Intraseasonal Oscillations and Interannual Variability of...

127
INTRASEASONAL OSCILLATIONS AND INTERANNUAL VARIABILITY OF THE INDIAN SUMMER MONSOON A thesis submitted for the award of the degree of Doctor of Philosophy in the Faculty of Engineering by R. S. AJAYA MOHAN Centre for Atmospheric and Oceanic Sciences Indian Institute of Science Bangalore 560 012 INDIA NOVEMBER 2001

Transcript of Intraseasonal Oscillations and Interannual Variability of...

Page 1: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

INTRASEASONAL OSCILLATIONSAND INTERANNUAL VARIABILITY OF

THE INDIAN SUMMER MONSOON

A thesis submitted for the award of the degree ofDoctor of Philosophy

in theFaculty of Engineering

by

R. S. AJAYA MOHAN

Centre for Atmospheric and Oceanic SciencesIndian Institute of Science

Bangalore 560 012INDIA

NOVEMBER 2001

Page 2: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Dedicated to My Parents

Page 3: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

”I looked forward to the coming of the monsoon and I became a watcher of skies,waiting to spot the heralds the preceded the attack. A few showers came. Oh! thatwas nothing, I was told; the monsoon has yet to come. Heavier rains followed, butI ignored them and waited for some extraordinary happening. While I waited I learntfrom various people that the monsoon had definitely come and established itself. Wherewas the pomp and circumstance and the glory of the attack, and the combat betweencloud and land, and the surging and lashing sea? Like a thief in the night the monsoonhad come to Bombay, as well it might have done in Allahabad or elsewhere. Anotherillusion gone”

Jawaharlal Nehru, The monsoon comes to Bombay, 1939

Page 4: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Contents

Acknowledgements i

Abstract iii

Acronyms viii

List of Figures ix

1 Introduction 11.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81.2 Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.2.1 Reanalysis Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91.2.2 NOAA Outgoing Long wave Radiation (OLR) Dataset . . . . . . . 101.2.3 Precipitation Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . 111.2.4 Statistics of Low Pressure Systems . . . . . . . . . . . . . . . . . . . 12

2 Basic Characteristics of Monsoon Intraseasonal Oscillations 132.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132.2 Propagation Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . 162.3 A Circulation Criterion for ’Active’ and ’Break’ Phases . . . . . . . . . . . 192.4 Mean Structure of ISOs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212.5 Meridional Bimodality of ISO Spatial Structure . . . . . . . . . . . . . . . . 302.6 Discussions and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3 Intraseasonal Oscillations and Interannual Variability of the Indian SummerMonsoon 343.1 A Common Spatial Mode of Intraseasonal and Interannual Variability . . 343.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon . 403.3 Interannual Variations of ISO Activity and Seasonal Mean Monsoon . . . 503.4 Discussions and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4 Estimate of Potential Predictability of Monthly and Seasonal Means in Tropicsfrom Observations 554.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.2 Estimation of Potential Predictability of Monthly Means . . . . . . . . . . . 58

4.2.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584.2.2 Estimation of ’Internal’ and ’External’ Interannual Variances . . . . 654.2.3 Potential Predictability of Monthly Means . . . . . . . . . . . . . . 68

4.3 Potential Predictability of Seasonal means . . . . . . . . . . . . . . . . . . . 764.4 Discussions and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . 82

Page 5: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.5 Appendix : Procedure for Estimating ’Climate Noise’ . . . . . . . . . . . . 85

5 Clustering of Synoptic Systems During the Indian Summer Monsoon by In-traseasonal Oscillations 875.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875.2 Wet and Dry Spells and Clustering of LPS . . . . . . . . . . . . . . . . . . . 895.3 Monsoon Intraseasonal Oscillation Index . . . . . . . . . . . . . . . . . . . 915.4 Clustering of Genesis of LPS by Intraseasonal Oscillations . . . . . . . . . 915.5 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

6 Conclusions 96

Bibliography 100

Page 6: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Acknowledgments

I am fortunate to have got a chance to work with Prof B. N. Goswami. The vastness

of his knowledge and his abundant enthusiasm to seek new results has always been

a source of encouragement to me. His critical appraisals and encouragement of inde-

pendent thought during discussions have contributed immensely to the course of this

thesis. It helped smoothen many hurdles. His calm and collective approach was invalu-

able in putting things in the right perspective. He shall always remain in my mind as a

model of an intelligent, enthusiastic and hardworking man-of-science.

Many thanks are due to the faculty in CAOS, especially Prof Sengupta and Prof

Srinivasan for many informal discussions and encouragement. The cheerful and help-

ful people in CAOS, supporting staff, project staff and students deserve appreciation.

Rama, Padma, Mohan, Raja and Shiva were always obliging despite their strenuous

work load. Thanks are due to Natraj, Rajasekhar and Srinivas for helping me in solving

system related problems. It was a pleasure to work with Rajendran, Sajani, Anagani,

Janakiraman, J V S Raju, Salil, Arindam, Chandru and Pallav. Informal discussions/chat

with Retish, Prince, Francis, Vinoj and Simi both within the department and in Raf-

fique’s Tea-kiosk is fondly acknowledged. No words can express my gratitude to Manu

as she has been a constant source of support and encouragement.

I would like to acknowledge interactions with Prof N. Balakrishnan for I have learned

a lot from him. His vision and cool and collected approach was indeed impressive. I

will be failing in my duty, if I do not thank Prof G. Padmanabhan - He taught me how

simple and straightforward a human being should be.

I am grateful to Prof M. Ghil, Dr M. Kimoto and Dr A. Robertson for providing

information regarding probability density estimation.

Thanks are due to the faculty in Department of Atmospheric Science, Cochin Uni-

versity of Science & Technology, especially Prof Mohan Kumar for encouraging me to

take up a research career.

Page 7: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Acknowledgements ii

I cherish the brief interaction with Bharataratna Dr A.P.J Abdul Kalam. His words

’Only strength respects strength’ gave me a fresh sense of enthusiasm to work for this

great country.

Working for Students’ council helped me understand problems of others and gave

me ample opportunity to interact with all kinds of people. This was also an oppor-

tunity to befriend many. I would like to acknowledge the support and friendship of

Saishankar, Rajkumar, Vasan, Ganesh, Guruprasad, Pratap Jayaprakash, John, Dhruba

and Suresh. Special thanks are due to Brar and Suma for, they offered good company.

Thanks are due to the the ’famous mallu gang’ as the friendship they offered is

matchless. Cheers to Venu, Anoop (simplan) and Bijoy. It is pleasure to have friends like

Anil (kunz), Hari (healy), Glomin (RC), Anil (aavi), Suresh (kumily), Manoj (neergosh),

Prabhu, Randhir (kuru), Baiju, Ajayan, Sameen, John, Vinay and Vinod. Thanks are

also due to Pappan and Pappy, Suresh (cobra) and Sridevi, Sunoj and Viji. Let me

acknowledge the support offered by Sriram.

Finally, my deepest sense of gratitude go to my parents and my family for their

goodwill and blessings. I am grateful to my father for sharing all my worries and hap-

piness. Without his constant encouragement this thesis would not have been possible.

Innumerable phone calls gave me a feeling that I am at home, away from home. I shall

always strive to rise up to his expectations.

Thanks to all the unknown faces that continue to develop and strive for free soft-

ware. Working with ”LINUX”, ”GrADS” and ”LATEX” and numerous other free soft-

ware made life easier in the pursuit of this thesis. Thanks are due to SERC for high

power computing.

Last but not least, let me thank Indian Institute of Science and Council of Scientific

and Industrial Research for providing financial support.

Page 8: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Abstract

Several modeling studies show that the predictability of the seasonal mean Indian

summer monsoon is limited due to a significant fraction of the interannual variability of

the seasonal mean being governed by internal chaotic dynamics. What causes the inter-

nal low frequency variations of the Indian summer monsoon? One possible candidate

is the monsoon intraseasonal oscillations (ISOs). Indian summer monsoon has vigorous

intraseasonal oscillations in the form of ’active’ and weak (or ’break’) spells of monsoon

rainfall within the summer monsoon season. These ’active’ and ’break’ spells of the

monsoon are associated with fluctuations of the tropical convergence zone. Temporally

ISOs of the Indian summer monsoon represent two preferred bands of periods, one be-

tween 10 and 20 days and the other between 30 and 60 days. As the separation between

the dominant ISO periods and the season is not large, the statistics of the ISOs could,

in principle, influence the seasonal mean monsoon and it’s interannual variability. To

the extent that the ISOs are intrinsically chaotic and unpredictable, the predictability of

the Indian summer monsoon would depend on relative contribution of the ISOs to the

seasonal mean compared to the more predictable externally forced component.

Therefore, it is of great importance to establish (a) whether there exits a physical

basis for monsoon ISOs to influence the seasonal mean. (b) Even if there exits a physi-

cal basis for the ISOs to influence the seasonal mean, is there an empirical evidence of

association between some statistics of the ISOs and interannual variability of the Indian

summer monsoon? (c) If such an association between monsoon ISOs and the seasonal

mean monsoon exits, it would be desirable to make quantitative estimate of the extent

to which ISOs influence the seasonal mean and its interannual variability. The primary

objectives of this study are to address these three issues using sufficiently long homoge-

neous daily circulation and convection data. Although spatial and temporal structures

of the monsoon ISOs have been examined extensively, the relationship between ISOs

and interannual variability has received little attention in the past. The existing litera-

ture on the subject is critically reviewed in Chapter 1.

Page 9: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Abstract iv

In an attempt to establish the physical basis for the ISOs to influence the seasonal

mean, we first examine the similarity between the spatial structure of the ISOs and the

seasonal mean. The large scale nature of the Indian summer monsoon ISOs and rela-

tionship between circulation and convection on this time scale are investigated using

42-years (1956-1997) daily circulation data from NCEP/NCAR reanalysis and satellite

derived outgoing long wave radiation data for the period 1974-1997. Traditionally, ’ac-

tive’ and ’break’ conditions or the dry and wet spells of the monsoon ISO are defined

based on continental precipitation. Arguing that the dry and wet spells are part of large

scale fluctuations associated with the ISO, a circulation based criterion is devised to de-

fine ’active’ and ’break’ monsoon conditions using zonal winds at 850 hPa over the Bay

of Bengal. Although the ISOs vary in intensity and period, it is shown that, the under-

lying spatial structure of a typical ISO cycle in circulation and convection is invariant

over the years and is constructed using a composite technique. Typical ISOs have large

scale horizontal structure similar to the seasonal mean and intensifies (weakens) the

mean flow during it’s ’active’ (’break’) phase. A typical ’active’ (’break’) phase is also

associated with enhanced (decreased) cyclonic low-level vorticity and convection and

anomalous upward (downward) motion in the northern position of the tropical con-

vergence zone (TCZ) and decreased (increased) convection and anomalous downward

(upward) motion in the southern position of the TCZ. The cycle evolves with a north-

ward propagation of the TCZ and convection from the southern to the northern position

of the TCZ. Thus the ISOs result in spinning up (or spinning down) of the large scale

mean monsoon circulation in it’s extreme phases. (Chapter 2)

A physical basis for ISOs to influence the seasonal mean and it’s interannual vari-

ability is established when it is shown that the intraseasonal and interannual variations

are governed by a common mode of spatial variability. The spatial pattern of standard

deviation of intraseasonal and interannual variability of low-level vorticity is shown to

be similar. The spatial pattern of the dominant mode of ISO variability of the low-level

winds is also shown to be similar to that of the interannual variability of the seasonal

mean winds. The similarity between the spatial patterns of the two variability indi-

cates that higher frequency of occurrence of ’active’ (’break’) conditions would result in

’stronger’ (’weaker’) than normal seasonal mean. This possibility is tested by calculat-

ing probability density function (PDF) of the ISO activity in the low-level vorticity repre-

sented by the two dominant empirical orthogonal functions (EOFs). The PDF estimates

for ’strong’ monsoon years and ’weak’ monsoon years are shown to be asymmetric in

Page 10: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Abstract v

both the cases. It is seen that the ’strong’ (’weak’) monsoon years are associated with

higher probability of occurrence of ’active’ (’break’) conditions. This result is further

supported by calculation of PDF of ISO activity from combined vorticity and outgoing

long wave radiation. This result, indicates that the frequency of intraseasonal pattern

determine the seasonal mean. As the ISOs are essentially chaotic, it raises an important

question on predictability of the Indian summer monsoon. (Chapter 3)

Having shown that the ISOs can influence the seasonal mean and its interannual

variability, the next objective is to make quantitative estimates of potential predictabil-

ity of the monsoon climate. A measure of potential predictability of the monthly and

seasonal means in a place could be obtained from the ratio of variances associated with

the ’external’ to the ’internal’ components. A method of separating the ’external’ com-

ponent arising from contributions from slowly varying boundary forcing from the ’in-

ternal’ components (e.g. intraseasonal oscillations) that determines the potential pre-

dictability of the monthly mean tropical climate is proposed. Based on 33 years of daily

low-level wind observations and 24 years of satellite observations of outgoing long

wave radiation, we show that the Indian monsoon climate is marginally predictable

on monthly time scales as the contribution from the boundary forcing in this region

is comparable to that from the internal dynamics. It is further shown that excluding

the Indian monsoon region, the predictable region is larger and predictability is higher

in the tropics during northern summer. Even though the boundary forced variance is

large during northern winter, the predictable region is smaller as the internal variance is

larger and covers a larger region during northern winter due to stronger intraseasonal

activity. It is also shown that most of the internal low frequency variability in the Indian

summer monsoon region arise from the ISOs. (Chapter 4)

An estimate of potential predictability for the Northern Hemisphere summer and

winter seasons in the tropics has also been made using an established method of esti-

mating ’climate noise’. Even on seasonal mean time scales, we show that the Indian

monsoon climate is only marginally predictable as the contribution of the boundary

forcing in this region is relatively low and that of the internal dynamics is relatively

large. (Chapter 4)

While the monsoon ISOs seem to lead to decrease in the predictability of monthly

or seasonal mean monsoon climate, it is possible that the same ISOs lead to extended

range prediction of spells of synoptic activity. We recall that the seasonal mean mon-

soon is strengthened in one phase of the ISOs (active phase) while it is weakened in

Page 11: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Abstract vi

another (’break’) phase of the monsoon. The main rain bearing system during the mon-

soon season are the Low Pressure Systems (LPS) consisting of lows and depressions.

Since the genesis of the LPS depends on the horizontal shear and low-level vorticity, it

is possible that more LPS form in the active phase relative to the break phase. In other

words, large scale circulation associated with the ISOs could modulate the frequency

of genesis of LPS. We examined how the LPS are modulated by the intraseasonal oscil-

lations. Using more than 40 years of LPS genesis statistics and daily circulation data,

here we show that the dry and wet spells are the result of clustering of lows and de-

pressions caused by modulation of the large scale monsoon flow by the intraseasonal

oscillations. The slow evolution of the ISOs may permit extended range prediction of

the ISO phases and through them dry and wet spells of the Indian summer monsoon

(Chapter 5). Major results and outstanding issues are discussed in Chapter 6.

Publications

1. B.N Goswami and R.S Ajaya Mohan, 2001: Intraseasonal Oscillations and Int

erannual Variability of Indian Summer Monsoon. J.Climate, 14, 1180 -1198.

2. B.N Goswami and R.S. Ajaya Mohan, 2001: Estimate of Predictability of Monthly

Means in Tropics from Observations. Curr.Sci., 80, 56-63.

3. R. S. Ajaya Mohan and B.N. Goswami, 2000: A Common spatial mode for in-

traseasonal and interannual variation and predictability of the Indian Summer

Monsoon. Curr.Sci., 79, 1106-1111.

4. B.N Goswami and R.S. Ajaya Mohan, 2001: Intra-seasonal Oscillations and pre-

dictability of the Indian summer monsoon. Proc.Ind.Nat.Sci.Aca., 67A (3), 369-383.

5. B.N Goswami, R.S Ajaya Mohan, Prince K Xavier and D. Sengupta 2001: Cluster-

ing of low-pressure systems during the Indian summer monsoon by Intraseasonal

Oscillations, Geophys.Res.Letts, 30, 1431, doi:10.1029/2002GL016734.

6. R. S. Ajaya Mohan and B. N. Goswami 2003: Potential predictability of the Asian

Summer Monsoon on Monthly and Seasonal Time Scales, Meteorol.Atmos.Phys.,

84, 83-100.

7. B.N Goswami and R.S Ajaya Mohan, 2005: Multi-scale interactions and pre-

dictability of the Indian summer monsoon, section 3 in Nonequillibrium Phenomena

Page 12: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Abstract vii

in Plasmas, Eds. Sharma, A. Surjalal, Kaw Predhiman, K, IX, 347p, ISBN: 1-4020-

3108-4, Springer, USA.

Page 13: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Acronyms

AGCM Atmospheric General Circulation Model

AMIP Atmospheric General Circulation Model Intercomparison Project

CEOF Combined Empirical Orthogonal Function

CMAP Climate prediction center Merged Analysis of Precipitation

DJF December-January-February

ENSO El Nino-Southern Oscillation

EOF Empirical Orthogonal Function

JJA June-July-August

JJAS June-July-August-September

IMD India Meteorological Department

IMR All India Monsoon Rainfall Index

ISOs Intraseasonal oscillations

LPS Low Pressure Systems

MISI Monsoon Intraseasonal Oscillation Index

MTV Monsoon Trough Vorticity

NCAR National Center for Atmospheric Research

NCEP National Centers for Environmental Prediction

NH Northern Hemisphere

OLR Outgoing Long wave Radiation

PC Principal Component

PDF Probability Density Function

SD Standard Deviation

TCZ Tropical Convergence Zone

U850 Low level zonal winds (850 hPa)

U200 Upper level zonal winds (200 hPa)

Z700 Geopotential height (700 hPa)

Page 14: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures

2.1 Climatological mean (JJAS) monsoon winds (ms−1) and precipitation (mm.day−1).

(a) 850 hPa vector winds, (b) Relative vorticity at 850 hPa (10−6s−1), (c) 200 hPa

vector winds, (d) Precipitation from Xie and Arkin [1997]. . . . . . . . . . . . . . 14

2.2 Some examples of raw time series of zonal winds at 850 hPa at a few selected

points during 1990. (Left panels) Daily zonal winds (ms−1) with the annual cy-

cle (annual and semi-annual harmonics, green lines). (Right panels) Anomalous

daily zonal winds (ms−1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.3 Examples of spectra of zonal winds and OLR for a typical year (1984) at a typical

point (90◦E, 10◦N). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.4 An example illustrating the horizontal scale and vertical structure of the domi-

nant ISO mode. (a) Lag-zero correlations of the 850 hPa 30-60 day filtered zonal

winds with respect to a reference point (85◦E, 10◦N). (b) Lag-zero correlations

between 30-60 day filtered zonal winds at 850 hPa and 200 hPa at each grid

point. Correlations are calculated between May 1 and October 31 of 1990. Cor-

relations exceeding 0.2 are significant at 95% confidence level. . . . . . . . . . . 17

2.5 (a) Correlations between 30-60 day filtered zonal winds at 850 hPa with respect

to that at a reference point (85◦E, 10◦N) at different lead/lags averaged over

(80◦E-90◦E) for 1990. (b) Same as (a) but for 30-60 day filtered OLR. Starting

contour is ±0.1 and contour interval is 0.2. . . . . . . . . . . . . . . . . . . . . . 18

2.6 (a) Correlations between 30-60 day filtered zonal winds at 850 hPa with respect

to that at a reference point (85◦E, 10◦N) at different lead/lags averaged over

(10◦N-20◦N) for 1990. (b) Same as (a) but for 30-60 day filtered OLR. Starting

contour is ±0.1 and contour interval is 0.2. . . . . . . . . . . . . . . . . . . . . 18

Page 15: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures x

2.7 (a) An example of 30-60 day filtered zonal winds for 1986 at a reference point

(90◦E, 15◦N). The thin horizontal lines correspond to +1 and -1 standard devi-

ations. ’Active’ (’break’) days are defined as days for which the filtered zonal

winds at the reference point are greater than +1 S.D (or less than -1 S.D). (b) 12-

year (1978-1989) mean precipitation difference (mm.day−1) between all ’active’

and ’break’ composites. Contours are ±(1, 3, 5, 7, 9, 11, 13, 15). . . . . . . . . . . 20

2.8 (a,b) Climatological mean composite vector wind anomalies (ms−1) at 850 hPa

corresponding to ’active’ and ’break’ conditions for the 30-60 day mode and

(c,d) associated relative vorticity (10−6s−1). The climatological mean composite

is calculated by averaging all ’active’ and ’break’ conditions occurring during

the 20-year period (1978-1997). Shading in the upper panels indicates regions

with anomalies significant above 90% confidence level. . . . . . . . . . . . . . . 21

2.9 Climatological mean composite vector wind anomalies (ms−1) corresponding

to ’active’ and ’break’ conditions for the 30-60 day mode (a,b) at 500 hPa and

(c,d) at 200 hPa. The climatological mean composite is calculated by averaging

all ’active’ and ’break’ conditions occurring during the 20-year period (1978-

1997). Shading indicates regions with anomalies significant above 90% confi-

dence level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.10 Climatological mean composite OLR anomalies (Wm−2) corresponding to ’ac-

tive’ and ’break’ conditions. ’Active’ and ’break’ composites are constructed

using unfiltered OLR anomalies and the same ’active’ and ’break’ dates defined

by 30-60 day filtered zonal wind anomalies as used in Figure 2.8. OLR anomalies

above 5 Wm−2 are significant above 90% confidence level. . . . . . . . . . . . . 23

2.11 Climatological mean composite pressure vertical velocity anomalies (ω) at 500

hPa (hPas−1). Again the same ’active’ and ’break’ dates chosen from 30-60 day

filtered zonal wind anomalies for the 20-year period (1978-1997) as used in Fig-

ure 2.8 and Figure 2.10 are used. . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.12 Climatological mean composite vector wind anomalies (ms−1) at 850 hPa and

associated relative vorticity (10−6s−1) corresponding to eight phases of evolu-

tion of the 30-60 day mode for the period 1979-1989. The phase-1 corresponds

to the days when the filtered zonal wind anomalies at the reference point is zero

and increasing toward positive values. . . . . . . . . . . . . . . . . . . . . . . . 27

Page 16: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures xi

2.13 Climatological mean composite OLR anomalies (Wm−2) corresponding to eight

phases of evolution of the 30-60 day mode for the period 1979-1997. Eight com-

posite phases are constructed using unfiltered OLR anomalies and the same

dates defined by 30-60 day filtered zonal winds as used in Figure 2.12. . . . . . . 28

2.14 (a,b) Climatological mean composite vector wind anomalies (ms−1) at 850 hPa

corresponding to ’active’ and ’break’ conditions for the 10-20 day mode and

(c,d) associated relative vorticity (10−6s−1). The climatological mean composite

is calculated by averaging all ’active’ and ’break’ conditions occurring during

the 20-year period (1978-1997). . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.15 Meridional bimodality of spatial structure of the dominant ISO. (a) Scatter plot

of daily 30-60 day filtered vorticity at 850 hPa (10−6s−1) over a northern band

(70◦E-100◦E, 12◦N-22◦N) and a southern band (70◦E-100◦E, 5◦S-10◦N) during

1 June to 30 September for 19 years (1979-1997). (b) Scatter plot of 30-60 day

filtered OLR anomalies (Wm−2) averaged over the northern TCZ (70◦E-100◦E,

12◦N-22◦N) and the southern TCZ (70◦E-100◦E, 0◦-12◦S) during 1 June to 30

September for 18 years (1979-1997, excluding 1994). . . . . . . . . . . . . . . . . 29

2.16 (a) Scatter plot of 30-60 day filtered relative vorticity at 850 hPa (10−6s−1) and

OLR (Wm−2) anomalies averaged over a box (85◦E-95◦E, 12◦N-22◦N) of the

northern TCZ during 1 May to 31 October for 19 years (1979-1997). (b) same

as (a) but averaged over a box (85◦E-95◦E, 0◦-12◦S) of the southern TCZ. . . . . . 31

3.1 Geographical distribution of intraseasonal and interannual activity. (a) Mean

standard deviation of ISO filtered relative vorticity (10−6s−1) at 850 hPa during

1 June to 30 September for 20 years (1978-1997). (b) Interannual standard devi-

ation of seasonal mean relative vorticity (JJAS, 10−6s−1) based on the same 20

years. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3.2 First EOF of the intraseasonal and interannual 850 hPa winds. (a) Intraseasonal

EOFs are calculated with ISO filtered winds for the summer months (1 June to

30 September) for a period of 20 years (1978-1997). (b) Interannual EOFs are

calculated with the seasonal mean (JJAS) winds for 40-year period (1958-1997).

Units of vector loading are arbitrary. (c) Relation between IMR and interannual

PC1. Filled bars indicate interannual PC1 and the unfilled bar represent IMR.

Both time series are normalized by their own standard deviation. Correlation

between the two time series is shown. . . . . . . . . . . . . . . . . . . . . . . . 38

Page 17: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures xii

3.3 Ratio between standard deviation of interannual variation of ISO activity and

interannual variation of the seasonal mean. (a) Relative vorticity at 850 hPa. (b)

OLR. Contours are (0.3, 0.4, 0.6, 0.8, 1.0). . . . . . . . . . . . . . . . . . . . . . . 39

3.4 First two EOFs of the daily ISO filtered 850 hPa vorticity from 1 June to 30

September. (a) EOF1 and (b) EOF2 for seven ’strong’ years (c) EOF1 and (d)

EOF2 for ten ’weak’ years (e) EOF1 and (f) EOF2 for ’all’ (20 years from 1978 to

1997) years. Arbitrary EOF loadings have been multiplied by a factor of 100. . . . 42

3.5 Evidence of change in regimes of ISOs during ’strong’ and ’weak’ monsoon

years. Illustrated are two-dimensional PDFs of the ISO state vector spanned by

two dominant EOFs of low-level vorticity. PDFs are calculated with principal

components normalized by their own standard deviation and taking the sum-

mer days (1 June to 30 September) for (a) 7 ’strong’ monsoon years (b) 10 ’weak’

monsoon years (c) 20 combined ’strong’, ’weak’ and ’normal’ years (1978-1997).

The smoothing parameter used is h=0.6 and PDFs are multiplied by a factor 100.

The first two EOFs (not shown) are different in ’strong’, ’weak’ and ’all’ years

but are related to ’active’ and ’break’ conditions. The origin of the plots corre-

sponds to a very weak state representing a transition between the two states (as

in the ’all’ case). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.6 Geographical patterns of the dominant regimes for low-level relative vorticity

(10−6s−1) shown in Figure 3.5. (a) ’strong’ monsoon years (b) ’weak monsoon

years (c) ’all’ years. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.7 The monsoon trough vorticity (MTV) and the Indian Monsoon Rainfall (IMR)

for a 40-year period (1958-1997). MTV is defined as the seasonal mean vorticity

(JJAS) averaged in the domain 40◦E-90◦E and 10◦N-30◦N. Both time series are

normalized by their own standard deviation. Correlation between the two time

series is shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.8 First two CEOFs of the daily ISO filtered 850 hPa vorticity and OLR from 1 June

to 30 September. (a) CEOF1 and (b) CEOF2 for six ’strong’ years (c) CEOF1 and

(d) CEOF2 for six ’weak’ years (e) CEOF1 and (f) CEOF2 for ’all’ (20 years from

1978 to 1997) years. Arbitrary EOF loadings have been multiplied by a factor of

100. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.9 Same as Figure 3.5 but based on the state vector defined by the first two com-

bined EOF of low-level vorticity and OLR. . . . . . . . . . . . . . . . . . . . . . 48

Page 18: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures xiii

3.10 Geographical patterns of the dominant regimes shown in Figure 3.9. (a) ’strong’

monsoon years (b) ’weak monsoon years (c) ’all’ years. OLR patterns are shown

as shaded contours (Wm−2) while the corresponding low-level vorticity are

shown in contours (10−6s−1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3.11 (a) Time series of ISO activity index (blue) and All India Monsoon Rainfall Index

(IMR, black) normalized by it’s own standard deviation for a 44-year period

(1954-1997). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.1 An illustration of variations of the annual cycle from year to year. The annual

cycle of zonal winds (ms−1) at 850 hPa at a point (80◦E, 5◦N) are shown for 20

years. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.2 First combined EOF of mean monthly ’external’ anomalies for the period Jan-

uary 1979 to December 1997 (228 months). (a) Zonal winds EOF at 850 hPa, (b)

OLR EOF and (c) PC1 (solid line) and Nino3 SST anomalies (dashed line). Both

the time series are normalized by their own standard deviation. Units of the

EOFs are arbitrary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.3 Time-longitude section of mean monthly ’external’ anomalies of zonal wind at

850 hPa (ms−1) and OLR (Wm−2) averaged around equator (5◦S-5◦N). . . . . . 64

4.4 Monthly variance of zonal winds (m2s−2) at 850 hPa based on 396 months for the

period January 1965 to December 1997. (a) Total variance (b) ’external’ variance

and (c) ’internal’ variance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.5 Same as Figure 4.4 but for OLR for the period January 1980 to December 1999

(240 months). Units, (Wm−2)2. . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.6 Estimates of ’F’ ratios for zonal winds at 850 hPa (a) for all northern hemisphere

summer months (June-July-August) and (b) for all northern hemisphere winter

months (December-January-February). . . . . . . . . . . . . . . . . . . . . . . . 69

4.7 The ’external’ variance of zonal winds at 850 hPa (m2s−2) during (a) NH sum-

mer months (JJA) and (b) NH winter months (DJF). . . . . . . . . . . . . . . . . 70

4.8 The ’internal’ variance of zonal winds at 850 hPa (m2s−2) during (a) NH summer

months (JJA) and (b) NH winter months (DJF). . . . . . . . . . . . . . . . . . . 70

4.9 Estimates of ’F’ ratios for zonal winds at 200 hPa (a) for all northern hemisphere

summer months (JJA) and (b) for all northern hemisphere winter months (DJF). . 72

4.10 Estimates of ’F’ ratios for OLR (a) for all northern hemisphere summer months

(JJA) and (b) for all northern hemisphere winter months (DJF). . . . . . . . . . . 72

Page 19: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures xiv

4.11 Estimates of ’F’ ratios for geopotential height at 700 hPa (a) for all northern hemi-

sphere summer months (JJA) and (b) for all northern hemisphere winter months

(DJF). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

4.12 The ’external’ variance of geopotential height at 700 hPa (gpm2) during (a) NH

summer months (JJA) and (b) NH winter months (DJF). . . . . . . . . . . . . . . 74

4.13 The ’internal’ variance of geopotential height at 700 hPa (gpm2) during (a) NH

summer months (JJA) and (b) NH winter months (DJF). . . . . . . . . . . . . . . 74

4.14 The ’internal’ variance of (a) zonal winds at 850 hPa (m2s−2) and (b) OLR (Wm−2)2

based on all months after removing the higher frequencies with period shorter

than 10 days. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.15 Estimates of ’F’ ratios for zonal winds at 850 hPa for (a) NH summer season

(JJA) (b) NH winter season (DJF). . . . . . . . . . . . . . . . . . . . . . . . . . . 77

4.16 Estimates of ’climate noise’ for zonal winds at 850 hPa for (a) NH summer sea-

son (JJA) (b) NH winter season (DJF). . . . . . . . . . . . . . . . . . . . . . . . 78

4.17 Estimates of ’F’ ratios for zonal winds at 200 hPa for (a) NH summer season

(JJA) (b) NH winter season (DJF). . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.18 Estimates of ’F’ ratios for OLR for (a) NH summer season (JJA) (b) NH

winter season (DJF). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.19 Estimates of ’F’ ratios for geopotential height at 700 hPa for (a) NH summer

season (JJA) (b) NH winter season (DJF). . . . . . . . . . . . . . . . . . . . . . . 81

4.20 Estimates of ’climate noise’ for geopotential height at 700 hPa for (a) NH sum-

mer season (JJA) (b) NH winter season (DJF). . . . . . . . . . . . . . . . . . . . 81

5.1 Genesis dates of LPS between 1 June and 30 September of all years during 1979

to 1993 over the monsoon trough as a function of normalized departure of pre-

cipitation over the trough from the seasonal mean. . . . . . . . . . . . . . . . . 89

5.2 Leading Empirical Orthogonal Functions ( (a) EOF1 & (b) EOF2) of 10-80 day

filtered wind anomalies (ms−1) at 850 hPa between June 1 and September 30 for

the period 1964-1973. (c) Normalized time series of PC1 and PC2 for ten years

(each year has 122 days). (d) Normalized Monsoon Intraseasonal Oscillation

Index (MISI) for 10 years. Periods of MISI > +1 (MISI< -1) correspond to active

(break) phases of the monsoon. It may be noted that positive (negative) phase

of MISI represents enhancement (weakening) of the EOF1 pattern. . . . . . . . . 90

Page 20: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

List of Figures xv

5.3 Histogram of genesis of synoptic events (lows & depressions) for the Indian

monsoon region (50◦E-100◦E, Eq-30◦N) during June to September for the period

1954-1993 as a function of normalized MISI. . . . . . . . . . . . . . . . . . . . . 92

5.4 Total (climatology+composite anomaly) relative vorticity (10−6s−1) at 850 hPa

during the (a) ’Active’ ISO phase (MISI > +1) and (b) ’Break’ ISO phase (MISI <

-1). Dark dots indicate the position of the genesis of the LPS during active and

break phases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

5.5 Composites based on active and break days as defined by the ISO index, MISI.

(a) ’Active’ minus ’Break’ composite wind anomalies (ms−1) and associated rel-

ative vorticity (10−6s−1) at 850 hPa during the 40 year period (1954-1993). Only

vectors significant at 95% confidence level are displayed. Positive contours are

shaded and negative contours are not shown. (b) ’Active’ minus ’Break’ com-

posite precipitation anomalies (mm.day−1) during 1979-1993. . . . . . . . . . . 94

Page 21: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 1

Introduction

The seasonal mean summer monsoon precipitation over the Indian continent and

neighbouring region is the lifeline of the agrarian economy of the region. A weak or late

monsoon can have disastrous consequences on productivity of the crops upon which

millions of people rely on for their sustenance [Swaminathan, 1987]. The monsoons,

which returns with remarkable regularity each summer provides rainfall needed for

60% of World’s population. The importance of Asian summer monsoon in the global cir-

culation and climate predictability is widely recognized. Monsoon displays substantial

interannual variability, which has profound socio-economic consequences. Prediction

of seasonal mean monsoon precipitation, therefore assumes great importance.

Statistical prediction of some gross features of the Indian summer monsoon (e.g. all

India monsoon rainfall, IMR) have been moderately successful during the decade of

80’s [Krishakumar et al., 1995], but have failed during the decade of 90’s. This is due

to the fact that correlation between IMR and many predictors undergo low frequency

decadal variation. The dynamical seasonal prediction of the Indian summer monsoon

has thus far remained elusive [Brankovic and Palmer, 2000]. However, dynamical sea-

sonal prediction of monsoons could be beneficial due to a variety of reasons.

• Ensembles of forecast can potentially yield information on probability of ’strong’

and ’weak’ monsoon.

• It can give more accurate information on regionality of anomalous rainfall and

circulation compared to statistical methods.

Hence, it is important to identify and understand the factors that may be limiting our

current level of predictability.

The predictability of the tropical climate (specially Indian summer monsoon), de-

pends on the relative contribution of ’external’ slowly varying boundary forcing and

Page 22: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 2

’internal’ dynamics (intraseasonal oscillations) to the interannual variability [Charney

and Shukla, 1981]. Following the seminal work of Charney and Shukla [1981] and Shukla

[1981], during the past two decades, it was shown that climate in large part of the tropics

is primarily determined by slowly varying sea surface temperature (SST) where poten-

tial for making dynamical forecasts several seasons in advance exists (e.g. Latif et al.

[1998]). However, during the same period it has been also recognized that there are

regions within the tropics, climate of which is not strongly governed by the anomalous

boundary conditions. The Indian summer monsoon is such a system [Brankovic and

Palmer, 1997; Webster et al., 1998; Goswami, 1998].

What limits the simulation and predictability of the Indian summer monsoon? Re-

search during that past decade has identified two possible explanations. The first is

that model errors in the mean monsoon simulations are still substantial enough that the

signal being sought is smaller than the systematic bias. Charney and Shukla [1981] sug-

gested that low frequency boundary forcing (e.g. sea surface temperature) predisposes

the monsoon system towards a dry or wet state. In other words anomalous bound-

ary conditions may provide potential predictability. If this is true, model simulations

should be able to capture interannual variability of the Indian summer monsoon and

hence could produce fairly good forecasts. But in reality, this is not the case as most

models find the simulation of mean monsoon precipitation extremely difficult and have

even greater difficulty in simulating the interannual variability of the Indian summer

monsoon rainfall [Sperber and Palmer, 1996; Gadgil and Sajani, 1998; Goswami, 1998]. If

the ’external’ slowly varying boundary forcing (e.g. sea surface temperature, soil mois-

ture etc) determine the predictability of monsoons, there is a clear need to improve the

model simulations, before any conclusive statements could be made about the dynam-

ical seasonal predictability of the Indian summer monsoon.

The second explanation involves the role of intraseasonal variability and the sugges-

tion that it introduces a chaotic element into the prediction of seasonal mean anomalies.

During the established phase of the monsoon, circulation pattern undergoes significant

variations associated with a pronounced northward excursion of the tropical conver-

gence zone (TCZ) which brings the monsoon intermittently from an ’active’ into an

inactive (’break’) phase over the continent. The change in precipitation distribution be-

tween ’active’ and ’break’ phases of monsoon is substantial [Webster et al., 1998] and it is

therefore quite possible that intraseasonal variability could have a significant influence

on the seasonal mean monsoon precipitation.

Page 23: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 3

The Indian summer monsoon has vigorous intraseasonal oscillations in the form of

’active’ and weak (or ’break’) spells of monsoon rainfall within the summer monsoon

season [Ramamurthy, 1969]. These ’active’ and ’break’ spells of the monsoon are associ-

ated with fluctuations of the tropical convergence zone (TCZ) [Yasunari, 1979, 1980, 1981;

Sikka and Gadgil, 1980]. The TCZ over the Indian monsoon region represents the as-

cending branch of the regional Hadley circulation. Intraseasonal oscillations (ISOs) are

essentially manifestation of fluctuations of the regional Hadley circulation. These fluc-

tuations initially seen in Indian station data [Keshavamurthy, 1973; Dakshinamurthy and

Keshavamurthy, 1976] were later shown to be related to coherent fluctuations of the re-

gional Hadley circulation [Krishnamurti and Subrahmanyam, 1982; Murakami et al., 1984;

Mehta and Krishnamurti, 1988; Hartmann and Michelson, 1989]. The ISOs of the Indian

summer monsoon have two preferred bands of periods [Krishnamurti and Bhalme, 1976;

Krishnamurti and Ardunay, 1980; Yasunari, 1980]. One band has periods between 10 and

20 days while the other band contains periods between 30 and 60 days. The 30-60 day

mode has a northward and eastward propagation over the monsoon region while the

10-20 day mode has a clear westward propagation and a weak northward propagation.

The seasonal summer mean monsoon precipitation (and associated circulation) is

a result of the shift of the seasonal mean position of the TCZ to about 25◦N during

boreal summer from a mean position south of the equator during boreal winter. The

seasonal summer mean (June-September, JJAS) precipitation distribution has a major

zone of large precipitation along the monsoon trough extending to the north Bay-of-

Bengal. There is also a secondary zone of precipitation maximum south of the equator

(between 0◦ and 10◦S) over the warm waters of the Indian Ocean. These two maxima

in seasonal mean precipitation represent two favored locations of the TCZ during the

summer monsoon season [Sikka and Gadgil, 1980; Goswami, 1994]. The ISOs are fluctua-

tions of the TCZ between the two favored locations within the monsoon season. In the

intraseasonal time scales, the TCZ form repeatedly over the ocean and moves north-

ward, persists for a while over the monsoon trough before decaying and regenerating

over the ocean. The tendency of the TCZ to persist over the monsoon trough results in

larger residence time over the continent leading to larger seasonal mean precipitation

over the land and a weaker one over the ocean. Therefore, there is a possibility that the

statistics of the ISOs influence the seasonal mean monsoon. If the ISOs indeed influence

the seasonal mean significantly, the part of the seasonal mean governed by the ISOs

would be unpredictable as ISOs are basically governed by internal dynamics [Webster,

Page 24: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 4

1983; Goswami and Shukla, 1984; Keshavamurthy et al., 1986] and are chaotic in nature. If

the ISOs do contribute significantly to the seasonal mean, the interannual variability of

the seasonal mean monsoon is expected to have a significant component arising from

internal dynamics. Several recent modeling studies show that indeed a significant frac-

tion of the interannual variability of the Indian summer monsoon may be governed by

internal dynamics [Harzallah and Sadourny, 1995; Rowell et al., 1995; Stern and Miyakoda,

1995; Goswami, 1998]. Most of the studies do not provide any insight regarding the ori-

gin of the internally generated interannual variability. Based on a series of sensitivity

studies with a GCM and a dynamical system model, Goswami [1997] indicates that the

modulation of the energetic intraseasonal oscillations by the annual cycle could give

rise to an internal quasi-biennial oscillation in the tropical atmosphere. These argu-

ments and the modeling studies set the stage to ask the question: Do the ISOs really

influence the seasonal mean monsoon? If so, how and to what extent?

Unfortunately, how and to what extent the ISOs influence the seasonal mean circu-

lation and precipitation has not been clearly established from observations. Not many

studies have actually addressed this question. Mehta and Krishnamurti [1988] examined

the interannual variability of the 30-50 day mode in the winds at 850 hPa and 200 hPa

for the period 1980 to 1984 using European Center for Medium Range Weather Forecasts

(ECMWF) operational analysis. They mainly examined the variations in the northward

propagation characteristics and did not attempt to relate these to the seasonal mean.

Hartmann and Michelson [1989] used 70 year (1901-1970) record of daily precipitation for

3700 stations distributed over whole India and created annual cycle of daily precipita-

tion at 1◦ × 1◦ blocks. They find statistically significant peak in the daily precipitation

around 40-50 day period over most of India south of 23◦N and the oscillations have a

northward propagation. This study did not address the interannual variability of the

ISOs. Singh and Kriplani [1990] and Singh et al. [1992] used long records of daily rainfall

data over the Indian continent and examined the 30-50 day oscillation. They found that

ISOs has largest amplitude over the western central India around 20◦N and can explain

upto 25% of 5-day averaged rainfall. They concluded that these oscillations have large

interannual variability in intensity and period and does not seem to be related with

overall performance of monsoon or phases of ENSO. They, however, could not come to

a clear conclusion regarding relationship between the ISOs and the interannual variabil-

ity of the Indian monsoon rainfall. Rao et al. [1990] used daily IR data from INSAT-1B for

the monsoon period of 1986 and 1987. After creating daily averages from three hourly

Page 25: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 5

observations, they calculated fractional cloud cover in each 2.5◦×2.5◦ boxes from the IR

brightness temperature. The fractional cloudiness shows a periodicity of 30-50 days in

both years up to 20◦N. Based on only two years, it was not possible to conclude much on

the interannual variability of this mode. Ahlquist et al. [1990] studied radiosonde obser-

vations at 12 Indian stations between 1951 and 1978. They examined ISOs with period

longer than 10 days but did not try to relate the ISOs with the interannual variability

of the monsoon. De and Natu [1994] examined upper wind data for six radiosonde sta-

tions at 850, 700, 500 and 300 hPa levels for the years 1979 to 1984 and 1987. They also

find that 30-50 day mode has considerable interannual variability and that the mode

becomes more significant during normal and excess rainfall years. Here, the sample

size is small to arrive at a robust result regarding interannual variability. Kondragunta

[1990] used daily OLR for the summer period form NOAA polar orbiting satellite for

eight years (1975 to 1983) and studied the interannual variability of the ISO over the

whole Asian region. He finds that intraseasonal oscillations occur on three time scales,

30-60 day, 10-20 day and less than 10 days. Fennessy and Shukla [1994] using GCM sim-

ulations of the Indian monsoon for 1988 and 1987 showed that the spatial structures of

the interannual variability of the seasonal mean and that of the intraseasonal variability

in their model simulations were quite similar. Ferranti et al. [1997] studied the relation-

ship between intraseasonal and interannual variability over the monsoon region using

data from five 10-year simulations of the ECMWF GCM differing only in their initial

conditions. They examined simulated precipitation and 850 hPa relative vorticity in

detail and showed that monsoon fluctuations within a season and between different

years have a common mode of variability with a bi-modal meridional structure in the

precipitation. While the common mode of variability is qualitatively consistent with

fluctuations of the TCZ in the two favored locations, their results suffer from some sys-

tematic errors inherent in the ECMWF GCM simulation of the Indian summer monsoon.

The model underestimates precipitation over the north Bay of Bengal and the monsoon

trough zone. This systematic error reflects in their interannual mode having appreciable

amplitude only east of 80◦E both in precipitation and low-level vorticity. Webster et al.

[1998] discusses mean circulation pattern at 850 hPa associated with ’active’ and ’break’

conditions based on ECMWF operational analysis for 14-year (1980-1993) and brings

out the large scale nature of these circulation anomalies. No attempt to relate these pat-

terns to the seasonal mean was, however, made. In another recent study Goswami et al.

[1998] studied daily surface winds from National Centers for Environmental Prediction

Page 26: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 6

(NCEP)/National Center for Atmospheric Research (NCAR) reanalysis for ten years

(1987-1996) and showed that the spatial structures of the intraseasonal mode and that

of the dominant interannual mode are strikingly similar.

Annamalai et al. [1999] examined the relationship between the intraseasonal oscilla-

tions and interannual variability using NCEP/NCAR reanalysis and ECMWF reanal-

ysis (ERA) for the period 1979-95. While the primary objective of the study was to

compare NCEP/NCAR reanalysis and ERA, they also identified a dominant mode of

intraseasonal variability which captures the active/break cycles of the monsoon. How-

ever, they could not clearly identify a common dominant mode that described intrasea-

sonal and interannual variability. They have tried to find the relationship between in-

terannual and intraseasonal variability by using a one dimensional probability density

function (PDF) of the principal component of the dominant ISO mode. Clear differ-

ence in probability of occurrence of ’active’/’break’ phases in two contrasting years

also could not be identified. Statistical significance of results could not be ascertained

due to small sample size.

Sperber et al. [2000] investigated the relationship between the relationship between

subseasonal and interannual variability of the Asian summer monsoon using 40 years

of NCEP/NCAR Reanalysis. They have confirmed that a common mode of variabil-

ity exists on subseasonal and interannual time scales. PDF of principal components

did not show any bimodality. Further they have shown that PDF is systematically and

significantly perturbed towards negative (positive) values in weak (strong) monsoon

years. However, they also mention that only a subset of subseasonal modes are sys-

tematically perturbed either by ENSO or in weak/strong monsoon years, suggesting

that predictability is likely to be limited by the chaotic, internal variability of the mon-

soon system. The PDF of the subseasonal modes are biased towards positive (negative)

side during strong (weak) years only if the low frequency interannual variations of the

seasonal mean are not removed.

Krishnamurthy and Shukla [2000] has used gridded rainfall dataset (1901-1970) to ana-

lyze the intraseasonal and interannual variability of the summer monsoon rainfall over

India. They have found that the nature of intraseasonal variability is not different dur-

ing the years of major droughts or major floods. They have also found that there is con-

siderable variability in the spatial patterns of the rainfall anomalies over India on both

daily and seasonal time scales. Their results indicate that the dominant mode (leading

EOF) of the daily rainfall anomalies has a spatial pattern different from the dominant

Page 27: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1 Introduction 7

mode of seasonal anomalies. The variances of the daily rainfall anomalies over India

are about 50-100 times larger than those of the seasonal rainfall anomalies. To relate

intraseasonal and interannual variability they have used a correlation analysis between

the daily and the seasonal anomalies and have found that there is a signature of the sea-

sonal anomaly pattern throughout the monsoon season. The frequency distributions of

the correlations involving daily anomalies that include the seasonal anomalies clearly

show a bias toward positive correlations and do not reveal any bimodality. The fre-

quency distribution do not show any bias, if the low frequency interannual variations

of the seasonal mean are removed.

Recently, Lawrence and Webster [2001] have examined the interannual variations of

the ISO using out-going long wave radiation (OLR) data for the period 1975-1997. By

developing an index representing seasonally averaged ISO activity, they have found

that summertime ISO activity exhibits an inverse relationship with Indian monsoon

strength. They concluded that the ISO activity is uncorrelated with any other leading

SST variability including the ENSO.

A conceptual model of how the ISOs influence the seasonal mean and interannual

variability of the Indian monsoon was proposed by Goswami [1994]. The conceptual

model is based on the similarity between the spatial structure of the dominant ISO mode

and that of the interannual variability. The seasonal summer mean (June-September,

JJAS) precipitation distribution has a major zone of large precipitation along the mon-

soon trough extending to the north Bay-of-Bengal (see Figure 2.1(d)) and a secondary

zone of precipitation maximum south of the equator (between 0◦ and 10◦S) over the

warm waters of the Indian Ocean. These two maxima in the seasonal mean precipitation

represent two favored locations of the TCZ during the summer monsoon season [Sikka

and Gadgil, 1980; Goswami, 1994]. The ISOs are fluctuations of the TCZ between the two

locations and repeated propagation from the southern to the northern position within

the monsoon season. During a typical ’active’ condition, the northern TCZ is stronger

and the southern one is weaker with stronger cyclonic vorticity and enhanced convec-

tion over the northern location with stronger anticyclonic vorticity and decreased con-

vection over the southern one. The situation reverses during a typical ’break’ condition.

Higher probability of occurrence of ’active’ like (’break’ like) conditions during a mon-

soon season could, therefore give rise to stronger (weaker) than normal seasonal mean

monsoon and precipitation. It may be noted that the ISOs are not purely sinusoidal

oscillations. Due to the broadband nature of their spectrum, the intensity as well as the

Page 28: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1.1 Objectives 8

duration of the ’active’ phases in a season could be different from those of the ’break’

phases. Moreover, the number of ’active’ and ’break’ spells within a monsoon season

(June 1 - September 30) may be different depending on the initial phase. These fac-

tors may lead to asymmetry in the probability density functions (PDF). Our conceptual

model is similar, to the one proposed by Palmer [1994]. However, in contrast to Palmer

[1994] who proposes that the asymmetry in the PDF is forced only by external forcing,

we claim that the asymmetry could arise even without external forcing.

1.1 Objectives

The background presented above lead us to the following conclusions. Low-frequency

(LF) large amplitude ISO with period around 10-20 days and 30-60 days are integral

part of Indian summer monsoon. Therefore, the phase, amplitude and period of these

ISOs can influence the seasonal mean monsoon. ISOs are driven by internal dynamics,

involving primarily feed back between organized convection and dynamics. A concep-

tual picture of the variability of these LF oscillations envisages competition between the

continental TCZ and oceanic TCZ.

While the considerations presented above are all plausible, there has been no reliable

quantitative estimate of how and to what extent ISO influence the seasonal mean and

it’s variability. As a result, the present study is undertaken with the following specific

objectives.

• The primary objective of the present study is to use sufficiently long daily obser-

vation to bring out how and to what extent the ISOs of the Indian monsoon affect

the seasonal mean and its interannual variability. The conceptual model proposed

above is used as the working hypothesis. The primary objective may be achieved

in two parts. Firstly, we bring out the underlying common spatial structure of

the dominant ISO in all years and compare it with the spatial structure of the sea-

sonal mean and interannual variability of the Indian monsoon. Secondly, attempt

is made to relate probability of occurrence of the ISO pattern to the interannual

variability of the seasonal mean. To achieve this goal a homogeneous data set for

a long enough period is essential so that the statistics of the ISOs and that of the

interannual variability of the seasonal mean could be reliably estimated. Many

earlier studies used data only for a short period as a result of which the interan-

nual variability of the circulation could not be reliably estimated. (Chapter 2 and

Page 29: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1.2 Datasets 9

Chapter 3)

• Having shown that the ISOs can influence the seasonal mean and its interannual

variability, the next objective is to make quantitative estimates of predictability of

the monsoon climate. A measure of potential predictability of the monthly and

seasonal means at a place could be obtained from the ratio of variances associated

with the external to the internal components. Using long homogeneous data sets,

attempts will be made to estimate the ’internal’ variability of monthly and sea-

sonal climate. The potential predictability of the Indian monsoon region will be

compared with that of other regions in the tropics. (Chapter4)

• The ISOs of the monsoon lead to strengthening (weakening) of the seasonal mean

monsoon in their active (break) phase. While this fact results in interannual vari-

ations of the seasonal mean monsoon at one end of the spectrum, it may modu-

late the statistics of the monsoon synoptic disturbances at another end. The main

rain bearing system during the monsoon season are Low Pressure Systems (LPS)

consisting of lows and depressions. Since the genesis of the LPS depends on the

horizontal shear and low-level vorticity, it is possible that more LPS may form in

active phase relative to the break phase. In other words, large scale circulation

associated with the ISOs could modulate the frequency of genesis of LPS. There-

fore one objective of our study will be to investigate how the synoptic events are

modulated by the ISOs. (Chapter 5)

1.2 Datasets

A brief description of the different datasets used in the study is provided below.

1.2.1 Reanalysis Data

The National Centers for Environmental Prediction/National Center for Atmospheric

Research (NCEP/NCAR) 40-year Reanalysis data is a research quality data set suitable

for weather and short-term climate research. The NCEP/NCAR Reanalysis Project uses

a Global Data Assimilation System (GDAS), along with the observations from 1957 to

the present to produce global meteorological fields through dynamically and thermo-

dynamically consistent interpolations to support the needs of the climate research com-

munity [Kalnay et al., 1996]. The project began in 1991 and involves the recovery and

quality control of historical land surface, ship, rawinsonde, aircraft, pibal, satellite and

Page 30: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1.2 Datasets 10

other data. These data are then assimilated with a GDAS that is kept unchanged over

the reanalysis period 1957-96, to avoid spurious climate jumps or trends. The project

uses a frozen state-of-the-art global data assimilation system and a data base as com-

plete as possible. The model used here is a T62 model (equivalent to a horizontal reso-

lution of 210 km) with 28 vertical levels. Thus, any output variable in the reanalysis is

a blend of observations and model. The fidelity of any variable to reality depends on

the accuracy and density of observations as well as on the performance of the analysis

scheme itself. The reliability of the parameters have been increased with the addition

of delayed observations, provided by different countries and organizations. Output

variables are classified into four categories; A, B, C and D; depending on the relative

influence of the observations and/or the model. A variable belongs to category ‘A’ (e.g.

wind, upper-air temperature) if it is strongly influenced by observed data and, hence,

reliable. The designation ‘B’ (e.g. SST, air temperature at 2 m, specific humidity at 2

m, relative humidity) indicates that although the variable is directly affected by obser-

vational data, the model also has a strong influence on it. A category ‘C’ variable (e.g.

cloud, precipitation, latent heat flux, sensible heat flux) indicates that there are no ob-

servations directly affecting the variable, so it is solely derived from the model. The

letter ‘D’ (e.g. ice concentration, plant resistance, land sea mask) represents a field that

is fixed from climatological values and does not depend on model.

This study uses daily averaged zonal (u) and meridional components (v) of reanaly-

sis winds at various pressure levels (surface, 850 hPa, 500 hPa and 200 hPa and geopo-

tential height at 700 hPa). To accurately estimate interannual variability, we have also

used monthly mean winds at the same levels for a 42-year period (1956-1997).

1.2.2 NOAA Outgoing Long wave Radiation (OLR) Dataset

One proxy for tropical rainfall measurements or deep tropical convection is the OLR

data at the top of the atmosphere. Twenty four years (1974-1997) years of daily averaged

OLR estimates from NOAA satellites [Salby et al., 1991; Gruber and Krueger, 1984] were

used in this study. Data gap periods (January 1974 to May 1974 and March 1978 to De-

cember 1978) were avoided. Missing values were owing to satellite problems, archival

problems or incomplete global coverage.These satellites uses AVHRR (Advanced Very

High Resolution Radiometer) which has unique characteristics of spectral response, im-

age geometry, frequency of spectral coverage and accessibility that makes it useful for

applications in oceanography and meteorology. OLR data are available in (2.5◦ × 2.5◦)

Page 31: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1.2 Datasets 11

latitude-longitude grid boxes.

Daily interpolated OLR data set (2.5◦ × 2.5◦) for the same period (1974-1997) were

also used for this study. This data set is taken from NOAA-CIRES Climate Diagnostics

Center (CDC), Boulder, USA, from their website at http://www.cdc.noaa.gov/. Data

gaps were filled with temporal and spatial interpolations; details of the interpolation

technique can be found in Liebmann and Smith [1996].

1.2.3 Precipitation Datasets

Monthly, pentad and daily precipitation datasets were used to substantiate some of

the results presented in this study. For monthly precipitation, Climate Prediction Center

Merged Analysis of Precipitation (CMAP) data [Xie and Arkin, 1997] were used. CMAP

is a gridded global monthly precipitation constructed on a 2.5◦ latitude-longitude grid

for a 17-year period from 1979 to 1995 by merging several kinds of information sources

with different characteristics. The sources include guage-observations, estimates of pre-

cipitation inferred from a variety of satellite observations. Satellite estimates come from

infrared as well as microwave sensors. First, the satellite estimates are combined using a

weighted average where weights are proportional to the estimated errors of the various

estimates. The weighted average is then merged with the guage observations. CMAP

provides very useful information for climate analysis and can be used to investigate

annual and interannual variability in large scale precipitation.

CMAP pentad data for a 15-year period from 1979-1993 were also used. The pentad

CMAP essentially uses the same algorithm and input data as monthly CMAP [Xie and

Arkin, 1997]. The version we use is based on a blend of guage data with satellite prod-

ucts including the GOES (Geostationary Operational Environmental Satellite) precipitation

index based on geostationary infrared data, Microwave Sounding Unit, OLR based pre-

cipitation index, SSM/I (Special Sensor Microwave/Imager) scattering and SSM/I emis-

sion. A detailed description of the pentad dataset is in preparation.

Daily gridded rainfall data over the Indian continent for a 12-year period from 1978

to 1989 was also utilized. The daily rainfall data was originally compiled by Singh et al.

[1992] at 2.5◦ latitude-longitude grids based on daily rainfall at 365 stations uniformly

distributed over the country. The original data reported in Singh et al. [1992] were later

extended to 1989. The version we use was regridded to 1.25◦× 1.25◦ latitude-longitude

boxes by Mike Fennessy of the Center for Ocean-Land-Atmosphere-Studies (COLA,

1999, personal communication).

Page 32: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

1.2 Datasets 12

1.2.4 Statistics of Low Pressure Systems

The dates and location of genesis of all lows, depressions and cyclones during April-

November for a 40-year period from 1954 to 1993 over the Indian monsoon region (50◦E-

100◦E, Eq-35◦N) were collected from various sources. For the Indian monsoon region,

data were based on reports from the India Meteorological Department (IMD). Data for

the first 30 years were taken from Mooley and Shukla’s [Mooley and Shukla, 1987, 1989]

compilation based on IMD’s daily weather reports. Data for the next 10 years (1984-

1993) were compiled from seasonal weather summaries published by IMD in Mausam.

For example, data for 1984 monsoon season can be found from [IMD, 1985].

For categorizing ’strong’ and ’weak’ monsoons we have used the All India Mon-

soon Rainfall index [Parthasarathy et al., 1994]. The IMR is constructed from a weighted

average of 306 stations spread over the whole of the Indian subcontinent.

Page 33: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 2

Basic Characteristics of MonsoonIntraseasonal Oscillations

In this chapter, the basic characteristics of intraseasonal oscillations of the Indian

summer monsoon are examined. The characteristics of monsoon ISOs such as their

horizontal and vertical structures and meridional and zonal propagation characteristics

have been previously studied extensively (see references cited in the Introduction). Our

objective here is not to repeat the results of the earlier studies. However, earlier studies

used limited number of years. As a result, it is not well established whether differ-

ent phases of the dominant ISO mode possess spatial patterns that are common to all

events. Our aim here is to bring out the underlying mean feature of the dominant ISO

mode that is invariant over the years. We have examined these characteristics of both

ISO modes for each year of the 20-year period (1978-1997). The general characteristics

of the 30-60 day mode and 10-20 day are consistent with most of the earlier studies.

Therefore, some of the important features are only briefly summarized here.

2.1 Methodology

The climatological summer mean (June-September, JJAS) circulation at lower and

upper atmosphere and associated mean low-level vorticity are shown in Figure 2.1(a,b,c).

The seasonal mean precipitation is shown in Figure 2.1(d). It may be noted that the neg-

ative mean vorticity between the equator and 10◦S (Figure 2.1(b)) represents a region

of cyclonic vorticity in the Southern Hemisphere and is coincident with the seasonal

precipitation maximum. The circulation, convection and precipitation in the monsoon

region are characterized by a strong seasonal cycle. An example of zonal winds at 850

hPa at a few selected points for 1990 is shown in Figure 2.2. The annual cycle is defined

by the sum of the annual and semi-annual harmonics (green solid lines in Figure 2.2).

Page 34: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.1 Methodology 14

Figure 2.1: Climatological mean (JJAS) monsoon winds (ms−1) and precipitation (mm.day−1).(a) 850 hPa vector winds, (b) Relative vorticity at 850 hPa (10−6s−1), (c) 200 hPa vector winds,(d) Precipitation from Xie and Arkin [1997].

The daily anomalies after removing the annual cycle are shown in the right panel. The

annual cycle, which is essentially driven by external conditions, has year to year varia-

tions that manifest in the interannual variations of the seasonal mean. In many studies,

daily anomalies are constructed by removing the climatological mean for each day from

the daily observations. In a particular year, the annual cycle may be significantly differ-

ent from the climatological mean annual cycle. This would introduce an additional bias

in the daily anomalies during the monsoon season. This bias can give rise to asymmetry

in the PDF of the ISOs that may not be intrinsic to the ISOs but may be related to the

external forcing changes. Since we are interested in the role of intraseasonal oscillations

in modifying the summer mean, we would like to avoid aliasing of any statistics of the

ISOs due to possible year to year variation of the annual cycle itself. This is achieved by

calculating the annual cycle for each year based on the data for that year alone and by

calculating the daily anomalies after removing the annual cycle of each year.

The intraseasonal oscillations are identified by estimating the spectra of zonal and

Page 35: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.1 Methodology 15

Figure 2.2: Some examples of raw time series of zonal winds at 850 hPa at a few selected pointsduring 1990. (Left panels) Daily zonal winds (ms−1) with the annual cycle (annual and semi-annual harmonics, green lines). (Right panels) Anomalous daily zonal winds (ms−1).

0 50 100 1500

0.5

1

1.5

2

Period (days)

Pow

er *

Fre

quen

cy Zonal wind

0 50 100 1500

50

100

150

Period (days)

Pow

er *

Fre

quen

cy OLR

Figure 2.3: Examples of spectra of zonal winds and OLR for a typical year (1984) at a typicalpoint (90◦E, 10◦N).

meridional winds as well as OLR anomalies. Power spectra are calculated from anomaly

time series between May 1 and October 31 (184 days) using Tukey lag window method

[Chatfield, 1980]. An example of spectra for zonal winds and OLR at a point in north

Bay-of-Bengal for 1984 is shown in Figure 2.3. This example shows two strong peaks,

one with period around 36 days and the other with period around 16 days. Similar,

power spectral estimates are made for each year and at all latitudes between 30◦S and

30◦N along a number of longitudes (e.g. 70◦E, 80◦E, 90◦E). From these estimates, the

most prevalent dominant periods are chosen. It is noted that the dominant periods

Page 36: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.2 Propagation Characteristics 16

Year Mode I Mode II1978 25 121979 42 171980 33 131981 42 161982 42 141983 25 141984 36 161985 33 121986 42 121987 30 101988 42 201989 40 141990 42 161991 42 201992 34 141993 32 141994 30 121995 42 201996 32 121997 42 12

Table 2.1: Period in days corresponding to the two peaks in the spectra for differentyears

found from the winds agree well with those found from OLR. The dominant periods

found in each year of the 20-year period (1978-1997) are listed in Table-I. The domi-

nant periodicity in each of the two bands show considerable variation from one year

to another. To study the detailed structure and characteristics of the two intraseasonal

oscillations Butterworth band pass filter [Murakami, 1979] with peak response around

the dominant periods are used.

2.2 Propagation Characteristics

The 30-60 day mode has a large horizontal scale (half wavelength of 70◦-80◦ lon-

gitude) as seen from the point-correlation map of the 30-60 day filtered zonal winds

with respect to those at a reference point (85◦E, 10◦N; Figure 2.4(a)). The mode has a

first baroclinic vertical structure close to the equator and over the Indian monsoon re-

gion as seen from correlations between 30-60 day filtered zonal winds at 850 hPa and

200 hPa (Figure 2.4(b)). The horizontal scale and vertical structure of the mode shown

in the example (Figure 2.4) is representative of other years. The 30-60 day mode is

Page 37: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.2 Propagation Characteristics 17

Figure 2.4: An example illustrating the horizontal scale and vertical structure of the dominantISO mode. (a) Lag-zero correlations of the 850 hPa 30-60 day filtered zonal winds with respectto a reference point (85◦E, 10◦N). (b) Lag-zero correlations between 30-60 day filtered zonalwinds at 850 hPa and 200 hPa at each grid point. Correlations are calculated between May 1 andOctober 31 of 1990. Correlations exceeding 0.2 are significant at 95% confidence level.

known to have a northward and eastward propagation in the Indian monsoon region

[Yasunari, 1979, 1980]. This is demonstrated in Figure 2.5, where lag correlation of the

30-60 day filtered zonal winds at 850 hPa and OLR with respect to the same fields at

a reference point (85◦E, 10◦N) for the year 1990 averaged over a longitude belt (80◦E-

90◦E) are plotted as a function of latitude. Northward propagation north of the equator

and a tendency for southward propagation south of the equator are seen in zonal winds

as well as in OLR. Webster et al. [1998] also referred to such northward propagation in

the northern hemisphere and southward propagation in the southern hemisphere of the

dominant ISO. The correlation represents an average picture of the 3-4 episodes of the

oscillation during the summer monsoon season. On individual episodes, it is seen that

not all episodes are associated with a clean northward propagation (not shown). The

character of the northward propagation (e.g. speed of propagation) also vary from one

year to another. On the average the 30-60 day mode has an eastward propagation in the

monsoon region (50◦E-110◦E). As in the case of northward propagation, the eastward

propagation may not be clear in each individual episode and has considerable year to

Page 38: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.2 Propagation Characteristics 18

Figure 2.5: (a) Correlations between 30-60 day filtered zonal winds at 850 hPa with respect tothat at a reference point (85◦E, 10◦N) at different lead/lags averaged over (80◦E-90◦E) for 1990.(b) Same as (a) but for 30-60 day filtered OLR. Starting contour is ±0.1 and contour interval is0.2.

Figure 2.6: (a) Correlations between 30-60 day filtered zonal winds at 850 hPa with respect tothat at a reference point (85◦E, 10◦N) at different lead/lags averaged over (10◦N-20◦N) for 1990.(b) Same as (a) but for 30-60 day filtered OLR. Starting contour is ±0.1 and contour interval is0.2.

Page 39: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.3 A Circulation Criterion for ’Active’ and ’Break’ Phases 19

year variability (Figure 2.6). The 10-20 day mode on the average has clear westward

propagation in the monsoon region. It is either stationary or northward propagating in

the meridional direction (figure not shown).

2.3 A Circulation Criterion for ’Active’ and ’Break’ Phases

Traditionally ’active’ and ’break’ monsoon conditions are defined based on a precip-

itation criterion [Ramamurthy, 1969]. Here, we propose a criterion to define ’active’ and

’break’ monsoon conditions based on a circulation index. Such a circulation based def-

inition of ’active’ and ’break’ monsoon may be useful for various purposes. During an

’active’ phase of the Indian monsoon, typically there is more precipitation over central

India and a stronger monsoon trough [Ramamurthy, 1969]. As a result we may expect

westerly zonal winds south of the monsoon trough to strengthen. Opposite is expected

during a ’break’ phase. With this consideration in mind, we propose a circulation based

definition of ’active’ and ’break’ monsoon conditions. A reference point just south of the

monsoon trough at (90◦E, 15◦N) is selected for this purpose and the 30-60 day filtered

zonal winds at 850 hPa are plotted (Figure 2.7(a)). The days for which the filtered zonal

winds at 850 hPa is greater than +1 standard deviation (as shown by the thin solid line,

i.e. stronger westerly anomalies) are considered as ’active’ days, while those for which

it is less than -1 standard deviation (i.e. stronger easterly anomalies) are considered as

’break’ days. The method of defining ’active’ and ’break’ conditions is somewhat simi-

lar to the one used by Webster et al. [1998] where they also used a zonal wind criterion

over the north Bay-of-Bengal but used a fixed cut off anomaly (+3 ms−1 or -3 ms−1)

to define ’active’ and ’break’ conditions. Our method of defining ’active’ and ’break’

also bears similarity with the one used by Krishnamurti and Subrahmanyam [1982] for the

year 1979 where they used filtered zonal winds at a point in ’Arabian Sea’ to define ’ac-

tive’ and ’break’ episodes. The identification of the ’active’ and ’break’ days is not very

sensitive to small changes in the position of the reference point. It can be noted that in

Figure 2.7(a) between June 1 and September 30 of this particular year, there were two

’active’ and three ’break’ episodes. The ’active’ and ’break’ days are thus identified for

all years.

To test whether our criterion for defining ’active’ and ’break’ monsoon conditions

is related to the traditional precipitation based criterion, we calculated daily precipita-

tion composites for all ’active’ and ’break’ days defined by the circulation criterion for

Page 40: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.3 A Circulation Criterion for ’Active’ and ’Break’ Phases 20

Figure 2.7: (a) An example of 30-60 day filtered zonal winds for 1986 at a reference point (90◦E,15◦N). The thin horizontal lines correspond to +1 and -1 standard deviations. ’Active’ (’break’)days are defined as days for which the filtered zonal winds at the reference point are greater than+1 S.D (or less than -1 S.D). (b) 12-year (1978-1989) mean precipitation difference (mm.day−1)between all ’active’ and ’break’ composites. Contours are ±(1, 3, 5, 7, 9, 11, 13, 15).

the period 1978 and 1989 during June 1 to September 30. This is the period for which

gridded daily rainfall data over India was available to us. The precipitation difference

between ’active’ and ’break’ composites is shown in Figure 2.7(b). It is clear that the

pattern of precipitation anomalies during ’active’ (’break’) conditions is identical to the

dominant empirical orthogonal function (EOF) of daily (or pentad) rainfall [Singh and

Kriplani, 1990; Krishnamurthy and Shukla, 2000] with an ’active’ monsoon condition be-

ing associated with enhancement of precipitation over most of continental India except

a small region in south eastern India and another in north eastern corner. Thus, the ’ac-

tive’ and ’break’ monsoon conditions defined by our circulation criterion captures the

Page 41: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 21

dominant mode of intraseasonal precipitation variability over the Indian continent and

hence are essentially same as those defined by traditional precipitation criterion.

As the low-level jet over Somali is also usually strengthened (weakened) during an

’active’ (’break’) condition, one could also select a reference point in Arabian Sea (e.g.

60◦E, 10◦N) and 850 hPa zonal wind to define the ISOs.

2.4 Mean Structure of ISOs

Figure 2.8: (a,b) Climatological mean composite vector wind anomalies (ms−1) at 850 hPa cor-responding to ’active’ and ’break’ conditions for the 30-60 day mode and (c,d) associated relativevorticity (10−6s−1). The climatological mean composite is calculated by averaging all ’active’and ’break’ conditions occurring during the 20-year period (1978-1997). Shading in the upperpanels indicates regions with anomalies significant above 90% confidence level.

In this section, the underlying common spatial patterns associated with different

phases of the dominant ISO modes are isolated. To obtain the mean spatial pattern

common to all episodes of the dominant ISO variability, we use 20-year data of circu-

lation and OLR (1978-1997). The phase composite technique [Murakami and Nakazawa,

Page 42: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 22

1985; Murakami et al., 1984] is followed to illustrate the common mode of evolution of

the oscillations. Having defined the ’active’ and ’break’ days as described in section

2.3, averaged vector wind anomalies at 850 hPa associated with all the ’active’ and

’break’ phases of the 30-60 day mode are calculated within a year. A climatological

mean composite ’active’ phase constructed by averaging ’active’ composites of all 20

years is shown in Figure 2.8 together with the associated composite relative vorticity

pattern. The composite of all ’break’ phases is also shown in Figure 2.8. These com-

posites (means) and similar composites to be described later are tested for statistical

significance using a Student t-test by using the inter-event variability as a measure of

standard error. Level of statistical significance is noted in some of these figures. The sig-

nificant coherent large wind anomalies that emerge after averaging over about eighty

’active’ (’break’) episodes over a period of twenty years, shows that all ’active’ (’break’)

Figure 2.9: Climatological mean composite vector wind anomalies (ms−1) corresponding to’active’ and ’break’ conditions for the 30-60 day mode (a,b) at 500 hPa and (c,d) at 200 hPa.The climatological mean composite is calculated by averaging all ’active’ and ’break’ conditionsoccurring during the 20-year period (1978-1997). Shading indicates regions with anomalies sig-nificant above 90% confidence level.

Page 43: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 23

phases possess a common spatial pattern of variability. The other important feature

that emerges from the composite is the large zonal scale of the circulation changes asso-

ciated with ’active’ (’break’) phases of the Indian monsoon extending from about 50◦E

to 120◦E. During the ’active’ phase, the mean monsoon circulation is strengthened and

the monsoon trough cyclonic vorticity is enhanced north of 10◦N (compare with Figure

2.1(a,b)). The anticyclonic vorticity is enhanced between equator and 10◦N and cy-

clonic vorticity is weakened in the southern hemisphere. The mean composite 850 hPa

wind anomalies corresponding to ’active’ and ’break’ conditions (Figure 2.8) is consis-

tent with the pattern shown in Webster et al. [1998]. The climatological mean composite

’active’ and ’break’ phase vector wind anomalies at 500 hPa and 200 hPa are shown

in Figure 2.9. At 500 hPa during ’active’ phase, the vector wind anomalies bear close

resemblance with those at 850 hPa with cross-equatorial flow and enhancement of mon-

soon trough vorticity. At 200 hPa, the vorticity anomalies over the monsoon trough have

become anticyclonic and the equatorial wind anomalies are generally out of phase with

those at 850 hPa consistent with a first baroclinic mode vertical structure for this mode.

The composite picture of ’active’ and ’break’ conditions described above is consis-

tent with a seesaw between the two favorable positions of the TCZ as mentioned in

the Introduction. If this scenario is correct, there should be enhanced convection in the

northern position and decreased convection in the southern position during an ’active’

Figure 2.10: Climatological mean composite OLR anomalies (Wm−2) corresponding to ’active’and ’break’ conditions. ’Active’ and ’break’ composites are constructed using unfiltered OLRanomalies and the same ’active’ and ’break’ dates defined by 30-60 day filtered zonal windanomalies as used in Figure 2.8. OLR anomalies above 5 Wm−2 are significant above 90% con-fidence level.

Page 44: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 24

Figure 2.11: Climatological mean composite pressure vertical velocity anomalies (ω) at 500 hPa(hPas−1). Again the same ’active’ and ’break’ dates chosen from 30-60 day filtered zonal windanomalies for the 20-year period (1978-1997) as used in Figure 2.8 and Figure 2.10 are used.

phase while it should be the other way round during a ’break’ phase. Figure 2.10 sup-

ports this conjecture where composite of unfiltered OLR anomalies for all ’active’ and

’break’ days are plotted. The ’active’ and ’break’ days used in the composite are exactly

the same days defined by the 30-60 day filtered zonal winds at the reference point as

in the circulation composite. Coherence of the OLR anomalies averaged over 20 years

of ’active’ and ’break’ conditions defined by the circulation criterion shows that there

is a close relationship between circulation and convection associated with ’active’ and

’break’ conditions. A notable feature of the composites is that the meridional seesaw

of the convection anomalies is consistent with the low-level vorticity anomalies. It is

also worth noting that even after averaging over approximately eighty active (break)

episodes, fluctuations of OLR anomalies up to ±15 Wm−2 is seen over the two pre-

ferred regions. This means that, notwithstanding some variation in the intensity and

mean position of the TCZ from one ’active’ (’break’) episode to another, there exists a

common mean position of the TCZ during a typical ’active’ (’break’) episode. During

individual years it is not unusual to see ±25 Wm−2 OLR anomalies over either of the

zones.

To put the dynamical link between low-level cyclonic vorticity and convection on a

stronger footing, climatological mean composites of unfiltered pressure vertical veloc-

ity (ω) anomalies at 500 hPa corresponding to the same ’active’ and ’break’ days over

the full 20-year period were also constructed. It is seen in Figure 2.11 that, enhanced

(decreased) convection seen in Figure 2.10 are clearly associated with upward (down-

Page 45: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 25

ward) motion in both ’active’ and ’break conditions. Location and spatial pattern of

the vertical velocity anomalies correspond well with those of the convection anoma-

lies. Thus, ’active’ (’break’) conditions are associated with a seesaw of the anomalous

regional Hadley circulation.

To illustrate the evolutionary character of the circulation anomalies associated with

the 30-60 day mode, composite vector wind anomalies at 850 hPa and associated rela-

tive vorticity corresponding to eight phases of evolution of the oscillation is shown in

Figure 2.12 for the period (1979-1989). The phase-1 corresponds to days when the fil-

tered zonal wind anomalies at the reference point (90◦E, 15◦N) is zero and increasing

towards positive direction. If T is the period, other phases are progressively T/8 days

apart. In this way phase-3 is our ’active’ phase while phase-7 is our ’break’ phase. The

transition of the vector wind anomalies from ’active’ to ’break’ phase is clear from this

figure. It is also seen that while the ’active’ (’break’) phase defined by us has the largest

horizontal scale, largest wind anomalies in the low-level jet region over the Arabian Sea

may occur about 14 period prior to our ’active’ (’break’) phase. The northward prop-

agation of the TCZ is also depicted clearly from the transition of vorticity in different

phases. Similarly the evolutionary character of OLR for the 30-60 day mode correspond-

ing to the eight phases is illustrated in Figure 2.13. The ’active’ and ’break’ days used

in the composite are exactly the same days defined by the 30-60 day zonal winds at the

reference point as in the circulation composite. Apart from illustrating the northward

propagation of the convection zones, the relationship between convection and vorticity

is revealed from these two figures (Figure 2.12 and Figure 2.13). The movement of the

convection anomalies seems to be in phase with the movement of the vorticity anoma-

lies. The wind anomalies in each phase appear to arise from a linear response to heating

associated with OLR anomalies. For example, the strong south westerly anomaly in the

Somalijet region in phase-2 is consistent with the fact that this phase is characterized by

more OLR anomalies over the continent. Similarly more zonal winds in phase-3 is due

to the fact that this phase is characterized by OLR anomalies over Bay of Bengal and

South China Sea.

The ’active’ and ’break’ composite are constructed for the 10-20 day mode following

a similar procedure. ’Active’ (’break’) conditions are now defined by the 10-20 day fil-

tered zonal winds at 850 hPa being greater than +1 standard deviation (less that -1 S.D)

at the same reference point south of the monsoon trough. Due to it’s shorter period, it is

possible to have 8-10 episodes of ’active’ or ’break’ conditions for this mode during the

Page 46: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 26

summer monsoon season. The climatological mean ’active’, ’break’ composite vector

wind anomalies for the 10-20 day mode based on the entire twenty year period at 850

hPa is shown in Figure 2.14 together with the corresponding relative vorticity. Most

important feature of this mode is that it has a much smaller horizontal scale, confined

mainly to the Bay-of-Bengal. ’Active’ (’break’) conditions are associated with a strong

cyclonic (anticyclonic) vortex at the north Bay-of-Bengal with an anticyclonic (cyclonic)

vortex south of it between 10◦N and the equator. Due to the localized character of the

10-20 day mode, it is unlikely to have a strong influence on the large scale mean cir-

culation. However, depending on the phase relationship between the two ISOs, the

strong cyclonic (anticyclonic) vorticity over the north Bay-of-Bengal associated with the

’active’ (’break) phase of the 10-20 day mode can enhance (weaken) the cyclonic vortic-

ity over the monsoon trough zone associated with the 30-60 day mode [Goswami et al.,

1998]. In this manner, it can indirectly contribute to the mean monsoon circulation.

Page 47: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 27

Figure 2.12: Climatological mean composite vector wind anomalies (ms−1) at 850 hPa andassociated relative vorticity (10−6s−1) corresponding to eight phases of evolution of the 30-60day mode for the period 1979-1989. The phase-1 corresponds to the days when the filtered zonalwind anomalies at the reference point is zero and increasing toward positive values.

Page 48: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 28

Figure 2.13: Climatological mean composite OLR anomalies (Wm−2) corresponding to eightphases of evolution of the 30-60 day mode for the period 1979-1997. Eight composite phases areconstructed using unfiltered OLR anomalies and the same dates defined by 30-60 day filteredzonal winds as used in Figure 2.12.

Page 49: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.4 Mean Structure of ISOs 29

Figure 2.14: (a,b) Climatological mean composite vector wind anomalies (ms−1) at 850 hPacorresponding to ’active’ and ’break’ conditions for the 10-20 day mode and (c,d) associatedrelative vorticity (10−6s−1). The climatological mean composite is calculated by averaging all’active’ and ’break’ conditions occurring during the 20-year period (1978-1997).

Figure 2.15: Meridional bimodality of spatial structure of the dominant ISO. (a) Scatter plot ofdaily 30-60 day filtered vorticity at 850 hPa (10−6s−1) over a northern band (70◦E-100◦E, 12◦N-22◦N) and a southern band (70◦E-100◦E, 5◦S-10◦N) during 1 June to 30 September for 19 years(1979-1997). (b) Scatter plot of 30-60 day filtered OLR anomalies (Wm−2) averaged over thenorthern TCZ (70◦E-100◦E, 12◦N-22◦N) and the southern TCZ (70◦E-100◦E, 0◦-12◦S) during 1June to 30 September for 18 years (1979-1997, excluding 1994).

Page 50: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.5 Meridional Bimodality of ISO Spatial Structure 30

2.5 Meridional Bimodality of ISO Spatial Structure

As we hypothesize that the basic period and northward propagation of the mon-

soon ISOs result from the competition between the two favored positions of the TCZ,

the existence of meridional bimodality in low-level circulation and convection would

vindicate our hypothesis. In section 2.4, we showed that the peak phase of the ISO (the

’active’ and ’break’ phases) are characterized by a meridional bimodal structure. In this

section, we demonstrate that the meridional bimodality is characteristic not only of the

peak phases but valid through the evolution of the ISO. The robustness of the merid-

ional bimodality of the low-level vorticity is illustrated in Figure 2.15(a). In this figure,

the daily filtered vorticity over a north band (12◦N-22◦N) and a south band (10◦N-5◦S)

averaged between (70◦E-100◦E) during the northern summer is shown as a scatter dia-

gram for 19 years. The southern belt is part of a larger zone of opposite vorticity. It is

rather striking to note that, the vorticity over the two regions tend to be out of phase on

most days. The correlation between the two time series shown in this figure supports

this conclusion. Based on 19 years of daily values during summer season (June 1 to

September 30, i.e. 122 x 19 days), this correlation is highly significant. One may argue

that the southern favored position of the TCZ is between equator and 10◦S rather than

5◦S and 10◦N. It may be noted that the whole region between 12◦S and 10◦N fluctuates

with anticyclonic (cyclonic) vorticity corresponding to cyclonic (anticyclonic) vorticity

in the monsoon trough zone (Figure 2.8(c,d)). Figure 2.15(b) illustrates that the convec-

tion over the two favored locations indeed tends to fluctuate out of phase with each

other. Figure 2.15(b) shows a scatter diagram of OLR anomalies averaged over the two

preferred locations during the northern summer (June 1 to September 30) for the 18

years (1979-1997, excluding 1994). The correlation between convection anomalies over

the two locations while not very high is highly significant. The fact that there is a bi-

modality of convection over the two locations was also evident in the ’active’/’break’

composites (Figure 2.10). To understand the reason of the scatter in Figure 2.15(b), we

examined ’active’ and ’break’ composites of OLR for individual years. It is found that

(figure not shown) that bimodality of convection is clear in every individual year. How-

ever, there is some shift (east-west or north-south) in the location of the maximum OLR

anomalies over the two bands from one year to another. As we have fixed the position

of the two bands in plotting Figure 2.15(b), the non-stationarity of the band from one

year to another gives rise to the scatter in the plot. The year 1994 was excluded from

Page 51: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.5 Meridional Bimodality of ISO Spatial Structure 31

the set (1979-1997) as it was found to be anomalous in the sense that the positions of the

northern and southern TCZ were appreciably different from the mean position used in

this scatter diagram.

In section 2.4, we pointed out that there is a close relationship between the low-

level vorticity and convective activity associated with the ISOs. The strength of this

relationship between the low-level relative vorticity and convection is illustrated in Fig-

ure 2.16(a,b). Here, we plot a scatter diagram of relative vorticity anomalies and OLR

anomalies averaged over the northern position of the TCZ (85◦E-95◦E, 12◦N-22◦N) and

the southern position of the TCZ (85◦E-95◦E, 0◦-12◦S). All 19-year data between May 1

to October 31 are used in these scatter plots. The relationship between the two is sig-

nificantly negative over the northern position while it is significantly positive over the

southern position of the TCZ. As the southern position of the TCZ falls in the southern

hemisphere, both the relationship show that cyclonic (anticyclonic) low-level vorticity is

significantly correlated with increase (decrease) of convective activity in both favorable

locations of the TCZ.

Figure 2.16: (a) Scatter plot of 30-60 day filtered relative vorticity at 850 hPa (10−6s−1) and OLR(Wm−2) anomalies averaged over a box (85◦E-95◦E, 12◦N-22◦N) of the northern TCZ during 1May to 31 October for 19 years (1979-1997). (b) same as (a) but averaged over a box (85◦E-95◦E,0◦-12◦S) of the southern TCZ.

Page 52: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.6 Discussions and Conclusions 32

2.6 Discussions and Conclusions

In this chapter, the basic characteristics of intraseasonal oscillations of the Indian

summer monsoon is examined. We present a conceptual model (chapter 1) to describe

how the ISOs influence the seasonal monsoon. It envisages the ISO arising out of fluc-

tuation of the tropical convergence zone (TCZ) between two favored regions, one over

the monsoon trough (northern TCZ) and other over the equatorial warm waters (south-

ern TCZ). In one extreme of the ISOs (’active’ phase), the TCZ resides over the north-

ern position strengthening the seasonal mean monsoon circulation, enhancing cyclonic

vorticity over the northern TCZ and enhancing convection (and precipitation) over that

location while suppressing convection over the southern position. In the other extreme

(’break’ phase) weakened large scale monsoon flow and weakened cyclonic vorticity

over the northern position keeps the northern position clear of convection and helps

enhance convection over the southern position. A higher probability of occurrence of

’active’ (’break’) conditions in a monsoon season could therefore result in a stronger

(weaker) than normal seasonal mean monsoon.

In order to bring out the influence of the ISOs on the seasonal mean, it is desirable

to separate the externally forced component of the seasonal mean from the internally

forced component. We expect the slowly varying external forcing to give rise to slow

and persistent changes and manifest in the interannual variation of the annual cycle.

Intraseasonal anomalies are constructed in our study by removing the annual cycle of

individual years (sum of annual and semiannual harmonics) from the observations. In

this manner, we have been able to separate the influence of the external forcing on the

ISOs. We believe that this procedure is important in bringing out the intrinsic role of

the ISOs. Some studies define ISO anomalies with respect to climatological daily mean

as annual cycle and hence the ISOs contain the effect of interannual variations of the

annual cycle. This may be one reason why results of some previous studies have been

inconclusive.

Our first objective is to find the mean large scale spatial pattern associated with the

ISOs and compare them with that of the seasonal mean pattern. For this purpose we

have evolved a ’circulation’ criterion for defining ’active’ and ’break’ monsoon condi-

tions. Large scale structure of the mean circulation anomalies associated with the ’ac-

tive’ and ’break’ conditions of the dominant ISO modes are then obtained by construct-

ing composite of filtered 30-60 day or 10-20 day circulation anomalies at all points for

Page 53: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

2.6 Discussions and Conclusions 33

all ’active’ and ’break’ days. Climatological mean of all such composites for individual

years is then constructed representing a spatial pattern of the ’active’ and ’break’ phases

that is invariant from year to year. Such climatological mean composites corresponding

to a typical ’active’ (’break’) condition of the 30-60 day mode is associated with a general

strengthening (weakening) of the large scale mean monsoon flow and strengthening

(weakening) of the monsoon trough. It is rather interesting that the circulation changes

are not confined only over the Indian region but extended all the way to east of 120◦E

(South China Sea). The enhanced low-level cyclonic (anticyclonic) vorticity in the north-

ern TCZ during an ’active’ (a ’break’) condition is associated with enhanced (decreased)

ascending motion leading to enhanced (decreased) convection over the northern TCZ

and decreased (enhanced) ascending motion and decreased (enhanced) convection over

the southern TCZ. In other words, the anomalous regional Hadley circulation has as-

cending motion over the northern TCZ and descending motion over the southern TCZ

during an ’active’ condition while the reverse is the case during a ’break’ condition. A

typical evolutionary cycle of the dominant ISO based on composite of circulation and

convection for 20 years (1978-1997) is also constructed that show repeated northward

propagation from the southern position to the northern position (monsoon trough).

On an average sense, the meridional bimodality of the spatial structure of the peak

phases of the dominant ISO is evident in the composites. We also show that there is a

seesaw of low-level vorticity between the two preferred locations of the TCZ on a day-

to-day basis. Whether it is the northern location or the southern location of the TCZ,

low-level anomalous cyclonic vorticity is associated with enhanced convection while

low-level anomalous anticyclonic vorticity is associated with decreased convection es-

tablishing a link between anomalous circulation and convection.

Page 54: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 3

Intraseasonal Oscillations and InterannualVariability of the Indian Summer Monsoon

In this chapter, the relationship between the intraseasonal oscillations and the inter-

annual variability of the seasonal mean Indian summer monsoon is investigated. An

attempt is made to arrive at some reliable conclusions through a series of detailed in-

vestigation of various aspects of the problem. Daily anomalies in a particular year is

defined as departure of observations from the annual cycle of that year and a Butter-

worth filter is used to isolate the ISO modes (see section 2.1). In order to include the

effect of both the ISO modes and keeping in mind their interannual variations in their

peak period, total intraseasonal activity is defined by a band pass filtered field with

peak response at 35 days and half responses at 15 days and 80 days respectively. For

all the calculations described below, these ISO filtered fields are used to bring out the

relationship between ISO and interannual variability.

3.1 A Common Spatial Mode of Intraseasonal and InterannualVariability

In the previous chapter (section 2.4), it was shown that the large scale structure of

the wind associated with the dominant ISO mode is quite similar to that of the seasonal

mean wind, strengthening and weakening the large scale flow during it’s ’active’ and

’break’ phases respectively. This similarity between the structure of the intraseasonal

variability and the seasonal mean flow provides the basis for our hypothesis that the

ISOs could influence the seasonal mean and it’s interannual variability. In this section,

we provide further evidence that the spatial structure of the intraseasonal variability

and the interannual variability are similar. The geographical distribution of the intrasea-

sonal activity and the interannual variability are compared in Figure 3.1. In this figure,

Page 55: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.1 A Common Spatial Mode of Intraseasonal and Interannual Variability 35

the standard deviation of ISO filtered 850 hPa relative vorticity averaged over the 20-

year period (1978-1997) and interannual standard deviation of the seasonal mean (JJAS)

relative vorticity based on the same 20-year period are shown. It may be noted that

mean amplitude of intraseasonal activity of this field is two to three times larger than

the interannual variation of the seasonal mean in most places. The similarity of the

geographical distribution of intraseasonal variability and interannual variability of 850

hPa relative vorticity is noteworthy. The correlation between the two patterns is 0.64

over the monsoon region (50◦E-100◦E, 20◦S-30◦N). Both the patterns are characterized

by strong activity in the two favored positions of the TCZ namely a northern position

around the monsoon trough and a southern position between the equator and 10◦S. Re-

gions of higher intraseasonal activity are also regions of larger interannual variability.

What we have shown so far (e.g. the composite, the similarity between the S.D of

ISO and interannual variability of the seasonal mean) are only indicative of a common

mode of variability. To bring out the common spatial pattern of intraseasonal and in-

terannual variability, the following procedure is adopted. An EOF analysis of the ISO

filtered 850 hPa winds from June 1 to September 30 for all 20 years (1978-1997) is carried

out. The first EOF explaining 14.8% of the total intraseasonal variance and represent-

ing the dominant ISO mode is shown in Figure 3.2(a). The dominant interannual mode

is obtained from an EOF analysis of seasonal mean (JJAS) 850 hPa winds for 40 years

(1958-1997). The first EOF explaining 16.8% variance of interannual variability of the In-

dian summer monsoon is shown in Figure 3.2(b). That the interannual EOF1 represents

dominant variability of the Indian summer monsoon is seen from the strong correlation

between PC1 and IMR (r=0.62) shown in Figure 3.2(c). Although there are some mi-

nor differences, the similarity between the spatial patterns of the dominant ISO mode

and the dominant interannual mode is noteworthy. The easterlies south of the equator,

the cross-equatorial flow east of 50◦E, the convergence of air mass from north-west and

south-west over the Arabian Sea around 10◦N, the monsoon trough, the anticyclonic

vortex around 75◦E, 5◦N are all common in both the patterns. The cross equatorial flow

near African coast around equator seen in the interannual mode (Figure 3.2(b)) is not

seen in the intraseasonal mode (Figure 3.2(a)). This is partly due to the fact that the

dominant ISO can not be entirely represented by a single EOF due to its northward

propagating character. The second dominant ISO mode is strongly correlated with the

first at a lag of about 10 days and has large loadings exactly in this region∗. Therefore,

∗The second ISO EOF is not shown here for brevity. But the reader can refer to Chapter 5, Figure 5.2

Page 56: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.1 A Common Spatial Mode of Intraseasonal and Interannual Variability 36

a common spatial pattern governs both the ISOs and the interannual variability, thus

linking the ISOs with the interannual variability of the Indian monsoon.

While the ISOs may have a common mode of spatial variability with the interannual

variations of the seasonal mean monsoon, they may not have appreciable influence on

the later unless the interannual variations of the ISO activity are significant. Therefore,

we estimate the amplitude of interannual variations of the ISO activity and compare

it with the amplitude of interannual variability of the seasonal mean. The standard

deviation of ISO filtered vorticity at 850 hPa and OLR between June 1 and September

30 is calculated each year at each grid point. The interannual standard deviation of

this intraseasonal standard deviation of each year is calculated based on 20 years (1978-

1997). The interannual standard deviation of seasonal mean (June-September) vorticity

at 850 hPa and OLR are separately calculated. The ratio between the standard deviation

of interannual variations of ISO activity and interannual variation of seasonal mean is

shown in Figure 3.3. It is seen that magnitude and pattern of ratio is similar for both

low-level vorticity and OLR. The equatorial belt (10◦S-10◦N) east of 100◦E is character-

ized by a low ratio, as the interannual variations are stronger in this region. In most

of the Indian monsoon regions the ratio ranges from 0.4 to 0.8. This indicates that the

variations of the ISO activity could account for 20% to 60% of interannual variability

of the seasonal mean in the Indian monsoon region. Thus, we can expect significant

modulation of the seasonal mean monsoon by the ISOs.

Page 57: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.1 A Common Spatial Mode of Intraseasonal and Interannual Variability 37

Figure 3.1: Geographical distribution of intraseasonal and interannual activity. (a) Mean stan-dard deviation of ISO filtered relative vorticity (10−6s−1) at 850 hPa during 1 June to 30 Septem-ber for 20 years (1978-1997). (b) Interannual standard deviation of seasonal mean relative vor-ticity (JJAS, 10−6s−1) based on the same 20 years.

Page 58: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.1 A Common Spatial Mode of Intraseasonal and Interannual Variability 38

Figure 3.2: First EOF of the intraseasonal and interannual 850 hPa winds. (a) IntraseasonalEOFs are calculated with ISO filtered winds for the summer months (1 June to 30 September)for a period of 20 years (1978-1997). (b) Interannual EOFs are calculated with the seasonal mean(JJAS) winds for 40-year period (1958-1997). Units of vector loading are arbitrary. (c) Relationbetween IMR and interannual PC1. Filled bars indicate interannual PC1 and the unfilled barrepresent IMR. Both time series are normalized by their own standard deviation. Correlationbetween the two time series is shown.

Page 59: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.1 A Common Spatial Mode of Intraseasonal and Interannual Variability 39

Figure 3.3: Ratio between standard deviation of interannual variation of ISO activity and inter-annual variation of the seasonal mean. (a) Relative vorticity at 850 hPa. (b) OLR. Contours are(0.3, 0.4, 0.6, 0.8, 1.0).

Page 60: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 40

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal MeanMonsoon

The fact that spatial pattern associated with the dominant ISO mode has a large spa-

tial scale similar to the interannual variability of the seasonal mean, provides the basis to

the idea that ISOs could influence the seasonal mean. For example, a higher frequency

of occurrence of ’active’ (’break’) conditions within the monsoon season could lead to

strengthening (weakening) of the seasonal mean resulting in a strong (weak) monsoon.

This essentially is the hypothesis proposed in our conceptual model (chapter1). To test

this hypothesis, we calculate probability density functions of the ISOs corresponding to

’strong’ and ’weak’ monsoon years. ’Strong’ and ’weak’ monsoon years are objectively

defined based on whether IMR is greater than 1 S.D or less than -1 S.D. To have enough

sample of such years we use daily data between 1956 and 1997. This period contains

seven strong years (1956, 1959, 1961, 1970, 1975, 1983 and 1988) and ten ’weak’ monsoon

years (1965, 1966, 1968, 1972, 1974, 1979, 1982, 1985, 1986, 1987). As mentioned earlier,

the spatial pattern of the ISOs involve a northward propagating component. As a result,

the evolutionary character of the ’active’ and ’break’ conditions cannot be described by

a single EOF. To estimate the PDF of the ISOs, therefore, it is necessary to include more

than one EOF. In the present study, we estimate the PDF of the ISOs using at least two

EOFs. To obtain the PDF for the ’strong’ (’weak’) years, daily ISO filtered 850 hPa vor-

ticity between June 1 and September 30 for all ’strong’, all ’weak’ and ’all’ (all 20 years

between 1978 and 1997 taken together) years are combined and an EOF analysis is car-

ried out in each case using singular value decomposition technique [Nigam and Shen,

1994]. The first two EOFs in each case are shown in Figure 3.4. It may be noted that the

first EOF in ’strong’ and ’weak’ cases for positive projection coefficients (PC’s) represent

’active’ like and ’break’ like conditions respectively. The PDF of the PC’s in the reduced

phase space defined by the first two EOFs explaining 17% of the total variance of the

’strong’ years (21% for ’weak’ and 15% for ’all’ years) are obtained using a Gaussian ker-

nel estimator [Kimoto and Ghil, 1993; Silverman, 1986] with a smoothing parameter large

enough to detect multi-modality with statistical significance. The smoothing parameter

(h) is selected from the minimum obtained from the least-squares cross validation tech-

nique [Kimoto and Ghil, 1993]. In our case, h usually varies between 0.6 and 0.8. In these

calculations both the PC’s are normalized by the temporal S.D of each of the PC’s.

Page 61: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 41

The two-dimensional PDF corresponding to ’strong’, ’weak’ and ’all’ years are shown

in Figure 3.5(a), (b) and (c) respectively. For ’strong’ and ’weak’ years the PDF of ISO ac-

tivity is clearly non-Gaussian while in the case of ’all’ years, it is a Gaussian. The spatial

pattern corresponding to the maxima of PDF in each case is constructed using appro-

priate normalization constants for the PC’s and the corresponding EOF1 and EOF2 pat-

terns. In the ’strong’ case we note that the two maxima have almost equal probability.

In the ’weak’ case there are three maxima of the PDF patterns while in the ’all’ case there

is only one maximum. As the number of ’strong’ and ’weak’ monsoon years included in

the PDF calculation are quite large, we expect the maxima of PDFs in Figure 3.5(a) and

Figure 3.5(b) to be robust. To test the statistical significance of these maxima, we created

1000 random sets of time series having same variance and autocorrelation at 1-day lag

equal to those of observed PC1 and PC2 and 2-D PDFs were calculated for each of them.

In Figure 3.5(a) and Figure 3.5(b), shading represents regions where the observed PDF

is significantly greater than the random ones with 90% confidence, i.e. 25 or fewer of

the random PDFs were larger than those shown in Figure 3.5(a) and Figure 3.5(b). The

maxima of the PDFs are found to be statistically significant in each case. In the ’strong’

and ’weak’ cases, we are primarily concerned with the statistical significance of the PDF

maxima representing deviation from Gaussian distribution. Since in the ’all’ case as the

PDF pattern is a Gaussian, similar significance test is not presented. In the ’strong’

case, the maximum with normalized PC1 close to 1 and PC2 close to zero represents a

strong ’active’ condition shown in Figure 3.6(a). The other maximum represents a very

weak ’break’ pattern. Although the two patterns have equal probability, strong ’active’

pattern would have the dominating influence on the seasonal mean. For the ’weak’

case, the maximum with both PC1 and PC2 close to zero represents a transition pattern.

Both the other maxima represent strong ’break’ conditions. One such ’break’ condition

with both PC1 and PC2 making approximately equal contributions is shown in Figure

3.6(b). On the other hand, the maximum of the PDF in ’all’ year case corresponds to a

transition pattern with insignificant vorticity associated with it is shown in Figure 3.6(c).

Page 62: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 42

Figure 3.4: First two EOFs of the daily ISO filtered 850 hPa vorticity from 1 June to 30 Septem-ber. (a) EOF1 and (b) EOF2 for seven ’strong’ years (c) EOF1 and (d) EOF2 for ten ’weak’ years(e) EOF1 and (f) EOF2 for ’all’ (20 years from 1978 to 1997) years. Arbitrary EOF loadings havebeen multiplied by a factor of 100.

Page 63: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 43

Figure 3.5: Evidence of change in regimes of ISOs during ’strong’ and ’weak’ monsoon years.Illustrated are two-dimensional PDFs of the ISO state vector spanned by two dominant EOFsof low-level vorticity. PDFs are calculated with principal components normalized by their ownstandard deviation and taking the summer days (1 June to 30 September) for (a) 7 ’strong’ mon-soon years (b) 10 ’weak’ monsoon years (c) 20 combined ’strong’, ’weak’ and ’normal’ years(1978-1997). The smoothing parameter used is h=0.6 and PDFs are multiplied by a factor 100.The first two EOFs (not shown) are different in ’strong’, ’weak’ and ’all’ years but are related to’active’ and ’break’ conditions. The origin of the plots corresponds to a very weak state repre-senting a transition between the two states (as in the ’all’ case).

Page 64: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 44

Figure 3.6: Geographical patterns of the dominant regimes for low-level relative vorticity(10−6s−1) shown in Figure 3.5. (a) ’strong’ monsoon years (b) ’weak monsoon years (c) ’all’years.

Page 65: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 45

We note that the seasonal mean low-level vorticity over the northern TCZ position

is cyclonic (see Figure 2.1). The cumulative effect of higher frequency of occurrence of

’active’ (’break’) conditions is expected to result in stronger (weaker) than normal cy-

clonic vorticity in this region. Since higher frequency of ’active’ (’break’) conditions are

associated with ’strong’ (’weak’) Indian monsoon, we can expect a strong relationship

between seasonal mean vorticity over the monsoon trough (northern position of TCZ)

and the strength of the Indian monsoon. This conjecture is tested in Figure 3.7 where we

plot the seasonal mean relative vorticity averaged over the monsoon trough and IMR

for the 40-year period (1958-1997). The correlation between the two time series is 0.74,

strongly supporting our conjecture.

Figure 3.7: The monsoon trough vorticity (MTV) and the Indian Monsoon Rainfall (IMR) fora 40-year period (1958-1997). MTV is defined as the seasonal mean vorticity (JJAS) averaged inthe domain 40◦E-90◦E and 10◦N-30◦N. Both time series are normalized by their own standarddeviation. Correlation between the two time series is shown.

It would be desirable to see if the conclusions derived from circulation alone will be

supported if convection is also included to describe ISOs. However, OLR data as proxy

for convection is available only from 1974 onwards. The period between 1974 and 1997

contains six ’weak’ monsoon years (1974, 1979, 1982, 1985, 1986, 1987) as described by

the criterion used earlier. However, the same criterion indicates only three ’strong’ mon-

soon years in this period. To enhance the sample size of the ’strong’ monsoon years, we

relaxed the objective criterion to include years for which IMR > 0.5 S.D. Based on the

relaxed criterion, six ’strong’ monsoon years (1975, 1978, 1983, 1988, 1990,1994) are se-

lected during this period. As in the previous case, a combined EOF analysis is carried

out for 850 hPa vorticity and OLR for ’strong’ (’weak’) years. The first two CEOFs in

each case are shown in Figure 3.8. It may be noted that the first CEOF in ’strong’ and

Page 66: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 46

’weak’ cases for positive projection coefficients (PC’s) represent ’active’ like and ’break’

like conditions respectively. The PDF is then calculated on the reduced phase space de-

fined by the first two CEOFs. Similarly, the PDF of CEOF for ’all’ years (all 20 years from

1978-1997) is also calculated. The PDFs for three different cases are shown in Figure 3.9.

It is clear that the PDFs are asymmetric for both ’strong’ and ’weak’ cases while it is a

Gaussian in ’all’ case. As in the earlier case, statistical significance for the observed PDFs

were carried out and regions of phase space where the observed PDFs are significantly

larger than the randomly generated ones with 90% confidence are shaded. Using appro-

priate normalization constants for the PC’s and corresponding EOF1 and EOF2 (Figure

3.8) the patterns representing maxima in PDF are calculated. In the ’strong’ case, the

most probable pattern corresponds to an ’active’ condition (Figure 3.10(a)). The other

maxima with much less probability represents a weak ’break’ condition (not shown).

For the ’weak’ case both maxima correspond to ’break’ conditions. The one with nor-

malized PC’s close to zero, however, represents a weak ’break’ condition (not shown)

while the other maxima in PDF represents a strong ’break’ condition (Figure 3.10(b)).

The most probable pattern in the ’all’ case (Figure 3.10(c)) corresponds to a very weak

pattern representing a transition between ’active’ and ’break’ patterns. Thus, even if

we take circulation and convection together, the strong (weak) monsoon appears to be

associated with a higher probability of occurrence of active (break) like conditions.

Page 67: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 47

Figure 3.8: First two CEOFs of the daily ISO filtered 850 hPa vorticity and OLR from 1 June to30 September. (a) CEOF1 and (b) CEOF2 for six ’strong’ years (c) CEOF1 and (d) CEOF2 for six’weak’ years (e) CEOF1 and (f) CEOF2 for ’all’ (20 years from 1978 to 1997) years. Arbitrary EOFloadings have been multiplied by a factor of 100.

Page 68: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 48

Figure 3.9: Same as Figure 3.5 but based on the state vector defined by the first two combinedEOF of low-level vorticity and OLR.

Page 69: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.2 Probability of ’Active’/’Break’ Conditions and Seasonal Mean Monsoon 49

Figure 3.10: Geographical patterns of the dominant regimes shown in Figure 3.9. (a) ’strong’monsoon years (b) ’weak monsoon years (c) ’all’ years. OLR patterns are shown as shadedcontours (Wm−2) while the corresponding low-level vorticity are shown in contours (10−6s−1).

Page 70: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.3 Interannual Variations of ISO Activity and Seasonal Mean Monsoon 50

3.3 Interannual Variations of ISO Activity and Seasonal MeanMonsoon

Instead of trying to relate the seasonal mean monsoon to statistics of ’active’ and

’break’ conditions, in this section we ask, is there a relationship between level of ISO

activity and seasonal mean Indian monsoon? Few previous studies have indicated evi-

dence to connect the interannual variations of ISO activity and seasonal mean monsoon

strength. Hendon et al. [1999] found that global ISO activity during boreal winter is in-

versely related to Australian monsoon strength. Recently Lawrence and Webster [2001]

found that summertime ISO activity exhibits an inverse relationship with the Indian

monsoon strength using OLR over a 22-year period (1975-97). However, they found

that the correlation between ISO activity and monsoon over Bay of Bengal is weak.

Thus the robustness of the negative relationship between ISO activity and Indian sum-

mer monsoon is not clear.

We have used circulation data for a 44-year period (1954-97) from NCEP/NCAR

reanalysis for this purpose. An index is defined to represent the strength of the ISO

activity. In the previous chapter (section 2.5), we have shown that the monsoon ISO is

characterized by a meridional bimodal structure. There is a seesaw of low-level vorticity

and convection between the two preferred locations of the TCZ on a day-to-day basis.

We have defined the ISO activity index as the variance associated with the low-level

ISO filtered vorticity averaged over one of the center of action of ISO activity (70◦E-

100◦E, 12◦N-22◦N) . The all India monsoon rainfall (IMR) represents the seasonal mean

monsoon strength. Figure 3.11 shows time series of the ISO activity index and IMR each

normalized by it’s own standard deviation for the 44 year period. The correlation of 0.32

is significant at the 5% level. Although this relationship is not very strong, it indicates

that interannual variations of ISO activity is positively related to the monsoon strength.

This result is consistent with Figure 3.1. This means that strong activity of intrasea-

sonal oscillations tends to correspond to seasons of above normal rainfall. This result is

in contrast with Lawrence and Webster [2001], as they found an inverse relationship be-

tween the strength of the ISO activity and IMR using OLR data for a period (1975-1997).

The correlation between the ISO activity index and IMR in this period (1975-1997) for

low-level vorticity is very low (r=0.09) whereas the correlation for the period 1954-1974

is high (r=0.59). This indicates that the relationship between ISO activity and IMR is

weakening in the recent decades. This demonstration of the two periods is rather arbi-

Page 71: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.4 Discussions and Conclusions 51

1955 1960 1965 1970 1975 1980 1985 1990 1995

−2

−1

0

1

2

r=0.32

r=0.09r=0.59

Figure 3.11: (a) Time series of ISO activity index (blue) and All India Monsoon Rainfall Index(IMR, black) normalized by it’s own standard deviation for a 44-year period (1954-1997).

trary. To examine the changing character of this correlation, a 21-year sliding window

correlation analysis was carried out between the two variables. It is seen that (figure not

shown) that the correlation remained about 0.6 until mid-seventies and then decreased

rapidly and remained low in eighties and nineties.

3.4 Discussions and Conclusions

The primary objective of this study is to investigate how and to what extent the

monsoon ISOs influence the seasonal mean and the interannual variability of the In-

dian summer monsoon. The underlying hypothesis is that the seasonal mean monsoon

has a component forced by internal dynamics in addition to a component forced by

external conditions such as slowly varying boundary forcings. This hypothesis can be

considered as an extension of the Charney and Shukla [1981] hypothesis that suggested

the interannual variation of Indian monsoon to be primarily forced by boundary forcing

at the earth’s surface. We propose that the part of the interannual variations of the sea-

sonal mean that is independent of external forcing arise from the changes in the statis-

tics of the intraseasonal oscillations of the Indian monsoon. As the ISOs are intrinsically

chaotic, the predictability of the seasonal mean Indian monsoon depends on the extent

to which the ISOs influence the seasonal mean relative to the externally forced compo-

nent. According to our conceptual model, the intraseasonal and interannual variations

of the Indian monsoon should be governed by a common mode of spatial variability.

In addition, if indeed the ISOs determine the ’strong’ and ’weak’ monsoons, the PDF

of the ISOs should have higher probability of occurrence of ’active’ conditions during

’strong’ monsoon years and ’break’ conditions during ’weak’ monsoon years. These

two elements of our hypothesis are rigorously tested using a sufficiently long record of

daily circulation and convection data.

Page 72: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.4 Discussions and Conclusions 52

Our first objective was to find the mean large scale spatial pattern associated with

the ISOs and compare them with that of the seasonal mean pattern. We have found

that spatial pattern of the ’active’ and ’break’ that is invariant from year to year and

is similar to the spatial structure of the seasonal mean (Chapter 2). The close resem-

blance between the spatial structure of the ’active’ and ’break’ composites and that of

the seasonal mean, indicates a similarity between the spatial structure of intraseasonal

and interannual variability. The spatial distribution of standard deviation of 850 hPa

vorticity associated with ISO variability and that of interannual variability of the sea-

sonal mean are shown to be closely similar (pattern correlation 0.64, Figure 3.1). That

the intraseasonal and interannual variations are governed by a common spatial mode

of variability (Figure 3.2) is seen from the notable similarity between the dominant EOF

of intraseasonal oscillations (based on 20 years of daily ISO filtered data during the

summer season) and the dominant EOF of the interannual variation of the seasonal

mean (based on 40 years data of seasonal mean data). In contrast to some recent stud-

ies [Annamalai et al., 1999; Sperber et al., 2000] where it is claimed that it is not possible

to describe the interannual variations of the Indian summer monsoon by a single EOF,

we show that the dominant EOF indeed represents interannual variations of the Indian

summer monsoon (correlation between IMR and PC1 is 0.62) if the domain is restricted

between 40◦E-100◦E and 20◦S-30◦N. Sperber et al. [2000] has found a common mode of

variability between intraseasonal and interannual variations in the third leading EOF.

If the domain of analysis included regions east of 100◦E, the ENSO related variation in

the eastern part of the domain dominates the first EOF and the interannual monsoon

variations may appear as the second or third EOF.

Next, we show that the interannual variations of the summer ISO variance has the

potential for significantly influencing (up to 20-60%) the interannual variations of the

seasonal mean. Then, we argue that it is not the amplitude of the ISO activity but the

asymmetry in the occurrence of the ’active’ and ’break’ conditions that affect the sea-

sonal mean. We investigate whether the frequency of occurrence of ’active’ and ’break’

conditions are distinctly different during ’strong’ (flood years) and ’weak’ (drought

years) monsoon years. For this purpose, a two-dimensional PDF estimation technique

is employed on the ISO filtered field. Daily low-level vorticity field between 1956 and

1997 is employed to include a large number of ’strong’ (seven) and ’weak’ (ten) mon-

soon years. This objective technique clearly shows that the PDFs are distinctly asym-

metric and different during ’strong’ and ’weak’ monsoon years and the most frequently

Page 73: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.4 Discussions and Conclusions 53

occurring pattern during ’strong’ (’weak’) monsoon years is the ’active’ (’break’) pat-

tern. On the other hand, if all years are linked together, the PDF is Gaussian with the

transition between ’active’ and ’break’ pattern being the most frequently occurring pat-

tern. Thus, the cumulative effect of the ’active’ condition during a ’strong’ monsoon

season lead to stronger that normal cyclonic vorticity in the north TCZ position and

stronger than normal seasonal mean. This conclusion is further supported by strong

correlation between seasonal mean vorticity over the northern TCZ position and IMR

(Figure 3.7). That the conclusions arrived at from the PDF of low-level vorticity are

robust is supported by PDF estimate of combined low-level vorticity and convection.

Using simultaneous convection and circulation data (1974-1997), combined EOF of the

low-level vorticity and OLR is carried out for all ’strong’ and ’weak’ years as well as all

the years taken together. This calculation also shows that the most frequent pattern dur-

ing a ’strong’ (”weak’) year is the ’active’ (’break’) pattern with enhanced (decreased)

cyclonic vorticity and negative (positive) OLR anomaly over the northern TCZ position.

Our results are consistent with the findings of Krishnamurthy and Shukla [2000] where

they examined daily rainfall over Indian continent for the period 1901-1970 and showed

that ’strong’ (’weak’) monsoon years are associated with ’active’ (’break’) conditions

(their Fig.12a). They define ISO anomalies with respect to a climatological mean sea-

sonal cycle. If they remove the ’seasonal mean anomaly’ (June 1-September 30) from

the anomalies they do not find a clear signal of skewness in the PDF. Their figure 13

which brings out the strengthening and weakening of the large scale monsoon flow by

the ISO is also consistent with our result. They argue that the active and break phases do

not change the character of the mean monsoon flow, but merely representing strength-

ening and weakening of the flow. However, if we look at the magnitude of change in

the mean in their figure, it accounts for 20-30% change of the mean during ’active’ and

’break’ conditions. Close to the equator the change could be even 100%. This we believe

represents significant change of the mean by ISO.

Both Krishnamurthy and Shukla [2000] and Sperber et al. [2000] agree with each other

and claim that the PDF of the ISO modes is biased towards positive or negative side

during strong and weak years only if the low frequency interannual variation of the

seasonal mean is not removed. Our results, differ with these studies in that we find

a distinct bias of the PDF towards ’active’ (’break’) conditions during strong (weak)

monsoon years even after removing the seasonal mean anomaly. Although we define

anomalies as departure of observations from annual cycle of individual years, the PDF

Page 74: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

3.4 Discussions and Conclusions 54

is calculated on the band pass filtered ISO anomaly. This procedure essentially removes

the seasonal mean. The main difference between ours and earlier studies is the use of

two EOFs to describe the dominant ISO mode in our study. Both the studies mentioned

above used only one EOF to describe the dominant ISO mode. Since the dominant ISO

mode has a prominent northward propagating character, it cannot be described by just

one EOF. We believe that this is the reason why the earlier studies failed to notice any

bias in the PDF after removing the seasonal mean.

Page 75: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 4

Estimate of Potential Predictability ofMonthly and Seasonal Means in Tropicsfrom Observations

4.1 Introduction

The predictability of weather (or the instantaneous state of the atmosphere) is lim-

ited to about two weeks [Lorenz, 1982] due to inherent instability and nonlinearity of the

system. The atmosphere, however, possesses significant low frequency variability. As

has been mentioned in Chapter 1, if the low frequency variations of the monthly and

seasonal means were entirely governed by scale interactions of the higher frequency

chaotic weather fluctuations, then the time averages will be no more predictable than

the weather disturbances themselves. However, it appears that a large fraction of the

low frequency variability, specially in the tropics, may be forced by slowly varying

boundary conditions such as the sea surface temperature (SST), soil moisture, snow

cover and sea-ice variations. Hence, the predictability of climate (e.g. space-time aver-

ages) is determined partly by chaotic internal processes and partly by slowly varying

boundary forcings. This understanding that anomalous boundary conditions (ABC)

may provide potential predictability has formed the scientific basis for deterministic

climate predictions [Charney and Shukla, 1981; Shukla, 1981, 1998]. Research during the

past decade has shown that the climate in large part of tropics is largely determined by

slowly varying SST forcing where potential for making dynamical forecast several sea-

sons in advance exists [Latif et al., 1998]. However, during the same period, we have also

learnt that there are regions within the tropics, climate of which is not strongly governed

by ABC. The Indian summer monsoon is such a system [Webster et al., 1998; Brankovic

and Palmer, 1997; Goswami, 1998]. The intraseasonal oscillations such as the east-ward

propagating Madden-Julian Oscillations (MJOs) and the north-ward propagating mon-

Page 76: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.1 Introduction 56

soon ISOs with period in the range of 30 to 60 days are quite vigorous in the tropics.

Both the MJOs as well as the monsoon ISOs are known to be driven by internal feed-

back between convection and dynamics. In addition to the scale interactions between

weather disturbances, time-averaging of the chaotic ISOs can also contribute to the low

frequency variability of monthly and seasonal means in the tropics. The nonlinear scale

interaction associated with the weather disturbances in the tropics is likely to be weak

as they are less vigorous compared to their counterpart in the extra-tropics. Therefore,

we envisage that most of the internal contribution to the low frequency variations in the

tropics may come from time averaged residual of the ISOs.

The total low frequency variance of any variable in a given region (σ2) could be

written as super-position of variance due to external forcing (σ2e ) and variance due to

internal processes (σ2i ).

σ2 = σ2e + σ2

i

Making unambiguous estimates of the ’internal’ and ’external’ components of vari-

ability from observations is rather difficult. Madden [1976, 1981], Madden and Shea [1978]

and Shea and Madden [1990] attempted to estimate the two variances in some extra-

tropical regions. They estimated synoptic scale internal variability from short time se-

ries (such as within a season) and extrapolated the power spectrum to lower frequencies

by assuming a white noise extension. This approach is simple but assumes that the low

frequency power spectrum would be white and it could be extrapolated from power

at higher frequencies. Shukla [1983] commented at length on Madden’s [1976] approach

and argued that the methodology used and assumptions made by Madden could have

overestimated the natural variability or ’climate noise’ and underestimated the poten-

tial predictability. Madden [1983] while disagreeing with Shukla that his method un-

derestimated the potential predictability agreed that there is considerable uncertainty

in separating the so called ’climate noise’ from the climate signal. Shukla and Gutzler

[1983] and Short and Cahalan [1983] used a more general low frequency extension of

the intraseasonal variance to estimate the level of ’climate noise’. Trenberth [1984a, b]

points out that these estimates depend crucially on the use of correct value of T, the

effective time between independent data. He pointed out that these studies may have

underestimated T by using negatively biased estimates of the lagged autocorrelations

by improperly removing the annual cycle and interannual variability.

Alternatively this ratio could be estimated using atmospheric general circulation

models (AGCM) from a long integration with observed boundary condition and an-

Page 77: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.1 Introduction 57

other long integration with fixed boundary condition [Goswami, 1998] or from an ensem-

ble of long integrations of the AGCM with the same observed boundary conditions but

differing only in the initial conditions [Stern and Miyakoda, 1995; Harzallah and Sadourny,

1995; Rowell et al., 1995]. Kumar and Hoerling [1995] estimated the ratio between the ex-

ternal and internal variability for the extra tropics using a large ensemble of long simu-

lations by an AGCM. Zweirs and Kharin [1998] have examined the interannual variability

and potential predictability of 850 hPa temperature, 500 hPa geopotential and 300 hPa

stream function simulated by AMIP models. They have found that there is a wide varia-

tion in the ability of the AGCMs to simulate observed interannual variability, both total

and weather noise component. Zheng et al. [2000] have proposed a method to estimate

potential predictability of seasonal means using monthly mean time series. Using this

technique they have estimated the potential predictability of surface temperature, 500

hPa geopotential height and 300 hPa winds. The potential predictability tends to be

high in the tropics and low in the extratropics as per their calculations.

Singh and Kriplani [1986] have estimated potential predictability of lower tropo-

spheric monsoon circulation and rainfall over India for JJA season. Daily 700 hPa

geopotential heights, mean sea level pressure and rainfall anomalies were used for the

study. They have found that potential predictability of seasonal lower tropospheric

fields is less over the monsoon trough, but it generally increases with decreasing lat-

itude. For rainfall, potential predictability is about 50% over the major parts of the

country. The reliability of the estimates of potential predictability in this study may be

affected by insufficient data length. The method of removing the annual cycles which

is important in these kind of analysis [Trenberth, 1984a] has not been outlined.

Sontakke et al. [2001] have estimated potential for long-range predictability of pre-

cipitation over the Indian sub-continent using precipitation data from 1901-1970. Their

study indicate that the climate noise is small compared to climate signal over the Indian

monsoon region. The F-ratios of JJAS precipitation ranges from 1.5 to 2.5, with high

values on the west coast of India. Although the F-ratios are not very high, it indicates

certain amount of potential predictability of the seasonal mean precipitation.

In all the studies of estimating potential predictability from observations mentioned

earlier, the total interannual variability (i.e the climate signal) is compared to the ’cli-

mate noise’. The so called ’climate signal’ actually contains contributions from the ’ex-

ternal’ forcing as well as the internal ’climate noise’. To the best of our knowledge, no

attempt has been made to separate the contributions from the ’external’ and the ’inter-

Page 78: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 58

nal’ components to the observed interannual variability. Here, we propose a method of

separation of interannual variances of monthly means associated with the slowly vary-

ing externally forced component and from the internally determined component. The

variances associated with the ’internal’ and ’external’ components are estimated. It is

also demonstrated that the ’external’ component separated by our method indeed rep-

resents the response of the tropical atmosphere to the slowly varying SST forcing. A

measure of potential predictability is defined as the ratio between the ’total’ (sum of

’external’ and ’internal’) and the ’internal’ components.

Primary objective of this chapter is to make a quantitative estimate of potential pre-

dictability of the Asian monsoon climate on monthly and seasonal time scales. Many

studies in the past [Madden, 1976, 1981; Madden and Shea, 1978; Shea and Madden, 1990;

Shukla and Gutzler, 1983; Short and Cahalan, 1983] have estimated potential predictability

of the extratropical climate from observations. Following the pioneering work of Char-

ney and Shukla [1981], some others (e.g. Singh and Kriplani 1986) also have attempted

to estimate the potential predictability of the Indian summer monsoon. Due to dif-

ferences in the methodology used and due to inhomogeneity of data used in different

studies, it has been difficult to arrive at an universal conclusion regarding the quan-

titative measure of predictability over different geographical locations in general and

the Indian monsoon region in particular. With the availability of long term record of

homogeneous atmospheric circulation data for over 40 years (e.g. from NCEP/NCAR

Reanalysis), it is now worthwhile to re-examine the quantitative measure of potential

predictability in the tropics. While potential predictability over the global tropical belt

will be estimated, the predictability of Asian monsoon region will be contrasted with

that over the other tropical regions. In particular, we shall try to assess the contribution

of the intraseasonal oscillations to the potential predictability.

4.2 Estimation of Potential Predictability of Monthly Means

4.2.1 Methodology

The main data used in this study are the daily low-level zonal winds (850 hPa), up-

per level zonal winds (200 hPa) and 700 hPa geopotential height from NCEP/NCAR

reanalysis for 33 years (1965-1997). Daily interpolated outgoing long wave radiation

from the NOAA polar orbiting satellites for 20 years (1980 to 1999) were also used. Our

methodology is based on the following premise. The anomalies associated with the

Page 79: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 59

synoptic and intraseasonal oscillations may be defined as the deviations from the an-

nual cycle. The annual cycle at any place can be defined by the sum of the first few

harmonics. In the present study, the annual cycle is defined as the sum of the first three

harmonics of daily data for each year. The annual cycle defined in this manner varies

from year to year. An example of such interannual variations of the annual cycle of

low-level zonal winds at a point over the Indian Ocean is shown in Figure 4.1. It is clear

that the annual cycle has significant year to year variations. We hypothesize that the

interannual variations of the annual cycle are essentially forced by the slowly varying

boundary forcing. The dominant slowly varying boundary forcing in the tropics is that

associated with the El Nino and Southern Oscillation (ENSO) related SST variations.

Since the time scale of variations of the boundary forcing is much longer (3-4 years to

decadal) than that of the annual cycle, it essentially modulates the annual cycle. Thus,

the interannual variations introduced by the ’external’ (slowly varying) forcing can be

estimated from the monthly means constructed from the deviations of the individual

annual cycles from the climatological mean annual cycle. The annual cycle of zonal

winds at 850 hPa and 200 hPa and geopotential height at 700 hPa for all years from

1965 to 1997 and those for OLR for all years from 1980 to 1999 are first calculated. From

the daily annual cycles, climatological mean daily annual cycles of each field is calcu-

lated. Monthly ’external’ anomalies are estimated as monthly means of deviations of

individual annual cycles from the climatological mean annual cycle.

If daily anomalies in a particular year is defined as the departure of daily obser-

vations from the annual cycle of that year, they represent the ’internal’ contribution

as the ’external’ component represented by the interannual variation of the annual cy-

cle is removed in this process. Thus, the monthly means of the daily anomalies con-

structed in this manner represent the ’internal’ component. This definition implies that

averaged over the whole year, the daily anomalies vanish. However, due to the in-

traseasonal oscillations, the monthly means are non-zero. Our definition of ’internal’

monthly anomaly implies that it is contributed primarily by the intraseasonal oscilla-

tions and any ’climate noise’ arising from higher frequency weather events is neglected.

The ’internal’ and ’external’ monthly mean anomalies calculated in this manner are sta-

tistically independent as the temporal correlation between the two is nearly zero every-

where (figure not shown).

Page 80: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 60

Figure 4.1: An illustration of variations of the annual cycle from year to year. The annual cycleof zonal winds (ms−1) at 850 hPa at a point (80◦E, 5◦N) are shown for 20 years.

Page 81: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 61

Let us define total monthly anomaly of any field (say, zonal wind ) as sum of monthly

anomalies associated with ’internal’ and ’external’ components.

UT (x, y, t) = UE(x, y, t) + UI(x, y, t)

where subscripts E and I refer to the ’external’ and the ’internal’ components. Squar-

ing both sides and summing over all months we can write the total variance to be

given by sum of variances associated with the ’internal’ and the ’external’ components,

namely

σ2T = σ2

E + σ2I ,

as the correlation between the ’internal’ and the ’external’ components is zero. The

total interannual variance may be estimated in two ways. The traditional way of calcu-

lating it is to construct monthly mean data from the raw daily data. Then construct a

climatological monthly mean annual cycle. Deviations of the monthly means from this

climatological monthly mean annual cycle are the total monthly mean anomalies. The

total interannual variance may be calculated from these total anomalies. Alternatively,

daily anomalies can be constructed with respect to the daily climatological mean an-

nual cycle. The monthly means obtained from these daily anomalies give us the total

monthly mean anomalies.

Let U(m,n) represent any field for the nth day of the mthth year, where n= 1,2...365;

m= 1,2...Y. The annual cycle (Ua(m,n)) is defined as the sum of the first three harmonics

of daily data.

To find external monthly anomalies:

Daily climatological mean of the annual cycles is defined as

Uca(n) =1Y

Y∑m=1

Ua(m,n) (4.1)

Daily ’external’ anomaly is defined as

U(m,n) = Ua(m,n)− Uca(n) (4.2)

Monthly mean of ’external’ anomalies

UE(m, k)k=1..12 =130

30∗k∑n=1+30∗(k−1)

U(m,n) (4.3)

Page 82: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 62

To find ’internal’ monthly anomalies:

Daily ’internal’ anomaly is defined as

U(m,n) = U(m,n)− Ua(m,n) (4.4)

Monthly mean of ’internal’ anomalies

UI(m, k)k=1..12 =130

30∗k∑n=1+30∗(k−1)

U(m, n) (4.5)

To find ’total’ monthly anomalies:

Daily climatological mean is defined as

Uc(n) =1Y

Y∑m=1

U(m, n) (4.6)

Daily ’total’ anomaly is defined as

UT (m,n) = U(m,n)− Uc(n) (4.7)

Monthly mean of daily anomalies

U′(m, k)k=1..12 =

130

30∗k∑n=1+30∗(k−1)

UT (m,n) (4.8)

To test our claim that the ’external’ anomalies as estimated by us are essentially

driven by slowly varying SST changes associated with the ENSO, we carried out a com-

bined EOF analysis of the monthly mean ’external’ anomalies of OLR and winds at 850

hPa. We have chosen the period between 1979 to 1997 for this analysis. The dominant

EOF explaining about 20 percent of the total variance is shown in Figure 4.2. The spatial

patterns of both OLR and low-level winds correspond well with the canonical patterns

associated with ENSO [Rasmusson and Carpenter, 1982; Wallace et al., 1998]. The prin-

cipal component for the dominant EOF, PC1 (normalized by its own temporal S.D) is

also shown in Figure 4.2 together with normalized Nino3 SST anomalies. The correla-

tion coefficient between PC1 and Nino3 (160◦W-90◦W, 5◦S-5◦N) SST anomalies is 0.84

indicating a strong link between the variability represented by the ’external’ component

and the ENSO.

Page 83: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 63

Figure 4.2: First combined EOF of mean monthly ’external’ anomalies for the period January1979 to December 1997 (228 months). (a) Zonal winds EOF at 850 hPa, (b) OLR EOF and (c) PC1(solid line) and Nino3 SST anomalies (dashed line). Both the time series are normalized by theirown standard deviation. Units of the EOFs are arbitrary.

Page 84: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 64

Figure 4.3: Time-longitude section of mean monthly ’external’ anomalies of zonal wind at 850hPa (ms−1) and OLR (Wm−2) averaged around equator (5◦S-5◦N).

Page 85: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 65

The second EOF and corresponding time coefficients (PC2) are not shown. However,

PC1 and PC2 are strongly correlated at a lag of about 6 months. This lag-correlation

together with the spatial patterns of the ’external’ component represent an east-ward

propagation of the anomalies, again characteristic of the ENSO anomalies. Therefore,

the ’external’ component separated here clearly represents the slow response of the at-

mosphere to the slowly varying SST forcing associated with the ENSO. Actual anoma-

lies of low-level winds and OLR along the equator associated with the slow external

forcing are shown in Figure 4.3. The magnitude of the anomalies during the warm

and cold events are similar to those known to be associated with typical warm or cold

phases of ENSO [Rasmusson and Carpenter, 1982] and the eastward propagation is also

clearly seen.

4.2.2 Estimation of ’Internal’ and ’External’ Interannual Variances

The total variance of monthly means as well as the ’internal’ and ’external’ com-

ponents of the variance of zonal winds at 850 hPa are calculated as described in the

previous section based on daily data for 33 years (1965-1997). The three variances are

shown in Figure 4.4. Similarly, the three variances for OLR are calculated based on

available 20 years of daily data (1980-1999). The OLR variances are shown in Figure

4.5. To start with, we note that the sum of the ’external’ and ’internal’ variances almost

exactly equals the total variances in all geographical locations in the tropics for both

the field. Secondly, it is clear from Figure 4.4(b) and Figure 4.5(b) that the geographi-

cal distribution of the ’external’ variances of low-level zonal winds as well as OLR has

the canonical pattern of the individual fields associated with the ENSO [Wallace et al.,

1998; Philander, 1990; Rasmusson and Wallace, 1983]. The ’external’ variance of U850 has

a major maximum centered around the dateline and a secondary maximum in the east-

ern equatorial Indian Ocean. Both the regions are known to be associated with large

zonal wind anomalies during peak ENSO phases. The major maximum on the ’exter-

nal’ variance of OLR is also centered around the dateline but has large extension to

the eastern Pacific. It is also noted that most of the appreciable ’external’ variance of

either OLR or U850 is confined between 10◦N and 10◦S, characteristic of the Walker re-

sponse associated with the ENSO. On the other hand, the ’internal’ variances of U850

have large amplitude (Figure 4.4(c)) in the ’monsoon’ regions of the tropics, namely the

Indian summer monsoon region, the South China Sea monsoon region and the Aus-

tralian monsoon region. We note that the ’internal’ variance is generally smaller than

Page 86: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 66

that of the ’external’ variance in the tropical Pacific. However, it could be comparable or

even larger than the ’external’ variance in the monsoon regions mentioned above. The

’internal’ variance associated with the OLR (Figure 4.5(c)) also have large amplitude in

same monsoon regions. In contrast to the ’external ’ variance, the ’internal’ variance is

not confined to the equatorial belt but extends even up to 30◦ latitude in the Indian and

Australian monsoon regions.

Figure 4.4: Monthly variance of zonal winds (m2s−2) at 850 hPa based on 396 months for theperiod January 1965 to December 1997. (a) Total variance (b) ’external’ variance and (c) ’internal’variance.

Page 87: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 67

Figure 4.5: Same as Figure 4.4 but for OLR for the period January 1980 to December 1999 (240months). Units, (Wm−2)2.

Page 88: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 68

4.2.3 Potential Predictability of Monthly Means

Ideally, the potential predictability of either monthly or seasonal means climate is

determined as the ratio between ’signal’ to ’noise’, the signal being the predictable ’ex-

ternal’ component while the ’noise’ being the ’internal’ unpredictable component. Since

it is normally difficult to separate the ’external’ component from the ’internal’ compo-

nent, usually potential predictability is defined as the ratio (F-ratio) between total vari-

ance (σ2) and climate noise (σ2i ). In finding the potential predictability of the monthly

means, since, we have separated the ’external’ and ’internal’ components, we can write

F =σ2

σ2i

=σ2

e

σ2i

+ 1.

Larger the value of this ratio compared to two, higher the predictability. The F-ratio

of two also signifies that the signal-to-noise ratio (i.e F-1) is equal to one and that half

of the observed interannual variability is potentially predictable. The monthly mean

climate may be considered marginally predictable if ’F’ is greater but of the order two.

If ’F’ is less than two, the climate would be unpredictable as the ’internal’ variability

exercises a dominating influence on the total monthly variability. Figure 4.6 represents

the geographical distribution of potential predictability for U850. The F-ratio for zonal

winds at 850 hPa for summer (JJA) months is shown in Figure 4.6(a), while for winter

(DJF) months are shown in Figure 4.6(b). Potential predictability is high wherever the

ENSO influence is large in the summer months (Figure 4.6(a)). These include equatorial

Pacific between 10◦S and 10◦N, equatorial Atlantic and equatorial Indian Ocean east

of 70◦E. Parts of Africa also indicate high predictability as this region is also known to

have strong influence of ENSO. It may be noted from Figure 4.6(a) and Figure 4.6(b)

that during the NH summer, not only the peak values of the ’F’ are higher than those

during northern winter, the area covered by ’F’ greater than two is much larger during

NH summer compared to that in NH winter. Thus, during NH winter the monthly

mean predictability not only decreases compared to that in NH summer, the predictable

region also shrinks. Over the Indian monsoon region ’F’ ratio ranges between 2 and 3

during NH winter and goes even below 2 during NH summer.

The qualitative difference in the predictability regimes during NH summer com-

pared to NH winter is probably not very surprising if we take into account the sea-

sonality of the ’external’ and the ’internal’ variances. As the ’external’ component of

the variance arises from a slowly varying signal (with time scales longer than a year),

Page 89: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 69

Figure 4.6: Estimates of ’F’ ratios for zonal winds at 850 hPa (a) for all northern hemi-sphere summer months (June-July-August) and (b) for all northern hemisphere winter months(December-January-February).

we do not expect much seasonality in the ’external’ variance. This is shown in Figure

4.7 for zonal winds at 850 hPa. Except that the maximum variance occurs in the west-

ern Pacific during NH summer compared to central Pacific during winter, the general

pattern of ’external’ variance is similar in the equatorial wave-guide during both the

seasons. The major difference between the ’external’ variance between the two seasons

occur in the central Pacific subtropics. This is due to the ENSO induced off equatorial

response being much stronger during the NH winter than in the NH summer. How-

ever, the ’internal’ variance has a pronounced seasonality (Figure 4.8). Barring Indian

monsoon region and a small region in the American monsoon region, the internal vari-

ability is very week throughout the equatorial wave-guide during NH summer. This

explains the larger magnitude and extension of ’F’ during NH summer (Figure 4.6(a)).

On the other hand, the ’internal’ variance during NH winter are quite strong from In-

dian Ocean to central Pacific, the maxima being over the Australian monsoon region

and the South Pacific Convergence Zone (SPCZ). The larger ’internal’ variability during

NH winter is consistent with the fact the ISO activity in tropics is strong during boreal

winter and spring and weak during boreal summer except over the Indian monsoon re-

gion [Madden and Julian, 1994; Wang and Rui, 1990]. Even though the ’external’ variance

Page 90: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 70

Figure 4.7: The ’external’ variance of zonal winds at 850 hPa (m2s−2) during (a) NH summermonths (JJA) and (b) NH winter months (DJF).

Figure 4.8: The ’internal’ variance of zonal winds at 850 hPa (m2s−2) during (a) NH summermonths (JJA) and (b) NH winter months (DJF).

Page 91: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 71

remains similar in magnitude and extent in winter compared to those in summer, the

’F’ ratio becomes smaller and the predictable region reduces to a smaller region in the

far eastern Pacific due to vigorous ’internal’ activity in Indian Ocean and central and

western Pacific.

The estimates of ’F’ ratios for zonal winds at 200 hPa for summer (JJA) months

is shown in Figure 4.9(a), while winter (DJF) months is shown in Figure 4.9(b). Po-

tential predictability is high wherever the ENSO influence is large (see Figure 4.9(a)).

These include equatorial Pacific between 10◦S and 10◦N, equatorial Atlantic and equa-

torial Indian Ocean. Parts of Africa indicate high predictability. Compared to low-level

winds potential predictability is generally high. Over the Indian sub-continent the ’F’

ratio ranges between 2 and 3. For NH winter months (Figure 4.9(b)) predictable region

shrinks compared to that of summer months. Maximum predictability is seen over the

central equatorial Pacific and equatorial Indian ocean. The region having ’F’ ratio be-

tween 2 and 3 occupies a larger region as compared to that during JJA.

Figure 4.10 represents the potential predictability for convection (or precipitation).

The estimates of ’F’ ratios for OLR for NH summer (JJA) months and NH winter (DJF)

months are shown in Figure 4.10(a) and Figure 4.10(b) respectively. It is seen from

Figure 4.10(a) that significant predictable region (e.g.’F’ ≥ 2) for convection (or precipi-

tation) is smaller than that for circulation. This region is mainly confined to the central

and eastern equatorial Pacific coincident with the core predictable region of ENSO in-

fluence. The geographical distribution of potential predictability for OLR for NH winter

months is shown in Figure 4.10(b). The predictable region gets confined to central and

east equatorial Pacific. The noteworthy feature is that over the Indian monsoon region,

’F’ ratios are less than two for convection. This indicates that the internal variability in

the Indian monsoon region is even stronger than the potentially predictable ’external’

component seriously limiting the potential predictability of the Indian summer mon-

soon.

The estimates of ’F’ ratios for geopotential height at 700 hPa for NH summer (JJA)

and NH winter (DJF) months are shown in Figure 4.11(a) and Figure 4.11(b) respec-

tively. In contrast to the other fields discussed earlier such as U850, U200 and OLR, the

geopotential field at 700 hPa does not show a major maximum only over the central

equatorial Pacific. The whole tropical belt (10◦S to 10◦N) shows high values of poten-

tial predictability and it is high in both the summer (Figure 4.11(a)) and winter (Figure

4.11(b)) months. During the summer months, southern India shows high potential pre-

Page 92: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 72

Figure 4.9: Estimates of ’F’ ratios for zonal winds at 200 hPa (a) for all northern hemispheresummer months (JJA) and (b) for all northern hemisphere winter months (DJF).

Figure 4.10: Estimates of ’F’ ratios for OLR (a) for all northern hemisphere summer months(JJA) and (b) for all northern hemisphere winter months (DJF).

Page 93: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 73

Figure 4.11: Estimates of ’F’ ratios for geopotential height at 700 hPa (a) for all northern hemi-sphere summer months (JJA) and (b) for all northern hemisphere winter months (DJF).

dictability while in the northern India and over the monsoon trough ’F’ ratio ranges

between 4 and 6. In the winter months also, ’F’ ratios are high in the tropical belt. Both

in summer and winter months ’F’ ratio becomes less between 20◦and 30◦latitudes.

Since the geographical distribution of potential predictability of geopotential height

is different from the other fields like zonal winds and convection, it might be interesting

to look into the external and internal variances separately. In order to highlight the

variance of the geopotential height in the tropics, the variances shown in Figure 4.12

and Figure 4.13 is restricted between 20◦S and 20◦N. This is because the variances of

geopotential height in the extratropics tend to be several times larger than those in the

tropics. The external variance of geopotential height is shown in Figure 4.12 for JJA

and DJF months. The variance associated with the external component is quite high

up to 120◦E though some parts of Africa is showing lower variance. East equatorial

Pacific also shows appreciable variance. For the winter months also, the variance up

to 120◦E is high. Over the Pacific, the peak shifts towards central Pacific. The spatial

pattern of external variance of Z700 appears to have a wave number two structure. This

may be associated with the externally forced interannual variations of divergent Walker

circulation.

Page 94: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 74

Figure 4.12: The ’external’ variance of geopotential height at 700 hPa (gpm2) during (a) NHsummer months (JJA) and (b) NH winter months (DJF).

Figure 4.13: The ’internal’ variance of geopotential height at 700 hPa (gpm2) during (a) NHsummer months (JJA) and (b) NH winter months (DJF).

Page 95: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.2 Estimation of Potential Predictability of Monthly Means 75

The geographical distribution of the variance of the geopotential height associated

with the internal component for the summer and winter months is shown in Figure

4.13. In the summer months, internal variance is low in the entire tropical belt. While

for the winter months internal variance values are nearly two times as high as those

in summer months. The seasonal variation of internal variance is consistent with the

observation that the intraseasonal oscillations in the equatorial region are stronger in

the boreal winter as compared to the boreal summer. Here too, the variance values are

high towards the midlatitudes (not shown in Figure 4.13). The high external variance

and the low internal variance in the tropics explain the high potential predictability in

the tropical belt for geopotential height (Figure 4.11).

What is responsible for the ’internal’ variability of the monthly means in the trop-

ics? The synoptic disturbances in the tropics are much less energetic than their extra-

tropical counterpart. Therefore, nonlinear interaction amongst the tropical synoptic dis-

turbances are unlikely to result in significant energy in the low frequency regime (e.g.

monthly and seasonal means). Moreover due to their higher frequency, the monthly

mean residuals from them are expected to be small. Therefore, the internal variability

that could influence tropical monthly means are the monsoon ISOs during NH summer

and the MJO in the other parts of the tropics. To test the correctness of this conjec-

ture, we calculate ’internal’ variance after removing the synoptic disturbances from the

daily anomalies. For this purpose, a Butterworth low-pass filter that keeps all periods

greater than 10 days and throws out all periods shorter than 10 days was applied on the

daily anomalies of all years after removing the annual cycle of each individual years.

Monthly mean anomalies, describing the ’internal’ component, are again calculated by

averaging the filtered anomalies over calendar months. The ’internal’ variance calcu-

lated from the monthly means of the filtered data has no contribution from the synoptic

variations and is solely contributed by the ISOs. The ’internal’ variance calculated in

this manner for U850 and OLR are shown in Figure 4.14. A comparison of Figure 4.14(a)

with Figure 4.4(c) and Figure 4.14(b) with Figure 4.5(c) reveals that removal of the con-

tribution of the synoptic disturbances from the daily data had no effect on the ’internal’

variance either in magnitude or in spatial distribution. This analysis establishes that the

’internal’ variability of the monthly means is entirely governed by the tropical ISOs.

Page 96: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 76

Figure 4.14: The ’internal’ variance of (a) zonal winds at 850 hPa (m2s−2) and (b) OLR (Wm−2)2

based on all months after removing the higher frequencies with period shorter than 10 days.

4.3 Potential Predictability of Seasonal means

In this section, we define climate by seasonal mean and examine potential pre-

dictability of seasonal mean climate. The ’climatic signal’ may arise from influences

truly external to the climate system or it may arise from slowly varying modes of the

entire climate system. An example of the latter is the El Nino and Southern Oscillation.

The day to day fluctuations or ’weather’ could give rise to variation of the seasonal

mean through scale interaction. In tropics, day to day fluctuations of weather is rather

weak, but the intraseasonal oscillations are strong. Hence the climate noise is mainly

contributed by the scale interaction between weather disturbances and the ISOs. Since

a season is significantly long compared to the typical time scale of the ISOs (30-60 days),

the ’climate noise’ arising due to the ISOs cannot be estimated by simple statistical av-

eraging (as we did in the case of monthly means) but may be estimated by some kind

of low frequency extension of high frequency spectrum. The focus of this section is to

find out whether there is significant difference between interannual variations of the

climatic states that can be distinguished from the climate noise.

Trenberth [1984a, b] has described a method to estimate the ’climate noise’ as the low

frequency extension of the high frequency component. We follow this method to find

Page 97: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 77

an estimate of potential predictability of seasonal mean in the tropics, for the Northern

Hemisphere summer and winter seasons. The methodology is explained in detail in

the Appendix (section 4.5). The potential predictability is defined as the ratio between

interannual variance of the seasonal means and the ’climate noise’. The potential pre-

dictability of NH summer and NH winter seasons for low-level zonal winds, upper

level zonal winds, OLR and geopotential height have been estimated. This part of the

our study is not quite new except that we make use of a long homogeneous data set and

that we focus on the potential predictability of the Indian monsoon region.

Figure 4.15: Estimates of ’F’ ratios for zonal winds at 850 hPa for (a) NH summer season (JJA)(b) NH winter season (DJF).

Figure 4.15 shows the geographical distribution of potential predictability for low-

level zonal winds (850 hPa) in NH summer and NH winter seasons. In NH summer,

regions where the ENSO influence is large shows high predictability. The potential pre-

dictability is maximum in the western equatorial Pacific, and is having an eastward ex-

tension over the central and eastern Pacific and equatorial Atlantic. Parts of Africa and

eastern equatorial Indian ocean also shows high potential predictability. In NH winter,

the maximum shifts towards central equatorial Pacific, but the pattern remains more or

less similar. It is noteworthy that the Indian monsoon region have potential predictabil-

ity values of the order of 1.5 in both the seasons which means that the monsoon climate

Page 98: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 78

is marginally predictable in the summer season. The ’climate noise’ associated with

U850 is shown in Figure 4.16. In the summer months, Asian monsoon region shows sig-

nificant ’internal’ variance. In the winter, variance maxima shifts towards the southern

equatorial Indian Ocean and the Australian monsoon region shows high ’internal’ vari-

ance. This indicate that the interannual variability of the intraseasonal oscillations in the

Indian monsoon region in the NH summer and Australian monsoon region in the NH

winter season is comparable to the predictable component, limiting the predictability

of the Indian and Australian monsoons.

Figure 4.16: Estimates of ’climate noise’ for zonal winds at 850 hPa for (a) NH summer season(JJA) (b) NH winter season (DJF).

Figure 4.17 shows the geographical distribution of potential predictability for upper

level zonal winds (200 hPa) in NH summer and NH winter seasons. Core predictable

regions like the equatorial Pacific, African region and equatorial Atlantic shows high

’F’ ratios both in the summer and winter seasons. Over the Indian monsoon region ’F’

ratio ranges between 1 and 3 in the summer months, while this ratio is between 2 and 4

in the winter months. Thus, the upper level winds during the Asian summer monsoon

are slightly more predictable than the low-level winds. Also south equatorial Indian

Ocean shows high predictability in the winter months for the upper level zonal winds.

Page 99: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 79

Figure 4.17: Estimates of ’F’ ratios for zonal winds at 200 hPa for (a) NH summer season (JJA)(b) NH winter season (DJF).

Figure 4.18: Estimates of ’F’ ratios for OLR for (a) NH summer season (JJA) (b) NHwinter season (DJF).

Page 100: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 80

Figure 4.18 shows the potential predictability distribution of convection (OLR) over

the tropics in NH summer and winter seasons. Predictable regions shrinks in the case

of convection compared to the large scale flow. In the summer season, western and

central equatorial Pacific shows high predictability. Some parts of Africa also come un-

der predictable regions. In the winter season, regions which have predominant ENSO

influence show high predictability. Seasonal mean climate in Indian monsoon region

is marginally predictable in the winter, but the ’F’ ratios are less than two in the sum-

mer season. The convection is even less predictable than low level winds during the

summer monsoon season.

Figure 4.19 shows the potential predictability distribution of geopotential height at

700 hPa over the tropics in NH summer and winter seasons. The ’F’ ratios in the equa-

torial wave-guide is quite high both in the summer and winter seasons. In the both the

seasons south equatorial Indian Ocean shows maximum predictability, though the ’F’

ratios are high in the winter. ’F’ ratios are low as we move up from 10◦latitude. Indian

region shows ’F’ ratios between 3 and 6 for geopotential height. Southern India shows

slightly higher ’F’ ratios. This is consistent with the earlier study done in the region

for the 700 hPa geopotential height [Singh and Kriplani, 1986]. The ’climate noise’ asso-

ciated with geopotential height is much less over the Indian monsoon region (Figure

4.20), compared to interannual variance in both the summer and winter months. This

explains, the high predictability associated with the geopotential height over the Indian

monsoon region.

Page 101: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.3 Potential Predictability of Seasonal means 81

Figure 4.19: Estimates of ’F’ ratios for geopotential height at 700 hPa for (a) NH summer season(JJA) (b) NH winter season (DJF).

Figure 4.20: Estimates of ’climate noise’ for geopotential height at 700 hPa for (a) NH summerseason (JJA) (b) NH winter season (DJF).

Page 102: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.4 Discussions and Conclusions 82

4.4 Discussions and Conclusions

In the present study, we attempt to determine the part of monthly and seasonal

mean climate variability governed by ’internal’ dynamics and that governed by ’exter-

nal’ slowly varying forcing from long daily observations. Potential predictability of the

climate (monthly and seasonal means) is defined as the ratio of the interannual variance

of the monthly or seasonal means and the ’internal’ unpredictable component. Four dif-

ferent fields (low-level zonal winds (850 hPa), upper level zonal winds (200 hPa), OLR

and geopotential height at 700 hPa) are used for this purpose. Daily U850, U200 and Z700

are taken from NCEP/NCAR Reanalysis for a period of 33 years (1965-1997). Daily

OLR for 20 years (1980-1999) are also used.

The monthly mean climate over the monsoon regions of the world appear to be

marginally predictable. But the ’F’ ratios ranges between 2 and 3 over the Indian mon-

soon region which is much less compared to that in other regions in the tropics. In many

recent studies, the difficulty in simulating and predicting the Indian summer monsoon

has been attributed to the role of the ISOs [Webster et al., 1998; Goswami, 1998, 1995].

In Goswami [1998], it was shown that the strength of the GCM simulated ENSO re-

sponse decreases as we reach the Indian Ocean and Indian monsoon region and the in-

ternal variability could compete with the externally forced variability in this region. The

present analysis shows, from observation that the internal variability in the Indian sum-

mer monsoon region is indeed comparable to the boundary forced variability. However

the fact that the F-ratio ranges between 2 and 3 indicates that the external forced pre-

dictable signal is slightly larger than the noise in some regions. Therefore, while de-

terministic prediction of the monthly mean summer monsoon climate may prove to be

difficult, there exists some hope of limited predictability coming from the boundary

forcing.

The other important result is that except over the Asian summer monsoon region,

the monthly mean climate during the boreal summer is more predictable over a much

larger region in the tropics than during boreal winter. As it is well known that the SST

signal associated with the ENSO tends to peak during NH winter, it appeared rather

strange that predictability should be weaker during this season. However, we show

that the weaker and limited predictability during boreal winter is due to stronger in-

ternal variability associated with stronger ISOs during winter while the amplitude of

the boundary forced variability remains similar to those in boreal summer. Thus, the

Page 103: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.4 Discussions and Conclusions 83

monthly mean tropical climate seems to be more predictable in NH summer compared

to NH winter over much of the tropical belt except in the Indian summer monsoon

region.

The predictability of the seasonal mean climate over the Indian monsoon region

also appear to be marginal. The ’F’ ratio which is a measure of potential predictability

is of the order of 1.5. As in the case of monthly mean climate, the Asian monsoon re-

gion is the region of lowest predictability of the seasonal climate during boreal summer.

Barring the Indian monsoon region, most of the regions in the equatorial wave guide

seem to be highly predictable. Equatorial Pacific are associated with higher predictabil-

ity values. Not surprisingly regions that come under the influence of ENSO have high

predictability.

As may be expected, the geographical distribution of potential predictability of

the monthly and seasonal mean climate bear similarity in all the fields. Comparison

between Figure 4.6(a) and Figure 4.15(a) reveal that the core predictable regions of

monthly mean climate in the summer months and that of the seasonal mean climate

in the summer season is the same for low-level zonal winds. Equatorial Pacific, equato-

rial Atlantic, south equatorial Indian Ocean and the African region seems to be highly

predictable in both the cases. Over the Indian monsoon region, ’F’ ratios are of the order

of two in the monthly mean climate, while the ratios of the order of 1.5 in the seasonal

mean. If we compare Figure 4.6(b) and Figure 4.15(b), it is clear that ’F’ ratios are much

larger in the central equatorial Pacific for the seasonal mean winter climate compared

to the monthly mean climate in the winter months. Some parts of Africa, equatorial In-

dian Ocean and equatorial Atlantic comes under predictable regions in both the cases.

Over the Indian monsoon region, ’F’ ratio is of the order of two in the monthly mean

climate, while it is of the order of 1.5 in the seasonal mean winter climate for low-level

zonal winds. Thus, it appears that the seasonal mean summer monsoon may be more

difficult to predict compared to the monthly means of monsoon during boreal summer.

It may be noted that, of the four fields used in this study, low-level zonal winds at

850 hPa, upper level zonal winds at 200 hPa and OLR shows some what similar charac-

teristics in both monthly and seasonal mean potential predictability. But the geographi-

cal distribution of potential predictability of geopotential height at 700 hPa shows high

potential predictability over almost the whole tropical belt. Within the tropics, the In-

dian summer monsoon region does show relatively lower potential predictability dur-

ing boreal summer compared to rest of the tropics (Figure 4.11(a) and Figure 4.19(a)).

Page 104: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.4 Discussions and Conclusions 84

However, the geopotential height seem to be predictable even over the Indian monsoon

region. The difference in the potential predictability of the geopotential height and

the circulation and convection fields is not surprising as the geopotential field is not

strongly coupled to circulation field as in the extratropics. In the tropics, the transient

disturbances (that give rise to internal variability) are driven not by available potential

energy associated with mean temperature gradient but by potential energy associated

with convection. That is why predictability is poorest for convection (OLR) and increas-

ingly higher for low level and upper level winds. Therefore, it is incorrect to conclude

that Indian monsoon is predictable by simply looking at the geopotential height field.

One need to look at the circulation, convection and precipitation fields to arrive at the

correct picture of predictability of the monsoon.

Page 105: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.5 Appendix : Procedure for Estimating ’Climate Noise’ 85

4.5 Appendix : Procedure for Estimating ’Climate Noise’

The day to day fluctuations or ’weather’ could give rise to variation of the seasonal

mean through scale interaction. This is often termed as ’climate noise’. This ’climate

noise’ has to be estimated from low frequency extension of the high frequency compo-

nent. We follow the method suggested by [Trenberth, 1984b] compute the ’climate noise’

of seasonal means.

First step is to remove the annual cycle. Daily climatological mean annual cycle for

the entire data has been found out for this purpose. The daily climatological mean is

smoothed using harmonic analysis. Daily anomalies are constructed with respect to the

smoothed daily climatological mean annual cycle. The daily anomalies are detrended

using a least squares linear fit.

Suppose that, the data base consists of N daily values (χi,j , i = 1, ..N, j = 1, ..J)

that make up the season for J years in which the mean and annual cycles have been

removed. The problem is to assess whether there is any significant climate variability

beyond that due to climatic noise.

For each year, mean χj may be computed

χj =1N

N∑i=1

χi,j (4.9)

and

S2m =

1J

J∑j=1

χ2j (4.10)

is the sample interannual variance; an unbiased estimate (ˆ) of the population interan-

nual variance which includes the effects of uncertainty in the overall mean, is therefore

σ2m =

J

J − 1S2

m =1

J − 1

J∑j=1

χ2j . (4.11)

The noise may be found as a low frequency extension of high frequency variability.

The intraseasonal sample variance for the jth year is

S2j =

1N

N∑i=1

(χi,j − χj)2. (4.12)

Therefore, the mean intraseasonal variance is

S2 =1J

J∑j=1

S2j (4.13)

Page 106: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

4.5 Appendix : Procedure for Estimating ’Climate Noise’ 86

An unbiased estimate of σ2 based solely upon the intraseasonal variances, is

σ2 =N

N − ToS2. (4.14)

The variance due to climatic noise σ2N is

σ2N =

σ2

Neff=

σ2To(N)N

(4.15)

where To is the time between independent values normalized by the sampling inter-

val. The effective number of independent observations Neff = N ∆TTo

where ∆T is the

sampling interval. Therefore from (4.14) and (4.15)

σ2N =

To

N − ToS2. (4.16)

To is not known apriori and is dependent upon autocorrelation, rL

To(N) = 1 + 2N∑

L=1

(1− L

N)rL. (4.17)

where rL is the autocorrelation with lag L of the data.

To find rL

CLj =1N

N∑i=L+1

(χi−L,j − χj)(χi,j − χj), (4.18a)

Sample autocorrelation at lag L is

rLj = CLj/Coj (4.18b)

and the overall autocorrelation, rL is

rL =1J

J∑j=1

rLj (4.18c)

One way to test whether there is any signal is to form the null hypothesis that there

is no signal. In that case σ2m and σ2

N are both independent estimates of interannual

variance. The former is based upon seasonal means, while the latter is based upon

intraseasonal variations. Consequently, the F ratio defined as

F =σ2

m

σ2N

(4.19)

is the ratio of the two estimated interannual variances and it should follow the F distri-

bution with J-1 and J(Neff − 1) degrees of freedom.

Page 107: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 5

Clustering of Synoptic Systems During theIndian Summer Monsoon by IntraseasonalOscillations

As shown in chapter 3 and 4, the monsoon ISOs has large spatial scale and results

in strengthening and weakening of the large scale monsoon flow in the extreme phases.

This results in strengthening and weakening of the shear of the zonal wind and low-

level vorticity over the monsoon trough. Since the higher frequency synoptic systems

arise from instability of the zonal flow, ISOs have the potential for modulating synoptic

activity during the monsoon season. In this chapter, we examine how higher frequency

synoptic systems are modulated by the intraseasonal oscillations. The motivation of this

study came from the fact that the slow evolution of ISOs may permit extended range

prediction of the ISO phases and through it probability of occurrence of wet and dry

spells of the monsoon.

5.1 Introduction

A prominent feature of the seasonal mean (June-September) Indian summer mon-

soon circulation is the monsoon trough (Figure 2.1a,b), an elongated semi-permanent cy-

clonic vortex in the lower atmosphere associated with low surface pressure extending

from Pakistan in the west to Myanmar in the east [Rao, 1976]. The summer monsoon is

punctuated by periods of abundant rainfall (’active’ or wet spells) and periods of scanty

rain (’break’ or dry spells) in the trough region. There are three or four active and break

spells each in a typical monsoon season. If long breaks occur in critical growth periods

of agricultural crops, they can lead to substantially reduced yields [Gadgil and Rao, 2000;

Lal et al., 1999]. Extended range prediction of the wet and dry spells of monsoon rain

could therefore be of immense benefit to Indian agriculture.

Page 108: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.1 Introduction 88

The main rain bearing weather systems over the monsoon trough region are syn-

optic scale low pressure systems with typical life time of 3-5 days and length scale of

about 2000 km. Monsoon LPS are called lows if the maximum wind speed is less than

8.5 ms−1 and has one closed isobar with the central pressure in the system being lower

than the surroundings by more than 2 hPa. The maximum wind speed in depressions is

between 8.5 ms−1 and 17 ms−1 and there are atleast two closed isobars, with 4 hPa pres-

sure drop associated with the system. Most depressions are born in the Bay of Bengal

and give copious rain as they move inland along the monsoon trough. Monsoon lows

and depressions arise as a result of dynamic instability energized by moist convection

[Shukla, 1978; Goswami et al., 1980; Mak, 1987]. Large meridional shear of the eastward

component of winds and high cyclonic vorticity at low levels in the monsoon trough

favor the growth of these instabilities. Intraseasonal oscillations of the Indian summer

monsoon have space and time scales that are distinct from those of synoptic systems.

ISOs have periods of 10-70 days, zonal scale of 8,000-10,000 km, and are associated

with repeated northward propagation of the tropical convergence zone from the south

equatorial Indian Ocean to the monsoon trough region [Sikka and Gadgil, 1980; Yasunari,

1979; Krishnamurti and Ardunay, 1980]. As the ISOs modulate the large scale monsoon

circulation, strengthening the low-level monsoon winds in one phase while weakening

them in the opposite phase (see Chapter 2), they have the potential to modulate synop-

tic activity. Although previous studies do indicate association of synoptic activity with

ISO regimes [Murakami et al., 1984, 1986; Yasunari, 1981], a comprehensive study of the

influence of ISOs on LPS genesis does not exist. Here, using daily circulation data and

LPS genesis data for 40 years we show that the wet and dry spells of the Indian summer

monsoon arise from space-time clustering of the LPS and that the clustering is caused

by modulation of the large scale monsoon circulation by ISOs. Our work implies that

the predictability of the timing of wet and dry spells is strongly tied to the predictability

of the slowly varying monsoon ISOs.

The dates and locations of genesis of all lows and depressions during June-September

of 1954-1993 over the Indian monsoon region (50◦E-100◦E, Eq-35◦N) are based on re-

ports of the India Meteorological Department (IMD). Data for the first 30 years (1954-

1983) are taken from Mooley and Shukla’s [Mooley and Shukla, 1989, 1987] compilation

based on IMD’s Daily Weather Reports; data for the next 10 years (1984-1993) are com-

piled from the Seasonal Weather Summaries published by IMD. For example weather

summary of 1984 monsoon season is from IMD [1985]. Circulation changes associated

Page 109: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.2 Wet and Dry Spells and Clustering of LPS 89

with monsoon ISOs are based on daily 850 hPa wind fields from NCEP/NCAR reanal-

ysis for the period 1954-1993. Rainfall data are based on five-day (pentad) Climate Pre-

diction Center Merged Analysis of Precipitation (CMAP) for fifteen years (1979-1993).

Anomalies are obtained by subtracting the annual cycle (sum of the mean, annual and

semiannual harmonics) from the daily (or pentad) observations for each year.

5.2 Wet and Dry Spells and Clustering of LPS

First, we demonstrate that the wet and dry spells of the monsoon rainfall arise

mainly from the time clustering of LPS. Pentad rainfall anomaly spatially averaged over

the two contiguous boxes (85◦E-95◦E, 12◦N-17◦N) and (70◦E-90◦E, 17◦N-22◦N) during

1979-1993 represents the rainfall over the monsoon trough, denoted by P. A zero value

of P corresponds to the seasonal mean rainfall over the trough (11.5mm/day). Peri-

ods of positive (negative) P correspond to wet (dry) spells. In Figure 5.1, we mark the

calendar dates of genesis of all LPS in the monsoon trough region between June and

September during 1979-1993 as a function of P normalized by its standard deviation

(4.5mm/day). More than two times as many LPS form during periods of positive P

(111systems) compared to periods of negative P (52 systems), clearly showing the close

association between the genesis of lows and depressions and timing of wet and dry

spells. We propose that this clustering of LPS is caused by modulation of the large scale

monsoon circulation by ISOs.

Figure 5.1: Genesis dates of LPS between 1 June and 30 September of all years during 1979 to1993 over the monsoon trough as a function of normalized departure of precipitation over thetrough from the seasonal mean.

Page 110: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.2 Wet and Dry Spells and Clustering of LPS 90

Figure 5.2: Leading Empirical Orthogonal Functions ( (a) EOF1 & (b) EOF2) of 10-80 day fil-tered wind anomalies (ms−1) at 850 hPa between June 1 and September 30 for the period 1964-1973. (c) Normalized time series of PC1 and PC2 for ten years (each year has 122 days). (d)Normalized Monsoon Intraseasonal Oscillation Index (MISI) for 10 years. Periods of MISI > +1(MISI< -1) correspond to active (break) phases of the monsoon. It may be noted that positive(negative) phase of MISI represents enhancement (weakening) of the EOF1 pattern.

Page 111: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.3 Monsoon Intraseasonal Oscillation Index 91

5.3 Monsoon Intraseasonal Oscillation Index

To establish that ISOs influence the genesis of LPS, we define a simple index that cap-

tures intraseasonal variability of circulation in the Indian monsoon region. The space

time evolution of ISOs may be described by the two leading empirical orthogonal func-

tions (EOF1 & EOF2) of 10-80 day band-pass filtered 850 hPa summer monsoon winds (1

June to 30 September for 1954-1963) in the region 40◦E-120◦E, 20◦S-30◦N. Together they

explain 25% of daily variance of the wind field; their principal components PC1 and

PC2 correlate strongly with a lag of about 9 days. The sum of the two EOFs represents

the northward propagating monsoon ISOs. We introduce the monsoon intraseasonal

oscillation index (MISI) based on the first two principal components of the wind field,

MISI(t) = PC1(t)+PC2(t-9).

The spatial structure of winds associated with these EOFs and their corresponding

principal components (PC1 and PC2) and MISI for 10 years (1964-1973) is shown in

Figure 5.2. The spatial structure of winds associated with EOF1 (Figure 5.2(a)) bears a

broad resemblance with the seasonal mean low-level circulation. The positive (nega-

tive) phase of MISI represents circulation anomalies that strengthen (weaken) the mean

monsoon winds (see Figure 2.1(a)) between 5◦N and 17◦N by upto 30%, thereby intensi-

fying (weakening) the monsoon trough. We say that the monsoon is in its active (break)

phase in periods when normalized MISI is greater than +1 (less than -1). The index

(MISI) for the remaining 30 years is constructed taking data for 10 years at a time.

5.4 Clustering of Genesis of LPS by Intraseasonal Oscillations

In order to bring out how genesis of LPS depend on the phase of the ISOs, the fre-

quency of occurrence of the LPS corresponding to the different ISO phases are counted.

ISO phases may be defined as bins of normalized MISI. Such a frequency distribution of

genesis of LPS as a function of the phase of monsoon ISOs during 1954-1993 is shown in

Figure 5.3. The total number of lows and depressions during this 40-year period is 503,

with a seasonal average of 12.5 LPS. We note that the number of depressions during

the last decade (1984-93) is lower than earlier decades [Mooley and Shukla, 1989], while

the number of lows is higher leaving the average number of LPS almost unchanged.

Out of the total 503 LPS, 320 occur in the positive phase of the ISO (positive MISI) and

183 in the negative phase. The enhanced low-level shear and cyclonic vorticity in the

Page 112: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.4 Clustering of Genesis of LPS by Intraseasonal Oscillations 92

−3 −2 −1 0 1 2 30

25

50

75

100

Normalised MISI

Fre

quen

cy

Figure 5.3: Histogram of genesis of synoptic events (lows & depressions) for the Indian mon-soon region (50◦E-100◦E, Eq-30◦N) during June to September for the period 1954-1993 as a func-tion of normalized MISI.

monsoon trough makes LPS genesis in the positive phase more probable. Figure 5.4

shows the total vorticity and locations of genesis of all LPS in the active (MISI > +1)

and break (MISI < -1) phases of the monsoon. The birth of an LPS is more than twice

as likely in the active phase (119 systems) than in the break phase (52 systems), with

dense clustering in the monsoon trough (Figure 5.4(a)). The total vorticity in the trough

region remains weakly cyclonic even during breaks (Figure 5.4(b)), and this explains

why some LPS form here even in the break phase. Relatively few LPS are born in the

southern Bay of Bengal in this phase although the cyclonic vorticity is high. This may be

because the large vertical shear of the winds in the southern region inhibits LPS genesis

[Rao, 1976] and partly because the boundary layer frictional convergence necessary for

cyclogenesis is less effective in this region as compared to the northern region. The cen-

tral result of the present study is that circulation changes associated with the monsoon

ISOs cause lows and depressions to cluster together in both time and space (Figures 5.3,

5.4). Mechanisms similar to the one proposed here seem to be responsible for the clus-

tering of tropical cyclones in the Gulf of Mexico [Maloney and Hartmann, 2000a], eastern

Pacific [Maloney and Hartmann, 2000b] and western Pacific Liebmann et al. [1994] through

modulation of large scale circulation by the Madden Julian Oscillations.

Finally, we show that intraseasonal fluctuations of cyclonic vorticity in the monsoon

trough are associated with coherent fluctuations in the large scale rainfall distribution.

We composite the anomaly winds over all active and break days based on MISI in the

period 1954-1993, and use these to create an active minus break composite of large scale

monsoon circulation and vorticity (Figure 5.5(a)).

Page 113: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.4 Clustering of Genesis of LPS by Intraseasonal Oscillations 93

Figure 5.4: Total (climatology+composite anomaly) relative vorticity (10−6s−1) at 850 hPa dur-ing the (a) ’Active’ ISO phase (MISI > +1) and (b) ’Break’ ISO phase (MISI < -1). Dark dotsindicate the position of the genesis of the LPS during active and break phases.

Page 114: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.4 Clustering of Genesis of LPS by Intraseasonal Oscillations 94

Figure 5.5: Composites based on active and break days as defined by the ISO index, MISI.(a) ’Active’ minus ’Break’ composite wind anomalies (ms−1) and associated relative vorticity(10−6s−1) at 850 hPa during the 40 year period (1954-1993). Only vectors significant at 95%confidence level are displayed. Positive contours are shaded and negative contours are notshown. (b) ’Active’ minus ’Break’ composite precipitation anomalies (mm.day−1) during 1979-1993.

Page 115: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

5.5 Summary and Conclusions 95

This composite represents intraseasonal changes in monsoon circulation as captured

by the two leading EOFs of the low-level winds. We also create the corresponding active

minus break rainfall composite based on CMAP rainfall anomalies using the same ac-

tive and break dates in the period 1979-1993 (Figure 5.5(b)). Enhanced cyclonic vorticity

in the monsoon trough region is accompanied by enhanced rainfall. The positive rain-

fall anomaly is mainly due to the larger number of lows and depressions formed in the

trough in the active phase of the monsoon compared to those during the break phase

(Figure 5.4). Decreased precipitation over the equatorial region and the rain shadow

region in southeast India are also evident in the active phase (Figure 5.5(b)). The spatial

pattern of the rainfall composite is consistent with the classical pattern of intraseasonal

monsoon rainfall variability seen in rain guage data over the continent [Singh and Kri-

plani, 1990; Krishnamurthy and Shukla, 2000].

5.5 Summary and Conclusions

The timing and duration of wet and dry spells of the summer monsoon have a strong

bearing on the agricultural production and water resources in the Indian subcontinent.

We show that the wet and dry spells are the result of space-time clustering of mon-

soon low pressure systems caused by modulation of the large scale monsoon flow by

intraseasonal oscillations. The ISOs alternately enhance and reduce horizontal shear

and cyclonic vorticity of low-level winds along the monsoon trough on time scales of

10-80 days. Genesis of LPS is twice as likely in periods when monsoon trough vorticity

is enhanced as compared to periods when it is reduced. There is also a spatial clustering

of LPS genesis, with a majority of LPS being born in north Bay of Bengal in periods of

enhanced monsoon trough vorticity.

Skillful statistical forecasts upto two to three weeks in advance have been demon-

strated for the slow evolution of the equatorially confined, eastward propagating MJO

[Lo and Hendon, 2000; Waliser et al., 1999; Mo, 2001]. We envisage that the slowly varying

monsoon ISOs will turn out to have similar predictability. Work in this direction might

lead to extended range prediction of the wet and dry spells of the Indian summer mon-

soon.

Page 116: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Chapter 6

Conclusions

Indian summer monsoon displays substantial interannual variability, which have

profound socio-economic consequences. Long range prediction of seasonal mean mon-

soon precipitation, therefore assumes great significance. Even though climate mod-

elling has made great progress in simulating and predicting the climate over several

tropical regions, dynamical prediction of seasonal mean monsoon precipitation how-

ever, remains as an extremely frustrating experience.

Within the summer monsoon season (June-September), the timing and duration of

the monsoon intraseasonal oscillations (wet and dry spells of the summer monsoon)

have a strong bearing on the agricultural production and water resources in the In-

dian subcontinent. Monsoon studies so far, has not clearly established whether the

occurrence of wet and dry spells of monsoon rainfall is due to some form of dynami-

cal instability of the mean monsoon flow, or a mere indicator of the formation, growth

and propagation of monsoon depressions, or due to low frequency chaotic intrasea-

sonal oscillations. Research during the past decades, has indicated the possible role of

intraseasonal oscillations as one of the reasons which limits the predictability of the sea-

sonal mean monsoon. In this study, we consider the intraseasonal oscillations (ISOs) as

the building block for Indian summer monsoon. We demonstrate how ISOs influence

the seasonal mean and limits its predictability in one hand while enhancing potential

predictability of the wet and dry spells of the monsoon by modulating the frequency of

occurrence of the synoptic events on the other. Some outstanding questions regarding

relationship between intraseasonal oscillations and interannual variability of the Indian

summer monsoon are addressed.

Page 117: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

6 Conclusions 97

• How could intraseasonal oscillations influence interannual variations of the In-

dian summer monsoon? Is there a common mode of variability between the in-

traseasonal and interannual variability of the Indian monsoon?

• Is there a distinct difference in the probability of occurrence of ’active’ and ’break’

phases’ in the strong and weak monsoon years?

• How much of the interannual variability of the Indian summer monsoon is gov-

erned by ’internal’ chaotic processes? How much of this ’internal’ low frequency

variability is contributed by the monsoon ISOs?

• Is there evidence from observation that occurrence of rain bearing monsoon syn-

optic systems (lows and depressions) are modulated by the intraseasonal oscilla-

tions?

First, we show that the underlying spatial pattern of the dominant intraseasonal

mode is invariant over the years and is similar to the spatial structure of the seasonal

mean monsoon. The dominant ISO is characterized by a meridional bimodal structure

with ascending (descending) motion and enhanced (decreased) convection over the

monsoon trough and descending (ascending) motion and decreased (enhanced) con-

vection over the oceanic TCZ in the ’active’ (’break’) phase. Thus extreme phases of the

dominant ISO mode (’active’ and ’break’ phases) are associated with general strength-

ening (weakening) of large scale mean monsoon flow leading to strengthening (weak-

ening) of the monsoon trough. Hence it is possible that, the statistics of ISO (phase,

amplitude) affect the seasonal mean monsoon. Then, we demonstrate that the intrasea-

sonal and interannual variations are governed by a common spatial mode of variability.

Further it is shown that probability of occurrence of the intraseasonal oscillations is re-

lated to the interannual variability of the seasonal mean. The frequency of occurrence

of ’active’ and ’break’ conditions are found to be distinctly different during ’strong’ and

’weak’ monsoon years. It is shown that the most frequent pattern during a ’strong’

(’weak’) monsoon year is the ’active’ (’break’) pattern with enhanced (decreased) cy-

clonic vorticity and convection over the monsoon trough. All these results lead to the

conclusion that monsoon ISOs modulate interannual variation of the Indian monsoon

in a significant way.

Having shown that the ISOs can influence the seasonal mean and its interannual

variability, we attempt to make quantitative estimates of potential predictability of mon-

Page 118: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

6 Conclusions 98

soon climate. Potential predictability is defined as the ratio between the interannual

variance of the the monthly or seasonal means and its internally forced ’climate noise’

component. We argue that the ISOs contribute mainly to the ’climate noise’ in the trop-

ics as the amplitude of the synoptic disturbances is rather small and are unlikely to

lead to much low frequency internal variability through nonlinear scale interactions.

For monthly climate, we propose a new method to separate the ’internal’ and ’external’

contribution to the interannual variability. For the seasonal climate, the internal ’cli-

mate noise’ is estimated using a method equivalent to low frequency extension of high

frequency spectrum as done in some previous studies. It is found that slowly varying

boundary forcing strongly govern the predictability of monthly or seasonal climate of

most of the tropical regions except the Indian monsoon region. Quantitative estimates

of potential predictability of monthly and seasonal climate reveal that the potential pre-

dictability of the Indian monsoon is much lower compared to the other regions in the

tropics. This is due to the fact that the influence of the internally forced component of

the seasonal mean is comparable to its externally forced counterpart in the Indian mon-

soon region. These estimates reveal that the Indian monsoon climate may be considered

only marginally predictable. We also find that the internally forced component of the

monthly/seasonal climate in the Indian monsoon region is due to the intraseasonal os-

cillations.

The monsoon ISO results in strengthening and weakening of the mean monsoon

flow in the extreme phases. The main rain bearing system during the monsoon season

are the Low Pressure Systems (LPS) consisting of lows and depressions. Since the gene-

sis of the LPS depends on the horizontal shear and low-level vorticity, it is possible that

more LPS form in the active phase relative to the break phase. In other words, large

scale circulation associated with the ISOs could modulate the frequency of genesis of

LPS. Using LPS genesis data for more that 40 years and corresponding circulation data

to describe the ISOs, we show that the dry and wet spells of the Indian monsoon are

caused by clustering of low pressures systems in space and time which is caused by

the modulation of the large scale monsoon flow by the intraseasonal oscillations. The

slow evolution of the ISO may permit extended range prediction of the ISO phases and

through them dry and wet spells of the Indian summer monsoon.

In this study, we have used long homogeneous data sets (30-40 years) to examine

the statistics of the ISO. Hence we hope that the results are reliable. Above results have

important implication on the seasonal mean monsoon prediction. While the monsoon

Page 119: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

6 Conclusions 99

ISOs seem to result in limiting the predictability of monthly or seasonal mean monsoon

climate, it is possible that the same ISOs lead to enhancing extended range prediction

of spells of synoptic activity. As demonstrated in the case of equatorial MJO, extended

range prediction of monsoon ISOs may be possible due to it’s slow evolution. Since

ISOs modulate the main rain bearing systems in the monsoon region, the prediction ISO

phase may lead to predicting the dry and wet spells of the Indian summer monsoon.

Studies in this direction will help in increasing the predictability of the Indian sum-

mer monsoon. Thus, ISOs appear to play a crucial role in determining predictability of

monsoon in different time scales. Hence, the success in predicting the Indian summer

monsoon rainfall would depend on the precise representation of ISOs in a dynamical

model.

Page 120: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography

Ahlquist, J., V. Mehta, A. Devanas, and T. Condo, Intraseasonal monsoon fluctuations

seen through 25 years of Indian radiosonde observations, Mausam, 41, 273–278, 1990.

Annamalai, H., J. M. Slingo, K. R. Sperber, and K. Hodges, The mean evolution and vari-

ability of the Asian summer monsoon: Comparison of ECMWF and NCEP-NCAR

Reanalyses, Mon. Wea. Rev., 127, 1157–1186, 1999.

Brankovic, C., and T. N. Palmer, Atmospheric seasonal predictability and estimates of

ensemble size, Mon. Wea. Rev., 125, 859–874, 1997.

Brankovic, C., and T. N. Palmer, Seasonal skill and predictability of ECMWF PROVOST

ensembles, Quart. J. Roy. Meteor. Soc., 126, 2035–2067, 2000.

Charney, J. G., and J. Shukla, Predictability of Monsoons, in Monsoon Dynamics, edited

by J. Lighthill and R. P. Pearce, pp. 99–108, Cambridge University Press, Cambridge,

1981.

Chatfield, C., The analysis of time series: An introduction, Chapman and Hall, London,

1980.

Dakshinamurthy, J., and R. N. Keshavamurthy, On oscillations of period around one

month in the Indian summer monsoon, Ind. J. Meteorol. Geophys., 27, 201–203, 1976.

De, U. S., and J. C. Natu, Low frequency oscillations in tropospheric winds during con-

trasting summer monsoon over India, Mausam, 45, 261–266, 1994.

Fennessy, M., and J. Shukla, Simulation and predictability of monsoons, in Proceedings

of the international conference on monsoon variability and prediction, WMO/TD 619, pp.

567–575, Trieste, 1994.

Ferranti, L., J. M. Slingo, T. N. Palmer, and B. J. Hoskins, Relations between interannual

Page 121: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 101

and intraseasonal monsoon variability as diagnosed from AMIP integrations, Quart.

J. Roy. Meteor. Soc., 123, 1323–1357, 1997.

Gadgil, S., and P. R. S. Rao, Famine strategies for a variable climate - A challenge., Curr.

Sci., 78, 1203–1215, 2000.

Gadgil, S., and S. Sajani, Monsoon precipitation in the AMIP runs, Climate Dyn., 14,

659–689, 1998.

Goswami, B. N., Dynamical predictability of seasonal monsoon rainfall: Problems and

prospects, Proc. Ind. Nat. Sci. Acad., 60A, 101–120, 1994.

Goswami, B. N., A multiscale interaction model for the origin of the tropospheric QBO,

J. Climate, 8, 524–534, 1995.

Goswami, B. N., Chaos and Predictability of the Indian summer monsoon, Pramana-

Journal of Physics, 48, 719–736, 1997.

Goswami, B. N., Interannual variations of Indian summer monsoon in a GCM: External

conditions versus internal feedbacks, J. Climate, 11, 501–522, 1998.

Goswami, B. N., and J. Shukla, Quasi-periodic oscillations in a symmetric general cir-

culation model, J. Atmos. Sci., 41, 20–37, 1984.

Goswami, B. N., R. N. Keshavamurthy, and V. Satyan, Role of barotropic-baroclinic

instability for the growth of monsoon depressions and mid-tropospheric cyclones,

Proc. Ind. Acad. Sci. (Earth & Planetary Sciences), 89, 79–97, 1980.

Goswami, B. N., D. Sengupta, and G. Sureshkumar, Intraseasonal oscillations and Inter-

annual variability of surface winds over the Indian monsoon region, Proc. Ind. Acad.

Sci. (Earth & Planetary Sciences), 107, 45–64, 1998.

Gruber, A., and A. F. Krueger, The status of the NOAA outgoing long-wave radiation

data set, Bull. Amer. Meteor. Soc., 65, 958–962, 1984.

Hartmann, D. L., and M. L. Michelson, Intraseasonal periodicities in Indian rainfall, J.

Atmos. Sci., 46, 2838–2862, 1989.

Harzallah, A., and R. Sadourny, Internal versus SST-forced atmospheric variability sim-

ulated by an atmospheric general circulation model, J. Climate, 8, 475–495, 1995.

Page 122: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 102

Hendon, H. H., C. Zhang, and J. D. Glick, Interannual variation of the Madden-Julian

oscillation during austral summer, J. Climate, 12, 2538–2550, 1999.

IMD, Monsoon Season (June-September, 1984), in Mausam, vol. 36, pp. 393–402, India

Meteorological Department, New Delhi, 1985.

Kalnay, E., et al., The NCEP/NCAR 40-year reanalysis project, Bull. Amer. Meteor. Soc.,

77, 437–471, 1996.

Keshavamurthy, R. N., Power spectra of large scale disturbances of the Indian south-

west monsoon, Ind. J. Meteorol. Geophys., 24, 117–124, 1973.

Keshavamurthy, R. N., S. V. Kasture, and V. Krishnakumar, 30-50 day oscillation of the

monsoon: a new theory, Beitr. Phys. Atmo., 59, 443–454, 1986.

Kimoto, M., and M. Ghil, Multiple flow regimes in the northern hemisphere winter. Part

I: methodology and hemispheric regimes, J. Atmos. Sci., 50, 2625–2643, 1993.

Kondragunta, C. R., On the intraseasonal variations of the Asiatic summer monsoon,

Mausam, 41, 11–20, 1990.

Krishakumar, K., M. K. Soman, and K. Rupakumar, Seasonal forecasting of Indian sum-

mer monsoon rainfall, Weather, 50, 449–467, 1995.

Krishnamurthy, V., and J. Shukla, Intraseasonal and interannual variability of rainfall

over India, J. Climate, 13, 4366–4377, 2000.

Krishnamurti, T. N., and P. Ardunay, The 10 to 20 day westward propagating mode and

’breaks’ in the monsoons, Tellus, 32, 15–26, 1980.

Krishnamurti, T. N., and H. N. Bhalme, Oscillations of monsoon system. Part I: Obser-

vational aspects, J. Atmos. Sci., 45, 1937–1954, 1976.

Krishnamurti, T. N., and D. Subrahmanyam, The 30-50 day mode at 850mb during

MONEX, J. Atmos. Sci., 39, 2088–2095, 1982.

Kumar, A., and M. P. Hoerling, Prospects and limitations of atmospheric GCM climate

predictions, Bull. Amer. Meteor. Soc., 76, 335–345, 1995.

Lal, M., K. K. Singh, G. Srinivasan, L. S. Rathore, D. Naidu, and C. N. Tripathy, Growth

and yield responses of soybean in Madhyapradesh, India to climate variability and

change, Agric. Forest Meterol., 93, 53–, 1999.

Page 123: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 103

Latif, M., D. Anderson, T. P. Barnett, M. Cane, R. Kleeman, A. Leetma, J. J. O’Brien,

A. Rosati, and E. K. Schneider, A Review of predictability and prediction of ENSO, J.

Geophys. Res., 103(C7), 14,375–14,393, 1998.

Lawrence, D. M., and P. J. Webster, Interannual variations of the intraseasonal oscillation

in the south Asian summer monsoon region, J. Climate, 14, 2910–2922, 2001.

Liebmann, B., and C. A. Smith, Description of a complete (interpolated) outgoing long-

wave radiation dataset, Bull. Amer. Meteor. Soc., 77, 1275–1277, 1996.

Liebmann, B., H. H. Hendon, and J. D. Glick, The relationship between tropical cyclones

of the Western Pacific and Indian Oceans and the Madden-Julian Oscillation, J. Meteor.

Soc. Japan, 72, 401–412, 1994.

Lo, F., and H. H. Hendon, Empirical extended-range prediction of the Madden-Julian

Oscillation, Mon. Wea. Rev., 128, 2528–2543, 2000.

Lorenz, E. N., Atmospheric predictability experiments with a large numerical model,

Tellus, 43, 505–513, 1982.

Madden, R. A., Estimates of natural variability of time averaged sea level pressure, Mon.

Wea. Rev., 104, 942–952, 1976.

Madden, R. A., A quantitative approach to long range prediction, J. Geophys. Res., 86,

9817–9825, 1981.

Madden, R. A., Reply, Mon. Wea. Rev., 111, 586–589, 1983.

Madden, R. A., and P. R. Julian, Observations of the 40-50 day tropical oscillation: A

Review, Mon. Wea. Rev., 122, 813–837, 1994.

Madden, R. A., and D. J. Shea, Estimates of natural variability of time averaged temper-

ature over the United States, Mon. Wea. Rev., 106, 1695–1703, 1978.

Mak, M., Synoptic-scale disturbances in the summer monsoon, in Monsoon Meteorology,

edited by C. P. Chang and T. N. Krishnamurti, pp. 435–460, Oxford University Press,

New York, 1987.

Maloney, E. D., and D. L. Hartmann, Modulation of hurricane activity in the Gulf of

Mexico by the Madden-Julian Oscillation, Science, 287, 2002–2004, 2000a.

Page 124: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 104

Maloney, E. D., and D. L. Hartmann, Modulation of Eastern North Pacific hurricanes by

the Madden-Julian Oscillation, J. Climate, 13, 1451–1460, 2000b.

Mehta, V. M., and T. N. Krishnamurti, Interannual Variability of 30-50 day wave motion,

J. Meteor. Soc. Japan, 66, 535–548, 1988.

Mo, C. K., Adoptive filtering and prediction of intraseasonal oscillations, Mon. Wea.

Rev., 129, 802–817, 2001.

Mooley, D. A., and J. Shukla, Tracks of low pressure systems which formed over India, ad-

joining countries, the Bay of Bengal and Arabian Sea in summer monsoon season during the

period 1888-1983, Center for Ocean Land Atmosphere Studies, Calverton, MD 20705,

USA, 1987, avaialble from J. Shukla, COLA, USA.

Mooley, D. A., and J. Shukla, Main features of the westward-moving low pressure sys-

tems which form over the Indian region during the summer monsoon season and

their relation to the monsoon rainfall, Mausam, 40, 137–152, 1989.

Murakami, M., Large scale aspects of deep convective activity over the GATE area, Mon.

Wea. Rev., 107, 994–1013, 1979.

Murakami, T., and T. Nakazawa, Tropical 45 day oscillations during 1979 northern

hemisphere summer, J. Atmos. Sci., 42, 1107–1122, 1985.

Murakami, T., T. Nakazawa, and J. He, On the 40-50 day oscillation during 1979 north-

ern hemisphere summer. Part I: Phase propagation, J. Meteor. Soc. Japan, 62, 440–468,

1984.

Murakami, T., L. X. Chen, and A. Xie, Relationship among seasonal cycles, low fre-

quency oscillations and transient disturbances as revealed from outgoing long wave

radiation, Mon. Wea. Rev., 114, 1456–1465, 1986.

Nigam, S., and H. S. Shen, Structure of Oceanic and Atmospheric low frequency vari-

ability over the tropical Pacific and Indian Oceans. Part I: COADS Observations, J.

Climate, 6, 657–676, 1994.

Palmer, T. N., Chaos and the predictability in forecasting monsoons, Proc. Ind. Nat. Sci.

Acad., 60A, 57–66, 1994.

Parthasarathy, B., A. A. Munot, and D. R. Kothawale, All India monthly and seasonal

rainfall series: 1871-1993, Theor. Appl. Climatol., 49, 217–224, 1994.

Page 125: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 105

Philander, S. G., Elnino, Lanina and the Southern Oscillation, Academic Press, New York,

1990.

Ramamurthy, K., Monsoon of India: Some aspects of ’break’ in the Indian South west monsoon

during July and August, India Meteorological Department, New Delhi, 1969, forecast-

ing Manual, Part IV.18.3.

Rao, A. V. R. K., A. K. Bohra, and V. Rajeswararao, On the 30-40 day oscillations in the

southwest monsoon. A satellite study, Mausam, 41, 51–58, 1990.

Rao, Y. P., Southwest Monsoon, India Meteorological Department, New Delhi, 1976, me-

teorological Monograph.

Rasmusson, E. M., and T. H. Carpenter, Variations in the tropical sea surface tempera-

ture and surface wind fields associated with Southern Oscillation/El Nino, Mon. Wea.

Rev., 110, 354–384, 1982.

Rasmusson, E. M., and J. M. Wallace, Meteorological aspects of the ElNino/Southern

Oscillation, Science, 222, 1195–1202, 1983.

Rowell, D., C. K. Folland, K. Maskell, and M. N. Ward, Variability of summer rainfall

over tropical North Africa (1906-1992): Observations and modeling, Quart. J. Roy.

Meteor. Soc., 121, 669–704, 1995.

Salby, M., H. Hudson, K. Woodberry, and K. Tavaka, Analysis of global cloud imagery

from multiple satellites, Bull. Amer. Meteor. Soc., 4, 467–479, 1991.

Shea, D. J., and R. A. Madden, Potential for long range prediction of monthly mean

surface temperature over north America, J. Climate, 3, 1444–1451, 1990.

Short, D. A., and R. F. Cahalan, Interannual variability and climate noise in satellite

observed long-wave radiation, Mon. Wea. Rev., 111, 572–577, 1983.

Shukla, J., CISK-barotropic-baroclinic instability and growth of monsoon depressions,

J. Atmos. Sci., 35, 495–508, 1978.

Shukla, J., Dynamical predictability of monthly means, J. Atmos. Sci., 38, 2547–2572,

1981.

Shukla, J., Comments on ’Natural variability and predictability’, Mon. Wea. Rev., 111,

581–585, 1983.

Page 126: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 106

Shukla, J., Predictability in the midst of chaos: a scientific basis for climate forecasting,

Science, 282, 728–731, 1998.

Shukla, J., and D. S. Gutzler, Interannual variability and predictability of 500 mb geopo-

tential heights over the northern hemisphere, Mon. Wea. Rev., 111, 1273–1279, 1983.

Sikka, D. R., and S. Gadgil, On the maximum cloud zone and the ITCZ over Indian

longitude during southwest monsoon, Mon. Wea. Rev., 108, 1840–1853, 1980.

Silverman, B. W., Density Estimation for Statistics and Data Analysis, Chapman and Hall,

London, 1986.

Singh, S. V., and R. H. Kriplani, Potential Predictability of lower-tropospheric monsoon

circulation and rainfall over India, Mon. Wea. Rev., 114, 758–763, 1986.

Singh, S. V., and R. H. Kriplani, Low frequency intraseasonal oscillations in Indian rain-

fall and outgoing long wave radiation, Mausam, 41, 217–222, 1990.

Singh, S. V., R. H. Kriplani, and D. R. Sikka, Interannual variability of the Madden-Julian

Oscillations in Indian summer monsoon rainfall, J. Climate, 5, 973–979, 1992.

Sontakke, N. A., D. J. Shea, R. A. Madden, and R. W. Katz, Potential for long-range

regional precipitation over India, Mausam, 52, 47–56, 2001.

Sperber, K. R., and T. N. Palmer, Interannual tropical rainfall variability in general

circulation model simulations associated with Atmospheric Model Intercomparison

Project, J. Climate, 9, 2727–2750, 1996.

Sperber, K. R., J. M. Slingo, and H. Annamalai, Predictability and the relationship be-

tween subseasonal and interannual variability during the Asian summer monsoons,

Quart. J. Roy. Meteor. Soc., 126, 2545–2574, 2000.

Stern, W., and K. Miyakoda, The feasibility of seasonal forecasts inferred from multiple

GCM simulations, J. Climate, 8, 1071–1085, 1995.

Swaminathan, M. S., Abnormal monsoons and economic consquences: the Indian ex-

periment, in Monsoons, edited by J. S. Fein and P. L. Stephens, pp. 121–134, Wiley and

Sons, New York, 1987.

Trenberth, K. E., Some effects of finite sample size and persistence on meteorological

statistics. Part I: Autocorrelations, Mon. Wea. Rev., 112, 2359–2368, 1984a.

Page 127: Intraseasonal Oscillations and Interannual Variability of ...web.uvic.ca/~ajayan/PDF/amthesis.pdf · 3 Intraseasonal Oscillations and Interannual Variability of the Indian Summer

Bibliography 107

Trenberth, K. E., Some effects of finite sample size and persistence on meteorological

statistics. Part II: Potential predictability, Mon. Wea. Rev., 112, 2369–2379, 1984b.

Waliser, D. E., C. Jones, C. Schemm, and N. E. Graham, A statistical extended-range

tropical forecast model based on the slow evolution of the Madden-Julian Oscillation,

J. Climate, 12, 1918–1939, 1999.

Wallace, J. M., E. M. Rasmusson, T. P. Mitchell, V. E. Kousky, E. S. Sarachik, and H. von

Storch, On the structure and evolution of ENSO-related climate variability in the trop-

ical Pacific, J. Geophys. Res., 103(C7), 14,241–14,259, 1998.

Wang, B., and H. Rui, Synoptic Climatology of Transient Tropical Intraseasonal Convec-

tion Anomalies: 1975-1985, Met. Atmos. Phys., 44, 43–61, 1990.

Webster, P. J., Mechanism of monsoon low-frequency variability: surface hydrological

effects, J. Atmos. Sci., 40, 2110–2124, 1983.

Webster, P. J., V. O. Magana, T. N. Palmer, J. Shuka, R. T. Tomas, M. Yanai, and T. Ya-

sunari, Monsoons: Processes, predictability and the prospects of prediction, J. Geo-

phys. Res., 103(C7), 14,451–14,510, 1998.

Xie, P., and P. A. Arkin, Global precipitation: A 17-year monthly analysis based on

guage observations, satellite estimates and numerical model outputs, Bull. Amer. Me-

teor. Soc., 78, 2539–2558, 1997.

Yasunari, T., Cloudiness fluctuation associated with the northern hemisphere summer

monsoon, J. Meteor. Soc. Japan, 57, 227–242, 1979.

Yasunari, T., A Quasi-stationary appearance of 30-40 day period in the cloudiness fluc-

tuation during summer monsoon over India, J. Meteor. Soc. Japan, 58, 225–229, 1980.

Yasunari, T., Structure of an Indian summer monsoon system with around 40-day pe-

riod, J. Meteor. Soc. Japan, 59, 336–354, 1981.

Zheng, X., H. Nakamura, and J. A. Renwick, Potential predictability of seasonal means

based on monthly time series of meteorological variables, J. Climate, 13, 2591–2604,

2000.

Zweirs, F., and V. V. Kharin, Intercomparison of interannual variability and predictabil-

ity in : an AMIP diagnostic subproject, Climate Dyn., 14, 517–528, 1998.