 %) (Fig. 17). California. According to Table 2 and related to what has been stated

15
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 17. Percentage of meteorological stations in the State of California with significant correlations (positive in green, negative in red) between AAO and rainfall.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

%) stand out. The latter is the one that most correlates with precipitation
(9.1 %) in the month of September. EPO and WPO are the most corre­
lated in October while in November neither value is worth noting.
Finally, in December, the teleconnection patterns that could have the
greatest effect on rainfall are PDO (17.0 %) and PNA (16.4 %).

3.2.1. Regionalisation
The results of regionalisation show four different areas (clusters) as
can be observed in Fig. 18. The percentage of correlation for each area is
presented in Fig. 19 and Table 2 (Supplementary material). In the North
of the State (Area 1 in yellow) rainfall is highly correlated with AAO
(68.5 %), NAO (84.9 %) and EPO (72.6 %) (Fig. 19). It is important that
both of the former are related to rainfall in February while the latter are
in October. The Area is made up of 66 meteorological stations. Related
to the area in blue (Area 2), it consists of 45 meteorological stations and
it extends from the centre of California up to Los Angeles and to the
southern areas of Sierra Nevada. In this area both AAO (84.3 %) and
WPO (69.8 %) correlate in April. In addition, here precipitation is
correlated with NAO (77.8 %) in May and to a lesser extent with PNA
(52.4 %) in June. If we move to the green one (Area 4), that goes from
Death Valley to r the entire Mojave Desert and up to Oceanside (34
meteorological stations), ENSO is highly correlated with precipitation in
January, February, and March, (62.5 %, 58.33 % and 79.8 % respec­
tively) and also, to a lesser extent, in May (37.5 %). Interestingly, this
area resembles the Mojave Climate Region (Abatzoglou et al., 2009).
The southern area in red (Area 3) deals with Imperial Valley, the
southern Santa Rosa Mountains and the Mojave Desert. This area is
made up of 20 meteorological stations and there are no clear tele­
connection patterns which show high correlation to precipitation.
Fig. 18. Regionalisation of California according to teleconnection patterns that Related to AO in February, NAO in March and PDO in December merely
correlate the most with precipitation from 1980 to 2019. (Yellow = Area 1, is 40 % is shown (Fig. 18).
made up of 66 meteorological stations; blue = Area 2, is made up of 45
meteorological stations; green = Area 4, consist of 34 meteorological stations
4. Conclusions
and red = Area 3 is made up of 20 meteorological stations). (For interpretation
of the references to colour in this figure legend, the reader is referred to the web
version of this article.) The main purposes of this research were: firstly, to determine pre­
cipitation trends in California from 1980 to 2019, on monthly, seasonal
and annual levels. In addition, precipitation trend contour maps were
above, precipitation in January seems to correlate mainly with ENSO,
designed for 17 levels. Moreover, statistically significant areas were also
while in February AAO and NAO are the most correlated patterns with
overlaid on the contour maps. Secondly, to clarify the relationship be­
precipitation (54.6 % and 61.8 % respectively). In that month, EPO and
tween precipitation and up to nine teleconnection patterns whose results
ENSO show correlation to a lesser extent. It is significant that only ENSO
were given on a monthly basis. Thirdly, to highlight areas where the
correlates in March (11 %). Moving on to April, AAO is the most
teleconnection patterns are correlated with precipitation and to what
correlated pattern (70.3 %) followed by EPO (18.2 %) and NAO (13.9
extent.
%), the latter is the most correlated in May (50.3 %) followed by EPO
Finally, in accordance with the results, some of the significant out­
(21.2 %). During the summer, in June for instance, PNA is the most
comes of this research are specified along these lines.
correlated to precipitation in 37.0 % of the meteorological stations. In
August, the correlation values obtained by EPO (11.5 %) and MJO (7.9

16
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 19. Histograms of the percentage of meteorological stations in each area with significant correlations between teleconnection patterns and precipitation.

– The global analysis of precipitation in California shows broad spatial remaining territories, mainly in the North, have shown a positive
and seasonal differences in trend directions throughout the whole trend (+0.3 mm year− 1).
State. During the period studied the highest positive values are found – One of the most significant findings to emerge from this study is that
in April (+0.28 mm year− 1) and December (+0.24 mm year− 1), in rainy seasons both in winter and spring, the positive rainfall trend
while in March (− 0.71 mm year− 1) and November (− 0.68 mm is located in the northern half of the State, while a decrease in rainfall
year− 1) there is a negative trend. Both increases and decreases are was brought to the fore in the southern half.
established mainly in mountainous areas and consequently, in those – The teleconnection patterns that correlate most with precipitation in
areas, the difference in monthly rainfall rises, showing that the California are North Atlantic Oscillation (NAO), El Niño-Southern
precipitation phenomena happens over shorter periods of time. Oscillation (ENSO), East Pacific Oscillation (EPO), West Pacific
Moreover, on a seasonal scale, autumn has shown a significant Oscillation (WPO) and Antarctic Oscillation (AAO).
decrease (− 0.38 mm year− 1) in rainfall trends. – Monthly precipitation is correlated with teleconnection patterns in
– The data reported here appears to support the assumption that on an April, May, July, and August, showing at least a little correlation
annual level, from Yreka, to Sacramento and San Joaquin Valley, with every pattern studied. In December precipitation correlates
over San Francisco to the rest of southern California a negative with Pacific Decadal Oscillation (PDO) and Pacific-North American
precipitation trend in presented (− 0.2 mm year− 1), while the (PNA).

17
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Fig. 19. (continued).

– California is a territory, based on what is observed in this study, future research related to teleconnection patterns and meteorological
where rainfall trends have shown great variability throughout the phenomena. In addition, it is important to research local physics,
State due to its orography, mainly in mountainous areas. In addition, topography, thermodynamics and also to investigate other potential
this precipitation might be influenced by the behaviour of atmo­ driving factors of change. Moreover, the results of this study have a
spheric teleconnection patterns, due to the fact that it is a territory broad implication for the future, for example, knowing about the past
where teleconnection patterns commonly take place. However, there and trying to predict how precipitation will behave on a more detailed
are four different areas where Area 4, described above, has shown scale, than that which has been investigated previously, to our knowl­
high correlation values with ENSO (El Niño-Southern Oscillation) edge. In this regard, this research can help other professionals, such as
and rainfall over the period studied, during January, February and environmental managers and politicians to take the necessary measures
March. In Area 2, AAO and WPO correlate the most in April while in the face of precipitation processes, whereabouts they should take
NAO and PNA are in May and June respectively. In addition, in Area them and with what urgency. Processes that, due to climate change, are
1 rainfall correlates highly with AAO (68.5 %), NAO (84.9 %) and expected to worsen in the course of time.
EPO (72.6 %) while in Area 3 there is not a clearly teleconnection
pattern with more than 40 % of stations correlated. Data Availability

Finally, the authors consider that these findings provide insights for The datasets generated and analysed during the current study are not

18
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

publicly available due to the fact that R package is in the process of being References.
published. However, the original data source can be consulted on htt
ps://wrcc.dri.edu/ and are available from the corresponding author Abatzoglou, J.T., 2011. Influence of the PNA on declining mountain snowpack in the
Western United States. Int. J. Climatol. 31, 1135–1142. https://doi.org/10.1002/
upon reasonable request. JOC.2137.
Abatzoglou, J.T., Redmond, K.T., Edwards, L.M., 2009. Classification of regional climate
variability in the state of California. J. Appl. Meteorol. Climatol. https://doi.org/
Author contributions
10.1175/2009JAMC2062.1.
Abatzoglou, J.T., Williams, A.P., 2016. Impact of anthropogenic climate change on
Ángel Penas and Sara del Río contributed to the study’s conception wildfire across western US forests. Proc. Natl. Acad. Sci. 113, 11770–11775. https://
and design. Material preparation, data collection and analysis were doi.org/10.1073/PNAS.1607171113.
Ahmad, I., Tang, D., Wang, T., Wang, M., Wagan, B., 2015. Precipitation trends over time
carried out by Ramón Álvarez Esteban. The first draft of the manuscript using Mann-Kendall and Spearman’s rho tests in Swat River Basin, Pakistan. https://
was written by Alejandro González Pérez and all the authors commented doi.org/10.1155/2015/431860.
on previous versions of the manuscript. All authors read and approved Allen, R.J., Anderson, R.G., 2018. 21st century California drought risk linked to model
fidelity of the El Niño teleconnection. NPJ Clim. Atmos. Sci. https://doi.org/
the final manuscript. 10.1038/s41612-018-0032-x.
Ambaum, M.H.P., Hoskins, B.J., Stephenson, D.B., 2001. Arctic Oscillation or North
Ethical approval and responsibilities of Authors Atlantic Oscillation? J. Clim. 14, 3495–3507. https://doi.org/10.1175/1520-0442
(2001)014<3495:AOONAO>2.0.CO;2.
Asarian, J.E., Walker, J.D., 2016. Long-term trends in streamflow and precipitation in
All authors in this research are aware of the submission and declare Northwest California and Southwest Oregon, 1953–2012. J. Am. Water Resour.
that: Assoc. 52, 241–261. https://doi.org/10.1111/1752-1688.12381.
Barber, X., Conesa, D., López-Quílez, A., Mayoral, A., Morales, J., Barber, A., 2017.
Bayesian hierarchical models for analysing the spatial distribution of bioclimatic
– The manuscript is not submitted to any other journal for simulta­ indices. Sort 41, 277–296. https://doi.org/10.2436/20.8080.02.60.
neous consideration. Beniston, M., 2007. Linking extreme climate events and economic impacts: Examples
from the Swiss Alps. Energy Policy 35, 5384–5392. https://doi.org/10.1016/j.
– The submitted work is original and has not been published elsewhere enpol.2006.01.032.
in any form or language. Beniston, M., van der Wiel, K., Bintanja, R., 2021. Contribution of climatic changes in
– Results of this research are presented clearly, honestly, and without mean and variability to monthly temperature and precipitation extremes. Commun.
Earth Environ. 2, 1–11. https://doi.org/10.1038/s43247-020-00077-4.
fabrication, falsification or inappropriate data manipulation. Berg, N., Hall, A., 2015. Increased interannual precipitation extremes over california
– All authors agreed with the content and that all give explicit consent under climate change. J. Clim. 28, 6324–6334. https://doi.org/10.1175/JCLI-D-14-
to submission and publication. 00624.1.
Blöschl, G., Hall, J., Viglione, A., Perdigão, R.A.P., Parajka, J., Merz, B., Lun, D.,
– This work does not include research including humans or animals.
Arheimer, B., Aronica, G.T., Bilibashi, A., Boháč, M., Bonacci, O., Borga, M.,
Čanjevac, I., Castellarin, A., Chirico, G.B., Claps, P., Frolova, N., Ganora, D.,
CRediT authorship contribution statement Gorbachova, L., Gül, A., Hannaford, J., Harrigan, S., Kireeva, M., Kiss, A., Kjeldsen,
T.R., Kohnová, S., Koskela, J.J., Ledvinka, O., Macdonald, N., Mavrova-Guirguinova,
M., Mediero, L., Merz, R., Molnar, P., Montanari, A., Murphy, C., Osuch, M.,
A. González-Pérez: Methodology, Software, Formal analysis, Ovcharuk, V., Radevski, I., Salinas, J.L., Sauquet, E., Šraj, M., Szolgay, J., Volpi, E.,
Investigation, Data curation, Writing – original draft, Writing – review & Wilson, D., Zaimi, K., Živković, N., 2019. Changing climate both increases and
decreases European river floods. Nature 2019 573:7772 573, 108–111. https://doi.
editing, Visualization, Funding acquisition. R. Álvarez-Esteban: Soft­
org/10.1038/s41586-019-1495-6.
ware, Validation, Formal analysis, Resources, Data curation, Visualiza­ Boé, J., Terray, L., 2008. A weather-type approach to analyzing winter precipitation in
tion, Writing – review & editing. A. Penas: Conceptualization, France: Twentieth-century trends and the role of anthropogenic forcing. J. Clim. 21,
Validation, Resources, Writing – review & editing, Visualization, Su­ 3118–3133. https://doi.org/10.1175/2007JCLI1796.1.
Brolley, J.M., 2007. Effects of ENSO, NAO (PVO), and PDO on Monthly Extreme
pervision. S. del Río: Conceptualization, Validation, Resources, Writing Temperature and Precipitation. Florida State University.
– review & editing, Visualization, Supervision, Project administration. Carvalho, L.M.V., 2020. Assessing precipitation trends in the Americas with historical
data: A review. Wiley Interdiscip. Rev. Clim. Change 11. https://doi.org/10.1002/
WCC.627.
Declaration of Competing Interest Carvalho, L.M.V., Jones, C., Ambrizzi, T., 2005. Opposite phases of the Antarctic
oscillation and relationships with intraseasonal to interannual activity in the tropics
during the austral summer. J. Clim. 18 https://doi.org/10.1175/JCLI-3284.1.
The authors declare the following financial interests/personal re­ Cayan, D.R., Dettinger, M.D., Diaz, H.F., Graham, N.E., 1998. Decadal variability of
lationships which may be considered as potential competing interests: precipitation over Western North America. J. Clim. 11, 3148–3166. https://doi.org/
10.1175/1520-0442(1998)011<3148:DVOPOW>2.0.CO;2.
Alejandro Gonzalez reports financial support was provided by Junta de Cayan, D.R., Tyree, M., Kunkel, K.E., Castro, C., Gershunov, A., Barsugli, J., Ray, A.J.,
Castilla y Leon Consejeria de Educacion. Overpeck, J., Anderson, M., Russell, J., Rajagopalan, B., Rangwala, I., Duffy, P.,
Barlow, M., 2013. Future climate: Projected average. In: Assessment of Climate
Change in the Southwest United States: A Report Prepared for the National Climate
Data availability
Assessment. https://doi.org/10.5822/978-1-61091-484-0_6.
Chang, E.K.M., Zheng, C., Lanigan, P., Yau, A.M.W., Neelin, J.D., 2015. Significant
Data will be made available on request. modulation of variability and projected change in California winter precipitation by
extratropical cyclone activity. Geophys. Res. Lett. 42 https://doi.org/10.1002/
2015GL064424.
Acknowledgements Cheng, R., Novak, L., Schneider, T., 2021. Predicting the interannual variability of
California’s total annual precipitation. Geophys. Res. Lett. https://doi.org/10.1029/
2020GL091465.
This paper was supported by the European Regional Development Fund Cheung, W., Houston, D., Schubert, J.E., Basolo, V., Feldman, D., Matthew, R.,
(ERDF) and the Junta de Castilla y León (JCyL). The grant was awarded Sanders, B.F., Karlin, B., Goodrich, K.A., Contreras, S.L., Luke, A., 2016. Integrating
to the first author and included in a Fellowship Scheme for a Doctoral resident digital sketch maps with expert knowledge to assess spatial knowledge of
flood risk: A case study of participatory mapping in Newport Beach, California. Appl.
Training Program: Orden de 12 de diciembre de 2019 de la Consejería de
Geogr. 74, 56–64. https://doi.org/10.1016/J.APGEOG.2016.07.006.
Educación (extracto publicado en el B.O.C. y L. n.◦ 245, de 23 de diciembre. Choi, K.S., Moon, I.J., 2012. Influence of the Western Pacific teleconnection pattern on
BDNS (Identifi.): 487971. The Authors would like to thank Ruth J.R. Western North Pacific tropical cyclone activity. Dyn. Atmos. Oceans 57, 1–16.
Winter for her advice on English terminology. https://doi.org/10.1016/J.DYNATMOCE.2012.04.002.
Cordero, E.C., Kessomkiat, W., Abatzoglou, J., Mauget, S.A., 2011. The identification of
distinct patterns in California temperature trends. Clim. Change. https://doi.org/
Appendix A. Supplementary data 10.1007/s10584-011-0023-y.
da Silva, R.M., Santos, C.A.G., Moreira, M., Corte-Real, J., Silva, V.C.L., Medeiros, I.C.,
2015. Rainfall and river flow trends using Mann-Kendall and Sen’s slope estimator
Supplementary data to this article can be found online at https://doi. statistical tests in the Cobres River basin. Nat. Hazards. https://doi.org/10.1007/
org/10.1016/j.jhydrol.2022.128211. s11069-015-1644-7.

19
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

Dai, P., Tan, B., 2017. The nature of the Arctic oscillation and diversity of the extreme França, G.B., Almeida, M.V., Neto, F.L.A., 2019. A Review of High Impact Weather
surface weather anomalies it generates. J. Clim. 30, 5563–5584. https://doi.org/ for Aviation Meteorology. Pure Appl. Geophys. 176 https://doi.org/10.1007/
10.1175/JCLI-D-16-0467.1. s00024-019-02168-6.
Dai, Y., Tan, B., 2019a. On the role of the eastern Pacific teleconnection in ENSO impacts Guo, R., Deser, C., Terray, L., Lehner, F., 2019. Human influence on winter precipitation
on wintertime weather over East Asia and North America. J. Clim. 32, 1217–1234. trends (1921–2015) over North America and Eurasia revealed by dynamical
https://doi.org/10.1175/JCLI-D-17-0789.1. adjustment. Geophys. Res. Lett. 46, 3426–3434. https://doi.org/10.1029/
Dai, Y., Tan, B., 2019b. Two types of the Western Pacific Pattern, their climate impacts, 2018GL081316.
and the ENSO modulations. J. Clim. 32, 823–841. https://doi.org/10.1175/jcli-d-17- He, M., Gautam, M., 2016. Variability and trends in precipitation, temperature and
0618.1. drought indices in the State of California. Hydrology. https://doi.org/10.3390/
Dasgupta, P., Metya, A., Naidu, C.V., Singh, M., Roxy, M.K., 2020. Exploring the long- hydrology3020014.
term changes in the Madden Julian Oscillation using machine learning. Sci. Rep. 10 Held, I.M., Soden, B.J., 2006. Robust responses of the hydrological cycle to global
https://doi.org/10.1038/S41598-020-75508-5. warming. J. Clim. 19, 5686–5699. https://doi.org/10.1175/JCLI3990.1.
Davenport, F.V., Burke, M., Diffenbaugh, N.S., 2021. Contribution of historical Hijmans, A.R.J., Phillips, S., Leathwick, J., Elith, J., Hijmans, M.R.J., 2021. Dismo:
precipitation change to US flood damages. Proc. Natl. Acad. Sci. U. S. A. 118 https:// species distribution modeling.
doi.org/10.1073/pnas.2017524118. Huang, X., Stevenson, S., Hall, A.D., 2020a. Future warming and intensification of
de Luis, M., González-Hidalgo, J.C., Longares, L.A., Štepánek, P., 2009. Seasonal precipitation extremes: A “Double Whammy” leading to increasing flood risk in
precipitation trends in the Mediterranean Iberian Peninsula in second half of 20th California. Geophys. Res. Lett. 47 https://doi.org/10.1029/2020GL088679.
century. Int. J. Climatol. 29, 1312–1323. https://doi.org/10.1002/JOC.1778. Huang, X., Swain, D.L., Hall, A.D., 2020b. Future precipitation increase from very high-
del Río, S., Herrero, L., Fraile, R., Penas, A., 2011. Spatial distribution of recent rainfall resolution ensemble downscaling of extreme atmospheric river storms in California.
trends in Spain (1961–2006). Int. J. Climatol. 31, 656–667. https://doi.org/ Sci. Adv. 6 https://doi.org/10.1126/sciadv.aba1323.
10.1002/JOC.2111. Hurrell, J.W., 1995. Decadal trends in the North Atlantic oscillation: Regional
del Río, S., Anjum Iqbal, M., Cano-Ortiz, A., Herrero, L., Hassan, A., Penas, A., 2013. temperatures and precipitation. Science 1979 (269), 676–679. https://doi.org/
Recent mean temperature trends in Pakistan and links with teleconnection patterns. 10.1126/science.269.5224.676.
Int. J. Climatol. 33, 277–290. https://doi.org/10.1002/joc.3423. Hurrell, J.W., Deser, C., 2010. North Atlantic climate variability: The role of the North
Dennison, P.E., Moritz, M.A., Taylor, R.S., 2008. Evaluating predictive models of critical Atlantic Oscillation. J. Mar. Syst. 79, 231–244. https://doi.org/10.1016/J.
live fuel moisture in the Santa Monica Mountains, California. Int. J. Wildland Fire JMARSYS.2009.11.002.
17, 18–27. https://doi.org/10.1071/WF07017. Jiang, M., Felzer, B.S., Sahagian, D., 2016. Predictability of precipitation over the
Díaz-Padilla, G., Sánchez-Cohen, I., Guajardo-Panes, R.A., del Ángel-Pérez, A.L., Ruíz- conterminous U.S. based on the CMIP5 multi-model ensemble. Sci. Rep. 6 https://
Corral, A., Medina-García, G., Ibarra-Castillo, D., 2011. Mapeo del Índice de Aridez y doi.org/10.1038/srep29962.
su Distribución Poblacional en México. Revista Chapingo Serie Ciencias Forestales y Jian-Qi, S., 2010. Possible impact of the boreal spring Antarctic oscillation on the North
del Ambiente XVII 267–275. https://doi.org/10.5154/r.rchscfa.2010.09.069. American SUMMER MONSOON. Atmos. Oceanic Sci. Lett. 3, 232–236. https://doi.
Diffenbaugh, N.S., Swain, D.L., Touma, D., 2015. Anthropogenic warming has increased org/10.1080/16742834.2010.11446870.
drought risk in California. Proc. Natl. Acad. Sci. https://doi.org/10.1073/ Jiménez-Quiroz, C., 2014. Indicadores Climáticos. Una Manera para identificar la
pnas.1422385112. variabilidad climática a escala global. Inapesca.
Dong, Q., Wang, W., Kunkel, K.E., Shao, Q., Xing, W., Wei, J., 2021. Heterogeneous Jones, C., 2000. Occurrence of extreme precipitation events in California and
response of global precipitation concentration to global warming. Int. J. Climatol. relationships with the Madden-Julian oscillation. J. Clim. 13, 3576–3587. https://
41, E2347–E2359. https://doi.org/10.1002/joc.6851. doi.org/10.1175/1520-0442(2000)013<3576:OOEPEI>2.0.CO;2.
Dong, L., Leung, L.R., 2021. Winter precipitation changes in California under global Kam, J., Sheffield, J., 2016. Increased drought and pluvial risk over California due to
warming: contributions of CO2, uniform SST warming, and SST change patterns. changing oceanic conditions. J. Clim. 29, 8269–8279. https://doi.org/10.1175/
Geophys. Res. Lett. https://doi.org/10.1029/2020GL091736. JCLI-D-15-0879.1.
Dong, L., Leung, L.R., Lu, J., Gao, Y., 2019. Contributions of extreme and non-extreme Karmeshu, N., 2015. Trend Detection in Annual Temperature & Precipitation using the
precipitation to California precipitation seasonality changes under warming. Mann Kendall Test – A Case Study to Assess Climate Change on Select States in the
Geophys. Res. Lett. 46, 13470–13478. https://doi.org/10.1029/2019GL084225. Northeastern United States. University of Pennsylvania.
Ehsani, M.R., Arevalo, J., Risanto, C.B., Javadian, M., Devine, C.J., Arabzadeh, A., Killam, D., Bui, A., LaDochy, S., Ramirez, P., Willis, J., Patzert, W., 2014. California
Venegas-Quiñones, H.L., Dell’oro, A.P., Behrangi, A., 2020. 2019–2020 Australia fire getting wetter to the North, drier to the South: natural variability or climate change?
and its relationship to hydroclimatological and vegetation variabilities. Water 2020, Climate. https://doi.org/10.3390/cli2030168.
Vol. 12, Page 3067 12, 3067. https://doi.org/10.3390/W12113067. Krivoruchko, K., 2012. Empirical Bayesian Kriging. ESRI Press Fall, pp. 6–10.
Environmental Systems Research Institute (ESRI), 2019. ARCGIS. Software. Kukal, M., Irmak, S., 2016. Long-term patterns of air temperatures, daily temperature
Exum, N.G., Betanzo, E., Schwab, K.J., Chen, T.Y.J., Guikema, S., Harvey, D.E., 2018. range, precipitation, grass-reference evapotranspiration and aridity index in the USA
Extreme precipitation, public health emergencies, and safe drinking water in the Great Plains: Part I. Spatial trends. J. Hydrol. https://doi.org/10.1016/j.
USA. Curr. Environ. Health Rep. 5:2 5, 305–315. https://doi.org/10.1007/S40572- jhydrol.2016.06.006.
018-0200-5. Lau, K.-M., Weng, H., 2002. Recurrent teleconnection patterns linking summertime
Fischer, E.M., Knutti, R., 2015. Anthropogenic contribution to global occurrence of precipitation variability over East Asia and North America. J. Meteorol. Soc. Jpn 80,
heavy-precipitation and high-temperature extremes. Nat. Clim. Change 5, 560–564. 1309–1324.
https://doi.org/10.1038/NCLIMATE2617. Lausier, A.M., Jain, S., 2018. Overlooked trends in observed global annual precipitation
Ge, Y., Gong, G., 2009. North American snow depth and climate teleconnection patterns. reveal underestimated risks. Sci. Rep. 8:1–7. https://doi.org/10.1038/s41598-018-
J. Clim. 22, 217–233. https://doi.org/10.1175/2008JCLI2124.1. 34993-5.
Gocic, M., Trajkovic, S., 2013. Analysis of changes in meteorological variables using Lee, Y.Y., Grotjahn, R., 2019. Evidence of specific MJO phase occurrence with
Mann-Kendall and Sen’s slope estimator statistical tests in Serbia. Global Planet. summertime California central valley extreme hot weather. Adv. Atmos. Sci. 36
Change 100, 172–182. https://doi.org/10.1016/j.gloplacha.2012.10.014. https://doi.org/10.1007/s00376-019-8167-1.
NASA Goddard Space Flight Center, 2018. UNDERSTANDING EARTH What’ s Up with Levizzani, V., Cattani, E., 2019. Satellite remote sensing of precipitation and the
Precipitation? 1–12. terrestrial water cycle in a changing climate. Remote Sens. https://doi.org/10.3390/
Goss, M., Swain, D.L., Abatzoglou, J.T., Sarhadi, A., Kolden, C.A., Williams, A.P., rs11192301.
Diffenbaugh, N.S., 2020. Climate change is increasing the likelihood of extreme Lin, Y.H., Hipps, L.E., Wang, S.Y.S., Yoon, J.H., 2017. Empirical and modeling analyses
autumn wildfire conditions across California. Environ. Res. Lett. 15, 094016. of the circulation influences on California precipitation deficits. Atmos. Sci. Lett. 18,
https://doi.org/10.1088/1748-9326/AB83A7. 19–28. https://doi.org/10.1002/ASL.719.
Gribov, A., Krivoruchko, K., 2020. Empirical Bayesian kriging implementation and Liu, Q., Shepherd, B., Li, C., 2020. Presiduals: An R package for residual analysis using
usage. Sci. Total Environ. 722 https://doi.org/10.1016/j.scitotenv.2020.137290. probability-scale residuals. J. Stat. Softw. 94, 1–27. https://doi.org/10.18637/jss.
Guan, B., Molotch, N.P., Waliser, D.E., Fetzer, E.J., Neiman, P.J., 2013. The 2010/2011 v094.i12.
snow season in California’s Sierra Nevada: Role of atmospheric rivers and modes of López-Moreno, J.I., Vicente-Serrano, S.M., Morán-Tejeda, E., Lorenzo-Lacruz, J.,
large-scale variability. Water Resour. Res. 49 https://doi.org/10.1002/wrcr.20537. Kenawy, A., Beniston, M., 2011. Effects of the North Atlantic Oscillation (NAO) on
Guirguis, K., Gershunov, A., Shulgina, T., Clemesha, R.E.S., Ralph, F.M., 2019. combined temperature and precipitation winter modes in the Mediterranean
Atmospheric rivers impacting Northern California and their modulation by a mountains: Observed relationships and projections for the 21st century. Global
variable climate. Clim. Dyn. 52, 6569–6583. https://doi.org/10.1007/s00382-018- Planet. Change 77, 62–76. https://doi.org/10.1016/J.GLOPLACHA.2011.03.003.
4532-5. Luković, J., Chiang, J.C.H., Blagojević, D., Sekulić, A., 2021. A later onset of the rainy
Gultepe, I., 2015. Mountain weather: Observation and modeling. Adv. Geophys. 56 season in California. Geophys. Res. Lett. https://doi.org/10.1029/2020GL090350.
https://doi.org/10.1016/bs.agph.2015.01.001. Luteyn, J.L., Hickman, J.C., 1993. The Jepson Manual: Higher Plants of California,
Gultepe, I., Rabin, R., Ware, R., Pavolonis, M., 2016. Light snow precipitation and effects Brittonia. https://doi.org/10.2307/2807611.
on weather and climate. Adv. Geophys. 57 https://doi.org/10.1016/bs. Mallakpour, I., AghaKouchak, A., Sadegh, M., 2019. Climate-induced changes in the risk
agph.2016.09.001. of hydrological failure of major dams in California. Geophys. Res. Lett. 46,
Gultepe, I., Heymsfield, A.J., Gallagher, M., Ickes, L., Baumgardner, D., 2017. Ice fog: the 2130–2139. https://doi.org/10.1029/2018GL081888.
current state of knowledge and future challenges. Meteorol. Monogr. 58 https://doi. Masson-Delmotte, V., Pirani, S.L., Connors, C., Péan, S., Berger, N., Caud, Y., Chen, L.,
org/10.1175/amsmonographs-d-17-0002.1. Goldfarb, M.I., Gomis, M., Huang, K., Leitzell, E., Lonnoy, J.B.R., 2021. Climate
Gultepe, I., Sharman, R., Williams, P.D., Zhou, B., Ellrod, G., Minnis, P., Trier, S., Change 2021: The Physical Science Basis. Contribution of Working Group I to the
Griffin, S., Yum, S.S., Gharabaghi, B., Feltz, W., Temimi, M., Pu, Z., Storer, L.N., Sixth Assessment Report of the Intergovernmental Panel on Climate Change, IPCC.
Kneringer, P., Weston, M.J., Chuang, H.Y., Thobois, L., Dimri, A.P., Dietz, S.J.,

20
A. González-Pérez et al. Journal of Hydrology 612 (2022) 128211

McAfee, S.A., 2014. Consistency and the lack thereof in pacific decadal oscillation mitigation: why haven’t we bent the global emissions curve? Annu. Rev. Environ.
impacts on North American Winter Climate. J. Clim. 27, 7410–7431. https://doi. Resour. https://doi.org/10.1146/annurev-environ-012220-011104.
org/10.1175/JCLI-D-14-00143.1. Sun, Q., Miao, C., Duan, Q., Ashouri, H., Sorooshian, S., Hsu, K.L., 2018. A review of
Meseguer-Ruiz, O., Sarricolea, P., 2017. Detección de inhomogeneidades en series de global precipitation data sets: data sources, estimation, and intercomparisons. Rev.
precipitación diaria en la región centro-sur de Chile. Interciencia 42, 242–249. Geophys. 56 https://doi.org/10.1002/2017RG000574.
Mohammad, P., Goswami, A., 2019. Temperature and precipitation trend over 139 major Swain, D.L., Horton, D.E., Singh, D., Diffenbaugh, N.S., 2016. Trends in atmospheric
Indian cities: An assessment over a century. Model. Earth Syst. Environ. 5, patterns conducive to seasonal precipitation and temperature extremes in California.
1481–1493. https://doi.org/10.1007/S40808-019-00642-7. Sci. Adv. https://doi.org/10.1126/sciadv.1501344.
Mosaffa, H., Sadeghi, M., Hayatbini, N., Gorooh, V.A., Asanjan, A.A., Nguyen, P., Thom, H.C.S., 1966. Some methods of climatological analysis.
Sorooshian, S., 2020. Spatiotemporal variations of precipitation over Iran using the Torres-Batlló, J., Martí-Cardona, B., 2020. Precipitation trends over the southern Andean
high-resolution and nearly four decades satellite-based PERSIANN-CDR dataset. Altiplano from 1981 to 2018. J. Hydrol. 590 https://doi.org/10.1016/j.
Remote Sens. 12 https://doi.org/10.3390/rs12101584. jhydrol.2020.125485.
Nardi, K.M., Baggett, C.F., Barnes, E.A., Maloney, E.D., Harnos, D.S., Ciasto, L.M., 2020. Trenberth, K.E., Dai, A., van der Schrier, G., Jones, P.D., Barichivich, J., Briffa, K.R.,
Skillful all-season s2s prediction of U.S. precipitation using the MJO and QBO. Sheffield, J., 2014. Global warming and changes in drought. Nat. Clim. Change 4,
Weather Forecast. 35, 2179–2198. https://doi.org/10.1175/WAF-D-19-0232.1. 17–22. https://doi.org/10.1038/nclimate2067.
Newman, M., Alexander, M.A., Ault, T.R., Cobb, K.M., Deser, C., di Lorenzo, E., Treppiedi, D., Cipolla, G., Francipane, A., Noto, L., v.,, 2021. Detecting precipitation
Mantua, N.J., Miller, A.J., Minobe, S., Nakamura, H., Schneider, N., Vimont, D.J., trend using a multiscale approach based on quantile regression over a Mediterranean
Phillips, A.S., Scott, J.D., Smith, C.A., 2016. The Pacific decadal oscillation, area. Int. J. Climatol. https://doi.org/10.1002/JOC.7161.
revisited. J. Clim. 29, 4399–4427. https://doi.org/10.1175/JCLI-D-15-0508.1. Velasco, E.M., Gurdak, J.J., Dickinson, J.E., Ferré, T.P.A., Corona, C.R., 2017.
Pathak, T., Maskey, M., Dahlberg, J., Kearns, F., Bali, K., Zaccaria, D., 2018. Climate Interannual to multidecadal climate forcings on groundwater resources of the U.S.
change trends and impacts on California agriculture: A detailed review. Agronomy. West Coast. J. Hydrol.: Reg. Stud. 11, 250–265. https://doi.org/10.1016/J.
https://doi.org/10.3390/agronomy8030025. EJRH.2015.11.018.
Pavia, E.G., Graef, F., Fuentes-Franco, R., 2016. Recent ENSO-PDO precipitation Walker, G.T., Bliss, E.W., 1932. World Weather V - NAO. Memoirs of the Royal
relationships in the Mediterranean California border region. Atmos. Sci. Lett. 17 Meteorological Society IV.
https://doi.org/10.1002/asl.656. Wallace, J.M., Gutzler, D.S., 1981. Teleconnections in the geopotential height field
Peña-Angulo, D., Gonzalez-Hidalgo, J.C., Sandonís, L., Beguería, S., Tomas-Burguera, M., during the Northern Hemisphere winter. Mon. Weather Rev. 109, 784–812. https://
López-Bustins, J.A., Lemus-Canovas, M., Martin-Vide, J., 2021. Seasonal doi.org/10.1175/1520-0493(1981)109<0784:TITGHF>2.0.CO;2.
temperature trends on the Spanish mainland: A secular study (1916–2015). Int. J. Wang, S., Anichowski, A., Tippett, M.K., Sobel, A.H., 2017. Seasonal noise versus
Climatol. 41 https://doi.org/10.1002/joc.7006. subseasonal signal: Forecasts of California precipitation during the unusual winters
Polade, S.D., Gershunov, A., Cayan, D.R., Dettinger, M.D., Pierce, D.W., 2017. of 2015–2016 and 2016–2017. Geophys. Res. Lett. 44 https://doi.org/10.1002/
Precipitation in a warming world: Assessing projected hydro-climate changes in 2017GL075052.
California and other Mediterranean climate regions. Sci. Rep. 7, 1–10. https://doi. Wang, L., Robertson, A.W., 2019. Week 3–4 predictability over the United States assessed
org/10.1038/s41598-017-11285-y. from two operational ensemble prediction systems. Clim. Dyn. 52, 5861–5875.
Powell, E.J., Keim, B.D., 2015. Trends in daily temperature and precipitation extremes https://doi.org/10.1007/s00382-018-4484-9.
for the southeastern United States: 1948–2012. J. Clim. 28 https://doi.org/10.1175/ Western Regional Climate Center. [WWW Document], 2020. URL https://wrcc.dri.
JCLI-D-14-00410.1. edu/Climate/summaries.php.
Ríos Cornejo, D., Penas, Á., Álvarez-Esteban, R., del Río, S., 2015. Links between Whan, K., Zwiers, F., 2016. The impact of ENSO and the NAO on extreme winter
teleconnection patterns and mean temperature in Spain. https://doi.org/10.1007/ precipitation in North America in observations and regional climate models. Clim.
s00704-014-1256-2. Dyn. 2016 48:5 48, 1401–1411. https://doi.org/10.1007/S00382-016-3148-X.
Rivas-Martínez, S., Rivas-Sáenz, S., Penas-Merino, A., 2011. Worldwide bioclimatic Wheeler, M.C., Hendon, H.H., 2004. An All-Season Real-Time Multivariate MJO Index:
classification system. Glob. Geobot. 1, 1–638. https://doi.org/10.5616/gg110001. Development of an Index for Monitoring and Prediction.
Roksvåg, T., Lutz, J., Grinde, L., Dyrrdal, A.V., Thorarinsdottir, T.L., 2021. Consistent Williams, A.P., Abatzoglou, J.T., Gershunov, A., Guzman-Morales, J., Bishop, D.A.,
intensity-duration-frequency curves by post-processing of estimated Bayesian Balch, J.K., Lettenmaier, D.P., 2019. Observed impacts of anthropogenic climate
posterior quantiles. J. Hydrol. 603, 127000. https://doi.org/10.1016/j. change on wildfire in California. Earth’s Future 7, 892–910. https://doi.org/
jhydrol.2021.127000. 10.1029/2019EF001210.
Sarricolea, P., Meseguer-Ruiz, Ó., Serrano-Notivoli, R., Soto, M.V., Martin-Vide, J., 2019. Wise, E.K., Wrzesien, M.L., Dannenberg, M.P., McGinnis, D.L., 2015. Cool-season
Trends of daily precipitation concentration in Central-Southern Chile. Atmos. Res. precipitation patterns associated with teleconnection interactions in the United
215, 85–98. https://doi.org/10.1016/j.atmosres.2018.09.005. States. J. Appl. Meteorol. Climatol. https://doi.org/10.1175/JAMC-D-14-0040.1.
Sayemuzzaman, M., Jha, M.K., 2014. Seasonal and annual precipitation time series trend Yu, B., Lin, H., Soulard, N., 2019. A comparison of North American surface temperature
analysis in North Carolina, United States. Atmos. Res. 137, 183–194. https://doi. and temperature extreme anomalies in association with various atmospheric
org/10.1016/j.atmosres.2013.10.012. teleconnection patterns. Atmosphere (Basel) 10, 172. https://doi.org/10.3390/
Sellars, S.L., Gao, X., Sorooshian, S., 2015. An object-oriented approach to investigate atmos10040172.
impacts of climate oscillations on precipitation: A western United States case study. Yuan, F., Liu, J., Berndtsson, R., Hao, Z., Cao, Q., Wang, H., Du, Y., An, D., 2020. Changes
J. Hydrometeorol. 16, 830–842. https://doi.org/10.1175/JHM-D-14-0101.1. in precipitation extremes over the source region of the Yellow River and its
Serra-Llobet, A., Conrad, E., Schaefer, K., 2016. Governing for integrated water and flood relationship with teleconnection patterns. Water (Switzerland) 12. https://doi.org/
risk management: Comparing top-down and bottom-up approaches in Spain and 10.3390/W12040978.
California. Water (Switzerland) 8. https://doi.org/10.3390/w8100445. Zhang, N., Pathak, T.B., Parker, L.E., Ostoja, S.M., 2021. Impacts of large-scale
Sheppard, P.R., Comrie, A.C., Packin, G.D., Angersbach, K., Hughes, M.K., 2002. The teleconnection indices on chill accumulation for specialty crops in California. Sci.
climate of the US Southwest. Clim. Res. 21, 219–238. https://doi.org/10.3354/ Total Environ. 791, 148025. https://doi.org/10.1016/J.SCITOTENV.2021.148025.
cr021219. Zheng, C., Chang, E.K.M., Kim, H.M., Zhang, M., Wang, W., 2018. Impacts of the
Song, X., Zhang, J., Zou, X., Zhang, C., AghaKouchak, A., Kong, F., 2019. Changes in Madden-Julian oscillation on storm-track activity, surface air temperature, and
precipitation extremes in the Beijing metropolitan area during 1960–2012. Atmos. precipitation over North America. J. Clim. 31 https://doi.org/10.1175/JCLI-D-17-
Res. https://doi.org/10.1016/j.atmosres.2019.02.006. 0534.1.
Stoddard, I., Anderson, K., Capstick, S., Carton, W., Depledge, J., Facer, K., Gough, C., Zhou, W., Yang, D., Xie, S.-P., Ma, J., 2020. Amplified Madden–Julian oscillation impacts
Hache, F., Hoolohan, C., Hultman, M., Hällström, N., Kartha, S., Klinsky, S., in the Pacific–North America region. Nat. Clim. Change, 10(7), 654–660. https://
Kuchler, M., Lövbrand, E., Nasiritousi, N., Newell, P., Peters, G.P., Sokona, Y., doi.org/10.1038/s41558-020-0814-0.
Stirling, A., Stilwell, M., Spash, C.L., Williams, M., 2021. Three decades of climate

21

You might also like