You are on page 1of 16

15 MARCH 2015 JIEN ET AL.

2459

The Influence of El Niño–Southern Oscillation on Tropical Cyclone Activity in


the Eastern North Pacific Basin

JERRY Y. JIEN, WILLIAM A. GOUGH, AND KEN BUTLER


Department of Physical and Environmental Sciences, University of Toronto at Scarborough, Scarborough,
Ontario, Canada

(Manuscript received 28 March 2014, in final form 14 November 2014)

ABSTRACT

The interannual variability of tropical cyclone (TC) activity due to El Niño–Southern Oscillation (ENSO)
in the main development region of the eastern North Pacific basin has received scant attention. Herein the
authors classify years of El Niño, La Niña, and neutral conditions using the multivariate ENSO index (MEI).
Storm measurements of the net tropical cyclone activity index and power dissipation index are used to
summarize the overall seasonal TC activity and TC intensity between 1971 and 2012. Both measures are found
to be statistically dependent on the ENSO phases in the basin’s main development region. However, when the
area is longitudinally divided, only the western portion of the development region experienced a significant
difference ( p , 0.05). Specifically, El Niño years are characterized by more frequent, more intense events
compared to La Niña conditions for this subregion. Correlation analyses on the relationships between the MEI
and both TC indices demonstrate correlations between ENSO and TC activity and intensity that are statis-
tically significant (p , 0.05) only in the western region. These relationships have the potential to improve the
short-term forecast of the local TC activity and intensity on a seasonal basis for public awareness and disaster
preparation.

1. Introduction Ocean shifts westward when transitioning to the La Niña


phase.
The El Niño–Southern Oscillation (ENSO) phe-
Given its feedback on the global atmospheric cir-
nomenon is arguably the most dominant control of
culation and its effect on worldwide climate anomalies
the interannual climate variability on a global scale
(Pielke and Landsea 1999; Landsea 2000), ENSO is
(Rasmusson and Carpenter 1982). Atmospheric vari-
also critical to the short-term prediction for a range of
ations in the ENSO cycle can be monitored by the
extreme weather events, including seasonal forecasts
mean sea level pressure difference between Tahiti and
of tropical cyclones (TCs) in multiple ocean basins
Darwin, Australia, known as the Southern Oscillation
(Camargo et al. 2007). In additional to its role as an
index (Ropelewski and Jones 1987). Oceanic anoma-
important contributor to seasonal TC activity (fre-
lies at the Niño-3 (58N–58S, 1608E–1508W) and Niño-
quency, intensity, and duration), the impact of ENSO
3.4 (58N–58S, 1708–1208W) regions of the equatorial
may vary depending on the strength of its signal and
Pacific Ocean during periods of warmer (colder) sea
the location of the TC development region (Pielke and
surface temperature are also applied to quantify the
Landsea 1999). In general, during the La Niña phase,
extent of El Niño (La Niña) events (Trenberth 1997;
an above-normal TC activity is observed in the south-
Wolter and Timlin 1998). Overall, the El Niño–induced
western Pacific (Nicholls 1979), the northwestern Pacific
atmospheric instability in the eastern equatorial Pacific
(Chan 1985), and North Atlantic (Gray 1984a) TC de-
velopment basins, while a greater number of storms
form during the El Niño phase in the central Pacific
basin, west of 1408W–1808 (Chu and Wang 1997).
Corresponding author address: Jerry Y. Jien, Department of The interannual fluctuation of the ENSO cycle acts as
Physical and Environmental Sciences, University of Toronto at
Scarborough, 1265 Military Trail, Scarborough ON M1C 1A4, a crucial control to regional storm intensity (Gray 1984b;
Canada. Bove et al. 1998). For instance, La Niña (El Niño) years
E-mail: jerry.jien@mail.utoronto.ca are linked to a greater (lower) number of intense TC

DOI: 10.1175/JCLI-D-14-00248.1

Ó 2015 American Meteorological Society


2460 JOURNAL OF CLIMATE VOLUME 28

landfall in the United States during seasons when more region of TC genesis, while more storms are distributed
(less) Atlantic storms develop (Klotzbach 2011). Since in the eastern region during the La Niña phase. Though
TC potential destruction power can be empirically re- Irwin and Davis (1999) conclude the seasonal total
lated and annually aggregated to its intensity (Emanuel storm count does not deviate between the ENSO
2005), the ENSO cycle plays an important role in de- phases, Collins (2007) suggests TC frequency of certain
termining the potential TC-afflicted societal damages storm category is sensitive to the alternation of El Niño
(Chu and Wang 1997; Pielke and Landsea 1999). Be- and La Niña events only in the western division of
cause more landfalling storms mature into hurricanes the ENP main development region (MDR; 108–208N,
on the eastern United States, the annual average eco- 858–1408W). Thus, this unequal east–west distribution
nomic damage (Hebert et al. 1997) caused by Atlantic of storm genesis contributes to such a difference of
TCs is exacerbated during La Niña years (Pielke and spatial sensitivity to environmental conditions (Collins
Landsea 1999). However, for the world’s most active and Mason 2000; Ralph and Gough 2009). In particular,
tropical storm genesis region when considered on a per thermodynamic influences at the western subdivision
unit area and time bases at the eastern North Pacific of MDR are argued to enforce such an ENSO-induced
(ENP) basin (Molinari et al. 2000), the relationship of difference of TC frequency (Collins 2007, 2010;
TC intensity with ENSO has not been explored. Typi- Klotzbach and Blake 2013).
cally, ENP storms dissipate over the open sea. However, To date, there remains a gap in the literature that
when they landfall, they not only affect continental statistically evaluates the relationship of TC climatology
North America but also possess the ability to inflict their in the ENP basin with ENSO. Instead of relying on
strengths on the Hawaiian Islands in the central Pacific seasonal storm count alone, an alternative option is to
basin. For instance, the formation of Iniki (1992) in the characterize metrics of seasonal TC activity and TC in-
ENP basin later reached the category-4 hurricane tensity through combining individual measures of storm
strength and was accounted for the financial loss of USD frequency, intensity, and duration. Though Collins and
2.5 billion in Hawaii, during the 1991/92 El Niño year Roache (2011) integrated these measurements into
(Chu and Wang 1997). metrics that better define and quantify seasonal storm
In the Pacific Ocean, where local signatures of El activity and intensity, a long-term record of these met-
Niño/La Niña conditions are evident, (Nicholls 1992; rics is needed to establish their relationships with ENSO
Trenberth 1997; Wolter and Timlin 1998), strong events. Following the works of Ralph and Gough (2009)
northwestern Pacific TCs, or typhoons, are suppressed and Collins and Mason (2000), our study region will
immediately after El Niño years (Chan 1985; Wang encompass MDR subdivisions to statistically illustrate
and Chan 2002; Chu 2004). Moreover, there is a gen- the spatial sensitivity of regional storm activity as well
eral consensus that the distribution of TC genesis area as the interannual variation due to ENSO influences.
in the northwestern Pacific basin has experienced Because globally stronger TCs tend to show a stronger
a southeastward drift during the peak typhoon season relationship with ENSO (Frank and Young 2007), sea-
of El Niño years relative to La Niña years (Chan 1985; sonal TC intensity will also be examined as a function of
Wang and Chan 2002). Thus, El Niño storms tend to ENSO conditions.
travel a greater distance (both westward and north-
ward) prior to experiencing unfavorable conditions to
dissipate. Continuing eastward in the North Pacific, we 2. Methods
anticipate that the interannual variation of the ENSO
a. Classification of ENSO events
cycle would also influence TC activity (frequency, in-
tensity, and duration) in the ENP basin. Specifically, we The multivariate ENSO index (MEI) was used to
hypothesize that ENP storm seasons during the El Niño identify ENSO cycles (Wolter and Timlin 1998). Com-
phase will be reflected by a heightened TC activity with pared to other ENSO indices, MEI is a more holistic
more intense TCs, while the La Niña phase will be more approach that reflects the coupled nature of the ocean–
subdued. atmosphere system through its incorporation of six
In the ENP basin where the influence of ENSO on predictors of sea level pressure, zonal and meridional
seasonal TC formation and development has remained surface winds, sea surface temperature, surface air
largely undetermined (Lupo et al. 2008), there is, how- temperature, and total cloud fraction. To overlap with
ever, strong evidence of ENSO modification on the the ENP hurricane season, El Niño (La Niña) events are
longitudinal shift in locations of TC origin and dissipa- defined by the 10 highest (lowest) years of sliding
tion (Irwin and Davis 1999). Specifically, during El Niño bimonthly MEI values, averaged from April–May to
events, more TCs tend to form west of the primary November–December, with the remaining 22 years
15 MARCH 2015 JIEN ET AL. 2461

FIG. 1. The main development region is longitudinally divided into EDR and WDR for eastern
North Pacific tropical cyclones. Dots represent areas of major urban centers.

defined as neutral years. Since this rank-based approach 1974). An overall measurement of the net tropical cy-
is less influenced by environmental conditions of a par- clone (NTC; Gray et al. 1994) activity index is defined as
ticular season, leading to an asymmetrical number of
El Niño and La Niña events, it is widely implemented in NTC 5 (%NS 1 %H 1 %MH 1 %NSD
not only MEI (Klotzbach 2012), but also anomalies of 1 %HD 1 %MHD)/6, (1)
other sea surface temperature–based ENSO indices
(Camargo and Sobel 2005). where each season’s percentage value is weighted
against the entire period mean (1981–2010) and is used
b. TC activity and TC intensity
for the six measures of seasonal activity: named storms
Historical TC track data, including the storm name (NS), named storm days (NSD), hurricanes (H), hurri-
and its 6-hourly geographic position by longitude and cane days (HD), intense hurricanes (IH), and intense
latitude, speed, and surface pressure, are obtained from hurricane days (IHD). In addition, TC intensity is
the northeastern and north-central Pacific hurricane measured through the power dissipation index (PDI;
database (HURDAT2) from the National Hurricane Emanuel 2005), defined as
Center. As in other works (e.g., Ralph and Gough ðt
2009), TC data after 1971, when the routine use of PDI 5 Vmax 3
dt , (2)
satellite to monitor storm development, were included. 0
Since the ENP storm formation is spatially unequally
distributed, data of TC activity is longitudinally divided where the maximum surface wind speed (Vmax) is sum-
at 1128W as the eastern (108–208N, 858–1128W) and med over each storm’s lifetime (t) every 6 h for all
western (108–208N, 1128–1408W) development regions storms of at least tropical storm strength. To make PDI
(EDR and WDR, respectively; Ralph and Gough more manageable, seasonal values are accumulated as
2009). During occurrences when EDR storms migrate annual aggregates and displayed at a scale of 107 m3 s23.
and establish their peak intensity in WDR, the tabu-
c. Statistical analyses
lation of storm count and duration is allocated ac-
cordingly, based on the 6-hourly storm-track positions Seasonal NTC and PDI values for the MDR are
(Collins and Roache 2011). stratified into the EDR and WDR (Fig. 1). To account
The seasonal distribution of storm frequency, in- for seasons with no WDR storms, time series for NTC
tensity, and duration during the 42-yr period was taken (PDI) are normalized through the square root (natural
into account when deriving the two empirically based log) transformation within MDR and its subdivisions.
TC metrics. Based on its surface wind speed, each TC is The Mann–Kendall test (Sprent 1989) is utilized to de-
categorized as a tropical/named storm (18 m s21), hur- termine if trends of the time series for both storm indices
ricane (33 m s21), and intense (or major) hurricane are statistically significant ( p , 0.05). The magnitude of
(50 m s21), which corresponds to categories 3, 4, or 5 on the trend is calculated through the Theil–Sen slope
a Saffir–Simpson hurricane intensity scale (Simpson (Helsel and Hirsch 1991).
2462 JOURNAL OF CLIMATE VOLUME 28

Seasonal NTC and PDI values are statistically com- TABLE 1. Years from 1971–2012 are classified by ENSO phases
pared to reveal any spatial and temporal effects. An based on the averages of the eight sliding bimonthly MEI values
from April–May to November–December. The MEI values are
analysis of variance (ANOVA) is applied to assess
provided in parentheses.
whether there is any statistically significant difference
among different groups of MDR storm measurements El Niño Neutral La Niña
(NTC and PDI). Seasonal storm measures are compared 1972 (11.50) 1976 (10.52) 1971 (21.33)
and grouped according to two conditions: 1) temporal 1982 (11.69) 1977 (10.76) 1973 (21.28)
classification of annual ENSO event (El Niño, La Niña, 1983 (10.96) 1978 (20.17) 1974 (20.86)
1987 (11.72) 1979 (10.61) 1975 (21.58)
or neutral years) and 2) spatial division of MDR into
1991 (11.02) 1980 (10.46) 1988 (21.11)
EDR and WDR. With each storm measurement strati- 1992 (10.97) 1981 (1.00) 1999 (20.83)
fied into groups of ENSO conditions and MDR sub- 1993 (11.12) 1984 (20.17) 2007 (20.70)
divisions, a two-way ANOVA is applied to test for both 1994 (10.95) 1985 (20.30) 2008 (20.40)
individual effects, as well as any interaction from the two 1997 (12.42) 1986 (10.76) 2010 (21.26)
2002 (10.88) 1989 (20.28) 2011 (20.60)
factors. An interactive effect from the two factors is ex-
1990 (10.31)
amined if relative differences due to ENSO influences 1995 (20.14)
are consistent at both EDR and WDR. Since there are 1996 (20.27)
three pairings (El Niño–La Niña, El Niño–neutral, and 1998 (20.07)
La Niña–neutral) of ENSO phases, a post hoc test using 2000 (20.29)
2001 (10.02)
Tukey’s method of multiple comparisons is applied to
2003 (10.27)
determine which ENSO pairing(s) is accounted for the 2004 (10.52)
different group means ( p , 0.05). 2005 (10.14)
A correlation analysis is conducted to measure the 2006 (10.73)
degree of relationship for both measures of TC activity 2009 (10.88)
2012 (10.48)
and TC intensity with the ENSO index of MEI. A
quantile–quantile (Q–Q) plot and the Shapiro–Wilk test
are applied to assess the normality of both storm indices
(Li) can be represented as Ii 5 PDIi/Li, where i indices
prior to the decision of selecting a parametric or non-
each individual storm. Hence, PDI1, PDI2, and PDI3 can
parametric testing for the correlation analysis. A Q–Q
be calculated when the storm intensity, duration, and
plot demonstrates the fitted distribution of seasonal
count are varied, respectively, according to each hurri-
storm measurements and displays the spread of data
cane season, while the other two terms are averaged
deviations from the respective line of normal distribu-
across all seasons of the entire 42-yr period (Camargo
tion. Data normality was improved through the best
and Sobel 2005). Correlations calculated with MEI and
parameter (lambda) estimated from the Box–Cox
PDI1, PDI2, and PDI3 can quantitatively determine and
transformation and quantitatively proven through the
separate the contributions of storm intensity, duration,
Shapiro–Wilk test. A least squares regression is used to
and frequency to PDI.
correlate MEI values with the two TC measures. A
Pearson coefficient of correlation r provides a degree of
association between NTC and PDI with MEI. The co- 3. Results
efficient of determination r 2 is applied to explain the
a. Classification of ENSO events
amount of variance that is captured by the linear re-
gression. The correlation analysis is extended when the Table 1 highlights the 10 highest (El Niño) and lowest
entire region is subsequently divided into EDR and (La Niña) MEI values of bimonthly averages from
WDR. Residuals of observed and predicted values from April–May to November–December, with an additional
linear regression are assessed for data independence. A 22 years classified as neutral years during the 1971–2012
Ljung–Box test is applied to examine the null hypothesis period. Our annual partitioning of ENSO phases is in
of residual independence in a correlogram (autocorre- general agreement with results of Lupo et al. (2008) and
lation plot). Irwin and Davis (1999). In particular, we show the
To distinguish the relative contribution of storm fre- strongest El Niño year (indicated by the highest MEI
quency, intensity, and duration to PDI, we derive three average) to occur in 1997, when climatological impacts
additional PDI-derived indices. Essentially, the PDI and extreme hydrometeorological events devastated the
time series is modified so that each of its surrogates only southwestern United States (Changnon 1999). Never-
varies with storm frequency, intensity, or duration only. theless, the choice of ENSO classification schemes and
The average intensity (Ii) of a storm during its lifetime the length of analysis period could have led to some
15 MARCH 2015 JIEN ET AL. 2463

FIG. 2. Time series for the seasonal (a) frequency (TS, H, and MH) and (b) duration (TSD, HD,
and MHD) in EDR from 1971 to 2012.

disagreement in the identification of ENSO events. For of storm count and lifetime (in days) are binned into
instance, using the sea surface temperature anomaly to storm strengths of tropical/named storm, hurricane, and
classify ENSO events, Lupo et al. (2008) deems 1992–94 major hurricane in EDR (Fig. 2) and WDR (Fig. 3).
as neutral years. However, our analysis shows 1992 as Both EDR and WDR experienced heightened storm
a strong El Niño year, associated with the greatest NTC activity in terms of the number of tropical storms de-
and PDI values reflected by the highest number (24) of veloped from the early 1980s to 1992 and 1993, re-
storm genesis with the longest duration spent at each spectively, when unusually high proportions of tropical
(tropical storm, hurricane, and major hurricane) stage. storms strengthened into major hurricanes. However,
In addition, the extension of our data from that of Lupo while it is followed by 1994 and 1999 when no EDR
et al. (2008) leads to the identification of three La Niña hurricanes grew into major hurricanes, no WDR storms
years (2008, 2010, and 2011). of even NS strength were found in 1996.
Overall, the entire region experiences an annual
b. Time series of storm activity and intensity
TC count of 15 storms. When MDR is subdivided,
In Fig. 1, the time series for TC measures of frequency, while similar seasonal frequencies of NS, H, and MH
intensity, and duration that define metrics of seasonal (Table 2) are shown between EDR and WDR, the
storm activity (NTC) and intensity (PDI) are summa- origin of a few WDR storms was pertained within
rized. For comparison purposes, the summary statistics WDR boundary. Instead, an average of four WDR
2464 JOURNAL OF CLIMATE VOLUME 28

FIG. 3. As in Fig. 2, but for WDR.

storms were initiated in EDR and migrated into storms of long duration, respectively (Figs. 4a and 5a).
WDR, while reaching their maximum intensities. Such a recent decrease in the ENP storm development is
Since these EDR-originated storms establish their consistent with findings from Schultz (2007), Lupo et al.
peak intensity in WDR, more than half of them (2008), and Wu et al. (2008), but opposite from that
strengthened into hurricanes or major hurricanes, experienced in adjacent regions of the Atlantic
thereby greatly extending the length of storm duration (Webster et al. 2005; Collins 2010) basin, marked by
during all three storm stages. consecutive (2004 and 2005) peaks of Atlantic hurricane
The time series of the seasonal overall TC activity seasons. Overall, this continuous increase of TC activity
(NTC; Fig. 4) and TC intensity (PDI; Fig. 5) are dis- (Steenhof and Gough 2008) and TC intensity (Wu et al.
played from 1971 to 2012. Although the Mann–Kendall 2008) in the Atlantic basin over a long time record is
test confirms the Theil–Sen slope values of MDR storm widely known to be enhanced by local environmental
measurements to be statistically insignificant, negative conditions of thermodynamic and dynamic factors
trends of NTC and PDI imply that there are decreasing (Emanuel 2005; Goldenberg and Shapiro 1996;
number of storms with long duration and fewer intense Goldenberg et al. 2001).
15 MARCH 2015 JIEN ET AL. 2465

TABLE 2. The TC activity in the ENP basin in the MDR, EDR,


and WDR from 1971 to 2012, while seasonal averages from 1981 to
2010 are incorporated into the derivation of NTC.

42-yr (1971–2012) Avg 30-yr (1981–2010) Avg


NTC measures MDR EDR WDR MDR EDR WDR
NS 15 8.0 7 15.4 8.0 7.5
NSD 60.4 25.2 35.3 62.1 25.6 36.5
H 8.5 4.3 4.2 8.4 4.1 4.3
HD 24.8 9.3 15.5 25.5 9.5 16.0
MH 3.9 2.0 2 3.9 1.9 2.1
MHD 7.0 2.4 4.5 7.5 2.5 5.0

When MDR is subdivided at 1128W (Ralph and


Gough 2009), while negative trends of both NTC
(Fig. 4b) and PDI (Fig. 5b) values are still maintained in
WDR, EDR storm measures have decreased at a slower
pace for NTC, with PDI experiencing virtually no
change through the 42-yr period. Though trends of the
regional contrast are determined as statistically in-
significant, there is growing evidence that WDR storm
activity is regulated by environmental controls that are
different from that in EDR (Collins and Mason 2000; FIG. 4. Time series charts for NTC at (a) MDR with horizontal
dashed lines showing values at the 25th and 75th percentiles and
Ralph and Gough 2009). when TC development region is subdivided into (b) EDR and
One such large-scale ocean–atmosphere climate WDR during 1971–2012. Dashed vertical lines are colored to dis-
phenomenon that can affect the interannual variability tinguish El Niño (red), La Niña (blue), and neutral (gray) years, to
of both NTC (Fig. 4) and PDI (Fig. 5) is the influence of be consistent with (a).
ENSO events. In general, El Niño years are identified
with greater values of storm indices derived from sea-
sonal TC activity. On the other hand, La Niña years are temporal classification of annual ENSO conditions and
associated with some of the least active TC seasons. spatial division of MDR (Tables 4 and 5). The F test
From Table 3, when seasonal mean values of six NTC confirms that temporal differences of NTC and PDI can
measures are binned to the appropriate ENSO events, be attributed to ENSO variation though we do not know
both the frequency and duration measures are consis- under which ENSO conditions are both indices differ-
tently highest during El Niño years, follow by neutral ent. Since the F test alone cannot detect exactly which
and La Niña years. However, as there is one year (1996) ENSO pairing(s) is attributed to a significant difference
without any WDR storm formation, direct spatial com- between both storm measures, Tukey’s method of
parison between the six NTC measures under ENSO multiple comparisons identifies NTC and PDI to be only
influences would violate fundamental assumptions (e.g., statistically different for pairings of El Niño–La Niña
data normality) of parametric statistics. Presumably, an and El Niño–neutral comparisons at p , 0.05 (Table 4).
ENSO-induced difference is greater in WDR where In previous work, though more El Niño storms became
drastic meteorological shifts resonate more with ENSO intense hurricanes, the analysis of monthly TC statistics
phase changes (Collins 2007). Although such an assess- does not yield a significantly greater TC activity during
ment of ENSO influences on ENP storms is only quali- El Niño years (Lupo et al. 2008). In contrast, our results
tative, statistical approaches attributing ENSO events to demonstrate that both storm measurements are signifi-
the regional TC activity and TC intensity using metrics cantly different under both ENSO pairings, which are
that include storm frequency, maximum sustained wind also acquainted to influence the mean genesis location
speed, and duration will follow in the next subsection. of ENP storms (Irwin and Davis 1999).
The second main effect of regional difference dem-
c. Statistical analyses
onstrates that only seasonal PDI values are statistically
different ( p , 0.05) between EDR and WDR. A follow-
1) TWO-WAY ANALYSIS OF VARIANCE
up regional comparison of TC measures using Tukey’s
The mean differences of the NTC and PDI are com- method is shown only as a confirmatory analysis since
pared using a two-way ANOVA under two factors: there are only two levels (EDR and WDR) of MDR
2466 JOURNAL OF CLIMATE VOLUME 28

TABLE 3. Mean seasonal (1971–2012) values of six NTC measures


subdivided into EDR and WDR during three ENSO conditions.

Region NTC measures El Niño La Niña Neutral


EDR NS 8.1 7.9 8.1
NSD 27.5 25.4 24.0
H 4.4 4.2 4.3
HD 10.8 8.9 8.8
MH 2.8 1.6 1.8
MHD 3.2 2.0 2.3

WDR NS 8.7 5.1 7.1


NSD 46.7 27.6 33.6
H 5.3 3.2 4.2
HD 22.7 11.4 14.0
MH 2.8 1.5 1.9
MHD 7.5 2.7 4.0

collectively bounded by all six NTC parameters’ 30-yr


averages (Table 2).
Since there is a combination of effects considered for
statistical comparison, an interaction effect of the re-
gional and ENSO-induced temporal (ENSO region;
Tables 4 and 5) factors on both NTC and PDI values is
assessed. An interactive outcome occurs when rankings
of relative NTC and PDI values based on ENSO con-
FIG. 5. As in Fig. 4, but for PDI (scale of 107 m3 s23). ditions shift in ranks between regions. Not only do our
results show that the average measures of both storm
indices are greatest during El Niño years, followed by
subdivisions. Regional variations of TC intensity (PDI)
neutral and La Niña years (Figs. 6a and 6b), but also,
due to ENSO events further confirm that MDR should
both MDR subdivisions maintained such relative
be viewed as two distinct development regions (Collins
ranks of storm metrics due to ENSO modification.
and Mason 2000; Ralph and Gough 2009). While the
This consistency of the result in the F test ( p . 0.05)
average WDR storm intensity (PDI) is noticeably
demonstrates the absence of an interaction effect of
higher during El Niño years, the mean difference of the
a combined ENSO region factor across NTC (Table 4)
overall TC activity (NTC) is not detected (Table 4),
and PDI (Table 5). However, Tukey’s multiple com-
possibly because of the derivation of seasonal NTC
parisons of the interaction effect demonstrate signifi-
values. The derivation of the NTC index is sensitive to
cant differences between NTC and PDI resulting from
the seasonal variability of relative measures of all TC
different ENSO comparisons that are not equally
components of the index to the respective 1981–2010
distributed within MDR subdivisions. In particular,
climatological means (Table 2). Because the seasonal
the differences of NTC and PDI values within the
(1981–2010) averages of storm duration measures are
ENSO pairing of El Niño–La Niña are maintained
distinctly higher in WDR, its seasonal NTC values
only in WDR.
during any given year can be comparably similar, sta-
tistically indifferent, to EDR values. For instance, 2009
2) CORRELATION ANALYSIS
saw the WDR storm duration (95 days) close to be 3
times more than that of EDR (36 days), with a compa- Relatively greater NTC and PDI values during El
rable difference for the total storm count. However, Niño years, followed by neutral and La Niña years sug-
owing to the fact that many EDR-originated storms gest that MEI values are positively linked to both in-
spend the majority of lifetimes in WDR, because long- dices. However, the use of the parametric correlation
term averages of all three NTC duration measures in analysis is deemed appropriate only when both storm
WDR are substantially greater, the regional difference metrics have been normalized. When TC data are bin-
of NTC values in 2009 is virtually the same. As such, ned into EDR and WDR, the data distribution of each
even if there is a substantial spatial difference of any storm measure varies. Visual inspections of Q–Q plots
individual NTC component in any given year, NTC is show that an absence of WDR storm activity may have
15 MARCH 2015 JIEN ET AL. 2467

TABLE 4. A two-way analysis of variance is tested for effects of ENSO and regional division on the NTC activity index (%).

Degrees of freedom Sum squared Mean squared F value Probability (.F )


ENSO 2 33 418 16 709.0 9.7441 0.000
Region 2 647 323.7 0.1888 0.828
ENSO region 4 3751 937.7 0.5469 0.702
Residuals 117 200 629 1714.8
Tukey multiple comparisons of means:
Diff p values
ENSO
El Niño–La Niña 44.889 0.000
El Niño–neutral 32.524 0.002
La Niña–neutral 212.365 0.367
Region
EDR–MDR 3.347 0.927
EDR–WDR 5.510 0.815
MDR–WDR 2.163 0.969
ENSO region
La Niña EDR–El Niño EDR 224.668 0.920
Neutral EDR–El Niño EDR 221.081 0.919
Neutral EDR–La Niña EDR 3.586 1.000
La Niña MDR–El Niño MDR 246.899 0.228
Neutral MDR–El Niño MDR 234.297 0.431
Neutral MDR–La Niña MDR 12.602 0.997
La Niña WDR–El Niño WDR 263.100 0.024
Neutral WDR–El Niño WDR 242.193 0.170
Neutral WDR–La Niña WDR 20.907 0.922

caused nonnormal distributions of both NTC (Fig. 7a) (Fig. 7c) and PDI (Fig. 7d) are closer to the line of
and PDI (Fig. 7b) indices. In particular, WDR storm normality. Under the null hypothesis that the sampled
development during 1996 went completely dormant data come from normal distributions, a subsequent
with no storms generated west of 1128W. After both Shapiro–Wilk normality test was applied to test against
datasets were transformed, distributions of both NTC the assumption of data normality. The results of the

TABLE 5. A two-way analysis of variance is tested for effects of ENSO and regional division on the PDI (107 m3 s23).

Degrees of freedom Sum squared Mean squared F value Probability (.F )


ENSO 2 1.846 0.923 7.962 0.001
Region 1 1.277 1.277 11.015 0.001
ENSO region 2 0.641 0.320 2.762 0.069
Residuals 78 9.045 0.116
Tukey multiple comparisons of means:
Diff p values
ENSO
El Niño–La Niña 0.403 0.001
El Niño–neutral 0.304 0.004
La Niña–neutral 20.099 0.528
Region
EDR–WDR 6.871 0.001
ENSO region
La Niña EDR–El Niño EDR 20.165 0.887
Neutral EDR–El Niño EDR 20.127 0.925
Neutral EDR–La Niña EDR 0.038 1.000
La Niña WDR–El Niño WDR 20.642 0.001
Neutral WDR–El Niño WDR 20.482 0.005
Neutral WDR–La Niña WDR 0.161 0.818
2468 JOURNAL OF CLIMATE VOLUME 28

hurricane season. A relatively higher r 2 in WDR in-


dicates a greater NTC and PDI variability that is
explained by MEI variability. Residuals (observed 2
predicted) for the WDR storm metrics are plotted in
Fig. 10; while, subsequent Ljung–Box test statistics (chi-
squared values) show that there is no statistical evidence
(p . 0.05) for nonzero autocorrelations.

4. Discussion
The influence of ENSO on different parameters of
MDR storm activity at the ENP is statistically addressed
in this study. Previously, a lack of such statistical evi-
dence in the basin could be attributed to inconsistent
classification schemes in distinguishing ENSO phases of
El Niño, La Niña, and neutral conditions, in addition to
the absence of reliable TC data on a longer time scale
(Collins 2007; Schultz 2007; Lupo et al. 2008). Though
maximum NTC and PDI values in 1992 could be related
to a greater midlevel moist static energy during the peak
(July–September) hurricane season (Wu and Chu 2007),
the recent 2007, 2008, and 2010 La Niña years are
FIG. 6. (a) The NTC activity index and (b) PDI (scale of
107 m3 s23) under 10 El Niño, La Niña, and neutral years are sum- marked as the three most meager TC years of the entire
marized as boxplots for MDR, WDR, and EDR. Green diamonds data analysis (1971–2012). NTC and PDI values for
represent NTC and PDI averages at each ENSO condition. Results these three years were well below seasonal averages of
of statistical evaluations of NTC and PDI differences are as in 97.56% and 1.20 3 107 m3 s23, respectively, over 10 La
Tables 4 and 5.
Niña years. In terms of the same storm measures of TC
activity and intensity, when MDR is subdivided, 2011 is
Shapiro–Wilk test calculate a W statistic that is used to found to exhibit the second least-active TC season of all
calculate the correlation of the sampled data after 10 La Niña years in WDR.
they have been ordered and standardized and what Alternative considerations of MDR divisions from
the samples would have been if they were drawn from Irwin and Davis (1999) and Collins and Mason (2000)
a normal distribution and ordered. Significance of the also have attributed to the detection of the difference of
Shapiro–Wilk test subjectively determines if both regional TC sensitivity to ENSO oscillation. The MDR
WDR storm metrics were normalized after data division at 1128W depicts a substantially greater TC in-
transformations. tensity in WDR (Ralph and Gough 2009), where many
A parametric testing of correlation analysis is used to EDR storms migrate and achieve their peak intensities
highlight the extent of NTC (Fig. 8) and PDI (Fig. 9) prior to dissipation. While there are more storms origi-
relationships with MEI. Time series of residuals nated from EDR, its overall TC activity and TC in-
(observed 2 correlated) their correlograms for the tensity are less influenced by environmental changes;
WDR storm metrics are plotted from 1971 to 2012 rather, it is WDR storms that are more responsive to
(Fig. 10). To test for the presence of serial independence ENSO-induced environmental influences (Collins and
of the residuals, subsequent Ljung–Box test statistics Mason 2000). Such a regional difference of storm sen-
(chi-squared values) show that there is no statistical sitivity to changes of the external environment is also
evidence ( p . 0.05) for nonzero autocorrelations. Pos- consistent with that of the original longitudinal division
itive associations for both MDR storm measurements of MDR storm frequency of at least tropical storm
with MEI are determined statistically significant for strength into eastern and western geographical bound-
PDI, not NTC (Table 6). However, when MDR is sub- aries (Collins and Mason 2000).
divided, such a positive linear relationship is only sig- Another important aspect is revealed when focusing
nificant in WDR ( p , 0.05). The presence of such on the correlation strengths of MEI with NTC and PDI
statistical significance confirms that the ENSO influence in MDR and within its subdivision. Overall, PDI is more
on TC activity is most pronounced in WDR, even sensitive to ENSO influences than NTC (Table 6).
though more EDR storms are generated during the local Among factors of storm intensity (PDI1), duration
15 MARCH 2015 JIEN ET AL. 2469

FIG. 7. The (Q–Q) plots for (a) NTC and (b) PDI (scale of 107 m3 s23) values in WDR have lines of normality
passing through the first and third quartiles. (c),(d) Data after transformation are presented. The Shapiro–Wilk test
statistic (W), with its p value, is provided before and after data transformations.

(PDI2), and count (PDI3) attributed to PDI variability may trigger the genesis of more storms and vice versa.
(Fig. 11), only storm intensity (PDI1) demonstrates In particular, fluctuating sea surface temperature could
a significant ( p , 0.05) effect on the correlation of PDI reinforce changes to the availability of atmospheric
with ENSO signal (Table 7). Such a result is contrary to moisture and dictate the seasonal variability of WDR
northwestern Pacific storms where Camargo and Sobel storm activity (Collins and Roache 2011). While the
(2005) found that only the duration variability attributes seasonal sea surface temperature in EDR is already
significantly to the ENSO signal on PDI. However, well above the critical threshold for storm formation
when MDR is subdivided, seasonal variations of all (Ralph and Gough 2009), fluctuation of the sea surface
three storm factors are significantly sensitive to the temperature–dependent MEI may be less crucial to the
ENSO signal on WDR storm intensity, while their in- region’s overall TC activity and storm intensification.
fluences are little to minimal in EDR. As the derivation These thermodynamic controls offer explanations to
of PDI is directly related to the total storm count, such why dramatic differences of the two indices appear
associations between (PDI1 versus MEI) show compa- only when MDR is subdivided.
rable strength with the relationship of PDI and MEI The fact that the ENP storm activity and intensity in
(Table 6). Overall, the subdued r 2 of ENSO with PDI, WDR are more sensitive to changes of ENSO condi-
and also NTC, suggests other local environmental forc- tions is geographically consistent with neighboring
ings are involved at explaining the annual variation of storm development regions. In the Southern Hemi-
seasonal TC measures. sphere, the strength of the correlation between south-
Changes to sea surface temperature due to ENSO western Pacific storm frequency and the Southern
modification can be linked to the atmospheric alter- Oscillation index gradually shifts in direction westward
ation of convective mechanisms for fueling storm de- from the 1708E longitude of boundary division (Basher
velopment (Collins 2007; Klotzbach and Blake 2013). and Zheng 1995). This means that in contrast to the ENP
One of the crucial indicators in determining ENSO basin, more South Pacific storms develop during the
signal (i.e., sea surface temperature) has been well La Niña phase. Similarly, at the adjacent northwestern
known for its local effect on the ENP storms (Ralph Pacific basin, there are more La Niña storms that reached
and Gough 2009). Since WDR ocean temperatures are the strength of a typhoon, equivalent to a hurricane, at its
closer to the 26.58C threshold for TC formation (Gray easternmost boundary near the International Date Line
1968), a warm anomaly arising from ENSO changes (Chan 1985; Wang and Chan 2002). Such a close proximity
2470 JOURNAL OF CLIMATE VOLUME 28

FIG. 8. Linear regressions of MEI values with data transformed FIG. 9. As in Fig. 8, but for PDI (scale of 107 m3 s23).
for NTC at (a) MDR, (b) EDR, and (c) WDR under three ENSO
conditions during 1971–2012. The Pearson coefficient of correla-
tion (r) and correlation of variation (r 2) are provided in Table 6. wind shear is observed over much of the MDR in 1992
during such a strong El Niño event, it could also play an
to the neighboring WDR could result into a greater important role at reinforcing WDR convection through
storm development, which is directly observed through the reduction of wind shear by enhancing the upper-level
the eastward extension of the monsoon trough crossing westerly wind and the lower-level easterly wind (Jones
the central Pacific boundary during extreme El Niño and Thorncroft 1998). When wind shear and other dy-
years (Clark and Chu 2002). namical factors are included in the seasonal genesis
Apart from its direct influences on moist static sta- parameter (Gray 1977), a greater number of El Niño
bility (Malkus and Riehl 1960), ENSO is also indirectly storms is observed to coexist with higher dynamical po-
acquainted with changes in the wind direction of the tential values than during La Niña conditions (Clark and
tropospheric profile. As demonstrated in the North Chu 2002).
Atlantic basin, ENSO is known to control wind shear in
other TC basins where such a dynamical factor is pri-
5. Conclusions
marily responsible for TC intensity (Gray 1984a; Jones
and Thorncroft 1998; Landsea 2000; Goldenberg et al. The longitudinal division of MDR has greatly facili-
2001) and is noted to be highly correlated with the pri- tated our understanding of environmental influences
mary ENP storm genesis region (Table 6). Because low at the ENP basin. We have statistically evaluated and
15 MARCH 2015 JIEN ET AL. 2471

FIG. 10. (left) Time series of residuals (observed 2 predicted) NTC values and (right) cor-
relograms (autocorrelation plots) are evaluated with the Ljung–Box test statistic (chi-squared
value), with its p value, based on the first 20 lags in WDR for (a) NTC and (b) PDI (scale of
107 m3 s23) during 1971–2012.

compared the impact of ENSO on TC activity and in- PDI, respectively. Accounting for the nonnormality of
tensity between the neighboring subdivisions of EDR storm data in WDR where fewer storms are observed
and WDR. In combination with the classification of seasonally, both TC indices have been transformed and
years of El Niño, La Niña, and neutral conditions, the tested for statistical differences between every pairwise
spatial and temporal influences on the ENP storm ac-
tivity and intensity are quantitatively resolved over a TABLE 6. Correlations of NTC and PDI with MEI are de-
42-yr period. Although the physical forcings accompa- termined as the Pearson coefficient of correlation (r) and the cor-
nying ENSO phase changes are not directly addressed, relation of determination (r 2) for Figs. 8 and 9. Boldface values are
previous studies have identified a combination of local statistically significant at p , 0.05.
thermodynamic and dynamic factors in speculating El NTC PDI
Niño–induced shifts of more intense storms with longer
MDR EDR WDR MDR EDR WDR
lifetimes.
Measures of the overall seasonal TC activity and in- r 0.28 0.12 0.35 0.32 0.17 0.35
r2 0.08 0.01 0.12 0.10 0.03 0.12
tensity are expressed empirically as indices of NTC and
2472 JOURNAL OF CLIMATE VOLUME 28

FIG. 11. Box-and-whisker plot summaries for data distributions of (a) PDI1, (b) PDI2, and
(c) PDI3 (scale of 107 m3 s23) in MDR, EDR, and WDR. Gray diamonds represent the regional
averages.

combination of three ENSO phases. Overall, both in- strength of correlations in WDR to be stronger for the
dices are only proven statistically different between El seasonal TC activity than TC intensity between 1971 and
Niño and La Niña years. When MDR is subdivided, such 2012. Statistical comparisons of regional TC indices at
a contrast is only maintained in WDR. This difference of different ENSO conditions validate previous findings
the regional sensitivity to ENSO is supported by results that the effects of environmental influences over ENP
of the correlation analysis, which demonstrates the storms would have been overlooked if MDR was not
15 MARCH 2015 JIEN ET AL. 2473

TABLE 7. Correlations of PDI1, PDI2, and PDI3 with MEI (May– Climate, 8, 1249–1260, doi:10.1175/1520-0442(1995)008,1249:
November) values in MDR, EDR, and WDR from 1971–2012. TCITSP.2.0.CO;2.
Boldface values are statistically significant at p , 0.05. Bove, M. C., J. B. Elsner, C. W. Landsea, X. Niu, and J. J. O’Brien,
1998: Effect of El Niño on U.S. landfalling hurricanes, re-
PDI1 PDI2 PDI3 visited. Bull. Amer. Meteor. Soc., 79, 2477–2482, doi:10.1175/
MDR 0.33 0.19 0.27 1520-0477(1998)079,2477:EOENOO.2.0.CO;2.
EDR 0.18 20.06 20.02 Camargo, S. J., and A. H. Sobel, 2005: Western north pacific
WDR 0.34 0.30 0.37 tropical cyclone intensity and ENSO. J. Climate, 18, 2996–
3006, doi:10.1175/JCLI3457.1.
——, K. A. Emanuel, and A. H. Sobel, 2007: Use of a genesis po-
tential index to diagnose ENSO effects on tropical cyclone
subdivided (Collins and Mason 2000; Ralph and Gough genesis. J. Climate, 20, 4819–4843, doi:10.1175/JCLI4282.1.
2009). Chan, J. C. L., 1985: Tropical cyclone activity in the northwest
Between storm parameters of frequency, intensity, Pacific in relation to the El Nino/Southern Oscillation
and duration, only the seasonal variability of TC in- phenomenon. Mon. Wea. Rev., 113, 599–606, doi:10.1175/
tensity contributes significantly to the ENSO signal on 1520-0493(1985)113,0599:TCAITN.2.0.CO;2.
Changnon, S. A., 1999: Impacts of 1997/98 El Niño generated
PDI. However, this independent analysis on the in- weather in the United States. Bull. Amer. Meteor. Soc.,
fluence of PDI due to ENSO-induced storm measure- 80, 1819–1827, doi:10.1175/1520-0477(1999)080,1819:
ments is also subject to regional variation. When IOENOG.2.0.CO;2.
comparing between MDR subdivisions, PDI is found to Chu, P.-S., 2004: ENSO and tropical cyclone activity. Hurricanes
be significantly sensitive to seasonal fluctuations for all and Typhoons: Past, Present, and Potential, R. J. Murnane and
K. B. Liu, Eds., Columbia University Press, 297–332.
WDR storm parameters, with storm count being rela- ——, and J. Wang, 1997: Tropical cyclone occurrences in the
tively more important. vicinity of Hawaii: Are the differences between El Niño
The findings generally underscore the societal impor- and non-El Niño years significant? J. Climate, 10, 2683–2689,
tance of ENSO on TC activity and intensity. Although doi:10.1175/1520-0442(1997)010,2683:TCOITV.2.0.CO;2.
most of the ENP storm development initiated in EDR, Clark, J. D., and P.-S. Chu, 2002: Interannual variation of tropical
cyclone activity over the central North Pacific. J. Meteor. Soc.
many storms did not reach maximum intensity until en-
Japan, 80, 403–418, doi:10.2151/jmsj.80.403.
tering WDR. During El Niño years, there are more WDR Collins, J. M., 2007: The relationship of ENSO and relative hu-
storms with higher intensity and longer lifetimes that de- midity to interannual variations of hurricane frequency in
velop than during La Niña years. A direct consequence is the North-East Pacific Ocean. Papers Appl. Geogr. Conf., 30,
that as the storms are transitioned from EDR; ENSO in- 324–333.
——, 2010: Contrasting high North-East Pacific tropical cyclone
fluences on seasonal WDR environmental conditions
activity. Southeast. Geogr., 50, 83–98, doi:10.1353/sgo.0.0069.
could fuel further storm development and prolong storm ——, and I. M. Mason, 2000: Local environmental conditions re-
tracks affecting islands in the central Pacific basin. Hence, lated to seasonal tropical cyclone activity in the Northeast
the determination and incorporation of ENSO as a pre- Pacific basin. Geophys. Res. Lett., 27, 3881–3884, doi:10.1029/
dictor for short-term, seasonal storm forecasts could prove 2000GL011614.
to be a practical tool that better anticipates the seasonal ——, and D. R. Roache, 2011: The 2009 hurricane season in the
eastern North Pacific basin: An analysis of environmental
outlook of WDR storm activity and intensity and the po- conditions. Mon. Wea. Rev., 139, 1673–1681, doi:10.1175/
tential TC-inflicted damages upon landfalls. Since Atlantic 2010MWR3538.1.
storm damage and the relationship with the oscillation of Emanuel, K. A., 2005: Increasing destructiveness of tropical cy-
ENSO phases have already been well documented clones over the past 30 years. Nature, 436, 686–688, doi:10.1038/
(Hebert et al. 1997; Pielke and Landsea 1999), future work nature03906.
Frank, W. M., and G. S. Young, 2007: The interannual variability of
on such a data archive should include additional input on
tropical cyclones. Mon. Wea. Rev., 135, 3587–3598, doi:10.1175/
the relationship of ENSO events with the societal cost MWR3435.1.
associated with ENP storm landfalls on the Pacific islands Goldenberg, S. B., and L. J. Shapiro, 1996: Physical mechanisms
and continental United States. for the association of El Niño and West African rainfall with
Atlantic major hurricane activity. J. Climate, 9, 1169–1187,
Acknowledgments. The authors would like to thank doi:10.1175/1520-0442(1996)009,1169:PMFTAO.2.0.CO;2.
——, C. W. Landsea, A. M. Mestas-Nuñez, and W. M. Gray, 2001:
the editor and anonymous reviewers for their valuable
The recent increase in Atlantic hurricane activity: Causes
feedback in greatly improving the paper. and implications. Science, 293, 474–479, doi:10.1126/
science.1060040.
REFERENCES Gray, W. M., 1968: Global view of the origin of tropical distur-
bances and storms. Mon. Wea. Rev., 96, 669–700, doi:10.1175/
Basher, R. E., and X. Zheng, 1995: Tropical cyclones in the 1520-0493(1968)096,0669:GVOTOO.2.0.CO;2.
Southwest Pacific: Spatial patterns and relationships to ——, 1977: Tropical cyclone genesis in the western North Pacific.
Southern Oscillation and sea surface temperature. J. J. Meteor. Soc. Japan, 55, 465–482.
2474 JOURNAL OF CLIMATE VOLUME 28

——, 1984a: Atlantic seasonal hurricane frequency. Part I: El Niño Nicholls, N., 1979: A possible method for predicting seasonal
and 30 mb quasi-biennial oscillation influences. Mon. Wea. tropical cyclone activity in the Australian region. Mon. Wea.
Rev., 112, 1649–1668, doi:10.1175/1520-0493(1984)112,1649: Rev., 107, 1221–1224, doi:10.1175/1520-0493(1979)107,1221:
ASHFPI.2.0.CO;2. APMFPS.2.0.CO;2.
——, 1984b: Atlantic seasonal hurricane frequency. Part II: ——, 1992: Recent performance of a method for forecasting Aus-
Forecasting its variability. Mon. Wea. Rev., 112, 1669–1683, tralian seasonal tropical cyclone activity. Aust. Meteor. Mag.,
doi:10.1175/1520-0493(1984)112,1669:ASHFPI.2.0.CO;2. 40, 105–110.
——, C. W. Landsea, P. W. Mielke Jr., and K. J. Berry, 1994: Pielke, R. A., Jr., and C. N. Landsea, 1999: La Niña, El Niño,
Predicting Atlantic basin seasonal tropical cyclone activity and Atlantic hurricane damages in the United States.
by 1 June. Wea. Forecasting, 9, 103–115, doi:10.1175/ Bull. Amer. Meteor. Soc., 80, 2027–2033, doi:10.1175/
1520-0434(1994)009,0103:PABSTC.2.0.CO;2. 1520-0477(1999)080,2027:LNAENO.2.0.CO;2.
Hebert, P. J., J. D. Jarrell, and M. Mayfield, 1997: The deadliest, Ralph, T. U., and W. A. Gough, 2009: The influence of sea-surface
costliest, and most intense United States hurricanes of this temperatures on Eastern North Pacific tropical cyclone
century (and other frequently requested hurricane facts). activity. Theor. Appl. Climatol., 95, 257–264, doi:10.1007/
NOAA Tech. Memo. NWS TPC-1, Miami, FL, 30 pp. s00704-008-0004-x.
[Available from National Hurricane Center, 11691 SW 17th Rasmusson, E. M., and T. H. Carpenter, 1982: Variations in
St., Miami, FL 33165.] tropical sea surface temperature and surface wind fields as-
Helsel, D. R., and R. M. Hirsch, 1991: Statistical methods in water sociated with the Southern Oscillation/El Niño. Mon. Wea.
resources. Techniques of Water-Resources Investigations of Rev., 110, 354–384, doi:10.1175/1520-0493(1982)110,0354:
the United States Geological Survey, USGS, 266 pp. VITSST.2.0.CO;2.
Irwin, R. P., III, and R. E. Davis, 1999: The relationship between Ropelewski, C. F., and P. D. Jones, 1987: An extension of the Tahiti–
the Southern Oscillation index and tropical cyclone tracks in Danvin Southern Oscillation index. Mon. Wea. Rev., 115, 2161–
the eastern North Pacific. Geophys. Res. Lett., 26, 2251–2254, 2165, doi:10.1175/1520-0493(1987)115,2161:AEOTTS.2.0.CO;2.
doi:10.1029/1999GL900533. Schultz, L., 2007: Some climatological aspects of tropical cyclones
Jones, C. G., and C. D. Thorncroft, 1998: The role of El Niño in the eastern North Pacific. Natl. Wea. Dig., 32, 45–54.
in Atlantic tropical cyclone activity. Weather, 53, 324–336, Simpson, R. H., 1974: The hurricane disaster potential scale.
doi:10.1002/j.1477-8696.1998.tb06409.x. Weatherwise, 27, 169–186, doi:10.1080/00431672.1974.9931702.
Klotzbach, P. J., 2011: El Niño–Southern Oscillation’s impact on Sprent, P., 1989: Applied Nonparametric Statistical Methods.
Atlantic basin hurricanes and U.S. landfalls. J. Climate, 24, Chapman and Hall, 259 pp.
1252–1263, doi:10.1175/2010JCLI3799.1. Steenhof, P. A., and W. A. Gough, 2008: The impact of tropical
——, 2012: El Niño–Southern Oscillation, the Madden–Julian sea surface temperatures on various measures of Atlantic
Oscillation and Atlantic basin tropical cyclone rapid in- tropical cyclone activity. Theor. Appl. Climatol., 92, 249–255,
tensification. J. Geophys. Res., 117, D14104, doi:10.1029/ doi:10.1007/s00704-007-0316-2.
2012JD017714. Trenberth, K. E., 1997: The definition of El Niño. Bull. Amer. Meteor.
——, and S. E. Blake, 2013: North-central Pacific tropical cyclones: Soc., 78, 2771–2777, doi:10.1175/1520-0477(1997)078,2771:
Impacts of El Niño–Southern Oscillation and the Madden– TDOENO.2.0.CO;2.
Julian oscillation. J. Climate, 26, 7720–7733, doi:10.1175/ Wang, B., and J. C. L. Chan, 2002: How strong ENSO events affect
JCLI-D-12-00809.1. tropical storm activity over the western North Pacific. J. Cli-
Landsea, C. W., 2000: El Niño/Southern Oscillation and the sea- mate, 15, 1643–1658, doi:10.1175/1520-0442(2002)015,1643:
sonal predictability of tropical cyclones. El Niño and the HSEEAT.2.0.CO;2.
Southern Oscillation: Multiscale Variability and Global and Webster, P. J., G. J. Holland, J. A. Curry, and H.-R. Chang, 2005:
Regional Impacts, H. F. Diaz and V. Markgraf, Eds., Cam- Changes in tropical cyclone number, duration, and intensity in
bridge University Press, 149–181. a warming environment. Science, 309, 1844–1846, doi:10.1126/
Lupo, A. R., T. K. Latham, T. H. Magill, J. V. Clark, C. J. Melick, science.1116448.
and P. S. Market, 2008: Interannual variability of hurricane Wolter, K., and M. S. Timlin, 1998: Measuring the strength of
activity in the Atlantic and East Pacific regions. Natl. Wea. ENSO events—How does 1997/98 rank? Weather, 53, 315–
Dig., 32, 119–133. 324, doi:10.1002/j.1477-8696.1998.tb06408.x.
Malkus, J. S., and H. Riehl, 1960: On the dynamics and energy Wu, L., B. Wang, and S. A. Braun, 2008: Implications of tropical
transformations in steady-state hurricanes. Tellus, 12A, 1–20, cyclone power dissipation index. Int. J. Climatol., 28, 727–731,
doi:10.1111/j.2153-3490.1960.tb01279.x. doi:10.1002/joc.1573.
Molinari, J., D. Vollaro, S. Skubis, and M. Dickinson, 2000: Origins Wu, P., and P.-S. Chu, 2007: Characteristics of tropical cyclone
and mechanisms of eastern Pacific tropical cyclogenesis: A activity over the eastern North Pacific: The extremely active
case study. Mon. Wea. Rev., 128, 125–139, doi:10.1175/ 1992 and the inactive 1977. Tellus, 59A, 444–454, doi:10.1111/
1520-0493(2000)128,0125:OAMOEP.2.0.CO;2. j.1600-0870.2007.00248.x.

You might also like