Вы находитесь на странице: 1из 18

Journal of Hydrology 556 (2018) 1153–1170

Contents lists available at ScienceDirect

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

Research papers

Multi-decadal 40- to 60-year cycles of precipitation variability in Chile


(South America) and their relationship to the AMO and PDO signals
Rodrigo Valdés-Pineda a,⇑, Julio Cañón b, Juan B. Valdés a
a
University of Arizona, Department of Hydrology and Atmospheric Sciences, 1133 E James E. Rogers Way, Tucson, AZ, USA
b
Grupo de Investigación en Gestión y Modelación Ambiental (GAIA), Facultad de Ingeniería, Universidad de Antioquia, SIU/UdeA, Calle 70 No. 52-21, Medellín, Colombia

a r t i c l e i n f o a b s t r a c t

Article history: High-frequency components of precipitation variability have been an important focus of study during the
Available online 1 February 2017 last few decades in Chile. Low-frequency variations, on the other hand, have received less attention, espe-
cially in association with multi-decadal cycles that can affect variability trends of precipitation in the long
Keywords: term. This study analyzes these low-frequency patterns of precipitation in Chile (>30 years), and their
Precipitation variability relationship to global Sea Surface Temperatures (SSTs), with special focus on associations with the
SSA Pacific Decadal Oscillation (PDO) and the Atlantic Multi-decadal Oscillation (AMO) indices. Singular
MSSA
Spectrum Analysis (SSA) and its Multi-Channel version (MSSA) techniques were applied to a dataset con-
Climate oscillations
PDO
taining long instrumental records of monthly precipitation aggregated yearly and seasonally. The rela-
AMO tionships between the low-frequency variability of precipitation and the PDO are significant to the
north of the country, whereas connections with the AMO are significant to the north and also to the
south. This is also evident from the global spatial correlation analysis of low-frequency precipitation
modes and SSTs, where the southernmost station shows a strong relationship with the Atlantic Ocean.
We conclude that a significant multi-decadal precipitation cycle between 40 and 60 years is evident at
the rain gauges located in the subtropical and extratropical regions of Chile. This low-frequency variabil-
ity seems to be largely linked to PDO and AMO modulation.
Ó 2017 Elsevier B.V. All rights reserved.

1. Introduction tree ring chronologies, the precipitation records, and the recon-
structions were mainly associated with QBO. The precipitation
In Chile, spectral analysis techniques have been used mainly to reconstructions also contained significant peaks in variance within
observe and model paleoclimate records. Most of these studies the frequency range of ENSO; however, their results did not reveal
have shown that a substantial part of climate variability seems to a clear ENSO modulation of precipitation variability over the east
be controlled by inter-annual to decadal-scale oscillations. These of the southern Andes.
oscillations include, the 2.5-year Quasi-Biennial Oscillation (QBO) In the Chilean Lake District (38–42°S), Fagel et al. (2008) applied
(see Graystone, 1959; Ebdon, 1975; Baldwin et al., 2001; Dean SSA to identify potential periodicities and reconstruct oscillatory
and Kemp, 2004), the 3- to 7-year El Niño Southern Oscillation components of varve thickness over the last centuries. The proce-
(ENSO) (Trenberth, 1984; Trenberth and Hoar, 1997;; Dawson dure analyzed several short and long annual laminated sediments
and O’Hare, 2000), and the 10- to 30-year Pacific Decadal Oscilla- of Lago Peyuhue (40°S and 72°W), which is located in a region
tion (PDO) (Mantua et al., 1997; Biondi et al., 2001; Deser et al., where precipitation is often driven by the southern westerlies.
2010). For instance, Villalba et al. (1998) applied Singular Spectrum The authors found that frequency bands of varve thickness are
Analysis (SSA) to develop long-term reconstructions (400 years) of mainly related to the periodicities of the QBO, ENSO, and PDO. In
seasonal and annual precipitation variations for northern Patago- addition, Le Quesne et al. (2009) developed a multi-proxy approach
nia (east of the Andes). The study used 22 precipitation records to analyze historical glacier variations in the Central Andes of Chile
(1912–1989) available between 39 and 43°S, and 27 chronologies and Argentina (33–36°S). They developed a tree-ring based multi-
of Austrocedrus chilensis collected along the eastern slopes of the variate model to reconstruct total annual precipitation of Santiago
Andean Patagonia. Oscillations observed in the power spectra of de Chile (1866–1999). The periodic behavior of the reconstructed
precipitation time series was evaluated by extracting the dominant
⇑ Corresponding author. oscillatory components using SSA. They found inter-annual oscilla-
E-mail address: rvaldes@email.arizona.edu (R. Valdés-Pineda). tory modes (2.1–3.4 yr.) that explained 18.4% of the variance, and

http://dx.doi.org/10.1016/j.jhydrol.2017.01.031
0022-1694/Ó 2017 Elsevier B.V. All rights reserved.
1154 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

decadal/inter-decadal modes (13–23 yr.) that explained 18.5% of 2. Methods


the variance. Interestingly, both the inter-annual and decadal/
inter-decadal signals of the El Niño 3.4 Index (extracted by SSA) 2.1. Study area
showed better correlations with precipitation records than those
observed for the extracted PDO signals. Continental Chile is located in the Pacific coast of South America
Aravena and Luckman (2009) analyzed the dominant spatial between latitudes 17°300 S and 56°300 S (about 4300 km), and is
and temporal patterns of a network of 23 homogenous instrumen- divided into 15 administrative regions. The geography of the coun-
tal precipitation records of southern South America (41–53°S), try is mainly dominated by steep mountainous terrains, with only
finding mostly two- to seven-year oscillations for the Patagonia around 20% of the continental territory being flat. The national ter-
Plains-Atlantic (43–50°S) and southern Patagonia (51–53°S), with ritory is composed of 101 main hydrologic basins, with more than
a strong biannual oscillation mode. Decadal oscillations 300 rivers flowing mainly from east to west. The Chilean territory
(12 years) were observed for the first leading mode of northwest- is also characterized by a wide variety of landscapes. From a geo-
ern (41–44°S) and central Patagonia (45–47°S). The regional morphological point of view, it is possible to distinguish four major
annual precipitation was proposed to be strongly influenced by geographical units in the country: Andes Mountains (east), Inter-
Antarctic circulation modes (i.e., the Antarctic Oscillation Index mediate Depression, Coastal Mountains, and Coastal plains (west)
or AAO); whereas ENSO influence on precipitation variability was (see Valdés-Pineda et al., 2014).
less evident. Northern Chile (17°300 –30°S) is a hyper-arid region under the
Rubio-Álvarez and McPhee (2010) studied the time series of persistent influence of the South East Pacific Anticyclone (SEPA),
annual and seasonal streamflow of 44 rivers in southern Chile, which blocks the westerlies at low levels of the atmosphere as
spanning the ecoregion between 34°S and 45°S (1952–2003). SSA far south as 27°S. The South American Monsoon System (SAMS)
was applied to the original time series of composite streamflow brings easterly moisture from Amazonia (austral summer) that
anomalies to obtain reconstructed time series and their peri- generates rain shadow effect over the western Andes and Atacama
odograms. Their results showed cycles of annual, seasonal, and Desert – i.e., annual accumulation increases from about
spatially averaged streamflows, covering a range between 2 and 20 mm year1 at 2700 m.a.s.l. to 300 mm year1 at 5000 m.a.s.l.
11 years. The quasi-biennial and quasi-quadrennial periods were (Miller, 1976; Aceituno et al., 1993; Garreaud et al., 2003;
associated with ENSO, while the six-year period was related to Houston, 2006; Minvielle and Garreaud, 2011). In central Chile
the North Pacific Sea Surface Temperatures (SSTs). The authors also (30–40°S) the strong seasonality of precipitation results from the
indicated that longer, near-decadal periods (of approximately 7, 8, wintertime retreat of SEPA, which allows stronger low-level west-
9, and 11 years), could be related to the PDO and AAO oscillations. erlies to carry frontal systems to the continent (Montecinos and
Urrutia et al. (2011) developed a 410-year annual streamflow Aceituno, 2003; Falvey and Garreaud, 2007). The annual accumula-
reconstruction for the Maule River watershed (35°S–36°300 S, with tion peaks during austral winter and increases southward from
a climate like that of the Mediterranean) using Austrocedrus chilen- around 100 mm at 30°S to nearly 2000 mm at 40°S (Quintana
sis tree ring chronologies. Their reconstruction revealed interan- and Aceituno, 2012). Due to the orographic effect of the Andes
nual, interdecadal, and multi-decadal oscillation modes that were Mountains, precipitation in central Chile also increases about twice
well correlated with climatic indices. Positive correlations were in the cross mountain direction (Falvey and Garreaud, 2007; Viale
observed with the El Niño 3.4 Index (correspondence between high and Garreaud, 2015). In southern and austral Chile (40°–56°300 S)
tropical SSTs and high streamflows); and negative correlations frontal cyclones drifting eastward along the oceanic storm tracks
were observed with the AAO, indicating that above-average are responsible for most of the precipitation and day-to-day
streamflows are significantly related to the negative phase of the weather variations (Quintana and Aceituno, 2012; Garreaud
AAO. This is the result of weakened sea level pressure (SLP) in et al., 2014; Valdés-Pineda et al., 2015). This translates into abun-
the mid-latitudes and intensified SLP in Antarctica. In this regard, dant precipitation accumulation in all seasons, reaching more than
one of the most important detected oscillatory modes was an 4000 mm year1 along the western Andes (Carrasco et al., 2002;
approximately 18-year cycle (30% of explained variance), possibly Villarroel, 2013). In general, precipitation accumulation in Chile
related to the 18.6-year cycle of the lunar tides. The other oscilla- tends to increase with latitude and altitude (Pizarro et al., 2012;
tory mode is a 50- to 60-year cycle, particularly significant after Quintana and Aceituno, 2012). The annual progression of this lati-
1850, which could possibly be associated with the Half- tudinal gradient shows perfect links with the interaction between
Gleissberg (70–100 years) solar magnetic cycle (see details in westerlies and SEPA (Fig. 1a). For instance, the band of westerlies
Peristykh and Damon, 2003), and the PDO. displaces southward during austral summer (strengthened SEPA)
Although several studies have applied SSA to determine (Fig. 1b and e), and moves northward during austral winter (weak-
oscillatory components in different regions of Chile, most of them ened SEPA) (Fig. 1c and d).
have been focused on the analysis of high- to mid-frequency These patterns of precipitation variability are originated by the
modes of precipitation, ranging particularly from quasi-biennial activity of SST in the equatorial Pacific, which indeed determines
to inter-decadal scales (2–30 years). Our study explores multi- modes of variation such as ENSO and the PDO. For instance,
decadal patterns of precipitation (defined here as those fluctuating research from the last few decades has established that inter-
over a period longer than 30 years) in Chile, by applying SSA and its annual precipitation variability in Chile is related to ENSO
Multi-Channel version (MSSA). These analyses were applied to a (Rutllant and Fuenzalida, 1991; Aceituno and Garreaud, 1995;
dataset of monthly series obtained from the longest instrumental Garreaud and Batisti, 1999; Montecinos and Aceituno, 2003;
records of precipitation in Chile, a span between 1866 and 2014. Garreaud et al., 2009), playing a major role during the austral win-
The degree of the relationship between low-frequency modes of ter (Valdés-Pineda et al., 2016). The decadal to inter-decadal vari-
precipitation variability and the ENSO, AAO, and PDO signals is also ability of precipitation in the country has been mainly associated
assessed. This study contributes to an increased knowledge of how with the PDO (Mantua et al., 1997; Montecinos and Aceituno,
possible unknown multi-decadal modes of SST or SLP in the Pacific 2003; Garreaud et al., 2009; Quintana and Aceituno, 2012). In
Basin and other atmospheric teleconnections could explain low- Chilean Patagonia another source of precipitation variability is
frequency variability of precipitation at an annual, seasonal, and the AAO, which is characterized by pressure anomalies of one sign
monthly scale. centered in the Antarctic and anomalies of the opposite sign on a
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1155

Fig. 1. Spatial distribution of Averaged Annual Precipitation accumulation (mm) for 238 ground stations (1979–2012) distributed in the country and 925-hPa zonal and
meridional wind (m/s) components (U925; V925). Source: Adopted from Valdés-Pineda et al. (2016).

circumglobal band, at about 40°–50°S (Garreaud et al., 2009; established in the United States (see Kunkel et al., 1999) and Eur-
Quintana and Aceituno, 2012; Villalba et al., 2012). ope (see Zolina et al., 2009).
The rain gauges located in northern regions were not included
2.2. Dataset: Precipitation Records, climate Indices, and gridded fields in the analysis because they have a significant number of zeros
in the monthly and annual accumulation records that disturb the
Since our interest was essentially focused in detecting low- implementation of the methods used in this study. Only one rain
frequency temporal variability (>30 years), only instrumental gauge containing less than 100 years’ of records (Puerto Aysen at
records of more than 100 years were used for SSA and MSSA anal- about 45°S) was included in the analysis (see Table 1) because it
ysis. The meteorological yearbooks of the Meteorological Direc- presents important information for decision-making about ongo-
torate of Chile (DMC) were reviewed to obtain the longest ing large-scale hydroelectricity projects. The records of Valdivia,
records of the country. The DMC’s database is digitally available which date back to 1853 (see Sanz Donaire, 2012), were not
only from 1920 to the present (http://164.77.222.61/climatolo- included in the analysis because from 1970 onwards another rain
gia/). Earlier digital records were acquired from the Global Histor- gauge located 15 km away was used to continue the suspended
ical Climatology Network database (GHCN-Monthly) of the U.S. records of Valdivia (National Directorate of Water Resources of
National Oceanic and Atmospheric Administration (NOAA). This Chile, DGA, personal communication, 2015).
database contains historical temperature, precipitation, and pres- Four climate indices and gridded data from NOAA Earth System
sure data for thousands of land stations worldwide (ftp://ftp. Research Laboratory (ESRL) Physical Sciences Division (PSD) were
ncdc.noaa.gov/pub/data/ghcn/v2/). After checking all available included in our analysis to verify their degree of relationship to
data, only 12 rain gauges, evenly distributed between 26° and low-frequency precipitation variability: (1) the monthly Southern
53°S, were selected (see Fig. 2). These records, spanning between Oscillation Index (SOI1) (1893–2014) obtained from NCAR-NOAA;
1866 and 2014, represent an important collection for southern (2) the monthly Antarctic Oscillation Index (AAO2) (1979–2014)
South America (Chile) since the current availability of digital cli-
mate data over many parts of the world is only restricted to the 1
SOI available at: http://www.cgd.ucar.edu/cas/catalog/climind/SOI.signal.ascii.
second half or the last third of the 20th century (Brunet and 2
AAO available at: http://www.cpc.ncep.noaa.gov/products/precip/CWlink/daily_
Jones, 2011). Other long instrumental records have already been ao_index/aao/monthly.aao.index.b79.current.ascii.table.
1156 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Fig. 2. Time series of annual accumulation of precipitation (mm) in Chile for records longer than 100 years.

Table 1
Sampling stations analyzed in the study area.

Map number Station/location Institution Longitude Latitude Region Period Years


1 Copiapó DGA 70.33 27.38 Atacama 1897–2014 118
2 La Serena DGA 71.26 29.91 Coquimbo 1869–2014 146
3 Ovalle DGA 71.20 30.60 Coquimbo 1912–2014 103
4 Los Andes DGA 70.60 32.84 Valparaiso 1907–2014 108
5 Santiago DMCNOAA 70.66 33.46 Metropolitana 1866–2014 149
6 Rancagua DGA-NOAA 70.73 34.17 O’Higgins 1910–2014 105
7 Talca DGA-NOAA 71.66 35.42 Maule 1907–2014 108
8 Concepcion NOAA 73.04 36.82 Biobio 1912–2014 103
9 Temuco NOAA 72.59 38.75 Araucanía 1912–2014 103
10 Puerto Montt NOAA 72.94 41.46 Los Lagos 1888–2014 138
11 Puerto Aysen DGA-NOAA 72.70 45.40 Aysen 1931–2014 84
12 Punta Arenas DGA-NOAA 70.88 53.13 Magallanes 1888–2014 127

acquired from NCEP-NOAA; (3) the monthly Pacific Decadal 2.3. Time series analysis of precipitation records
Oscillation (PDO3) (1900–2014) obtained from the University of
Washington; and (4) the Atlantic Multidecadal Oscillation (AMO4) Total annual accumulation at each raingauge was obtained by
(1856–2014) obtained from NCEP-NOAA. For further research and aggregating all monthly accumulations of each calendar year. The
discussion we incorporated SST, precipitable water (PW), and zonal seasonal series for dry (Oct.-Mar.) and wet (Apr.-Sept.) periods
winds (ZW) gridded datasets as part of our study. were split according to the Chilean hydrologic year. These time ser-
ies were then standardized in order to give equal weight to all sta-
tions and avoid absolute-value bias (Wilks, 2011). Once all the time
3
PDO available at: http://jisao.washington.edu/pdo/PDO.latest. series were organized, we applied time-series analysis (TSA) tech-
4
AMO available at: http://www.esrl.noaa.gov/psd/data/timeseries/AMO/. niques to investigate temporal patterns of precipitation. TSA is
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1157

commonly used to detect all essential features of the records, i.e., 2.4.2. Multi-Channel Singular Spectrum Analysis (MSSA)
normality, stationarity, homogeneity, presence/absence of trends, MSSA approach was useful for obtaining a common structure of
and persistence (Machiwal and Jha, 2012; Machiwal et al., 2015). several series of precipitation and climatic oscillations (see i.e.
We also used the Autocorrelation Function (ACF) at different lags Cañón et al., 2007), and also for establishing causality between
(Box and Jenkins, 1970; Box et al., 1994), to determine the degree the different series. The absence of causality between time series
of temporal dependency among precipitation observations at (i.e. X and Y) implies that the knowledge of some predictor (Y ser-
annual, seasonal, and monthly scales. The ACF was also useful to ies) does not improve the quality of forecasts (Hassani et al., 2010),
identify patterns in terms of trends, seasonality, and periodicity while knowledge of some climate oscillation (X series) does not
(i.e. Huesca et al., 2015). improve the quality of precipitation predictions. MSSA is a natural
extension of SSA for analyzing multivariate time series. The algo-
2.4. Low-frequency modes of precipitation variability rithm is almost the same as in the univariate mode; however, in
this approach the Henkel or Trajectory Matrix of the multivariate
2.4.1. Singular Spectrum Analysis (SSA) time series consists of stacked trajectory matrices of separate time
Taking into account the consistent length of the period of series. MSSA also was applied to all rain gauge records in combina-
records, we searched for significant low-frequency cycles that tion only with the SOI, PDO, and AMO (1912–2014) since they have
could explain precipitation variability at a multi-decadal scale. In much longer records than AAO (see Section 2.2). The Grand Data
this regard, SSA was applied to each individual annual, seasonal, Matrix that was assembled as the Hankel Matrix (see Partington,
and monthly series of precipitation and the climate indices: the 1988; Makowsky and Labai, 2015) was finally arranged containing
SOI, AAO, PDO and AMO. The details and applications of SSA to records of precipitation and climate indices. To be consistent with
hydroclimatic series can be found in the last two decades of liter- SSA analysis, the L parameter was again fixed to one-half of the
ature (i.e. Weare and Nasstrom, 1982; Broomhead and King, 1986a, time series length (L = N/2). This method has gained global popu-
b; Fraedrich, 1986; Vautard and Ghil, 1989; Ghil and Vautard, larity during the last decade. For instance, Krishnamurthy and
1991; Ghil et al., 2002; Marques et al., 2006; H. Wang et al., Misra (2010) applied MSSA to a climatic series of daily outgoing
2015; R. Wang et al., 2015); therefore, they are not described in longwave radiation in the South American Monsoon System.
this study.5 The approach is essentially a principal component (PC) Recently, Zotov et al. (2015) applied MSSA to filter data from the
analysis in the time domain which allows for the study of separated Gravity Recovery and Climate Experiment (GRACE) and separate
spectral contributions of the main oscillatory modes (see Vautard its principal components (PCs) at different periodicities over the
and Ghil, 1989; Ghil et al., 2002; Golyandina et al., 2001; Hassani 15 largest river basins of Russia. Despite this, no studies in Chile
et al., 2009, 2015; Hassani and Thomakos, 2010; Unnikrishnan and have used this multi-channel version of the analysis, at least using
Jothiprakash, 2015; Silva and Hassani, 2015). climate series. However, MSSA has many applications and it is
There are two important parameters to take into account when especially popular in analyzing and forecasting economic and
using SSA to extract each source signal, i.e., a suitable selection of financial time series (Patterson et al., 2011; Hassani et al., 2013;
the window length (L) for constructing the trajectory matrix, and Hassani and Mahmoudvand, 2013), and for determining spatio-
the number of components or eigenvalues (r) to be retained (H. temporal patterns using climate series (i.e. Krishnamurthy and
Wang et al., 2015; R. Wang et al., 2015). The range of the window Stan, 2015; Krishnamurthy and Krishnamurthy, 2015).
length L is usually in 2 6 L 6 N/2 (H. Wang et al., 2015; R. Wang
et al., 2015). A discussion about the L parameter is provided by 2.5. Cross correlations between multi-decadal modes of precipitation,
Elsner and Tsonis (2013) who remarked that choosing L = N/4 is climate indices and gridded fields
a common practice. Others authors suggest that L should be large
enough, but not larger than N/2 (see Golyandina et al., 2001; Mah- The large number of interactions comprising the complex
moudvand and ZokaeiWe, 2012; Golyandina and Zhigljavsky, ocean-atmospheric system can be determined by analyzing the
2013). In general, large values of L allow longer period oscillations cross-correlations between climate signals, in order to better diag-
to be resolved, but choosing an L that is too large leaves too few nose and understand the whole system. Because of its simplicity,
observations from which to estimate the covariance matrix of the traditional cross-correlation analysis (TCA) has become the most
L variables. widely used method for this purpose (Yuan et al., 2015). Especially
In this study, the L parameter was fixed to one-half of the time in climatology research, TCA is used in various fields, including
series length, since it has been shown to be a suitable size for the dynamic diagnosing and climate forecasting (Yuan et al., 2015).
decomposition stage of low frequency patterns (Mahmoudvand In accordance with this, we analyzed the degree of the relationship
and ZokaeiWe, 2012). For instance, the L parameter for a 100- between precipitation variability and climate indices by imple-
year series was 50 years. Additionally, in order to quantify the peri- menting traditional (and lagged) cross-correlation analysis (see
odicities of the cycles associated to low-frequency variability, the details in Bendat and Piersol, 2011). For instance, the reconstructed
first 10 leading modes (PCs) were retained and their significances low-frequency modes of precipitation at each rain gauge (RPC), cal-
calculated according to North (1982) (Fig. 3a). Then the periodici- culated from either SSA or MSSA, and the modes of climate indices
ties of each PC were separately calculated using periodogram anal- were used as predictors of annual, seasonal, and monthly accumu-
ysis as described in Ghil et al. (2002) and Chatfield (2013) (Fig. 3b). lation variability. The same analysis was performed to derive sig-
Those series with largest periodicities (number of years or months) nificant correlation patterns between the low-frequency modes
as represented by the highest significant spectral peaks were of precipitation and SST, and also extended in the discussion sec-
selected as low-frequency cycles for further analysis. The Multi tion to determine how this multi-decadal signal, mostly observed
Taper Method (MTM) was additionally applied to verify results in the ocean, can be easily detected at some atmospheric levels
from periodogram analysis (see Rao and Hamed, 2003). This of PW and ZW.
method reduces the variance of spectral estimates by using a small The idea of applying time-lagged cross-correlation analysis was
set of tapers rather than the unique data taper or spectral window to identify significant lagged predictors influencing the low-
used by classical methods as Fourier analysis (Ghil et al., 2002; frequency precipitation variability in the country. This analysis
Golyandina and Zhigljavsky, 2013). was conducted through the use of lagged scatterplots. One attri-
bute of the lagged scatterplots is that they can show if the cross-
5
See Appendix for more details about SSA and MSSA. correlation is representative of the majority of the data or if it is
1158 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Fig. 3. (a) Eigenvalue spectrum for the Trajectory (Henkel) matrix of annual time series of annual precipitation in Copiapó for the period 1897–2014 (L = 59 years). The error
bars for the eigenvalues were defined according to North et al. (1982). (b) Periodogram calculated for the First Reconstructed Principal Component (RPC1) of Copiapó (27°S
Lat.).

driven by one or more outliers (Chatfield, 2013). When the time be associated to QBO and ENSO (see Fig. 4a) as stated in previous
series is completely random, and the sample size is large, the cross studies for different regions of Chile (Villalba et al., 1998; Fagel
lagged-correlation coefficient is approximately normally dis- et al., 2008; Le Quesne et al., 2009). In addition, 11.7% of these
tributed with the mean 0 and the variance 1/N. Therefore the sig- annual leading modes revealed oscillations between 9 and
nificance for the cross-correlation analysis was calculated 30 years, which are expected to be connected primarily with the
assuming that the approximate critical level of correlation at 95% decadal to inter-decadal modulation of the PDO (Fagel et al.,
pffiffiffiffi
significance (a = 0.05) is r95 ¼ 0 þ 2= N; where N is the sample 2008; Masiokas et al., 2010). The remaining 9.2% of the annual
size. According to this criterion, the required level of correlation leading modes revealed multi-decadal oscillations (>30 years) that
for significance becomes very small in a large sample size can be linked to the multi-decadal signal of the PDO, as it can vary
(Chatfield, 2013). irregularly from inter-annual to inter-decadal time scales (Mantua
et al., 1997; Deser, 2010; Dai, 2013; Dong and Dai, 2015). No stud-
ies have reported a possible influence of the AMO signal in Chile;
3. Results and discussion however, as revealed in the following sections of this study the
multi-decadal modulation of the North Atlantic Ocean also influ-
3.1. Leading modes of precipitation and explained variance ences low-frequency precipitation variability in Chile. In addition,
as indicated by Urrutia et al. (2011), the Half-Gleissberg solar mag-
Most of the significant low-frequency oscillatory components of netic cycle could be the external forcing influencing the multi-
precipitation variability were identified within the first five leading decadal variability of ocean-atmosphere processes. The results
modes of annual, seasonal, and monthly precipitation. The sum of found at an annual scale agree with those observed for monthly
these five leading modes can explain annual precipitation variance precipitation (Fig. 4b), and they are also consistent during wet
ranging from 20.3% (Santiago) to 51% (Puerto Aysen), with an aver- (Fig. 4c) and dry (Fig. 4d) seasons.
age of 31.2% when considering all the rain gauges analyzed
(Table 2). These explained variances are slightly smaller than those 3.3. Phase-space representation
calculated by Aravena and Luckman (2009) for the first four lead-
ing modes of annual regional averaged indices of precipitation in Most of these frequencies can be explained by looking at the
the Chilean Patagonia (41–53°S, between 1961–2000); but this is phase diagrams that represent the trajectories of coupled PCs
expected, given the different periods and aggregations analyzed. and are commonly found within the first few leading modes of
For instance, the limited length of the PC series used by Aravena variability (Fig. 5). This is designed to seek not periodicities, but
and Luckman (40 years) only allowed for the identification of fre- rather repeated trajectories in the phase space (Kim et al., 2015).
quency oscillations <20 years, which limited the ability to detect The PCs are related in time as they trace out oscillations in the
the signal of multidecadal climatic variations such as the PDO or precipitation time series. The low- and high-frequency cycles of
AMO. precipitation are characterized by different trajectories. Multi-
decadal oscillations (>30 years), for instance, are represented by
3.2. Periodicities of the leading modes circular trajectories. High-frequency cycles are represented by dif-
ferent geometries depending on the periodicities of the signal i.e.,
The periodicities calculated for the leading modes of annual, triangles (3-yr cycle) and pentagons (5-yr cycle). Well-defined
seasonal, and monthly precipitation ranged mainly from biennial trajectories in the phase-space are revealed at the annual scale
to multi-decadal oscillations. For instance, 79.1% of the first five since the noise level is smaller relative to the precipitation signal
leading modes of annual precipitation (see Table 2), are dominated and vice versa. The irregularities observed in some plots can be
by high-frequency oscillations (2–9 years) which are expected to described mostly by inter annual variability (Hannachi, 2004).
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1159

Table 2
Explained variance (%) of the first five leading modes of annual precipitation obtained by SSA (L = N/2).

Station/location RPC1 RPC2 RPC3 RPC4 RPC5 Total Variance (%)


% Years % Years % Years % Years % Years
1. Copiapó 10.34(S) 60 10.18(S) 60 5.93(S) 12 5.61(S) 11 4.04 10 36.1
2. La Serena 4.68 73 4.49 2.3 4.38 2.3 4.31 2.3 3.68 2.8 21.5
3. Ovalle 6.43 3.9 5.96 3.9 5.12 2.5 4.91 2.5 4.31 5 26.7
4. Los Andes 6.76 2.5 6.66 2.5 5.57 3.9 5.45 3.9 5.08 2.3 29.5
5. Santiago 4.48 3.8 4.47 3.8 3.83 2.4 3.81 4.4 3.75 4.4 20.3
6. Rancagua 6.36 2.5 5.68 2.5 5.67 2.4 5.46 2.3 4.95 10.7 28.1
7. Talca 6.02 2.2 5.44 2.2 5.12 10.7 5.1 10.7 4.24 6.4 25.9
8. Concepcion 6.54(S) 3 6.38(S) 3 4.68 6.7 4.64 6.7 4.32 2.5 26.6
9. Temuco 8.94(S) 51 5.83(S) 6 5.72(S) 6 5.46(S) 4.5 5.42(S) 4.5 31.4
10. Puerto Montt 9.15 54 7.98 54 6.72 4.5 6.54 4.5 5.0 3.8 35.4
11. Puerto Aysen 14.69(S) 42 14.5(S) 42 8.17(S) 42 7.11(S) 14 6.51(S) 8.4 51.0
12. Punta Arenas 13.49(S) 42.7 11.85(S) 42.7 5.95(S) 7.1 5.66(S) 7.1 4.58(S) 2.56 41.5

(S) Represent significant modes calculated according to North (1982).

Fig. 4. Annual, monthly, and seasonal (wet and dry) histograms of the periodicities extracted for the first five leading modes of precipitation at each raingauge. The blue boxes
highlight high-frequency oscillations as those related to QBO and ENSO (2–9 years). The green boxes are low-frequency oscillations likely associated to the PDO (10–
30 years). The red boxes represent multi-decadal oscillations, the focus of analysis in this study. (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)

3.4. Low-frequency modes of precipitation at those rain gauges located along the westerlies wind belt located
between 40° and 55°S. As a result, the RPC1 of annual accumula-
3.4.1. Annual and monthly scales tion in Puerto Aysen revealed a marked 42-year cycle that
The low-frequency modes of annual and monthly precipitation explained 14.3% of the variance. Punta Arenas showed two signif-
(>30 years) explain the largest variance at those rain gauges icant leading modes (RPC1 and RPC2) that explain 20.5% of the
located in the northern and southern regions of the country total variance and evolve in a cycle fluctuating between 42 and
(Fig. 6). For instance, the RPC1 of annual precipitation variability 64 years. In these regions, high-frequency precipitation is largely
in Copiapó, located in northern Chile at around 27°S, reveals a sig- influenced by the Southern Annular Mode (SAM) or AAO, in
nificant semi-centennial cycle fluctuating between 54 and 64 years combination with ENSO (Aravena and Luckman, 2009; Elbert
that explains 10.3% of the total variance. This rain gauge is placed et al., 2011). However, the regime is quite variable, both in space
in the southern part of the Atacama Desert, where precipitation is and time, because of the topography of the Andes and changes in
largely controlled by the sub-tropical high-pressure belt the strength and latitudinal position of the southern Westerlies
(15°–30°S) (Houston, 2006). The upwelling cold Peruvian Current (Garreaud et al., 2009). For the same region of southern
inhibits the moisture capacity of onshore winds (below 1000 m.a. South-America, Aravena and Luckman (2009) detected frequencies
s.l), and the proximity of the Andean Cordillera restricts moisture of fewer than 15 years, observing decadal variations (12 years) in
advection from the east (Houston, 2006). Therefore, winter precip- northwestern (41–44°S) and central Patagonia (45–47°S); and
itation originates largely from northerly and easterly moving fron- 2–7 year cycles with a strong biennial oscillation mode in the
tal systems originating from the Pacific (Vuille and Ammann, Patagonia Plains-Atlantic (43–50°S) and southern Patagonia
1997). Surprisingly, this multi-decadal pattern was also significant (51–53°S).
1160 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Fig. 5. Phase-space representation of different eigencouples representing trajectories for low and high-frequency oscillations in long rain gauge records of Chile. The
explained variance of each eigencouple is represented as percentage and the periodicity of the signal is stated in parentheses.

Fig. 6. Reconstructed leading modes of annual and monthly precipitation.

3.4.2. Wet and dry seasons pendent. Thus, the positive (negative) phase of the Southern Oscilla-
The rain gauges shown in Fig. 6 contain a notable multi-decadal tion and a strengthened (weak) subtropical high tend to be associated
cycle, which is also apparent at seasonal scale. The low-frequency with an increased (reduced) baroclinicity at mid-latitudes in the
RPC1 of those precipitation records is statistically significant in both southeast Pacific (Quintana and Aceituno, 2012). Our results agree
wet (Apr.-Sept.) and dry (Oct.-Mar.) seasons, being slightly stronger with those found by Núñez et al. (2013) who showed that changes
during the wet periods (Fig. 7). The RPC1 observed for those rain in precipitation in central Chile (30–36°S) are not strongly related
gauges located in central Chile revealed mostly high-frequency to shifts between the warm and cold phases of the PDO.
oscillations (2–9 years), which cannot be attributed to the PDO or
AMO modulation. For instance, a significant biennial cycle was 3.4.3. Climate indices
revealed for the RPC1 of Santiago during dry seasons, which suggests The low-frequency signals of the SOI and AAO did not share
high ENSO-modulation (Rutllant and Fuenzalida, 1991; Aceituno similarities with the low-frequency cycles of precipitation
and Garreaud, 1995; Garreaud and Batisti, 1999; Montecinos and observed in the extreme arid (north) and cold (south) regions of
Aceituno, 2003; Valdés-Pineda et al., 2016). The tropical and extra- Chile. This pattern is consistent at annual, seasonal, and monthly
tropical factors affecting precipitation in central Chile are not inde- scales. The periodicities calculated for the low-frequency signal
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1161

Fig. 7. Reconstructed low-frequency oscillations for the leading modes of precipitation during wet (Apr.-Sept.) and dry (Oct.-Mar.) seasons.

(RPC1 + RPC2) of mean annual SOI (Fig. 8a) ranged between 5.8 and between the multi decadal modes of precipitation and climate
6.8 years, with a total explained variance of 5.9% (Fig. 8b and c); indices using MSSA is detailed in the following section.
and the low-frequency component of the mean annual AAO ranged
between 5.1 and 6.0 years with an explained variance of 14.7%. 3.5. Multi-channel analysis of precipitation variability
These results are consistent with those described in Section 1 of
this paper i.e., studies in which ENSO and AAO activity are strongly Compared to the results from using SSA, the determination of
modulated by high-frequency modes of variability, which are not common oscillatory signals between precipitation and climate
the focus of this study. indices using MSSA was associated with lower variances of the
The first two reconstructed leading modes (RPC1 and RPC2) of leading modes at all temporal scales. This is due to the fact that
the PDO (Fig. 8j) reveal a significant multi-decadal 51- to 58-year precipitation is easily influenced by local factors, which translates
cycle with a total explained variance of 33.2% (Fig. 8k and l) which into more complex spatio-temporal variability within the matrix
is consistent at annual, seasonal, and monthly scales. This lower containing the time series (Antunes et al., 2006; Gomes, 2011).
frequency component of the PDO, with its characteristic multi- The first three low-frequency modes (L = 52 years) associated with
decadal timescale of about 60 years (d’Orgeville and Peltier, the matrix – which includes all rain gauges, the SOI, PDO, and AMO
2007), has indicated regime shifts associated to SSTs that can range (1912–2014) – revealed significant explained variances of 4.7%
between 50 and 80 years (see i.e. Minobe, 1997; Lee et al., 2013; (RPC1), 3.6% (RPC2), and 3.5% (RPC3). The low values observed in
Steinman et al., 2015). A 20-year quasi-periodic oscillation these multi-decadal modes are in central Chile, which is mostly
observed in RPC5 and RPC6 is also suggested in previous studies dominated by high-frequency ENSO modulation. Weighted cross-
(Minobe, 1997; d’Orgeville and Peltier, 2007), as well as a quasi- correlations between normalized multi-decadal components of
30- to 35-year cycle observed in the RPC9, which is not significant precipitation and climate indices confirm positive relationships
under this SSA (L = N/2). This cycle, however, could be significantly between pairs of multi-decadal modes extracted from those rain
related to inter-decadal low-frequency variability in north-central gauges located in central Chile. These normalized modes are posi-
Chile as indicated by Núñez et al. (2013). tively correlated with the PDO and negatively correlated with the
The existence of multi-decadal cycles was first evinced by SOI and AMO (Fig. 9a), suggesting possible large-scale multi-
Schlesinger and Ramankutty (1994), who detected a 65- to 70- decadal patterns that affect precipitation variability in central
year cycle for global mean temperature records. In addition, SST Chile. The multi-decadal cycles, however, are much stronger and
records of 11 globally distributed geographical regions showed that more significant in the extreme regions of the country. Therefore,
this multi-decadal 50- to 90-year cycle is related to the first Empir- they constitute a primary basis upon which to explain the past,
ical Orthogonal Function(EOF)of SST anomalies in the North Atlantic present, and future periodicities of extreme hydrologic events in
region, known as the Atlantic Multi-decadal Oscillation (AMO) those regions of Chile.
(Knight et al., 2005; Justino and Peltier, 2005; Trenberth and Shea, When only Copiapó and Punta Arenas are considered in the
2006). Specifically, positive (negative) phases of the AMO coincide analysis with the SOI, PDO, and AMO (L = 52 years), the explained
with warmer (colder) North Atlantic SSTs (Enfield et al., 2001; variances doubled compared to the previous results, and the
McCarthy et al., 2015). The leading low-frequency modes of AMO multi-decadal modes are easily found within the first two signifi-
were calculated to determine possible unrevealed linkages between cant components (Fig. 9b). In this regard, the periodicity of the
the multi-decadal oscillations of precipitation observed over the PDO signal indicates a clear multi-decadal cycle that evolves in an
extreme arid and cold regions of Chile. Results reveal that RPC1 almost identical phase to the multi-decadal cycle of precipitation
+ RPC2 of AMO (Fig. 8m) evolve in a significant multi-decadal 50- extracted for the Atacama Region’s rain gauge (Fig. 9c). The
to 80-year cycle, which explains more than 50% of the total variance amplitude of the multi-decadal PDO signal is greater than the ampli-
(Fig. 8n and o). The determination of common oscillatory signals tude of the AMO and SOI. Therefore, these three large-scale patterns
1162 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Fig. 8. First two reconstructed components (first column); eigenvalues with phase diagram (second column); and power spectra of first two reconstructed components (third
column) of annually averaged time series of Copiapó (1897–2014), Punta Arenas (1888–2014), the SOI (1866–2014), the PDO (1900–2014), and the AMO (1860–2014).
Eigenvalue error bars are defined according to North et al. (1982). The black line represents the original time series of Copiapó, Punta Arenas, the SOI, PDO, and AMO, and the
red line represents the low-frequency RPC1 + RPC2 obtained from SSA with L = N/2. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

should have different levels of influence over long-term precipita- dent (10-year-lagged), since it shows a weaker linkage when
tion variability in arid (i.e., Copiapó) and cold (i.e., Punta Arenas) outlining the trajectory of both components (Fig. 10c). The AMO
regions of Chile. For instance, if both ENSO and the PDO are in the signal indicates phase shifts of about 25 years, which are weaker
same phase, it is suggested that El Niño/La Niña impacts can be than the multi-decadal mode of Punta Arenas (Fig. 10d). Moreover,
intensified. Conversely, if ENSO and the PDO are out of phase, they the phase-space representation of both rain gauges did reveal a
may offset one another, preventing the impacts of a strong ENSO. relationship to a small degree, and the most notable pattern is
essentially that of Copiapó offsetting Punta Arenas by about
3.6. Phase-space of multi-decadal modes 12 years (Fig. 10e and f).
Keeping in mind these previous relationships, it is important to
The evolution between each pair of multi-decadal cycles of pre- add that on the timescale of the PDO series (1900–2014), AMO is
cipitation and climate indices can be defined as a dynamic system. shown to lag the PDO by 17 years (Fig. 10g) and to lead the
This system is described by trajectories such as those represented PDO by approximately 13 years (Fig. 10h), and. These latter results
by phase-space diagrams (see i.e. Sivakumar et al., 2002; Nolte, are in agreement with those found by d’Orgeville and Peltier
2010; Kim et al., 2015). These trajectories reveal significant offsets (2007), who stated that this relationship between the AMO and
i.e., the trajectory representing the relationship between each pair the 60-year component of the PDO are signatures of the same oscil-
of multi-decadal modes showed that the PDO (RPC1 + RPC2) is sig- lation cycle.
nificantly but slightly out of phase (4-year-lagged) with the series
of Copiapó (Fig. 10a). Surprisingly, the multi-decadal AMO signal 3.7. Spatial significances of multi-decadal modes
(RPC1 + RPC2) shows a strong negative (20-year-lagged) offset
from the same multi-decadal component (Fig. 10b). For Punta Are- The projection of the multi-decadal signals of the PDO and AMO
nas (RPC1 + RPC2) the positive-lagged PDO relationship is less evi- (MSSA, L = 52 years) onto grids of global SSTs revealed unprece-
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1163

Fig. 9. (a) Matrix of w-correlations for the 15 reconstructed components of annual precipitation, the SOI, PDO, and AMO. (b) Eigenvalue spectrum for the Trajectory (Henkel)
matrix of annually averaged time series of Copiapó, Punta Arenas, the SOI, PDO, and AMO, obtained from MSSA analyses. The error bars for the eigenvalues were defined
according to North et al. (1982). (c) Multi-decadal modes (RPC1 + RPC2) of Copiapó, Punta Arenas, the SOI, PDO, and AMO.

Fig. 10. Phase-space diagrams representing the trajectory between the multi-decadal (low-frequency) modes of annual precipitation in Copiapó (27.4°S) and Punta Arenas
(53.1°S), and the multi-decadal signals the PDO and AMO.

dented ocean-climate linkages for Chile. We observed that the previous results, this spatial dipole was also found (at all temporal
multi-decadal PDO signal projected onto global SSTs resulted in scales) when we projected the multi-decadal mode of Copiapó onto
two significant bands concentrated over low (strong pattern) and global SSTs (Fig. 11b). The spatial distribution of this signal showed
high (weak pattern) latitudes of South America; this suggests that to be significantly stronger during dry conditions (see Appendix).
large-scale precipitation variability at annual and seasonal scales The multi-decadal warming over the North Atlantic Ocean denotes
(wet and dry conditions) is driven by a significant multi-decadal global AMO-like significant patterns. This low-frequency signal,
dipole warming over the Pacific Ocean (Fig. 11a). As expected from which is significantly strong over the North Atlantic, contrasts with
1164 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

a weakened and not significant pattern over the eastern Equatorial signals constitute an additional and important element to be con-
and subtropical Pacific Ocean (Fig. 11c). These weaker conditions sidered when predicting future impacts associated with extreme
over the eastern Pacific basin are not evident over the western hydrological events such as floods and droughts. For example, the
Pacific, which reveals significant patterns concentrated over low extreme flood events that occurred during March and July of 2015
latitudes that can be associated with enhanced precipitation condi- in the Atacama Desert (Wilcox et al., 2016), which resulted in sev-
tions over northeast Australia (Sun et al., 2015) and Southeast Asia eral deaths and economic loses, were generated by two different
(i.e., Indonesia, the Philippines, China, Japan, etc.) (see Lu et al., ocean-atmosphere processes that occurred during the high phase
2006; Zhang and Delworth, 2006; Li et al., 2015). This pattern also of AMO. The first event that occurred at the end of the dry season
reveals two elongated significant bands of AMO that extend (end of March) was recently defined by Barrett et al. (2016) as a cut-
towards mid- and high latitudes of the eastern Pacific at both off low with anomalous heights that moved east and was directed
hemispheres. This ocean-connection (warm AMO) means more ashore near 25°S (see Fig. C1 in the Appendix) by circulation around
precipitation along the northwestern United States (see Enfield a very strong ridge centered near 60°S. This phenomenon coincided
et al., 2001; McCarthy et al., 2015), but it also shows clear influ- with warmer temperatures in central Chile and anomalously high
ences over the multi-decadal signal of annual precipitation found PW content (20 mm above climatological norms) over the Peruvian
for Punta Arenas (Fig. 11d). This annual signal also reveals a con- Bight. North-westerlies picked up higher PW over an area of the
siderable connection with the South Atlantic Ocean, which is sig- subtropical eastern Pacific Ocean (similar to that observed in
nificant but weakened during wet and dry conditions (see Fig. 11) that had warmed rapidly during the prior week, and was
Figs. B1 and B2 in the Appendix). transported over the northern regions of Chile. The second event
occurred during the rainiest month of the wet season (July) where
moisture from westerlies reached the region as a frontal system due
4. Discussion
to a weakened SEPA (northward displacement). This finding coin-
cides with the significant positive correlation of AMO with PW at
4.1. Multi-decadal cycles of precipitation variability
500 mb over the north and south regions of Chile during the wet
season (Apr.-Sep.). In the North of Chile the correlation concen-
The emergence of an approximately multi-decadal oscillation
trates over the area spanned by the SEPA, but in the South it appears
has been investigated in different areas of the earth sciences
as an elongated band extended from New Zealand towards south-
(Klyashtorin et al., 2009). For instance, long tide gauge records
ern Chile. This correlation pattern is also significantly positive along
show a significant oscillation with a period of around 60 years
the North-East Pacific coastline and reaches countries all along
for the majority of the tide gauges examined during the 20th cen-
South America, Central America, Southern Mexico and the United
tury in every ocean basin of the world (Chambers et al., 2012). In
States (Fig. 12a). It is important to mention that, during the high-
general these multi-decadal oscillations are characterized by 30-
phase of AMO, the PW reduces in the Pacific Basin and produces
year and 50- to 70-year periodic cycles (Klyashtorin and
neutral to negative anomalies of PW in Central Chile; however,
Lyubushin, 2007; Klyashtorin et al., 2009). In this study, we
the subtropical and extratropical regions of Chile reveal positive
extracted significant multi-decadal components of precipitation
anomalies of PW, which is part of a pattern extended all over the
that explain the largest variances at those rain gauges located in
American Continent. On the other hand, during the dry season
the extreme northern (arid) and southern (cold) regions of Chile.
(Oct.-Mar.) this pattern of positive correlations is still observed in
The periodicities calculated for these multi-decadal components
the coastline of northern Chile but no longer in southern Chile
indicate cycles ranging between 40 and 64 years. These cannot
(Fig. 12b). According to these findings, both March and July events
be associated with the modes of the SOI (AAO), since their compo-
could be potentially associated to the high phase of AMO (probably
nents (at least SOI) show a weak multi-decadal signal, with a
intensified by 2015–2016 ENSO), which is strongly associated to a
source of variability mostly dominated by strong high-frequency
warmer ocean (Bozkurt et al., 2016), higher precipitable water
cycles (<10 years). The strong high-frequency cycles introduce
amounts and positive ZW anomalies at 500 mb (Fig. 13a and b) in
more noise into the precipitation records because of intensification
northern and southern regions of Chile. Such events must be also
or weakening of precipitation contributions. This led to a much
related to the interaction between different phases of ENSO, the
more difficult identification of multi-decadal cycles of precipita-
PDO, and the AMO which can intensify or diminish the impacts of
tion in central Chile and, as a result, most of the rain gauges pro-
each other according to their phases (see i.e. Wang et al., 2013).
duced no significant outcomes. This relationship also remained
When both ENSO and the PDO are in the same phase, it is expected
consistent and significant during wet and dry seasons.
that El Niño (La Niña) impacts may be magnified generating posi-
On the other hand, this large-scale temporal variability of pre-
tive (negative) precipitation anomalies over Chile. Accordingly,
cipitation seems to be highly associated with the multi-decadal
we hypothesize that this reinforcing was also evident during the
features observed over the Pacific (PDO) and Atlantic (AMO)
recent positive phases of both ENSO and AMO, producing enhanced
Oceans (i.e. Enfield et al., 2001; McCabe et al., 2004; McCarthy
positive anomalies of precipitation in Chile and other regions of the
et al., 2015). In fact, these multi-decadal signals are significantly
World. The application of this knowledge will potentially benefit
correlated with those records of rain gauges located in the subtrop-
the predictive capacity of future extreme hydrometeorological
ical and extratropical regions of Chile. The PDO signal is stronger to
events in semiarid basins of northern Chile (ungauged) where times
the north of the country (Copiapó), whereas the AMO signal is
of concentration are low and lead times are essentially required to
strong in both the north and the south of the country (Punta Are-
minimize impacts on the society.
nas). Positive SST anomalies are observed during the positive phase
of both oceanic oscillations. This pattern is evident from the global
4.3. Ongoing and future research
spatial correlation of low-frequency precipitation modes and SSTs.
The recent opposition between the warm (positive) phase of the
4.2. Multi-decadal oscillations and extreme hydrometeorological AMO and a substantially negative-trending (cold phase) of the
events multi-decadal PDO signal (or Pacific Multi-decadal Oscillation,
PMO) was apparently producing the slowdown or false pause of
These unprecedented spatial linkages between the multi- warming temperatures of the past decade (Steinman et al., 2015).
decadal precipitation variability over Chile and the PDO and AMO This global warming hiatus has been also described as the result
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1165

Fig. 11. Multi-decadal modes of annual signals of (a) the PDO; (b) Copiapó (27.4°S); (c) the AMO; and (d) Punta Arenas (53.1°S), projected onto SSTs (see Hirahara et al.,
2014). The green marker is the Copiapó rain gauge, and the yellow marker is the Punta Arenas rain gauge. Colored patterns represent significant correlation coefficients
calculated by the large sample approximation method. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)

Fig. 12. Correlation map for the relationship between the raw signal of AMO and Fig. 13. Correlation map for the relationship between the raw AMO signal and
the amount of PW at 500 mb. Blue colors represent positive correlations and red zonal wind speeds at 500 mb. Blue colors represent positive correlations and red
colors negative correlations. (For interpretation of the references to colour in this colors negative correlations. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.) figure legend, the reader is referred to the web version of this article.)
1166 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

of interactions between a secular warming tendency due in a large Appendix A. Singular Spectrum Analysis (SSA)
part to the buildup of anthropogenic CO2 concentrations, and an
internally generated cooling due to a cool phase of a quasi-60- Traditionally the EOF method finds a combination of variables
year oscillatory variability that is closely associated with the which explains most of the variability, usually observing high cor-
AMO and PDO (Yao et al., 2015). Despite the hiatus periods that relation in space. However, the significant correlation in time is
have appeared in surface air temperature records, other compo- not taken into account. Auto- and cross-correlation in time are
nents of the climate system associated with warming have contin- very useful for prediction purposes and also for building stochas-
ued at multi-decadal scales. These include displacements in the tic time series models. An extension of the traditional EOF
mean latitudinal position of the Intertropical Convergence Zone method is the Extended Empirical Orthogonal Functions (EEOFs),
(ITCZ) that vary on a similar cycle to the AMO (Knudsen et al., which is a technique to deal not only with spatial but also with
2011), and likely linkages with external forces related to astronom- temporal correlations observed in climate data. The method was
ical dynamics (see i.e. Scafetta, 2010; Knudsen et al., 2014). These first introduced by Weare and Nasstrom (1982), who applied it
interrelations must be clarified in future studies in order to to the 300-mb Relative Vorticity. A similar approach was devel-
improve forecasting schemes and climate scenarios for near- and oped to deal with the dynamical reconstruction of low order
long-term impact assessment of extreme hydrological events. chaotic systems by Broomhead and King (1986a,b) who called it
These studies will need to include the effect of the multi-decadal Singular System Analysis (SSA) in the one-dimensional case.
climate variability on the hydrological regime of Chilean water- Fraedrich (1986) proposed the name of the method as Singular
sheds as a way to improve water governance schemes. Spectrum Analysis (SSA), and a modified version named
Multichannel-SSA (M-SSA). Recently, Patterson et al. (2011)
developed the non-parametric method of Multivariate Singular
5. Conclusions
Spectrum Analysis (MSSA) for multi-vintage data. This approach
has proven much more flexible than the standard methods of
Multi-decadal cycles of precipitation in Chile are significant at
modelling, which involve at least one of the restrictive assump-
annual, seasonal, and monthly scales. The high-frequency compo-
tions of linearity, normality, and stationarity.
nent associated with precipitation records of central Chile is mostly
Over the last few decades, SSA has been considered a powerful
related to the QBO, ENSO, and AAO. The ENSO (represented by SOI)
technique in time series analysis, and it has been developed and
and the AAO did not explain the multi-decadal precipitation vari-
applied to many practical problems. SSA has been described as a
ability, since their frequencies are lower than those observed for
very useful tool that can be used for solving the following prob-
the multi-decadal leading modes of precipitation variability. The
lems: finding trends of different resolution; smoothing; extraction
strong influence of the PDO and AMO signals is significant over
of seasonality components; simultaneous extraction of cycles with
the multi-decadal precipitation variability of the extreme regions
small and large periods; extraction of periodicities with varying
of Chile i.e., those located in low and high latitudes. Specifically,
amplitudes; simultaneous extraction of complex trends and peri-
the PDO correlations are evident in the north of the country,
odicities; finding structure in short time series; and change-point
whereas correlations with the AMO are strong both in the north
detection. (Vautard and Ghil, 1989; Ghil et al., 2002; Golyandina
and in the south of the country. For instance, the raw signal of
et al., 2001; Hassani et al., 2009; Hassani and Thomakos, 2010).
AMO revealed to be significantly correlated with SST, PW and zonal
The SSA decomposes the original time series into a sum of small
winds (both measured at 500 mb), suggesting that this multi-
numbers of interpretable components, such as a slowly varying
decadal variability is not only observed in the ocean but also in tro-
trend, an oscillatory component, and noise. The basic SSA method
pospheric processes, since the positive phase of AMO is also related
consists of two complementary stages, Decomposition and Recon-
to higher PW and positive anomalies of zonal winds at 500 mb.
struction. Each stage includes two separate steps. At the first stage,
Based on our results, we conclude that the PDO and AMO are two
the series are decomposed to obtain the Trajectory Matrix; at the
of the most important factors influencing multi-decadal precipita-
second stage the noise-free series are reconstructed for forecasting
tion variability in these northern and southern regions of Chile.
new data points. A detailed description of the SSA technique is
The result of the relationships between warm and cold phases
given in several publications. For more explanations and compar-
of the ENSO, AAO, PDO, and AMO must be researched further to
ison with other time series analysis techniques, refer to the refer-
determine the physical factors or patterns that explain this
ences provided in this article.
multi-decadal mode of variability in Chile and other regions of
In general terms, the whole procedure of the SSA technique
the World. The possible connection between this multi-decadal
depends upon two parameters:
pattern and external forcing also must be examined in future
studies. Gridded data and assembled models can be used as part
I. The lagged window length (L).
of a future integrated approach to draw new conclusions about
II. The number of needed singular values for reconstruction (r).
the physical origin and interrelations of these signals and these
conclusions can be used to improve water governance schemes
An improper choice of parameters L and r may yield incomplete
at catchment scale.
reconstruction and misleading results in forecasting. Recent results
obtained by Mahmoudvand and Zokaei (2012), show that the
Acknowledgments choice of L close to one-half (N/2) of the time series length is a suit-
able choice for the decomposition stage in most cases of SSA. How-
The authors acknowledge the Dirección General de Aguas de ever, in the case of climatic indices the oscillations reported in
Chile and the Dirección Meteorológica de Chile for providing the previous studies may be used as an appropriate lagged window
data used in this study. The first author received support from length for analysis purposes.
the Comisión Nacional de Investigación Científica y Tecnológica
de Chile (CONICYT) to carry out this research. The second author
received partial support from the NSF-USAID program PEER cycle Appendix B. Spatial significances of multi-decadal modes of wet
1 (project 31) and Universidad de Antioquia project PI12-1-03 to and dry seasons
participate in this research. These contributions are greatly
acknowledged.
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1167

Fig. B1. Multi-decadal modes of wet signal of (a) Copiapó (27.4°S); (b) PDO; (c) Punta Arenas (53.1°S), and (d) AMO; projected onto gridded SSTs (see Hirahara et al., 2014) for
1912–2014. Green marker is the Copiapó rain gauge and the yellow marker is the Punta Arenas raingauge. Colored patterns represent significant correlation coefficients
calculated by the large sample approximation method. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. B2. Multi-decadal modes of dry signals of (a) Copiapó (27.4°S); (b) PDO; (c) Punta Arenas (53.1°S), and (d) AMO; projected onto gridded Sea Surface Temperatures (see
Hirahara et al., 2014) for 1912–2014. Green-marker is Copiapó raingauge and yellow-marker is Punta Arenas raingauge. Colored patterns represent significant correlation coefficients
calculated by the large sample approximation method. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
1168 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Appendix C. Extreme event of March 2015 in the Atacama Chambers, D.P., Merrifield, M.A., Nerem, R.S., 2012. Is there a 60-year oscillation in
global mean sea level? Geophys. Res. Lett. 39 (18).
Desert Chile Chatfield, C., 2013. The Analysis of Time Series: An Introduction. CRC Press.
Dai, A., 2013. The influence of the inter-decadal Pacific oscillation on US
precipitation during 1923–2010. Clim. Dyn. 41, 633–646.
Dean, J.M., Kemp, A.E.S., 2004. A 2100 year BP record of the Pacific Decadal
Oscillation, El Niño Southern Oscillation and Quasi-Biennial Oscillation in
marine production and fluvial input from Saanich Inlet, British Columbia.
Palaeogeogr., Palaeoclimatol., Palaeoecol. 213, 207–229.
Deser, C., Alexander, M.A., Xie, S., Phillips, A.S., 2010. Sea surface temperature
variability: patterns and mechanisms. Ann. Rev. Mar. Sci. 2 (1), 115–143. http://
dx.doi.org/10.1146/annurev-marine-120408-151453. PMID 21141660.
Dong, B., Dai, A., 2015. The influence of the Interdecadal Pacific Oscillation on
Temperature and Precipitation over the Globe. Clim. Dyn., 1–15
d’Orgeville, M., Peltier, W.R., 2007. On the Pacific decadal oscillation and the
Atlantic multidecadal oscillation: might they be related? Geophys. Res. Lett. 34
(23).
Dawson, A.G., O’Hare, G., 2000. Ocean-atmosphere circulation and global climate:
the El Niño-Southern Oscillation. Geography 85, 193–208.
Ebdon, R.A., 1975. The quasi-biennial oscillation and its association with
tropospheric circulation patterns. Met. Mag. 104, 282–297.
Elbert, J., Grosjean, M., von Gunten, L., Urrutia, R., Fischer, D., Wartenburger, R.,
Ariztegui, D., Fujak, M., Hamann, Y., 2011. Quantitative high-resolution winter
(JJA) precipitation reconstruction from varved sediments of Lago Plomo 47 S,
Patagonian Andes, AD 1530–2002. The Holocene, 0959683611425547.
Elsner, J.B., Tsonis, A.A., 2013. Singular Spectrum Analysis: A New Tool in Time
Series Analysis. Springer Science & Business Media.
Enfield, D.B., Mestas- Núñez, A.M., Trimble, P.J., 2001. The Atlantic Multidecadal
Oscillation and its relationship to rainfall and river flows in the continental U.S.
Geophys. Res. Lett. 28, 2077–2080.
Fagel, N., Boës, X., Loutre, M.F., 2008. Climate oscillations evidenced by spectral
analysis of Southern Chilean lacustrine sediments: the assessment of ENSO over
the last 600 years. J. Paleolimnol. 39 (2), 253–266.
Falvey, M., Garreaud, R., 2007. Wintertime precipitation episodes in Central Chile:
Fig. C1. Scheme of a trapped cutoff-low in the Pacific in front of the Atacama Region associated meteorological conditions and orographic influences. J.
and centered about 25°S. The phenomenon triggered the transport of moisture from Hydrometeorol. 8 (2), 171–193.
the Peruvian Bight and Northern Chile towards the Atacama Desert Region Source: Fraedrich, K., 1986. Estimating the dimensions of weather and climate attractors. J.
Own source, adapted from Di Liberto (2015). Atmos. Sci. 43 (5), 419–432.
Garreaud, R., Batisti, D., 1999. Inter-annual (ENSO) and inter-decadal (ENSO-like)
variability of the Southern Hemisphere tropospheric circulation. J. Clim. 12,
2113–2123 (PDF).
References Garreaud, R., Vuille, M., Clement, A.C., 2003. The climate of the Altiplano: observed
current conditions and mechanisms of past changes. Palaeogeogr.,
Aceituno, P., Fuenzalida, H., Rosenblüth, B., 1993. Climate along the extratropical Palaeoclimatol., Palaeoecol 194 (1), 5–22. Chicago.
west coast of South America. In: Mooney, H.A., Fuentes, E.R., Kronberg, B.I. Garreaud, R., Vuille, M., Compagnucci, R., Marengo, J., 2009. Present-day South
(Eds.), Earth System Responses to Global Change: Contrast between North and American climate. Palaeogeogr. Palaeoclimatol. Palaeoecol. 281 (3–4), 180–195.
South America. Academic Press, New York, pp. 61–69. http://dx.doi.org/10.1016/j.palaeo.2007.10.032.
Aceituno, P., Garreaud, R., 1995. Impacto de los fenómenos el Niño y la Niña en el Garreaud, R.D., Gabriela Nicora, M., Bürgesser, R.E., Ávila, E.E., 2014. Lightning in
régimen pluviométrico Andino. Revista Chilena de Ingeniería Hidráulica 9, 12– western Patagonia. J. Geophys. Res.: Atmos. 119 (8), 4471–4485.
20. Graystone, P., 1959. Meteorological office discussion on tropical meteorology. Met.
Antunes, S., Pires, O., Rocha, A., 2006. Detecting spatio-temporal precipitation Mag. 88, 117.
variability in Portugal using multichannel singular spectral analysis. Int. J. Ghil, M., Vautard, R., 1991. Interdecadal oscillations and the warming trend in
Climatol. 26 (15), 2199–2212. global temperature time series. Nature 350 (6316), 324–327.
Aravena, J.C., Luckman, B.H., 2009. Spatio-temporal rainfall patterns in Southern Ghil, M., Allen, M.R., Dettinger, M.D., Ide, K., Kondrashov, D., Mann, M.E., Robertson,
South America. Int. J. Climatol. 29 (14), 2106–2120. http://dx.doi.org/10.1002/ A.W., Saunders, A., Tian, Y., Varadi, F., Yiou, P., 2002. Advanced spectral methods
joc.1761. for climatic time series. Rev. Geophys. 40 (1). 3-1.
Baldwin, M.P., Gray, L.J., Dunkerton, T.J., Hamilton, K., Haynes, P.H., Randel Holton, J. Golyandina, N., Nekrutkin, V., Zhigljavsky, A.A., 2001. Analysis of Time Series
R., Alexander, M.J., Hirota, I., Horinouchi, T., Jones, D.B.A., Kinnersley, J.S., Structure: SSA and Related Techniques. CRC Press.
Marquardt, C., Sato, K., Takahashi, M., 2001. The quasi-biennial oscillation. Rev. Golyandina, N., Zhigljavsky, A., 2013. Singular Spectrum Analysis for Time Series.
Geophys. 39 (2), 179–229. Springer Briefs in Statistics. Springer.
Barrett, B.S., Campos, D.A., Veloso, J.V., Rondanelli, R., 2016. Extreme temperature Gomes, P.T., 2011. Interannual oscillations in winter rainfall over Europe: Iberia
and precipitation events in March 2015 in central and northern Chile. J. study case. Finisterra: Revista portuguesa de geografia 46 (91), 27–45.
Geophys. Res.: Atmos. 121 (9), 4563–4580. Hannachi, A., 2004. A Primer for EOF Analysis of Climate Data. Department of
Bendat, J.S., Piersol, A.G., 2011. Random Data: Analysis and Measurement Meteorology, University of Reading. 1–33.
Procedures, vol. 729. John Wiley & Sons. Hassani, H., Heravi, S., Zhigljavsky, A., 2009. Forecasting European industrial
Biondi, F., Gershunov, A., Cayan, D.R., 2001. North Pacific decadal climate variability production with singular spectrum analysis. Int. J. Forecast. 25 (1), 103–118.
since 1661. J. Clim. 14 (1), 5–10. Hassani, H., Thomakos, D., 2010. A review on singular spectrum analysis for
Box, G., Jenkins, G., 1970. Time Series Analysis: Forecasting and Control. Holden- economic and financial time series. Stat. Its Interface 3 (3), 377–397.
Day, San Francisco. Hassani, H., Zhigljavsky, A., Patterson, K., Soofi, A., 2010. A Comprehensive Causality
Box, G.E.P., Jenkins, G.M., Reinsel, G.C., 1994. Time Series Analysis: Forecasting and Test based on the Singular Spectrum Analysis. Causality in Science, Oxford
Control. Prentice Hall, Englewood Cliff, N.J.. University Press.
Bozkurt, D., Rondanelli, R., Garreaud, R., Arriagada, A., 2016. Impact of Warmer Hassani, H., Heravi, S., Zhigljavsky, A., 2013. Forecasting UK industrial production
Eastern Tropical Pacific SST on the March 2015 Atacama Floods. Mon. Weather with multivariate singular spectrum analysis. J. Forecasting 32 (5), 395–408.
Rev. 144 (11), 4441–4460. Hassani, H., Mahmoudvand, R., 2013. Multivariate singular spectrum analysis: a
Broomhead, D.S., King, G.P., 1986a. Extracting qualitative dynamics from general view and new vector forecasting approach. Int. J. Energy Stat. 1 (01),
experimental data. Physica D 20 (2), 217–236. 55–83.
Broomhead, D.S., King, G.P., 1986b. On the qualitative analysis of experimental Hassani, H., Webster, A., Silva, E.S., Heravi, S., 2015. Forecasting US tourist arrivals
dynamical systems. Nonlinear Phenomena Chaos 113, 114. using optimal singular spectrum analysis. Tourism Manage. 46, 322–335.
Brunet, M., Jones, P., 2011. Data rescue initiatives: bringing historical climate data Hirahara, S., Ishii, M., Fukuda, Y., 2014. Centennial-scale sea surface temperature
into the 21st century. Clim. Res. 47 (1), 29. analysis and its uncertainty. J. Clim. 27, 57–75. http://dx.doi.org/10.1175/JCLI-
Cañón, J., González, J., Valdés, J.B., 2007. Precipitation in the Colorado River Basin D-12-00837.1.
and its low frequency associations with PDO and ENSO signals. J. Hydrol. 333 Houston, J., 2006. Variability of precipitation in the Atacama Desert: its causes and
(2), 252–264. hydrological impact. Int. J. Climatol. 26, 2181–2198. http://dx.doi.org/10.1002/
Carrasco, J.F., Casassa, G., Rivera, A., 2002. Meteorological and climatological aspects joc.1359.
of the Southern Patagonia Icefield. In: The Patagonian Icefields. Springer, US, pp. Huesca, M., Merino-de-Miguel, S., Eklundh, L., Litago, J., Cicuéndez, V., Rodríguez-
29–41. Rastrero, M., Ustin, S., Palacios-Orueta, A., 2015. Ecosystem functional
R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170 1169

assessment based on the ‘‘optical type” concept and self-similarity patterns: An Núñez, J., Rivera, D., Oyarzún, R., Arumí, J.L., 2013. Influence of Pacific Ocean
application using MODIS-NDVI time series autocorrelation. Int. J. Appl. Earth multidecadal variability on the distributional properties of hydrological
Observ. Geoinformation. variables in north-central Chile. J. Hydrol. 501, 227–240.
Justino, F., Peltier, W.R., 2005. The glacial North Atlantic Oscillation. Geophys. Res. Partington, J.R, 1988. An introduction to Hankel operators. LMS Student Texts 13.
Lett. 32 (21). Cambridge University Press. ISBN 0-521-36791-3.
Kim, S., Kim, Y., Lee, J., Kim, H.S., 2015. Identifying and evaluating chaotic behavior Patterson, K., Hassani, H., Heravi, S., Zhigljavsky, A., 2011. Multivariate singular
in hydro-meteorological processes. Adv. Meteorol. spectrum analysis for forecasting revisions to real-time data. J. Appl. Stat. 38
Klyashtorin, L.B., Lyubushin, A.A., 2007. Cyclic Climate Changes and Fish (10), 2183–2211.
Productivity. VNIRO (All Russian Institute Fisheries and Oceanography) Peristykh, A.N., Damon, P.E., 2003. Persistence of the Gleissberg 88-year solar cycle
Publishing, Moscow. 224. over the last 12,000 years: evidence from cosmogenic isotopes. J. Geophys. Res.
Klyashtorin, L.B., Borisov, V., Lyubushin, A., 2009. Cyclic changes of climate and 108 (A1), 1003.
major commercial stocks of the Barents Sea. Mar. Biol. Res. 5 (1), 4–17. Pizarro, R., Valdés-Pineda, R., García-Chevesich, P., Vallejos, C., Sangüesa, C.,
Knight, J.R., Allan, R.J., Folland, C.K., Vellinga, M., Mann, M.E., 2005. A signature of Morales, C., Balocchi, F., Abarza, A., Fuentes, R., 2012. Latitudinal analysis of
persistent natural thermohaline circulation cycles in observed climate. rainfall intensity and mean annual precipitation in Chile. Chilean J. Agric. Res.
Geophys. Res. Lett. 32 (20). 72 (2), 252–261.
Knudsen, M.F., Seidenkrantz, M.S., Jacobsen, B.H., Kuijpers, A., 2011. Tracking the Quintana, J., Aceituno, P., 2012. Changes in the rainfall regime along the
Atlantic Multidecadal Oscillation through the last 8,000 years. Nat. Commun. 2, extratropical west coast of South America (Chile): 30–43°S.Atmósfera 25 (1),
178. 1–22.
Knudsen, M.F., Jacobsen, B.H., Seidenkrantz, M.S., Olsen, J., 2014. Evidence for Rao, A.R., Hamed, K., 2003. Multi-taper method of analysis of periodicities in
external forcing of the Atlantic Multidecadal Oscillation since termination of the hydrologic data. J. Hydrol. 279 (1), 125–143.
Little Ice Age. Nat. Commun. 5. Rubio-Álvarez, E., McPhee, J., 2010. Patterns of spatial and temporal variability in
Krishnamurthy, V., Misra, V., 2010. Observed ENSO teleconnections with the South streamflow records in south central Chile in the period 1952–2003. Water
American monsoon system. Atmos. Sci. Lett. 11 (1), 7–12. Resour. Res. 46 (5).
Krishnamurthy, V., Stan, C., 2015. Simulation of the South American climate by a Rutllant, José., Fuenzalida, Humberto., 1991. Synoptic aspects of the Central Chile
coupled model with super-parameterized convection. Clim. Dyn. 44 (9–10), Rainfall variability associated with the Southern Oscillation’. Int. J. Climatol. 11,
2369–2382. 63–76.
Krishnamurthy, L., Krishnamurthy, V., 2015. Teleconnections of Indian monsoon Sanz Donaire, J.J., 2012. Las series anuales de precipitación más largas de Chile:
rainfall with AMO and Atlantic tripole. Clim. Dyn., 1–17 estudio y enseñanzas. Estudios Geográficos 73 (273), 625–656.
Kunkel, K.E., Andsager, K., Easterling, D.R., 1999. Long-term trends in extreme Scafetta, N., 2010. Empirical evidence for a celestial origin of the climate oscillations
precipitation events over the conterminous United States and Canada. J. Clim. and its implications. J. Atmos. Solar Terr. Phys. 72 (13), 951–970.
12 (8), 2515–2527. http://dx.doi.org/10.1175/1520-0442(1999) 012 2515: Schlesinger, M.E., Ramankutty, N., 1994. An oscillation in the global climate system
LTTIEP 2.0.CO;2. of period 65–70 years. Nature 367 (6465), 723–726.
Le Quesne, C., Acuña, C., Boninsegna, J.A., Rivera, A., Barichivich, J., 2009. Long-term Silva, E.S., Hassani, H., 2015. On the use of singular spectrum analysis for forecasting
glacier variations in the Central Andes of Argentina and Chile, inferred from US trade before, during and after the 2008 recession. Int. Econ. 141, 34–49.
historical records and tree-ring reconstructed precipitation. Palaeogeogr. Sivakumar, B., Jayawardena, A.W., Fernando, T.M.K.G., 2002. River flow forecasting:
Palaeoclimatol. Palaeoecol. 281 (3), 334–344. use of phase-space reconstruction and artificial neural networks approaches. J.
Lee, T., Ouarda, T.B.M.J., Li, J., 2013. An orchestrated climate song from the Pacific Hydrol. 265 (1), 225–245.
and Atlantic Oceans and its implication on climatological processes. Int. J. Steinman, B.A., Mann, M.E., Miller, S.K., 2015. Atlantic and Pacific multidecadal
Climatol. 33 (4), 1015–1020. oscillations and Northern Hemisphere temperatures. Science 347 (6225), 988–
Li, S., Jing, Y., Luo, F., 2015. The potential connection between China surface air 991.
temperature and the Atlantic Multidecadal Oscillation (AMO) in the pre- Sun, C., Li, J., Feng, J., Xie, F., 2015. A decadal-scale teleconnection between the North
industrial period. Sci. China Earth Sci., 1–13 Atlantic Oscillation and subtropical eastern Australian rainfall. J. Clim. 28 (3),
Lu, R., Dong, B., Ding, H., 2006. Impact of the Atlantic Multidecadal Oscillation on the 1074–1092.
Asian summer monsoon. Geophys. Res. Lett. 33 (24). Trenberth, K.E., 1984. Signal versus noise in the Southern Oscillation. Mon. Wea.
Mahmoudvand, R., Zokaei, M., 2012. On the singular values of the Hankel matrix Rev. 112, 326–332.
with application in singular spectrum analysis. Chilean J. Stat. 3 (1), 43–56. Trenberth, K.E., Hoar, T.J., 1997. El Niño and climate change. Geophys. Res. Lett. 24
Machiwal, D., Jha, M.K., 2012. Hydrologic Time Series Analysis: Theory and Practice. (23), 3057–3060.
Springer Science & Business Media. Trenberth, K.E., Shea, D.J., 2006. Atlantic hurricanes and natural variability in 2005.
Machiwal, D., Kumar, S., Dayal, D., 2015. Characterizing rainfall of hot arid region by Geophys. Res. Lett. 33, 12.
using time-series modeling and sustainability approaches: a case study from Unnikrishnan, P., Jothiprakash, V., 2015. Extraction of nonlinear rainfall trends using
Gujarat, India. Theor. Appl. Climatol., 1–15 singular spectrum analysis. J. Hydrol. Eng. 0. 05015007.
Makowsky, J.A., Labai, N., 2015. Hankel Matrices: From Words to Graphs. In Urrutia, R.B., Lara, A., Villalba, R., Christie, D.A., Le Quesne, C., Cuq, A., 2011.
Language and Automata Theory and Applications. Springer International Multicentury tree ring reconstruction of annual streamflow for the Maule River
Publishing, pp. 47–55. watershed in south central Chile. Water Resour. Res. 47 (6).
Mantua, N.J., Hare, S., Zhang, Y., Wallace, J.M., Francis, R.C., 1997. A Pacific Valdés-Pineda, R., Pizarro, R., García-Chevesich, P., Valdés, J.B., Olivares, C., Vera, M.,
interdecadal climate oscillation with impacts on salmon production. Bull. Am. Balocchi, F., Pérez, F., Vallejos, C., Fuentes, R., Abarza, A., Helwig, B., 2014. Water
Meteorol. Soc. 78, 1069–1079. governance in Chile: availability, management and climate change. J. Hydrol.
Marques, C.A.F., Ferreira, J.A., Rocha, A., Castanheira, J.M., Melo-Gonçalves, P., Vaz, 519, 2538–2567. http://dx.doi.org/10.1016/j.jhydrol.2014.04.016.
N., Dias, J.M., 2006. Singular spectrum analysis and forecasting of hydrological Valdés-Pineda, R., Pizarro, R., Valdés, J.B., Carrasco, J.F., García-Chevesich, P.,
time series. Phys. Chem. Earth, Parts A/B/C 31 (18), 1172–1179. Olivares, C., 2015. Spatio-temporal trends of precipitation, its aggressiveness
Masiokas, M.H., Villalba, R., Luckman, B.H., Mauget, S., 2010. Intra-to multidecadal and concentration, along the Pacific coast of South America (36°–49° S). Hydrol.
variations of snowpack and streamflow records in the Andes of Chile and Sci. J. http://dx.doi.org/10.1080/02626667.2015.1085989.
Argentina between 30 and 37 S. J. Hydrometeorol. 11 (3), 822–831. Valdés-Pineda, R., Valdés, J.B., Díaz, H.F., Pizarro-Tapia, R., 2016. Analysis of spatio-
McCabe, G.J., Palecki, M.A., Betancourt, J.L., 2004. Pacific and Atlantic Ocean temporal changes in annual and seasonal precipitation variability in South
influences on multidecadal drought frequency in the United States. Proc. Natl. America-Chile and related ocean–atmosphere circulation patterns.
Acad. Sci. 101 (12), 4136–4141. International Journal of Climatology 36 (8), 2979–3001.
McCarthy, G.D., Haigh, I.D., Hirschi, J.J.M., Grist, J.P., Smeed, D.A., 2015. Ocean impact Vautard, R., Ghil, M., 1989. Singular spectrum analysis in nonlinear dynamics, with
on decadal Atlantic climate variability revealed by sea-level observations. applications to paleoclimatic time series. Physica D 35 (3), 395–424.
Nature 521 (7553), 508–510. Viale, M., Garreaud, R., 2015. Orographic effects of the subtropical and extratropical
Miller, A., 1976. The climate of Chile. In: Schwerdtfeger, W. (Ed.), Climates of Central Andes on upwind precipitating clouds. J. Geophys. Res. Atmos. 120. http://dx.
and South America. Elsevier, Amsterdam, pp. 113–145. doi.org/10.1002/2014JD023014.
Minobe, S., 1997. A 50–70 year climatic oscillation over the North Pacific and North Villalba, R., Cook, E.R., Jacoby, G.C., D’Arrigo, R.D., Veblen, T.T., Jones, P.D., 1998.
America. Geophys. Res. Lett. 24, 683–686. Tree-ring based reconstructions of northern Patagonia precipitation since AD
Minvielle, M., Garreaud, R.D., 2011. Projecting rainfall changes over the South 1600. The Holocene 8 (6), 659–674.
American Altiplano. J. Clim. 24 (17), 4577–4583. Villalba, R., Lara, A., Masiokas, M.H., Urrutia, R.B., Luckman, B.H., Marshall, G.J.,
Montecinos, A., Aceituno, P., 2003. Seasonality of the ENSO-related rainfall Mundo, I.A., Christie, D.A., Cook, E.R., Neukom, R., Allen, K., Fenwick, P.,
variability in central Chile and associated circulation anomalies. J. Clim. 16 Boninsegna, J.A., Srur, A.M., Morales, M.S., Araneo, D., Palmer, J.G., Cuq, E.,
(2), 281–296. Aravena, J.C., Holz, A., Le Quesne, C., 2012. Unusual Southern Hemisphere tree
Di Liberto, T., 2015. Flooding in the Atacama Desert: How did that happen?. NOAA – growth patterns induced by changes in the Southern Annular Mode. Nat. Geosci.
National Oceanic and Atmospheric Administration. Available at: <https:// 5, 793–798.
www.climate.gov/news-features/event-tracker/flooding-atacama-desert-how- Villarroel, C., 2013. Eventos extremos de precipitación y temperatura en Chile:
did-happen> (accessed Sept 14th, 2015). Proyecciones para fines del siglo XXI. Available at: <http://www.repositorio.
Nolte, D.D., 2010. The tangled tale of phase space. Phys. Today. uchile.cl/handle/2250/114066>.
North, G.R. et al., 1982. Sampling errors in the estimation of empirical orthogonal Vuille, M., Ammann, C., 1997. Regional snowfall patterns in the high, arid Andes. In:
functions. Mon. Weather Rev. 110 (7), 699–706. Climatic Change at High Elevation Sites. Springer, Netherlands, pp. 181–191.
1170 R. Valdés-Pineda et al. / Journal of Hydrology 556 (2018) 1153–1170

Wang, B., Liu, J., Kim, H.J., Webster, P.J., Yim, S.Y., Xiang, B., 2013. Northern Yao, S.L., Huang, G., Wu, R.G., Qu, X., 2015. The global warming hiatus—a natural
Hemisphere summer monsoon intensified by mega-El Niño/southern product of interactions of a secular warming trend and a multi-decadal
oscillation and Atlantic multidecadal oscillation. Proc. Natl. Acad. Sci. 110 oscillation. Theoret. Appl. Climatol., 1–12
(14), 5347–5352. Yuan, N., Fu, Z., Zhang, H., Piao, L., Xoplaki, E., Luterbacher, J., 2015. Detrended
Wang, H., Sankarasubramanian, A., Ranjithan, R.S., 2015. Understanding the low- partial-cross-correlation analysis: a new method for analyzing correlations in
frequency variability in hydroclimatic attributes over the southeastern US. J. complex system. Sci. Rep. 5.
Hydrol. 521, 170–181. Zhang, R., Delworth, T.L., 2006. Impact of Atlantic multidecadal oscillations on
Wang, R., Ma, H.G., Liu, G.Q., Zuo, D.G., 2015. Selection of window length for singular India/Sahel rainfall and Atlantic hurricanes. Geophys. Res. Lett. 33 (17).
spectrum analysis. J. Franklin Inst. 352 (4), 1541–1560. Zolina, O., Simmer, C., Belyaev, K., Kapala, A., Gulev, S., 2009. Improving estimates of
Weare, B.C., Nasstrom, J.S., 1982. Examples of extended empirical orthogonal heavy and extreme precipitation using daily records from European rain gauges.
function analyses. Mon. Weather Rev. 110 (6), 481–485. J. Hydrometeorol. 10 (3), 701–716.
Wilcox, A.C., Escauriaza, C., Agredano, R., Mignot, E., Zuazo, V., Otárola, S., Mao, L., Zotov, L.V., Shum, C.K., Frolova, N.L., 2015. Gravity Changes over Russian River
2016. An integrated analysis of the March 2015 Atacama floods. Geophys. Res. Basins from GRACE. In: Planetary Exploration and Science: Recent Results and
Lett. 43 (15), 8035–8043. Advances. Springer, Berlin Heidelberg, pp. 45–59.
Wilks, D.S., 2011. Statistical Methods in the Atmospheric Sciences. Elsevier
Academic Press, California, USA.

Вам также может понравиться