State of the Climate
in 2005
WMO-No. 1015
State of the Climate
in 2005
K. A. Shein, 82 Ed.
Contributing Editors
A. M. Waple, H. J. Diamond, and J. M. Levy
This publication was adapted, with permission, from the “State of the Climate for 2005”, published as
a supplement to the Bulletin of the American Meteorological Society, Volume 87, Number 6, June 2006,
available online at dx.doi.org/10.1175/BAMS-87-6-Shein.
WMO-No. 1015
© 2007, World Meteorological Organization
ISBN 92-63-11015
NOTE
The designations employed and the presentation of material in this publication do not imply the expression of any opinion whatsoever
on the part of the Secretariat of the World Meteorological Organization concerning the legal status of any country, territory, city or
area, or of its authorities, or concerning the delimitations of its frontiers or boundaries.
CONTENTS
Foreword ........................................................................................................................................................................ v
Abstract .......................................................................................................................................................................... vi
Affiliations ..................................................................................................................................................................... vii
1. Introduction ............................................................................................................................................................... 1
2. Global climate ........................................................................................................................................................... 5
A. Overview ................................................................................................................................................................ 5
B. Global temperature .............................................................................................................................................. 5
I) Surface temperature ......................................................................................................................................... 5
II) Upper-air tropospheric temperatures ........................................................................................................... 6
C. Hydrologic cycle ................................................................................................................................................... 8
I) Global precipitation .......................................................................................................................................... 8
II) Snow ............................................................................................................................................................... 11
D. Trace gases ......................................................................................................................................................... 12
I) Carbon dioxide ............................................................................................................................................... 12
II) Methane ......................................................................................................................................................... 13
III) Carbon monoxide ......................................................................................................................................... 14
IV) Nitrous oxide and sulphur hexafluoride ................................................................................................... 14
V) Halocarbons ................................................................................................................................................... 14
3. Global oceans .......................................................................................................................................................... 17
A. Overview ............................................................................................................................................................. 17
B. Temperature ........................................................................................................................................................ 17
I) SSTs ................................................................................................................................................................. 17
II) Heat content ................................................................................................................................................... 18
III) Heat fluxes .................................................................................................................................................... 19
C. Circulation ............................................................................................................................................................. 20
I) Surface currents .............................................................................................................................................. 20
II) Thermohaline circulation ............................................................................................................................... 22
D. Sea level ............................................................................................................................................................... 23
E. Ocean carbon ..................................................................................................................................................... 24
4. The tropics ............................................................................................................................................................... 27
A. Overview .............................................................................................................................................................. 27
B. El Niño–Southern Oscillation ............................................................................................................................ 27
I) Overview ........................................................................................................................................................... 27
II) The Madden–Julian oscillation, Kelvin wave activity, and atmospheric circulation ............................. 28
C. Tropical cyclones ................................................................................................................................................. 29
I) Seasonal activity overview ............................................................................................................................ 29
II) Atlantic basin .................................................................................................................................................. 30
III) East Pacific basin .......................................................................................................................................... 33
IV) Western North Pacific basin ........................................................................................................................ 35
V) Indian Ocean basins ...................................................................................................................................... 36
VI) South Pacific basins ..................................................................................................................................... 39
D. Pacific intertropical convergence zone ............................................................................................................ 42
5. The poles .................................................................................................................................................................. 43
A. Overview .............................................................................................................................................................. 43
B. Arctic ..................................................................................................................................................................... 43
I) Atmosphere ...................................................................................................................................................... 43
II) Arctic Ocean ................................................................................................................................................... 45
III) Sea ice cover .................................................................................................................................................. 46
IV) Land ................................................................................................................................................................ 48
V) Greenland ....................................................................................................................................................... 49
iii
C. Antarctic .............................................................................................................................................................. 51
I) Atmospheric circulation ................................................................................................................................ 51
II) Temperature .................................................................................................................................................. 51
III) Sea ice ............................................................................................................................................................ 52
IV) Stratospheric ozone ...................................................................................................................................... 52
6. Regional climates .................................................................................................................................................... 53
A. Overview .............................................................................................................................................................. 53
B. Africa ................................................................................................................................................................... 53
I) Eastern Africa .................................................................................................................................................. 53
II) Northern Africa ............................................................................................................................................... 55
III) Southern Africa ............................................................................................................................................. 56
IV) Western Africa ............................................................................................................................................... 56
C. North America ...................................................................................................................................................... 58
I) Canada .............................................................................................................................................................. 58
II) United States of America .............................................................................................................................. 60
III) Mexico ............................................................................................................................................................ 64
D. Central America and the Caribbean .................................................................................................................. 65
I) Temperature .................................................................................................................................................... 65
II) Precipitation ................................................................................................................................................... 66
III) Notable events ............................................................................................................................................. 66
E. South America ..................................................................................................................................................... 67
I) Overview .......................................................................................................................................................... 67
II) Northern South America and the Southern Caribbean ............................................................................. 67
III) Tropical South America East of the Andes ................................................................................................ 68
IV) Tropical South America West of the Andes ............................................................................................... 69
V) Southern South America .............................................................................................................................. 69
F. Asia ........................................................................................................................................................................ 70
I) Russia .............................................................................................................................................................. 70
II) China ............................................................................................................................................................... 72
III) Southeast Asia ............................................................................................................................................. 73
IV) India and Southern Asia .............................................................................................................................. 75
V) Southwestern Asia ....................................................................................................................................... 76
G. Europe ................................................................................................................................................................. 78
I) Overview ......................................................................................................................................................... 78
II) Central and Eastern Europe ......................................................................................................................... 78
III) Fennoscandinavia, Iceland, and Greenland ............................................................................................... 79
IV) Central northern Europe .............................................................................................................................. 80
V) Northwestern Europe ................................................................................................................................... 81
VI) Iberia .............................................................................................................................................................. 82
VII) Mediterranean and Southern Europe ........................................................................................................ 84
VIII) Southeastern Europe ................................................................................................................................. 85
H. Oceania ............................................................................................................................................................... 85
I) Australia ........................................................................................................................................................... 85
II) New Zealand ................................................................................................................................................... 88
III) South Pacific Islands ..................................................................................................................................... 90
7. Seasonal summaries ............................................................................................................................................... 93
Acknowledgements .................................................................................................................................................... 97
Appendix: contributors and reviewers ..................................................................................................................... 98
References ..................................................................................................................................................................... 99
iv
FOREWORD
As the United Nations specialized agency for weather,
water and climate, part of the World Meteorological
Organization’s (WMO) role is to coordinate the
exchange of timely, regular and updated information on the status of the global climate system with
its Members, international agencies and global
and Regional Climate Centres. In particular, the
Organization interacts with the National Meteorological
and Hydrological Services (NMHSs) and climate
institutions within the six WMO Regions to collect,
analyse and produce national and regional climate
monitoring reports on major climate events affecting
their countries and regions.
In December of each year, WMO issues an annual press
release on the state of the climate system, summarizing global temperature trends, regional temperature
anomalies and major climate events that have occurred
during the outgoing year. Early in the following year,
WMO expands that information in a brochure on the
status of the global climate system, featuring updated
information, graphs, maps and charts. This brochure
is aimed at Members, international agencies, research
institutions and other users interested in receiving
synthesized information.
With a view to improve the Organization’s climate
monitoring strategy, the WMO Executive Council,
at its fifty-eight session (Geneva, 2006), examined
three recommendations of the Commission for
Climatology: first, that the WMO Global Climate
System Review be replaced by the annual State of
the Climate report published by the Bulletin of the
American Meteorological Society (BAMS); second,
that WMO take part in producing this publication
to help achieve balanced geographical coverage;
third, to make it available to all NMHSs, especially
those of developing countries, as well as to provide
assistance in translating the publication into other
languages. The WMO Secretariat has proceeded
immediately to implement this strategy for the 2005
edition. On behalf of WMO and its Members, I wish
to thank the United States National Oceanic and
Atmospheric Administration/National Climate Data
Center (NOAA/NCDC) for its outstanding work in
editing and producing the report.
As indicated in the report, all years of the new century rank among the 10 warmest years of the observational period, including 2005. The 2005 hurricane
season saw some of the most destructive hurricanes
on record. Disastrous floods and landslides, due to
extreme precipitation events, were also reported
worldwide. Prolonged drought conditions continued
to affect many parts of the world without geographical exception. These extreme events also led to many
fatalities and considerable economic loss, with social
impacts on the developing and in particular on the
least developed countries.
The timely and authoritative climate statements, assessments and reviews and their historical perspective
provide crucial information on the state of the climate,
which helps to address the challenges related to climate variability and change, since accurate and timely
weather-, climate- and water-related products and
services are prerequisites to the successful formulation
and implementation of adaptive response policies and
measures, especially to climate extremes.
WMO programmes, and in particular the World Climate
Programme (WCP), help to build capacities by promoting
the development of comprehensive climate data management systems, ensuring that high-quality climate
data are readily available to WMO Members, and assist
them and the relevant international organizations in
furthering applications to maintain public safety, health
and welfare, alleviate poverty and promote sustainable
development. In addition, WMO co-sponsors major
international observing and research programmes and
projects, such as the World Climate Research Programme
(WCRP), the Global Climate Observing System (GCOS)
and the International Polar Year 2007–2008 (IPY), to help
improve our understanding of the basic processes of
climate variability and climate change.
(M. Jarraud)
Secretary-General
v
ABSTRACT
K. A. SHEIN82
The State of the Climate 20 05 report
summarizes global and regional climate
c o nditions and places them, where
possible, in the context of historical records.
Descriptions and analyses of notable
climatic anomalies, both global and regional,
also are presented.
According to the Smith and Reynolds global
land and ocean surface temperature data
set in use at the NOAA National Climatic
Data Center (NCDC), the globally averaged
annual mean surface temperature in 2005
was the warmest since the inception of
consistent temperature observations in
1880. Unlike the previous record positive
anomaly of 1998 (+0.50°C), the 2005 global
anomaly of 0.53°C above the 1961–1990
mean occurred in the absence of a strong
El Niño signal. The record ranking of 2005
was corroborated by a data set maintained at NASA, while United Kingdom
archives placed 2005 second behind 1998.
However, statistically, the 2005 global
te m p e r a t u re a n o m a l y c o u l d n o t b e
differentiated from either 1998 or any of
the past four years. The majority of the
top 10 warmest years on record have
occurred in the past decade, and 2005
continues a marked upward trend in globally
averaged temperature since the mid-1970s.
Lower-tropospheric temperature was the
second warmest on record, with northern
polar regions the warmest at 1.3°C above
the 1979–1998 mean.
Unlike air temperatures, globally averaged
precipitation was near normal relative to the
1961–1990 period mean value. The global
2005 anomaly was just -0.87 mm. Over the
past 25 years, only seven years have had
above-normal precipitation. Additionally,
in 20 0 5 , only September– November
experienced a positive anomaly. Northern
vi
Hemisphere snow cover ex tent was
0.9 million km2 below the 36-year average
(fifth lowest) and Arctic sea ice extent was
record lowest in all months of 2005 except
May, resulting in a record lowest annual
average Arctic sea ice extent for the year
and continuing a roughly 8% yr-1 decline in
ice extent.
Carbon dioxide (CO2 ) concentrations rose
to a global average of 378.9 parts per
million (ppm), about 2 ppm over the value
from 2004. This record CO2 concentration
in 2005 continues a trend toward increased
atmospheric CO 2 since the preindustrial
era values of around 280 ppm. The globally
averaged methane (CH4 ) concentration in
2005 was 1774.8 parts per billion (ppb),
or 2.8 ppb less than in 2004. Stratospheric
ozone over Antarctica reached a minimum
of 110 Dobson units (DU) on 29 September.
This represented the 10th lowest minimum
level in the 20 years of measurement of
stratospheric ozone.
In the global ocean, sea level was above the
1993–2001 base period mean and rose at a
rate of 2.9 ±0.4 mm yr-1. The largest positive
anomalies were in the tropics and Southern
Hemisphere. Globally averaged sea surface
temperature (SST) also was above normal
in 2005 (relative to the 1971–2002 mean),
reflecting the general warming trend in SST
observed since 1971. In the tropics, only a
weak warm phase of El Niño materialized,
but dissipated by March. A relatively active
Madden–Julian oscillation (MJO) resulted in
the disruption of normal convective patterns
in the tropical Pacific and generated several
Kelvin waves in the oceanic mixed layer.
In the Atlantic Ocean basin, there was
re c o rd tro p i c al s to r m a c ti v i t y, w i th
27* named storms (15 hurricanes). Three
became category 5 storms on the Saffir–
Simpson scale, and Hurricane Wilma set
a new record for the lowest pressure
( 882 hPa) recorded in the basin. Both
Hurricanes Stan and Katrina had exceptional
death tolls, and Katrina became the costliest storm on record. Below-normal tropical
storm activity in several other basins resulted
in near-normal conditions globally in 2005.
Regionally, annual and monthly averaged
temperatures were above normal across
most of the world. Australia experienced its
warmest year on record, as well as its hottest April. For both Russia and Mexico, 2005
was the second warmest year on record.
Intermittent and delayed monsoons in
Africa and East Asia resulted in belownormal precipitation in many areas. Drought
continued in much of the Greater Horn
of Africa and developed in the central
United States. Record severe drought
occurred over both the Iberian Peninsula
and western Amazonia in 2005. In the
Amazon, river levels dropped by as much
as 11 m between May and September.
Conversely, heavy snows early in 2005
combined with a warm boreal spring to
generate widespread flooding in areas
of southwest Asia. Canada experienced
its wettest year on record in 2005, with
flooding in Alberta, Manitoba, and Ontario.
In July, the South Asian monsoon delivered
a record 944.2 mm of precipitation over
24 hours to areas around Mumbai, India.
* Tropical cyclone counts in this report do not reflect
the 10 April 2006 identification, by the NOA A
National Hurricane Center, of a 28th (subtropical)
storm in the Atlantic basin. Please visit www.nhc.
noaa.gov for further information.
AFFILIATIONS
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
(ALPHABETICAL BY AUTHOR)
Christine Achberger, Institutionen för Geovetenskaper,
Göteborgs Universitet, Gothenburg, Sweden
Peter Ambenje, Kenya Meteorological Department,
Nairobi, Kenya
Anthony Arguez, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
Molly O. Baringer, NOAA/OAR Atlantic Oceanographic
and Meteorological Laboratory, Miami, Florida
Gerald D. Bell, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
Michael A. Bell, International Research Institute for
Climate and Society, Palisades, New York
Mario Bidegain, University of the Republic, Montevideo,
Uruguay
Eric Blake, NOAA/NWS/NCEP National Hurricane Center,
Miami, Florida
Mark A. Bourassa, COAPS, Florida State University,
Tallahassee, Florida
Jason E. Box, Byrd Polar Research Center, Ohio State
University, Columbus, Ohio
Olga N. Bulygina, All-Russian Research Institute of
Hydro-meteorological Information, Obninsk, Russia
Stuart M. Burgess, National Institute of Water and
Atmospheric Research, Wellington, New Zealand
José Luis Camacho, Centro Internacional para la
Investigación del Fenómeno de El Niño, Guayaquil,
Ecuador
Suzana J. Camargo, International Research Institute for
Climate and Society, Earth Institute at Columbia
University, Palisades, New York
Muthuvel Chelliah, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
Deliang Chen, Institutionen för Geovetenskaper,
Göteborgs Universitet, Gothenburg, Sweden
John C. Christy, University of Alabama at Huntsville,
Huntsville, Alabama
Miguel Cortez Vázquez, Servicio Meteorológico Nacional,
Mexico City, Mexico
Howard J. Diamond, NOAA/NESDIS National Climatic
Data Center, Silver Spring, Maryland
Geoff S. Dutton, NOAA/OAR Earth Systems Research
Laboratory, Boulder, Colorado
James W. Elkins, NOAA/OAR Earth Systems Research
Laboratory, Boulder, Colorado
Richard A. Feely, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
Gao Ge, National Climate Center, China Meteorological
Administration, Beijing, China
Ricardo F. García-Herrera, Universidad Complutense
Madrid, Madrid, Spain
Jean-Claude Gascard, Université Pierre et Marie Curie,
Paris, France
Stephen Gill, NOAA National Ocean Service, Silver
Spring, Maryland
Tracey Gill, South African Weather Service, Pretoria,
South Africa
28. Karin L. Gleason, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
29. Stanley B. Goldenberg, NOAA/OAR Atlantic
Oceanographic Meteorological Laboratory, Miami,
Florida
30. Gustavo Goni, NOAA/OAR Atlantic Oceanographic and
Meteorological Laboratory, Miami, Florida
31. Emily K. Grover-Kopec, International Research Institute for
Climate and Society, Palisades, New York
32. Michael S. Halpert, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
33. Paul Hughes, COAPS, Florida State University,
Tallahassee, Florida
34. John E. Janowiak, NOAA/NWS Climate Prediction Center,
Camp Springs, Maryland
35. Gregory C. Johnson, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
36. Khadija Kabidi, Direction de la Météorologie Nationale,
Rabat, Morocco
37. Michael Karcher, Alfred Wegener Institute, Bremerhaven,
Germany
38. John J. Kennedy, Hadley Centre for Climate Prediction
and Research, Met Office, Exeter, United Kingdom
39. Chris Kocot, Environment Canada, Ottawa, Ontario,
Canada
40. Natalia N. Korshunova, All-Russian Research Institute of
Hydrometeorological Information, Obninsk, Russia
41. Mahbobeh Khoshkam, I.R. of the Iran Meteorological
Organization (IRIMO), Tehran, Iran
42. K. Rupa Kumar, Indian Institute of Tropical Meteorology,
Pune, India
43. Willem A. Landman, South African Weather Service,
Pretoria, South Africa
44. Chris W. Landsea, NOAA/NWS/NCEP National Hurricane
Center, Miami, Florida
45. Jay H. Lawrimore, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
46. David H. Levinson, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
47. Joel M. Levy, NOAA/OAR Climate Program Office, Silver
Spring, Maryland
48. Rick Lumpkin, NOAA/OAR Atlantic Oceanographic and
Meteorological Laboratory, Miami, Florida
49. John M. Lyman, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
50. José A. Marengo, CPTEC/INPE, Centre for Weather
Forecasts and Climate Studies, São Paulo, Brazil
51. Rodney Martínez, Centro Internacional para la
Investigación del Fenómeno de El Niño, Guayaquil,
Ecuador
52. Jim Maslanik, University of Colorado, Boulder, Colorado
53. Michael J. McPhaden, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
54. Christopher S. Meinen, NOAA/OAR Atlantic Oceanographic
and Meteorological Laboratory, Miami, Florida
55. Matthew J. Menne, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
vii
56. Mark A. Merrifield, University of Hawaii at Manoa,
Honolulu, Hawaii
57. Gary T. Mitchum, University of South Florida,
St. Petersburg, Florida
58. Kingtse C. Mo, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
59. A. Brett Mullan, National Institute of Water and
Atmospheric Research, Ltd., Wellington, New Zealand
60. Laban A. Ogallo, IGAD Climate Prediction and
Applications Centre, Nairobi, Kenya
61. Christopher Oludhe, University of Nairobi, Nairobi, Kenya
62. James E. Overland, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
63. José Daniel Pabón, Universidad Nacional de Colombia,
Bogotá, Colombia
64. Daniel Paredes, Universidad Complutense Madrid,
Madrid, Spain
65. Richard Pasch, NOAA/NWS/NCEP National Hurricane
Center, Miami, Florida
66. Donald K. Perovich, ERDC Cold Regions Research and
Engineering Laboratory, Hanover, New Hampshire
67. David Philips, Meteorological Service of Canada,
Environment Canada, Ottawa, Ontario, Canada
68. Andrey Proshutinsky, Woods Hole Oceanographic
Institute, Woods Hole, Massachusetts
69. Richard W. Reynolds, NOAA/NESDIS National Climatic
Data Center, Asheville, North Carolina
70. Fatemeh Rahimzadeh, Atmospheric Science and
Meteorological Research Center (ASMERC),
Tehran, Iran
71. Madhavan Rajeevan, National Climate Centre, India
Meteorological Department, Pune, India
72. Vyacheslav N. Razuvaev, All-Russian Research Institute of
Hydrometeorological Information, Obninsk, Russia
73. Ren Fumin, National Climate Center, China
Meteorological Administration, Beijing, China
74. Jacqueline A. Richter-Menge, ERDC Cold Regions Research
and Engineering Laboratory, Hanover, New Hampshire
75. David A. Robinson, Rutgers–The State University of New
Jersey, New Brunswick, New Jersey
76. Jeremy Rolph, COAPS, Florida State University,
Tallahassee, Florida
77. Vladimir E. Romanovsky, University of Alaska, Fairbanks,
Fairbanks, Alaska
78. Matilde Rusticucci, Universidad de Buenos Aires, Buenos
Aires, Argentina
79. Christopher L. Sabine, NOAA/OAR Pacific Marine
Environmental Laboratory, Seattle, Washington
80. M. James Salinger, National Institute of Water and
Atmospheric Research, Ltd., Newmarket, Auckland,
New Zealand
81. Russell C. Schnell, NOAA/ESRL Global Monitoring
Division, Boulder, Colorado
82. Karsten A. Shein, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
83. Alexander I. Shiklomanov, University of New Hampshire,
Durham, New Hampshire
viii
84. Shawn R. Smith, COAPS, Florida State University,
Tallahassee, Florida
85. Wassila M. Thiaw, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
86. Ricardo M. Trigo, CGUL, Universidade de Lisboa, Lisbon,
Portugal
87. Donald Walker, University of Alaska, Fairbanks,
Fairbanks, Alaska
88. Rik Wanninkhof, NOAA/OAR Atlantic Oceanographic and
Meteorological Laboratory, Miami, Florida
89. Anne M. Waple, NOAA/NESDIS National Climatic Data
Center (STG, Inc.), Asheville, North Carolina
90. Andrew B. Watkins, Bureau of Meteorology, Melbourne,
Australia
91. Robert A. Weller, Woods Hole Oceanographic Institution,
Woods Hole, Massachusetts
92. Robert Whitewood, Environment Canada, Ottawa,
Ontario, Canada
93. Joshua K. Willis, NASA Jet Propulsion Laboratory,
Pasadena, California
94. David B. Wuertz, NOAA/NESDIS National Climatic Data
Center, Asheville, North Carolina
95. Pingping Xie, NOAA/NWS/NCEP Climate Prediction
Center, Camp Springs, Maryland
96. Lisan Yu, Woods Hole Oceanographic Institution, Woods
Hole, Massachusetts
Corresponding author address: Dr. Karsten Shein,
Climate Monitoring Branch, National Climatic Data Center,
NOAA/NESDIS, Asheville, NC 28801
E-mail:
[email protected]
1. INTRODUCTION
K. A. SHEIN82
The year 2005 was a year of weather records.
Unfortunately, many of these records came with a
record price. At the December 2005 United Nations
Climate Change Conference [Conference of the Parties
to the Convention (COP) 11 & COP/Meeting of the
Parties to the Protocol (MOP) 1] in Montreal, Quebec,
Canada, the Munich Re Foundation (reinsurance)
noted that preliminary estimates of global economic
losses in which the weather was a contributing factor
exceeded US$ 200 billion. This easily tops the previous
record of US$ 175 billion in losses in 1995, and makes
2005 the costliest weather year on record [Source:
Inter-Press Service and United Nations Framework
Convention on Climate Change (UNFCCC)]. Of these
losses, approximately US$ 185 billion are attributed
to windstorms (e.g., tropical cyclones), of which
US$ 125 billion were due to Hurricane Katrina, the costliest hurricane in recorded history (Munich Re 2006).
In addition, the several global temperature data sets
currently used by various institutions to estimate globally averaged annual temperatures were in agreement
that 2005 was one of the warmest years in the historical
record. The Smith and Reynolds (2005) surface (land
and ocean) temperature data set in use at the National
Oceanic and Atmospheric Administration (NOAA)
National Climatic Data Center (NCDC) places 2005 as
the warmest on record, although the 2005 anomaly
was statistically indistinguishable from the previous
record warmth of 1998. This record 2005 temperature
is made even more remarkable given that it occurred
in the absence of a strong El Niño anomaly.
This special supplement to BAMS presents a discussion
and analysis of the global climate system for 2005, and
analyzes some of the more notable regional climatic
events that had impacts on society. The purpose of
the State of the Climate series of publications is to
summarize the climate conditions of the past calendar year and to put those conditions into a historical
perspective, both globally and regionally.
Overall, this is the 16th annual State of the Climate
report (known as the Climate Assessment until 2001)
and the 10th year that the article has appeared as part
of BAMS. However, this is the first year that the State
of the Climate is appearing as a special supplement to
BAMS. For the past six years, NOAA’s NCDC has taken
the lead in the document’s development and production. However, this effort is truly international, with
contributions from scientists from numerous institutions and organizations around the world special
effort has been made to acknowledge all contributors,
and authorship has been noted through citation by
individual sections as well as in the list of authors and
the acknowledgements in the appendix. Furthermore,
we acknowledge the important contribution of the
World Meteorological Organization (WMO) in helping
to identify and encouraging the participation of
authors from regions previously underrepresented
in this publication.
It should be noted that, given the complexity and variability of the global climate system, it is impossible
to provide comprehensive coverage of all aspects of
the observed annual climate in a document of this
length. However, the authors, editors, and contributors to the State of the Climate have made every effort
to address the most important aspects and events
related to the climate of 2005 and have attempted
to convey these to a broad audience. Additionally,
data-gathering efforts, quality control, and analysis
continue long after year end. Thus, although the
information presented in the State of the Climate
in 2005 reflects the most current data available as
of early 2006, values should be considered open to
updating as data sets are refined.
Each year, the scope of this publication is broadened
with the discussion of additional climatic variables,
introduction of new or unusual topics, and expansion
of coverage of regional climate summaries. Included
this year is an in-depth analysis of the record Atlantic
basin tropical storm season, the addition of a section
on tropical convergence zones, and the improved
coverage of the oceans through close collaboration
with authors of the Annual Report on the State of
the Oceans (NOAA OCO 2006).
1
The following is an executive summary that highlights
many of the most important topics and statistics of
the climate of 2005.
•
•
SECTION 2: GLOBAL CLIMATE
•
•
•
•
Globally averaged mean annual air temperature
in 2005 slightly exceeded the previous record
heat of 1998, making 2005 the warmest year on
record. Monthly average surface air temperatures
were above normal in all 12 months.
The globally averaged annual air temperature
in 2005 was 0.62°C above the 1880–2004 mean
(0.53°C above the 1961–1990 mean), while 1998
was +0.59°C (+0.50°C) according to the Smith
and Reynolds (2005) data set in use at NOAA/
NCDC. Comparatively, other global temperature
data sets obtained slightly different values and
rankings for 2005; however, no data set was able
to identify a statistically significant difference
between the temperatures of the two years.
Globally, precipitation was at or near mean
annual totals over land areas, but snow cover
was below average across much of the Northern Hemisphere (NH). The NH annual snow
cover extent averaged 24.7 million km2 in 2005,
0.9 million less than average.
The year was the second warmest on record
for lower tropospheric temperature, with
polar regions being the warmest on record.
Similar warm anomalies were reported for
the middle–upper troposphere. High-latitude
lower-stratospheric temperatures were very
cold in the NH but were warmer than average
over the Southern Hemisphere (SH).
•
Ocean currents were near to slightly stronger than normal in 2005, while thermohaline
circulation, as measured in the Florida Straits,
was near the long-term mean.
Sea levels, based on tide gauges and satellite
altimetry, were generally above average over
most of the global ocean (1993–2001 base) and
were consistent with long-term increases of
2.9 ±0.4 mm yr -1. The highest positive anomalies
were in the tropics and SH.
Carbon inventories may be increasing in
the Pacific basin at about twice the rate of
the Atlantic.
SECTION 4: THE TROPICS
•
•
The 2004/2005 El Niño did not materialize
beyond a weak warm phase, which largely
ended by February. Convection was suppressed
across the equatorial Pacific for much of the
year. Two active phases of the Madden–Julian
oscillation (MJO) generated Kelvin waves
that contributed to intra-annual oscillations in
equatorial Pacific SST.
The tropical cyclone season was extremely
active in the North Atlantic but was below
normal in several other basins. There was an
above-average number (103) of named storms
globally in 2005, but the number of hurricanes/
typhoons/cyclones (53) was below average.
The number of major storms (28) was slightly
above average. The Atlantic basin had record
tropical activity, as well as several recordsetting storms [e.g., lowest central pressure
(Hurricane Wilma), most category 5 storms in
a season (3), and most northeasterly genesis
(Hurricane Vince)].
SECTION 3: GLOBAL OCEANS
SECTION 5: THE POLES
•
2
Globally averaged sea surface temperature
(SST) was above normal in 2005 (1971–2002
base), as measured by ship and buoy in
situ data as well as Advanced Very High
Resolution Radiometer (AVHRR) satellite remote
sensing. Such positive anomalies reflect a
continuance of the general warming trend seen
in SST since 1971. As with 2004, high-latitude
locations in the North Atlantic and North Pacific
experienced the greatest positive departures.
Furthermore, the areal extent of July–August
SST greater than 28°C in the tropical North
Atlantic increased from 2004 to 2005.
•
•
•
In the Arctic, annually averaged surface air
temperature remained above the twentiethcentury mean, although it was cooler than in
the past two years.
The Arctic Oscillation (AO) index was slightly
negative in 2005, consistent with low index
values since the mid-1990s.
In the Arctic Ocean, the heat content of the
Beaufort Gyre increased (the result of a twofold
increase in Atlantic layer water temperature),
and the centre of freshwater shifted toward
Canada and intensified.
•
•
A few strong April storms generated flash
floods in parts of the region. Meanwhile,
western Africa experienced its second
wettest rainy season since 1994, although
a few areas were drier than normal. The
heavy precipitation and flooding fostered a
widespread cholera epidemic. Northern Africa
experienced 0.25°–1.5°C above-normal average
temperatures in 2005, but began the year with
record cold temperatures in places. With a few
exceptions, North African precipitation was
generally below normal. Similarly, southern
Africa also experienced warmer-than-normal
annual average temperatures, but precipitation
was near average after a dry start to the year.
Record minimum NH sea ice extents were
observed in every month of 2005 except May.
This continues a substantial negative trend in
NH sea ice extent since 1979.
In 2005, Arctic tundra greenness, as measured
by the Normalized Difference Vegetation
Index (NDVI), continued a marked trend toward
greener conditions. This was coupled with a
general increase in total annual discharge from
large Eurasian pan-Arctic rivers and an increase
in permafrost temperatures over the past
several decades.
SECTION 6: REGIONAL CLIMATES
A number of significant climatic conditions affected various regions in 2005 (Figure 1.1). This section expands
upon the global coverage of previous sections by
summarizing and discussing the climatic conditions
and notable events that occurred in many of the world’s
geographic regions.
•
AFRICA: Patchy and sporadic rainfall was common during the rainy seasons in the Greater
Horn region, resulting in persistent drought
over much of the region throughout the year.
•
NORTH AMERICA: In general, North America
was warmer and wetter than normal in 2005.
Canadian high latitudes experienced the greatest positive temperature anomalies (some
near record), as well as substantially abovenormal precipitation. This year was the wettest in
recorded history for Canada, which included
several widespread flood events in Manitoba,
Alberta, and Ontario. Furthermore, Canadian
Arctic sea ice extent dropped to its record
Figure 1.1. Geographical distribution of notable climate anomalies and events occurring around the planet in 2005 [Source: NOAA/NCDC;
online at www.ncdc.noaa.gov/oa/climate/research/2005/ann/ann05.html]
3
delayed by about 10 days and was weaker
than normal. As a result, precipitation in 2005
was below normal over much of continental
Southeast Asia, with temperatures slightly
above normal. Above-normal rainfall was
experienced by many of the Southeast Asian
islands. In southern Asia, severe cold started
the year but summer heatwaves took their
toll, and late-year winter conditions hampered
relief efforts related to the 8 October
earthquake in Pakistan. The region experienced
a variable and delayed monsoon season,
with south and west India receiving abundant
rainfall while other regions were below
normal. One storm deposited 944.2 mm of
rainfall over 24 hours on Mumbai (Bombay).
In southwest Asia, well-above-normal annual
average temperatures were observed and
annual precipitation was slightly below
normal, although some regions experienced
record snowfall or well-above-normal seasonal
precipitation.
lowest level, continuing the decline of roughly
8% decade-1 since the 1970s. The contiguous
United States recorded its seventh warmest year
on record, reinforcing the warming trend of the
past 30 years. Unlike Canada, the United States
experienced near-normal precipitation, with
drought conditions in central regions and excessive precipitation in the northeast and southwest.
The United States was struck by several major
hurricanes, including Katrina and Rita, which
resulted in losses over US$ 125 billion and well
over 1,000 fatalities (Munich Re 2006). A record
area of the United States also was impacted
by wildfires. Mexico observed above-normal
precipitation, partially due to the active tropical season (e.g., Hurricanes Stan and Wilma),
and had its second warmest year on record.
Wilma was Mexico’s most powerful landfalling
hurricane on record.
•
•
•
4
CENTRAL AMERICA: Annual mean temperatures
were slightly above normal across Central America
and the Caribbean in 2005, and conditions were
generally drier than the long-term average
(1979–2000). Cuban drought conditions eased,
with above-normal conditions in eastern parts
of the island. The region experienced heavy
damage and a high death toll from the tropical
storm season. Hurricane Stan brought torrential
rain to Guatemala, and roughly 1,500 fatalities were
reported in association with the storm.
SOUTH AMERICA: Generally below-normal
precipitation occurred across most of South
America, except in the west and southwest.
Eastern South America experienced abovenormal temperatures, while western regions
were below normal. Western Amazonia recorded
its worst drought in 40 years.
ASIA: Russia observed its second warmest
year on record, with some areas in northeastern Russia up to 10°C above normal
in January, and Siberia had its warmest
October in 65 years. In China, the annual average
temperature was slightly above the 1971–2000
mean (the ninth consecutive warmer-thannormal year) and precipitation was 17.7 mm
above normal. An above-normal eight tropical
systems struck China in 2005, impacting
millions of people and causing large economic
losses. The Southeast Asian monsoon was
•
EUROPE: Annually averaged air temperatures
over Europe were slightly above normal
(1961–1990), except in parts of the southeast.
The United Kingdom and northern Scandinavia
experienced exceptional warmth. Precipitation
was generally above normal in Eastern Europe
and below normal in western regions.
Southwest Europe had well-below-average
precipitation, with severe drought across the
Iberian Peninsula. Several strong extratropical
cyclones affected Scandinavia, and flooding
occurred in several eastern European countries.
•
OCEANIA: For Australia, 2005 was the hottest
year on record and temperatures were above
normal across much of the region. Australia
also recorded extremely dry conditions during
the first half of the year but rebounded in the
second half. In general, precipitation was below
average for the year. New Zealand experienced
above-normal temperature in 2005. Spatially
variable precipitation ranged from near-record
deficits to flooding rains, but on average the
country observed slightly below-normal totals
for the year. Over the South Pacific, temperatures
also were above normal, and precipitation
was spatially variable but averaged near
normal. Five strong tropical storms impacted the
region, causing damage on several islands.
2. GLOBAL CLIMATE
A.
A. M. WAPLE,89 ED.
OVERVIEW
The year 2005 was notable for its global warmth,
both at the surface and throughout the troposphere.
Globally, surface temperature remained above average
in all 12 months and reached a record high value for
the year. This anomalous warmth is part of a longterm warming trend of approximately 0.7°C century-1
since 1900, and a rate of increase almost three times
as great since 1976. This section provides a summary
of tropospheric and surface global temperatures,
and outlines the differences between the three main
surface data sets used to calculate global temperature.
Also, below-average stratospheric temperatures in
2005, consistent with a stratospheric cooling trend,
are discussed.
Record high monthly global averages were observed
in April, May, June and September 2005 (Figure 2.2).
Figure 2.2 indicates that global surface temperature values in 2005 were sustained at levels near the
1998 values, but without the influence of a strong
El Niño–Southern Oscillation (ENSO) warm event like
that which occurred in 1997/1998.
Global land surface temperatures ranked highest
on record according to the NOAA/NCDC record,
while sea surface temperatures ranked third highest
(Figure 2.3), behind 1998 and 2003. Many regions
across the globe recorded temperatures well in excess
of the 1961–1990 mean (Figure 2.4). In some areas,
most notably throughout much of the North Atlantic
basin, the average for 2005 exceeded the 90th or
98th percentiles of the mean annual temperature
Global average precipitation in 2005 was near the
long-term mean, with some regions wetter and some
drier than normal. Snow cover was below average
across much of the Northern Hemisphere for the
year, consistent with the anomalously warm surface
temperatures. Global carbon dioxide (CO2) concentration rose in 2005 by about 2 parts per million (ppm)
to 378.9 ppm, which is an increase slightly above the
+1.6 ppm yr -1 observed since 1980.
B.
GLOBAL TEMPERATURE
I)
SURFACE TEMPERATURE M. J. MENNE55
The global annual average surface temperature
in 2005 was at or near record high levels according
to analysis conducted independently at institutions
in the United Kingdom (Hadley Centre of the
Met Office and the Climate Research Unit of the
University of East Anglia) and in the United States
[NOAA’s National Climatic Data Center and the
National Aeronautics and Space Administration’s
(NASA’s) Goddard Institute for Space Studies (GISS)].
As shown in Figure 2.1, the value for 2005 ranks as
highest on record according to the NOAA and NASA
analyses and second highest, behind 1998, according
to the Met Office’s Hadley Centre/University of East
Anglia’s Climate Research Unit analysis. However,
when uncertainties related to estimating the global
mean are considered, the two highest annual values
are effectively indistinguishable from one another.
Based on the NOAA/NCDC record, the rise in global
surface temperatures since 1900 is 0.66°C, when
calculated as a linear trend.
Figure 2.1. Global annual surface temperature departures
(°C) from the 1961 to 1990 average; the 95% confidence
limits for the annual global estimates are shown (black error
bars). [Sources: NOAA/NCDC; The Hadley Centre for Climate
Prediction and Research and the Climate Research Unit of
the University of East Anglia; and NASA GISS]
Figure 2.2. Serial monthly surface temperature anomalies (°C)
relative to a 1961–1990 base period, based on Quayle et al.
(1999) [Source: NOAA/NCDC]
5
II)
Figure 2.3. Sea surface and land surface temperature
anomalies (°C) with respect to the 1961–1990 mean
[Source: NOAA/NCDC]
distribution (Figure 2.5) estimated using 1961–1990
observations (Horton et al. 2001). These anomalies
are discussed in more detail in the tropical cyclones
section (see section 4C). Averaged separately by
hemisphere, 2005 surface temperatures rank as second highest in the NH and as sixth highest in the SH
according to the NOAA/NCDC archive.
Regionally, 2005 temperatures were highest on record
in Australia. High average temperatures were also
observed across much of Canada and Siberia. An
intense heatwave occurred across India, Pakistan, and
Bangladesh in May and June. Conversely, cold conditions were experienced across much of Europe and
North Africa in February. These events are discussed in
detail in the regional climate sections (see section 6).
UPPER-AIR TROPOSPHERIC TEMPERATURES
J. C. CHRISTY17
The temperature variations of three broad atmospheric layers, the low to middle troposphere
(LT: surface–300 hPa), the middle troposphere to lower
stratosphere (MT: surface–70 hPa), and the upper
troposphere to lower stratosphere (LS: 150–20 hPa)
are presented. Products from two radiosonde-based
data sets—Radiosonde Atmospheric Temperature
Products for Analysis of Climate (RATPAC; Free et al.
2005) based on 85 stations, and Hadley Atmospheric
Temperatures (HadAT2; Thorne et al. 2005) based on
about 650 stations—are included in the data. Satellite
products are of the LT, MT, and LS from the University
of Alabama in Huntsville (UAH; Christy et al. 2003)
and of the MT and LS from Remote Sensing Systems
(RSS; Mears et al. 2003).
The annual LT temperature anomaly for 2005 was
second warmest (tied for second in UAH) since either
radiosonde (1958) or satellite (1979) records began
(Figure 2.6). The warmest calendar year remains 19981.
The anomaly correlations are extremely high among
the data sets, but the linear trend reveals slight differences since 1979 (Table 2.1). The largest trends in all
data sets are found in the northern third of the globe.
A visual interpretation of long-term change suggests
1
Note that global and tropical tropospheric temperatures tend to lag surface
temperatures by about six months, meaning the phase of the ENSO warmth
was shifted further into 1998 for upper air relative to that which was measured
at the surface.
Figure 2.4. Geographic distribution of surface air temperature anomalies (°C) in 2005 relative to the 1961–1990 mean [Source: NOAA/NCDC]
6
a relatively strong increase in global temperatures
around 1978 and another shift to warmer temperatures associated with the 1997/1998 ENSO. The base
period for the plot is 1979–1998, and since 1977, no
seasonal anomaly has dipped below -0.4°C. Since
1998, only one data set has dipped below -0.1°C.
stratosphere. The 2005 global annual anomaly was
near those of 1995, 1996, and 2000 in the satellite
records (Figure 2.8), but was not the coldest. Linear
trends are more variable among the data sets, but all
indicate significant cooling over the period (Table 2.1).
However, since about 1995, global trends have been
near zero (Figure 2.8).
The northern polar region (60°–85°N) experienced
its warmest annual reading in the period of record
(+1.3°C). No large-scale region had an annual mean
negative anomaly relative to the 1979–1998 base
period, although smaller areas did.
The MT layer (Figure 2.7), which includes some stratospheric influence, presents a similar picture. However,
while 2005 was the second warmest globally for the
two radiosonde data sets, it was fourth and fifth
warmest in the RSS and UAH satellite data sets,
respectively. Linear trends in MT are more negative
than LT, because the cooling stratosphere exerts an
influence on MT (Table 2.1).
Figure 2.5. Land surface temperature anomalies (°C) based on
The quasi-biennial oscillation (QBO) was in the easterly
phase in 2005 and contributed to a tropical (20°N–
20°S) anomaly that was the coldest on record in the
HadCRUT3 expressed as percentiles of modified two-parameter gamma
distributions fit to annual data for 1961–1990 according to Horton et al.
(2001) [Source: Hadley Centre for Climate Prediction and Research]
Table 2.1. Linear trends (°C decade -1) from 1979 to 2005 of global and tropical (20°S–20°N) anomalies for the three temperature products
(1958–2005 trends in parentheses)
RATPAC
HADAT2
UAH
RSS
Global LT
+0.15 (+0.15)
+0.17 (+0.15)
+0.13*
Tropical LT
+0.11 (+0.13)
+0.09 (+0.13)
+0.07
Global MT
+0.04 (+0.08)
+0.06 (+0.09)
+0.05
+0.14
Tropical MT
+0.02 (+0.08)
+0.01 (+0.09)
+0.06
+0.15
Global LS
-0.71 (-0.42)
-0.57 (-0.36)
-0.45
-0.32
Tropical LS
-0.75 (-0.45)
-0.60 (-0.34)
-0.41
-0.29
* When subsampled at 50°–85°N to more closely represent the geographical extent of HadAT2, UAH LT “global” trend is +0.15°C decade-1.
7
Figure 2.6. Seasonal anomalies (°C) of global average
lower-tropospheric layer temperature: HadAT2, UAH,
and RSS depict the temperature of layers representing
the microwave brightness temperature weighting
functions (roughly surface to 300 hPa, peaking around
700 hPa), while RATPAC depicts the 850–300-hPa layer
mean temperature.
Figure 2.7. Seasonal anomalies (°C) of global average
tropospheric temperature; this layer includes a portion
of the lower stratosphere, representing a microwave
brightness temperature weighting function that
starts at the surface, peaks in the midtroposphere,
and diminishes rapidly above 100 hPa.
Figure 2.8. Seasonal anomalies (°C) of global lowerstratospheric temperature; this layer begins around
120 hPa and diminishes above 20 hPa. RATPAC data
depict anomalies of the 100–50-hPa layer.
8
High-latitude LS anomalies were exceptionally cold
in a broad area centred over Greenland, but were
warmer than average over eastern Antarctica. The
coldest large-scale monthly anomalies occurred over
the NH polar region in January and February (-6°C,
60°–85°N), while the warmest monthly anomalies
(+2°C) appeared over the SH polar region in October
and November. This marks the fourth year in a row
where anomalies of the SH polar region were above
average (1979–1998) for austral spring in satellite
records (see section 5C).
C.
HYDROLOGIC CYCLE
I)
GLOBAL PRECIPITATION
(i)
Land D. H. LEVINSON,46 J. H. LAWRIMORE,45 AND
D. B. WUERTZ94
Global precipitation anomalies determined from landbased gauges were analyzed on both an annual and a
seasonal basis using data from the Global Historical
Climatology Network (GHCN; Peterson and Vose
1997). Anomalies over the period of 1901–2005 were
determined from the GHCN data set with respect to the
1961–1990 mean using those stations with a minimum
of 20 years of data during the base period (Peterson
and Vose 1997; Vose et al. 1992). Global precipitation
was average in 2005, with an annual anomaly less
than 1 mm (-0.87 mm) below the 1961–1990 mean
(Figure 2.9a).
Over the past two-and-a-half decades, global precipitation has been generally below the long-term
mean, with above-average precipitation anomalies in
only 7 of the last 25 years. This multidecadal period
of below-normal precipitation and anomalously
dry conditions began during the early 1980s, and
has continued through the 1990s into the present
decade. The peak in this dry period appears to have
occurred in 1992, corresponding with a multiyear
El Niño event. Previous studies (i.e., Ropelewski and
Halpert 1987) have shown that variability associated
with ENSO influences large-scale precipitation patterns
in the tropics and midlatitudes, both of which contribute
to generally drier conditions at high latitudes.
Regardless of the potential causes of the multidecadal
dry period, much of the observed signal in global
precipitation anomalies appears to be seasonally
dependent (Figures 2.9b–e). In 2005, negative anomalies
were observed in three of the four seasons, with only
the boreal autumn [September–November (SON)]
having above-average precipitation for the year. During
this extended dry period, below-normal precipitation
occurred primarily during two seasons [March–May
(MAM) and June–August (JJA)], with the boreal
summer having the longest continuous period of
drier-than-normal conditions, extending from the late
1980s through 2005.
For the first half of the twentieth century, and extending into the 1960s, much of the tropics in the NH were
dominated by wet precipitation anomalies, while
midlatitude regions in both hemispheres were drier
than normal (Figure 2.10). It is likely that the period of
dry anomalies in the northern high-latitudes during
1900–1940 was due to a lack of precipitation data, as
well as a systematic undercatch of snow and solid precipitation. Since the 1960s, there has been an extended
period of dry anomalies in the tropics, particularly in
the NH. In recent years, including 2005, high-latitude
regions of the NH have been wetter than normal, with
a multiyear wet period north of 60°N.
Significant precipitation anomalies were observed at
many long-term monitoring locations over the past
year (Figure 2.11), but several regions stand out as
being either significantly wetter or drier than normal.
Of special note were the large dry anomalies in parts
of East Asia, particularly across southern Japan,
eastern China, and the Korean Peninsula, in part due to
below-normal number of landfalling tropical cyclones
(see section 4C). Other regions with significant dry
anomalies included much of Australia, South Africa
and coastal areas of the Gulf of Guinea, the Iberian
Peninsula and France, and a large portion of the central
United States, extending from the Gulf of Mexico to
the Great Lakes. Locations significantly wetter than
average in 2005 included most of Scandinavia, Senegal
and the Atlantic coast of West Africa, Venezuela and
Colombia, and a large portion of the Caribbean basin.
Further details on these and other regional precipitation anomalies can be found in section 6.
(ii)
Oceans P. XIE95 AND J. E. JANOWIAK34
Real-time monitoring of global oceanic precipitation
is routinely conducted at NOAA’s Climate Prediction
Center (CPC) with the use of the gauge–satellite merged
Climate Anomaly Monitoring System (CAMS)–outgoing
longwave radiation (OLR) Precipitation Index (OPI)
data set (Janowiak and Xie 1999). By combining the
Figure 2.9. Time series of annual and seasonal global
precipitation anomalies over the period 1901–2005, with
observations from the GHCN data set: (a) annual,
(b) December–February (DJF), (c) MAM, (d) JJA, and (e)
SON; the precipitation anomalies were calculated in mm with
respect to the 1961–1990 base period mean: green bars =
positive anomalies, yellow bars = negative anomalies,
and red bar = 2005 anomaly. In addition, the black line in
each time series denotes the smoothed annual or seasonal
values using a 13-point binomial filter.
Figure 2.10. Hovmoeller plot of the percentage departure from
1961 to 1990 means of GHCN global annual precipitation,
with zonal means determined over 5° latitude bands and
covering the period 1900–2005; a 13-point binomial filter
was applied to each zonal time series, with green and blue
shades corresponding to wet anomalies and red and brown
shades corresponding to dry anomalies. Gray shading in
the early twentieth century is due to a lack of data in highlatitude regions.
9
anomalies occur over the central and eastern Pacific
(Ropelewski and Halpert 1989; Xie and Arkin 1997).
Figure 2.11. Precipitation anomalies (mm) relative
to a 1970–2000 base period from the gauge–satellite
merged CAMS–OPI precipitation data set
(Janowiak and Xie 1999)
gauge observations of precipitation collected and
archived by CPC via CAMS (Ropelewski et al. 1985)
with the satellite-based OPI (Xie and Arkin 1998),
CAMS–OPI provides monthly precipitation estimates
over global land and ocean on a real-time basis.
Global oceanic precipitation during 2005 is characterized by rainbands associated with the intertropical
convergence zone (ITCZ), South Pacific convergence
zone (SPCZ), and the midlatitude oceanic storm tracks
(Figure 2.12). Mean precipitation over the entire global
ocean during 2005 was 2.840 mm day-1, equivalent
to a freshwater influx of 1036.6 kg m -2 . Maximum
annual precipitation rates of over 10 mm day-1 were
observed during the year over the tropical western
Pacific where the ITCZ merges with the SPCZ (see
section 4D). Meanwhile, relatively light precipitation
occurred over several oceanic dry zones of the southeast
Pacific, northeast Pacific off the coast of the southwest
United States, southeast Atlantic, tropical North Atlantic
near western Africa, and the eastern Indian Ocean.
The distribution of precipitation anomalies during
2005 indicates a dipole pattern of wet and dry
anomalies over the western and eastern tropical
Pacifi c, respectively (Figure 2.11). Although weak
El Niño conditions prevailed over the tropical Pacific
in the second half of 2004 and continued into early
2005 (Lyon and Barnston 2005; see also section 4B),
enhanced precipitation was limited mostly to the
tropical Pacific west of the date line. This pattern is
different from that typical of a medium or strong
El Niño event when large positive precipitation
10
Enhanced convection, and attendant above-normal
precipitation, was first observed in late 2004 over the
tropical western Pacific north of Indonesia (Figure 2.13).
The positive precipitation anomaly intensified as it
moved eastward, and reached its maximum intensity
during January–March 2005. Such large intraseasonal
variations present in Figure 2.13 are associated with the
strong MJO (Madden and Julian 1971) activity that was
observed during that period (Climate Prediction Center
2005). The positive precipitation anomaly over the
tropical western Pacific gradually shifted westward
during the second half of 2005 as the coupled ocean–
atmosphere system evolved toward a weak La Niña.
A substantial positive precipitation anomaly was
observed over regions of strong hurricane activity in
the Atlantic basin, especially the Caribbean Sea (see
section 4C). The Atlantic ITCZ was located slightly north
of its climatological latitude, which is reflected by two
parallel bands of positive and negative precipitation
anomalies over the tropical Atlantic (Figure 2.11).
Substantially depressed precipitation over the northwestern Atlantic was also noticeable during 2005.
Examination of monthly precipitation and atmospheric
circulation fields suggest that the bulk of the negative
precipitation anomaly over the region was the result
of below-normal winter storm activity in the oceanic
storm track, which is a climatological feature over
the western North Atlantic during the cool season.
This reduced storminess was associated with strong
Figure 2.12. Precipitation totals (mm) for 2005 from the
gauge–satellite merged CAMS–OPI precipitation data set
(Janowiak and Xie 1999)
anticyclonic blocking activity over the high latitudes
that was observed during the winter season.
II)
SNOW D. A. ROBINSON75
Annual snow cover extent (SCE) over NH lands
averaged 24.7 million km2 in 2005. This is 0.9 million km2
less than the 36-year average and ranks 2005 as having
the 32nd most extensive cover of record (Table 2.2).
This evaluation includes snow over the continents,
including the Greenland Ice Sheet. The SCE in
2005 ranged from 47.6 million km2 in February to
2.3 million km2 in August. Monthly snow extent values
are calculated at the Rutgers Global Snow Laboratory
from weekly SCE maps produced by NOAA meteorologists, who rely primarily on daily visible satellite
imagery to construct the maps.
Hemispheric SCE was only above the long-term
mean in February, March, and December; thus,
the 12-month running means of NH extent were
below the long-term average throughout the year
(Figure 2.14). This has almost exclusively been the situation since the late 1980s. Eurasian SCE was somewhat
Figure 2.13. Time–longitude section of precipitation
anomaly (mm; 1979–1995 base period) averaged over
the tropical Pacific (5°S–5°N) as observed by the
Global Precipitation Climatology Project (GPCP)
pentad precipitation data set (Xie et al. 2003)
Table 2.2. Monthly and annual climatological information for NH and continental snow extent between November 1966 and December 2005.
Included are the numbers of years with data used in the calculations, means, standard deviations, 2005 values, and rankings.
Areas are in millions of km2. 1968, 1969, and 1971 have 1, 5, and 3 missing months, respectively, and thus are not included in the
annual (Ann) calculations. North America includes Greenland.
YEARS
MEAN
STD DEV
2005
2005 NH
RANK
EURASIA
RANK
NORTH
AMERICAN
RANK
Jan
39
46.9
1.5
46.7
21
27
15
Feb
39
45.9
1.9
47.6
9
3
33
Mar
39
41.0
1.9
41.3
18
12
25
Apr
39
31.4
1.7
29.9
31
20
38
May
39
20.5
1.9
17.6
38
36
38
Jun
38
11.0
2.1
9.0
31
31
30
Jul
36
4.9
1.4
3.0
36
33
35
Aug
37
3.5
1.0
2.3
37
33
37
Sep
37
5.7
1.0
4.8
31
33
21
Oct
38
18.3
2.6
16.9
31
24
34
Nov
40
34.1
2.1
32.2
32
30
31
Dec
40
43.5
1.8
44.5
10
11
13
Ann
36
25.6
1.0
24.7
32
25
35
11
below the long-term average in 2005 and ranked as the
25th most extensive cover of the satellite era. North
American SCE was much below average, ranking 35th
(Figures 2.15 and 2.16).
Figure 2.14. Anomalies of monthly snow cover extent
over Northern Hemisphere lands (including Greenland)
between Nov 1966 and Dec 2005, calculated from
NOAA snow maps; also shown are 12-month running
mean anomalies of hemispheric snow extent, plotted
on the seventh month of a given interval. Mean
hemispheric snow extent is 25.6 million km2 for the
full period of record. Monthly means for the period of
record are used for nine missing months between
1968 and 1971 in order to create a continuous series
of running means. Missing months fall between
June and October; no winter months are missing.
As is common, 2005 SCE showed significant spatial
and temporal variability. For instance, hemispheric
SCE ranked in the top 10 in February and December,
while totals ranked in the lowest 10 from April through
November. February’s NH anomaly was positive due to
the Eurasian extent being the third largest on record,
despite North America ranking 33rd. December’s high
ranking was the result of extensive cover over both
continents; this followed November, when both ranked
quite low. Spring snow cover continues to be less
extensive in the second half of the satellite record than
in the first half. Lower-than-average North American
SCE was first noted in February, while well-belowaverage Eurasian cover did not occur until May.
Over the contiguous United States, end-of-winter
SCE was very low after February, while in Alaska,
May SCE was at a record low. Both regions began the
boreal fall snow season slowly, especially in Alaska,
where SCE was at a record low in September and only
increased to second lowest in October and third lowest in November, resulting in a very long snow-free
season for 2005 in Alaska.
Figure 2.15. Same as Figure 2.14, except for Eurasia
Figure 2.16. Same as Figure 2.14, except for North
America (including Greenland)
12
Maps depicting daily, weekly, and monthly conditions, daily and monthly anomalies, and monthly
climatologies for the entire period of record may
be viewed at the Rutgers Global Snow Laboratory
Web site (available online at http://climate.rutgers.
edu/snowcover).
D.
TRACE GASES
I)
CARBON DIOXIDE R. C. SCHNELL81
Carbon dioxide emitted from natural and anthropogenic
(i.e., fossil fuel combustion) sources is partitioned into
three mobile reservoirs: atmosphere, oceans, and the
terrestrial biosphere. The result of increased fossil
fuel combustion has been that atmospheric CO2 has
increased from about 280 ppm (parts in 106 by dry-air
mole fraction) at the start of the Industrial Revolution
to about 380 ppm today (Figure 2.17a). Roughly half the
emitted CO2 remains in the atmosphere and the remainder has gone into two “sinks”–oceans and the land
biosphere (which includes plants and soil carbon).
The present rate of anthropogenic carbon emission
to the atmosphere is nearly 7 Pg yr -1 (Pg = 1015 g).
During the 1990s, net uptake by the oceans was estimated at 1.7 ±0.5 Pg yr -1, and by the land biosphere as
1.4 ±0.7 Pg yr -1 (Prentice et al. 2001). The gross atmosphere–ocean and atmosphere–terrestrial biosphere
(i.e., photosynthesis and respiration) fluxes are on
the order of 100 Pg yr -1. Interannual variations in the
atmospheric increase of CO2 (Figure 2.17a; based on
Conway et al. 1994) are not attributed to variations in
fossil fuel emissions, but rather to small changes in
these net fluxes. Most attempts to explain the interannual variability of the atmospheric CO2 increase
have focused on short-term climate fluctuations
(e.g., ENSO and post–Mt. Pinatubo cooling), but the
mechanisms, especially the role of the terrestrial
biosphere, are poorly understood. To date, about 5%
of conventional fossil fuels have been combusted.
If combustion were stopped today, it is estimated that
after a few hundred years, 15% of the total carbon
emitted would remain in the atmosphere, and the
remainder would be in the oceans.
In 2005, the globally averaged atmospheric CO2 mole
fraction was 378.9 ppm, just over a 2 ppm increase
from 2004. This continues the steady upward trend
in this abundant and long-lasting greenhouse
gas. Since 1900, atmospheric CO 2 has increased
84 ppm (22 %), with an average annual increase of
1.6 ppm since 1980.
In calculating the global mean mole fractions of CO2 and
other trace gases, the number of stations varies through
time and may be different for each gas at any point in the
period of record. Each record is temporally smoothed,
with evenly spaced values selected from these temporal
fits to generate new fits as a function of latitude (ϕ). Values
extracted from the latitude fit at a spacing of sin(ϕ) = 0.05
(i.e., at equal atmospheric volumes) define a matrix of
species value as a function of time and latitude. These
are then used to calculate global averages at weekly and
annual time resolution (Conway et al. 1994).
II)
METHANE R. C. SCHNELL81
Methane’s (CH4 ) contribution to anthropogenic radiative forcing, including direct and indirect effects, is
about 0.7 W m-2, or roughly half that of CO2. Also,
changes in the burden of CH4 feed back into atmospheric chemistry, affecting the concentrations of
hydroxyl (OH) and ozone (O3 ). The increase in CH4
since the preindustrial era is responsible for about half
Figure 2.17. Trace gas mole fractions (black symbols)
determined from samples collected at the NOAA ESRL
Mauna Loa Observatory (MLO) for (a) CO2 (courtesy:
T. J. Conway, NOAA); current trends at MLO and
globally averaged are available online at www.cmdl.
noaa.gov/ccgg/trends/; (b) CH4 (courtesy:
E. J. Dlugokencky, NOAA); and (c) CO (courtesy:
P. C. Novelli, NOAA). In all panels, the solid blue line is
the deseasonalized trend and the red line is a smooth
curve fitted to the black symbols. More plots can be
found online at www.cmdl.noaa.gov/ccgg.
of the estimated increase in background tropospheric
O3 during that time. Changes in OH concentration
affect the lifetimes of other greenhouse gases such
as hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs).
High-precision measurements of atmospheric CH4
provide climate modelers with current and past rates
of CH4 increase, and they are also useful in constraining
the CH4 budget. During 20 years of measurements,
CH4 has increased, but the rate of increase has slowed
in recent years (Figure 2.17b). A large increase in 1998
was probably the result of climatic conditions that
resulted in increased emissions from wetlands and
biomass burning. Measurements of CH4 at Mauna
Loa, Hawaii, remained nearly constant from 1999
to 2002 (Dlugokencky et al. 2003). In 2003, CH 4
increased by about 5 ppb (parts in 109 by dry-air mole
fraction), primarily due to increases in the Northern
Hemisphere. This was followed by a small decrease
in 2004, and little change from those levels in 2005.
Globally averaged CH4 in 2005 was 1774.8 ppb, which
represented a decrease of 2.8 ppb from 2004.
13
III)
CARBON MONOXIDE R. C. SCHNELL81
Unlike CO2 and CH4, carbon monoxide (CO) does not
strongly absorb terrestrial infrared radiation, but does
impact climate through its chemistry. The chemistry of
CO affects OH (which influences the lifetimes of CH4
and HFCs) and tropospheric O3 (itself a greenhouse
gas), so emissions of CO can be considered equivalent
to emissions of CH4 (Prather 1996). Current emissions
of CO may contribute more to radiative forcing over
decadal time scales than emissions of anthropogenic
nitrous oxide (N2O; Daniel and Solomon 1998).
Carbon monoxide mole fractions from Mauna Loa
(Figure 2.17c, symbols) show little trend over the
measurement period (updated from Novelli et al.
2003). Superimposed on the flat trend is a significant
increase during 1997 and 1998, which was probably the
result of tropical (Langenfelds et al. 2002) and boreal
biomass burning (Kasischke et al. 2000). Because the
lifetime of CO is relatively short (a few months), the
anomaly quickly disappeared and CO quickly returned
to pre-1997 levels. Carbon monoxide levels in 2005 were
comparable to those found in the early 2000s. The globally
averaged CO mole fraction in 2005 was 83.5 ppb, very
near the average of the past five years. Since 1991, little
trend in globally averaged CO has been observed.
IV)
NITROUS OXIDE AND SULPHUR HEXAFLUORIDE
J. W. ELKINS21 AND G. S. DUTTON20
Atmospheric N2 O and sulphur hexafluoride (SF 6 )
are present in lower concentrations than CO2, but
the radiative forcing of each is far greater. Nitrous
oxide is the third strongest greenhouse gas, while
each SF6 molecule is 22,200 times more effective
as an infrared absorber than one CO2 molecule and
has an atmospheric lifetime of between 500 and
3,200 years. The concentration of both species has
grown at a linear rate, N2O at 0.76 ppb yr -1 (0.25% yr -1)
since 1978 (Figure 2.18a) and SF6 at a rate of 0.22 ppt
(parts in 1012 by dry-air mole fraction) yr -1 (~5% yr -1)
since 1996 (Figure 2.18b).
The concentration of 320 ppb N2O in 2005 has added
a radiative forcing of around 0.17 W m -2 over the
preindustrial N2O concentration of around 270 ppb
(Figure 2.18a). Atmospheric N2O is also a major source
of stratospheric nitric oxide (NO), a compound that
helps to catalytically destroy stratospheric O3.
The atmospheric concentration of SF6 has grown due to
its use as an electrical insulator for power transmission
throughout the world. Its global mean concentration was
5.75 ppt at the end of 2005 (Figure 2.18b). While total radiative forcing of SF6 from preindustrial times to the present
is relatively small (0.003 W m-2), its long atmospheric
lifetime, high atmospheric growth rate and high global
warming potential are a concern for the future.
V)
HALOCARBONS R. C. SCHNELL81
Concern over stratospheric ozone depletion has restricted
or eliminated production of many halocarbons. The phaseout of human-produced halocarbons was the result of the
1987 Montreal Protocol on Substances that Deplete the
Ozone Layer. As a result of these efforts, mixing ratios of
many ozone-depleting gases have been declining at Earth’s
surface in recent years; this decline continued in 2005
(Figure 2.19). NOAA/Earth System Research Laboratory
(ESRL)/Global Monitoring Division (GMD) measurements
from around the globe show that tropospheric mixing
ratios of CFC-12, the longest lived and most abundant
human-made ozone-depleting gas in the atmosphere,
peaked within the last few years (Figure 2.19a).
Those data also show that mixing ratios of some
halogenated gases continue to increase globally
Figure 2.18. (a) Global atmospheric N2O from
the NOAA/ESRL in situ and flask network
(red) has increased with a linear growth rate
(blue) of 0.76 ppb yr -1 from 1978 to the end
of 2005. (b) Global atmospheric SF 6 from the
NOAA/ESRL in situ and flask network (red)
has increased at a linear rate (blue) of
0.22 ppt yr -1 from 1996 to the end of 2005.
14
Figure 2.19. (a)–(d) Changes in global mean tropospheric mixing ratios (ppt) of the most abundant CFCs, HCFCs, HFCs, chlorinated solvents
and brominated gases, calculated from atmospheric measurements made at remote sites in both the NH and SH
[Source: NOAA/ESRL/GMD cooperative air sampling network]; (e) secular changes in atmospheric equivalent chlorine (EECl; ppb);
(f) recent changes in effective equivalent stratospheric chlorine (EESC) observed by the NOAA/GMD global network relative to
the secular changes observed in the past. EESC is derived from EECl by adding three years to represent the lag associated with mixing
air from the troposphere to the middle stratosphere (updated from Montzka et al. 2003a,b). [Courtesy: S. A. Montzka, J. H. Butler,
T. Thompson, D. Mondeel, and J. W. Elkins, NOAA/CMDL]
(Figures 2.19a–d). The most rapid increases are in
HCFCs and HFCs, which are chemicals commonly
used as replacements for chlorofluorocarbons (CFCs),
halons and other ozone-depleting gases. Although
HCFCs contain chlorine (Cl) and deplete O3 with a
reduced efficiency compared to CFCs, HFCs do not
participate in O3-destroying reactions.
Increases in HCFCs have slowed notably in recent
years. By mid-2005, the Cl in the three most abundant
HCFCs amounted to 217 ppt, or 8.0% of all Cl carried
by long-lived halocarbons. Mixing ratios of HFC-134a,
the most abundant HFC in the global background atmosphere, increased nonlinearly in the 1990s. From 2001
through 2005, however, it has increased in the global
troposphere at a fairly constant linear rate of 4.2 ppt
yr -1. Concern over increases in HFCs is largely due to
their efficiency as absorbers of infrared radiation.
The influence of these disparate trends on future
levels of stratospheric ozone can be gauged roughly
from a sum of Cl and bromine (Br) in long-lived halocarbons, provided the enhanced efficiency for Br to
destroy ozone is considered. This sum is expressed
as effective equivalent chlorine (EECl; Figure 2.19e),
and is derived from surface-based measurements.
EECl provides an estimate of the O3-depleting power
of the atmosphere a few years in the future, when air
at Earth’s surface will have become mixed into the
stratosphere.
Observations indicate that the EECl content of the
lower atmosphere has declined at a mean rate of
26 ppt yr -1 since its peak in 1994. Scenarios projecting future halocarbon mixing ratios have been
derived elsewhere based upon full compliance with
the fully amended and revised Montreal Protocol
and our understanding of atmospheric lifetimes of
these gases (Montzka et al. 2003b). These analyses suggest that it will take 40–50 years for EECl
to decline to the levels present in 1980, before O 3
depletion was first observed. This 1980 level is
notable, given that one might expect nearly full
recovery of stratospheric ozone once the atmospheric EECl returns to this level. The time frame
for O 3 recovery will depend upon other factors as
well, such as stratospheric temperatures and atmospheric aerosol loading. Nonetheless, the declines
in EECl from 1994 to the present time represent a
significant drop in the atmospheric EECl burden.
As of 2005, EECl had declined 20% of the way back
to the 1980 level (Figure 2.19f).
Changes in the direct radiative influence of longlived halocarbons can be estimated from observed
changes in atmospheric mixing ratios with knowledge of trace-gas radiative efficiencies. Such an
analysis suggests that the direct radiative forcing
of these gases was still increasing in 2005, though
at a much slower rate than observed from 1970 to
1990.
15
3. GLOBAL OCEANS
A.
J. M. LEVY47 AND K. A. SHEIN82, EDS.
OVERVIEW K. A. SHEIN82
Recent decades have seen a marked increase in our
knowledge of the coupled ocean–atmosphere system,
although as is discussed, it is only through a number of
relatively new or planned ocean-monitoring activities
that our understanding of these interactions, such
as the role of the oceans in the anomalously active
2005 Atlantic tropical storm season, can be improved.
Notable characteristics of the oceans in 2005 included
above-normal sea surface temperatures and heat
losses that were generally below normal over the
global basins. Globally, ocean circulation was near
to slightly stronger than normal in 2005, and average
sea levels continued their rise for another year. Also,
there is preliminary information that anthropogenic
carbon inventories may be increasing in the Pacific
at about twice the rate of the Atlantic.
B.
TEMPERATURE
I)
SSTs R. W. REYNOLDS69
regarding whether changes in SSTs can imply
statistically significant changes in hurricanes
(Webster et al. 2005) or not (Trenber th 2005),
but all of the reports state that SSTs are not the
only variable affecting hurricanes. Because of the
changes in the northern tropical Atlantic SSTs from
2004 to 2005, time series are shown for the North
Atlantic (10°–30°N) and Gulf of Mexico (Figure 3.2).
The tropical Atlantic region shows overall warm
anomalies with irregular variability on 2- to 5-year
periods, which does not strongly correlate with
the seasonal cycle or ENSO. However, a relatively
strong maximum anomaly occurs in the summer
of 2005. The time series from the Gulf of Mexico
appears to be noisier without any noticeable signal.
Goldenberg et al. (2001) state that local SSTs greater
than 26.5°C are needed to generate hurricanes. A closer
examination of the 2004 and 2005 tropical Atlantic
anomalies shows that areas exceeding 26.5°C are
similar in the corresponding months for 2004 and 2005
(Figure 3.3). However, the areal extent of SST greater
than 28°C is larger in July and August 2005 than in
July and August 2004.
Annually averaged SSTs for both 2004 and 2005
(Figure 3.1) are derived from monthly fields interpolated from the Reynolds et al. (2002) weekly optimum
interpolation (OI) analyses. The analysis uses ship and
buoy in situ SST data as well as satellite SST retrievals from the infrared (IR) AVHRR, and anomalies are
departures from the 1971–2002 base period mean
(Xue et al. 2003).
The impression is that the anomalies are primarily
positive and are part of the overall warming trend
of global SSTs since 1971. Comparisons of the two
years show two important changes: the first in middle
latitudes (40°–66°N) and the second in the tropical
North Atlantic (0°–30°N). In the midlatitude regions,
there is a tendency for the positive anomalies to
increase from 2004 to 2005 in the eastern Pacific and
western Atlantic. Both regions show an overall warm
anomaly for the period of 1995–2005, although variability in the Pacifi c is less than the Atlantic. There
is also evidence of regular summer warming in the
Pacific region from 2000 to 2005. This signal is only
clearly evident in the Atlantic region in 2003, and is
the oceanic response to that summer’s European
heatwave.
During 2005, there was a record number of strong
Atlantic hurricanes. There has been discussion
Figure 3.1. Annual SST anomalies (°C) for (top) 2004 and
(bottom) 2005, computed relative to a 1971–2000 base period
17
II)
HEAT CONTENT G. C. JOHNSON,35 J. M. LYMAN,49 AND
J. K. WILLIS93
Ocean storage and transport of heat and freshwater,
and their variations, are intrinsic to many aspects of climate, including El Niño, the North Atlantic Oscillation
(NAO), the global water cycle, hurricane seasons
and global change (e.g., Levitus et al. 2005; Hansen
et al. 2005). Regional studies of decadal freshwater
variability are possible in well-sampled regions like
the North Atlantic (e.g., Curry and Mauritzen 2005),
but in situ ocean salinity data are too sparse and
their reporting is too delayed for a global 2005 perspective of ocean freshwater storage. However, the
rapidly maturing Argo Project array of profiling floats
measuring temperature and salinity (Roemmich et
al. 2004) is remedying this situation.
Figure 3.2. Time series (January 1995–December
2005) of monthly SST anomalies (°C) for the (top)
tropical North Atlantic and (bottom) the Gulf of Mexico
Figure 3.3. Tropical monthly SST anomalies (°C) for July,
August, and September (left) 2004 and (right) 2005;
contours show the corresponding monthly total (anomaly
+ climatology) SST isotherms for 26°, 27°, and 28°C.
18
Here, we discuss an estimate of the upper-ocean
(0–750 m) heat content anomaly (OHCA) for 2005
produced by combining in situ ocean temperature
data and real-time satellite altimetry data collected
from 1 January to 31 October 2005 following the
techniques of Willis et al. (2004). The 2005 combined
OHCA map, when compared to a 1993–2002 baseline
(Figure 3.4, top), shows eddy and meander variability
down to 100-km mapping scales. There is a great deal
of small-scale spatial variability associated with the
western boundary currents in every gyre, as well as
the Antarctic Circumpolar Current.
Large-scale patterns are also evident in the OHCA.
The combined OHCA map for 2005 is high in the
subpolar North Atlantic and low in the subtropical
North Atlantic, consistent with a decreased strength
of the North Atlantic Current. This pattern is probably related to decadal changes in the NAO index
(e.g., Curry and McCartney 2001). The NAO index
was lower in 2005 than during the baseline period,
and has generally trended lower from 1993 through
2005. For the most part, the tropics in 2005 have
only slightly higher OHCA than average, reflecting the lack of a pronounced El Niño or La Niña in
2005. In 2005, OHCA is high throughout the South
Pacific and South Atlantic Oceans in a belt located
north of the Antarctic Circumpolar Current. This
change may be related to changes in the wind-stress
field associated with an increase in the Antarctic
Oscillation.
The difference in combined OHCA maps between
2005 and 2004 (Figure 3.4, bottom) illustrates the
large year-to-year variability in ocean heat storage,
with changes reaching or exceeding the equivalent of
an 80 W m-2 surface flux. Ocean advection probably
plays a large role in many of these changes. Such
differences between the two years clearly show the
influence of eddies and meanders, but there are also
contributions from some of the aforementioned largerscale patterns in the subtropics and subpolar regions.
The decrease of OHCA in the central equatorial Pacific
between 2005 and 2004 probably reflects the transition from weak El Niño to more normal conditions.
Finally, given the strong 2005 hurricane season and
the potential link between hurricane intensity and
warm ocean waters (e.g., Emanuel 2005), the large
increases in OHCA around Florida and the Gulf of
Mexico are also of interest.
III)
HEAT FLUXES L. YU96 AND R. A. WELLER91
Oceanic latent heat flux (LHF) is heat energy extracted
from the ocean by the evaporation of surface water,
while sensible heat flux (SHF) is heat energy transferred
by conduction and convection at the air–sea interface.
These two fluxes vary with near–sea surface circulation, humidity and temperature, and influence weather
and climate processes. The global change of the two
fluxes in 2005 is examined here using daily analyzed
fields produced by the Objectively Analyzed Air–Sea
Fluxes (OAFlux) project. The resulting flux fields were
validated against more than 100 flux buoy measurements acquired by the Woods Hole Oceanographic
Institution (WHOI) Upper Ocean Processes group
(Moyer and Weller 1997) and by the Pilot Research
Moored Array in the Tropical Atlantic (PIRATA) and
Tropical Atmosphere Ocean/Triangle Trans-Ocean
Buoy Network (TAO/TRITON) moored buoy arrays in
the tropical Atlantic and the equatorial Pacific Oceans
(Yu et al. 2004).
The 2005 annual mean field of LHF plus SHF over
the global oceans (Figure 3.5, top) shows that large
oceanic latent and sensible heat losses occurred over
the regions associated with three major Western
Boundary Currents (WBCs): the Kuroshio, Gulf Stream
and African Agulhas Currents. In these regions, nearsurface vertical gradients of humidity and temperature
are largest, wind speeds are greatest in the respective
hemisphere’s fall and winter seasons, and the cold season variability of LHF plus SHF dominates the annual
mean pattern. On a year-to-year basis, variability of
the oceanic heat losses in the three WBC regions is
also largest, with magnitudes reaching or exceeding
Figure 3.4. Combined satellite altimeter and in situ
ocean temperature data upper-ocean (0–750 m) heat
content anomaly (OHCA) (J m-2) map for (top) 2005
relative to a 1993–2002 base period, following Willis
et al. (2004), and (bottom) the difference between 2005
and 2004 OHCA maps expressed as a surface heat flux
equivalent (W m-2)
Figure 3.5. (top) Annual mean latent plus sensible
heat fluxes (W m-2 ) in 2005; sign is defined as upward
(downward) positive (negative); (bottom)
differences between 2004 and 2005 annual mean latent
plus sensible heat fluxes
19
40 W m-2. This is clearly shown from the difference
map of the 2004 and 2005 annual mean LHF plus SHF
(Figure 3.5, bottom). Interestingly, the signs of the
anomalies associated with different WBC systems
are different. For example, the total oceanic heat loss
was enhanced (large positive anomalies) from 2004
to 2005 over the Kuroshio region, but reduced (large
negative anomalies) over the Gulf Stream region.
Influence of eddy-scale structures is evident in
the 2-year difference map (Figure 3.4, bottom).
Nevertheless, the change in the LHF plus SHF field
from 2004 to 2005 is large scale—anomalies are
primarily negative over the global basins, suggesting that the oceanic heat loss was overall reduced
in 2005 (Figure 3.6, right). A persistent, long-term
increase in the globally averaged annual mean LHF
plus SHF is particularly notable, although that trend
appears to have decreased in recent years. It is not
clear yet whether or not the reduction in oceanic
heat loss for 2005 is a trend change or merely a
shor t-term anomaly. Additionally, the globally
averaged LHF plus SHF has increased by about
10 W m-2 in the past 20 year, and the magnitude of the
variability is dominated by that of the LHF. The
mean SHF value is about one order smaller than
that of LHF, and the change in SHF is also small
(< 2 W m -2 ) over the entire analysis period.
Areal averages of LHF plus SHF variability in the
Kuroshio and Gulf Stream regions clearly show a
large upward trend in both (Figure 3.6, left). However,
unlike the global averages, the regional averages
show strong interannual fluctuations. Whereas the
upward trend in global averages begins in 1981 (start
of the analysis record) and has flattened in recent
years, the boundary current regional averages show
a trend toward larger values only starting in the
early 1990s but remaining positive to the present.
Furthermore, the slope of the trend in these two
regions is twice as great as that of the global
averages. Overall, the regional oceanic heat loss
has been enhanced by about 20 W m -2 over the past
two decades.
C.
CIRCULATION
I)
SURFACE CURRENTS R. LUMPKIN48 AND G. GONI30
Near-surface currents are measured in situ by
satellite-tracked drifting buoys and by acoustic pointmeasuring meters on the Autonomous Temperature
Line Acquisition System (ATLAS) moorings. In 2005,
the drifter array reached its target goal of 1250 drifters
worldwide, becoming the first fully realized component
of the Global Ocean Observing System. During 2005,
surface currents were well sampled globally, except in
the far northern Pacific, the southwest Pacific between
20° and 40°S, 150°E to the date line, the Arabian Basin
of the Indian Ocean, and the extreme Southern Ocean
south of 55°S.
A climatology of monthly mean currents was computed
from all available drifter observations from 1994 to
2004, using the methodology of Lumpkin and Garraffo
(2005). Anomalous currents were calculated with
respect to this climatology (Figure 3.7).
Figure 3.6. (left) Year-to-year variations
of globally averaged annual mean latent
heat flux (red), sensible heat flux (blue),
and latent plus sensible heat flux (black);
(right) year-to-year variations of the annual
mean latent heat plus sensible heat fluxes
averaged over the regions of the Gulf Stream
[(25°–45°N, 85°–50°W), red] and Kuroshio
[(20°–40°N, 120°–150°E), black]
20
(i)
Indo-Pacific basins
Annual mean anomalies were most prominent in the
tropical Pacific Ocean (Figure 3.7). Westward anomalies of nearly 20 cm s-1 were observed on the equator
between 120°W and the date line. Weaker anomalies of
5–10 cm s-1 were seen in the North Equatorial Current
(NEC) region (10°–20°N). Drifter observations did not
indicate anomalously strong eastward currents in the
Kuroshio Extension or North Pacific Current, conflicting
with the simple hypothesis of a more-intense-thanaverage wind-driven gyre.
The strongest intraseasonal anomalies were observed
in early 2005 in the western and central tropical Pacific,
associated with Kelvin wave activity driven by intraseasonal (MJO) wind fluctuations (cf. Eisenman
et al. 2005). In January (Figure 3.8), very strong
westward anomalies were measured in the northern branch of the South Equatorial Current (nSEC).
The nSEC at 0°–5°S, 160°W–170°E was 80–100 cm s-1
westward, compared to a mean January speed of 40–60
cm s-1. During February, a dramatic reversal was seen
at 6°–12°S, 155°W–180° where several drifters moved
eastward at 50–100 cm s -1. Drifters suggested the
passage of a second Kelvin wave in March and April,
when strong westward, then eastward, anomalies
were seen west of the date line (Figure 3.8). This was
corroborated by observations at the TAO mooring at
0°, 170°W. The previously noted NEC anomalies were
first observed in February. Westward anomalies in the
South Equatorial Current (SEC) at 0°–6°S appeared in
April. Both the NEC and SEC continued to flow anomalously quickly through the remainder of the year.
Figure 3.7. 2005 mean anomalies (cm s-1) from 1994–2004
surface current climatology
(ii)
Atlantic Basin
The North Atlantic subtropical gyre, benchmarked
by transport through the Florida Straits and Yucatan
Channel, was close to its decadal mean strength
during 2005. The Western Boundary Current transport
through the Florida Straits, measured by cable voltage,
averaged 31.4 ±1.2 Sv (1 Sv is 10 6 m 3 s-1) during
2005 (C. Meinen 2006, personal communication), not
significantly different from the 1982–2005 mean of
32.1 Sv. Transport through the Yucatan Channel,
estimated from altimetric sea level anomaly
variations calibrated by in situ transects, indicated a
transport within 1 Sv of the long-term mean. In the
North Equatorial Current region (5°–15°N, 30°–50°W),
westward anomalies of 10 cm s-1 were observed. Drifter
deployments during recent French research cruises
seeded the Gulf of Guinea region heavily for the first
time, and anomalies here were large, but may reflect a
poorly defined climatology for the period of 1994–2004.
Thus, it is difficult to tell from these data what role
anomalous advection may have played in the development of the unusually large cold tongue during 2005.
(iii) Southern Ocean
Drifters did not reveal large-scale anomalies in the
strength of the Antarctic Circumpolar Current (Figure 3.7).
From August to December, 18 drifters passed south of
Cape Horn. These drifters indicated that the flow entering
the Drake Passage was 8 ±9 cm s-1, much weaker than
the mean speed of 23 ±11 cm s-1 here. This anomaly was
most prominent during August–September. Four drifters passed through the region during February–May,
Figure 3.8. January–April 2005 drifter trajectories in the
tropical Pacific; arrowheads indicate direction; colour
indicates zonal speed anomaly (m s-1, positive eastward).
21
Figure 3.9. Monthly geostrophic transport
anomalies (Sv) of the Agulhas Current at
28°W derived from satellite altimetry
observations
measuring speeds of 27 ±14 cm s-1 (very close to normal). Altimetric estimates of geostrophic transports
suggest that the 2005 annual mean of the geostrophic
transport remained slightly lower than its historical
value, and significantly lower (5–10 Sv) than during
2004 (Figure 3.9).
II)
THERMOHALINE CIRCULATION M. O. BARINGER4 AND
C. S. MEINEN54
The component of the ocean circulation associated
most with variability in heat redistribution is the
meridional overturning circulation (MOC), also called
the thermohaline circulation. The MOC is a global
circulation cell wherein high-latitude surface waters
are cooled and become denser, and in certain locations this dense water sinks and flows toward the
tropics. In tropical and subtropical regions around
the world these waters warm, become less dense
and return to the surface to flow back toward higher
latitudes, while transporting a signifi cant amount
of heat. The primary locations where deep convection occurs are in the northern North Atlantic and
in the subpolar ocean around Antarctica, while the
upwelling of new surface waters is spread broadly
around the globe.
Variations in the strength of the overturning circulations are directly related to variations in net poleward
heat transport. Current best estimates for steady-state
global mass and heat transport can be found in the
analyses of Ganachaud and Wunsch (2003; see also
Talley 2003). For example, the North Pacific overturning cell carries about 0.5 PW northward, but most
of this heat transport is associated with water mass
transformations in the upper layers of the ocean, while
the North Atlantic carries a much larger transport of
1.2 PW northward, most of which is carried in the
top-to-bottom MOC system. Typically, it is this deep
circulation cell that is described as the MOC, although
other oceans and water masses are important for the
redistribution of heat.
22
The Florida Current contains most of the upper limb
of the MOC as it flows through the Florida Straits in
the North Atlantic, with a smaller contribution being
carried by the Antilles Current east of the Bahamas.
Fluctuations in the Florida Current have shown a clear
negative correlation with the atmospheric phenomenon
known as the NAO; however, while the NAO has been
trending toward its negative extreme over the past
20 years, the Florida Current transport shows no such
long-term trend through 2005 (Figure 3.10). The annual
mean transport observed in 2005 (31.4 Sv) falls just
within the lowest quartile of historical annual mean
transports from the cable. However, this transport is
well within 1 std dev of the long-term mean of 32.1 Sv,
and given the statistical standard error (1 Sv) of the
mean for the year, 2005 cannot be considered anomalous in terms of the Florida Current transport.
In the near future, a new NOAA-funded system for
monitoring the Deep Western Boundary Current, coupled
with a new international program for monitoring basinwide circulation [Meridional Overturning Circulation
Heat-Transport and Heat-Flux Array (MOCHA)] will
allow for a greater degree of certainty in statements
regarding variations in the integrated, basin-wide MOC
Figure 3.10. Florida Current transport (Sv; red solid)
as measured by the NOAA-funded submarine cable
across the Florida Straits, along with the NAO index
(blue dashed) produced by NOAA/NCEP
circulations and the time scales on which they vary.
Bryden et al. (2005), postulated a 30% reduction in
the MOC transport between the 1950s and the present day; however, that analysis was based on a very
limited data set. Other data over the past few decades,
such as the moored observations of the Deep Western
Boundary Current at the southeast Newfoundland
Rise in the early 1990s and early 2000s, showed no
indication of such a significant trend in the MOC
(F. Schott 2005, personal communication). In contrast,
Koltermann et al. (1999) showed the large variability of
the MOC that they concluded was related to the strength
of Labrador Sea Water production with larger (smaller)
MOC transport corresponding to less (more) Labrador
Sea Water export. More recently, these data have been
reanalyzed to formally test the hypothesis that the MOC
circulation was steady. Lumpkin et al. (2006, manuscript
submitted to J. Mar. Res.) found that a steady MOC over
the same time period could not be ruled out, based on
the uncertainty in determining the barotropic circulation. Also, if Bryden et al. (2005) are correct in that the
thermohaline circulation has been reduced over the past
50 years by 30%, this represents a much higher rate of
change than that predicted in coupled climate model
simulations (e.g., Schmittner et al. 2005).
Changes in heat transport on the order of 30%, such
as that postulated by Bryden et al. (2005), could
have substantial impacts on climate, and time series
observations are clearly necessary in order to put onetime hydrographic sections in temporal context. The
programs in place in 2005 are an excellent first step
toward the development of an integrated measurement
system, including repeated hydrographic sections and
moored observations that can play a significant role
in measuring and monitoring the MOC system over
a broader range of locations than would be feasible
with moored instrumentation alone. However, much
work remains before it will be possible to state that the
Atlantic MOC system, much less the global overturning
circulation system, is truly being monitored.
D.
SEA LEVEL M. A. MERRIFIELD,56 S. GILL26 AND
G. T. MITCHUM57
Data on sea level were obtained from both tide gauges
and satellite altimetry [Ocean Topography Experiment
(TOPEX)/Poseidon/Jason]. Tide gauges have long
observed sea level, and these instruments still provide the long-term context for understanding climate
variations. Since 1992, however, satellite altimetry
Figure 3.11. Deviation of 2005 annual average sea level (cm)
from 1993 to 2001 mean, as measured by (coloured shading)
satellite altimeter and (filled circles) tide gauges [Source:
University of Hawaii Sea Level Center]
has provided global views of the sea surface height
field. Sea level anomalies in 2005 were computed relative to 1993–2001 mean values (Figure 3.11). Coastal
and island sea level deviations, as measured by tide
gauges, were generally consistent with the deep ocean
patterns measured by satellite altimeters. A notable
exception was the east coast of North America, where
coastal tide gauge sea levels were higher than normal,
whereas nearby ocean values from altimetry were
lower than normal.
The deviations shown in Figure 3.11 were above average over most of the global ocean. Since at least 1993
(the time span of TOPEX/Poseidon/Jason altimeter
measurements), global sea levels have been rising
at a linear rate of 2.9 ±0.4 mm yr -1 (Leuliette et al.
2004; Cazenave and Nerem 2004; see also information
online at http://sealevel.colorado.edu/ ). The general
increase in globally averaged sea level during 2005 was
consistent with this longer trend. At any particular
point, however, the rate of sea level change can be very
different from the globally averaged rate. A notable
example is along the coasts of North America, where
sea level has fallen in recent years. In the North Pacific,
this is presumably associated with a trend in equatorward winds near the coast, which favor upwelling.
Recent sea level trends appear to be determined by the
polarity of the Pacific Decadal Oscillation (Mantua et
al. 1997; Cummins et al. 2005; see information online
at http://tao.atmos.washington.edu/pdo/), and it seems
likely that the North Pacific sea level trend pattern will
reverse sign, resulting in rising sea levels near the
North American coast.
23
O ther regional pat terns are also evident. In
general, levels were higher than average in the
tropics and the Southern Hemisphere. Regions
of lower-than-average sea level occurred in the
midlatitude North Atlantic, the North Pacifi c, the
tropical Indian Ocean, and off the east and west coasts
of Australia. In many instances, the sea level deviations
can be linked directly to anomalous surface winds
(Figure 3.12). A striking example is a cyclonic wind
anomaly centred over the Nor th Atlantic. This
anomaly corresponds to a weakening of the Bermuda
high, and possibly to a corresponding weakening of
the subtropical oceanic gyre. This would account for
the aforementioned lowered sea level in the open
ocean, and the higher-than-average tide gauge sea
levels along the North American coast.
Relative to the previous year, 2005 saw a noticeable
increase in sea level in the region of the western
tropical Pacifi c north of the equator and east of the
Philippines. This pattern developed in response to
the increased local anticyclonic wind forcing in 2005
relative to 2004. Otherwise, sea level changes in the
tropical oceans were generally weak, consistent with
the relatively weak ENSO variability experienced in
2005 (see section 4B).
In addition to these large-scale, low-frequency
patterns, extreme sea levels were also examined.
NOAA has developed a new exceedance probability
product from long-term tide station records using
a generalized extreme value analysis approach
(Zervas et al. 2005). The remarkably active hurricane
season of 2005, in terms of number and severity of
storms, produced unusually extreme water levels on
the east and Gulf coasts of the United States. The
eastern Pacifi c was also subject to extreme water
levels associated with several tropical cyclones
(see section 4C).
E.
Because there were no ocean carbon measurements prior to the mid-1800s, the anthropogenic CO2
component of the tot al dis solved inorganic
carbon (DIC) concentration was estimated using a
back-calculation technique based on the current
understanding of the physical and biological
contributions to the measured DIC (Gruber et al.
1996; Sabine et al. 2004). As a consequence, it is
implicitly assumed that the ocean circulation and
biological processes were in steady state over the
industrial era. Although this work provided our
best assessment of the state of the ocean in the
mid-1990s, these studies are unable to establish
temporal and spatial scales of variability, or the
temporal evolution of the ocean carbon cycle.
To address questions of decadal variability and
temporal evolution, the US Climate Variability and
OCEAN CARBON C. L. SABINE,79 R. A. FEELY,22 AND
R. WANNINKHOF88
The paucity of carbon measurements in the ocean
currently impedes our ability to generate annual
assessments of global anomalies. However, our
understanding of the global ocean carbon cycle
has greatly improved over the last decade based
on accumulated data and modeling activities.
Our best estimates of the current distributions of
natural and anthropogenic carbon in the ocean
24
stem from the collaborative efforts of two international programmes—the World Ocean Circulation
Experiment (WOCE) and the Joint Global Ocean
Flux Study (JGOFS) —to conduct an extensive survey of the chemical and physical properties of the
global ocean in the 1990s (Feely et al. 2001; Wallace
2001). An analysis of more than 70,000 carbon
measurements from this survey found that the
ocean accumulated approximately 118 Pg of carbon between 1800 and 1994 (Sabine et al. 2004).
This accumulation accounts for 48% of the CO 2
released from fossil fuel combustion over this same
time period.
Figure 3.12. Same as Figure 3.11, excluding tide gauge
deviations and including the 2005 deviation of surface winds
(m2 s-2 ) [Source: NCEP]
Predictability (CLIVAR) CO 2 Repeat Hydrography
Program has identifi ed 19 hydrographic sections
distributed around the global ocean that will be
re-occupied every 5–10 years. The programme began
in 2003 with three cruises in the North Atlantic that
were repeats of cruises in the 1990s. Each year, one
to three cruises are run in different locations, with
the goal of completing the first global resurvey by
2012. In 2005, cruises were run in the South Atlantic
and South Pacifi c Oceans, but these data will take
several years to finalize and thoroughly examine.
Analysis of the initial repeat lines over this past
year has indicated that several biogeochemical parameters are changing with time ( Feely
et al. 20 05 ). For example, changes of -10 to
+30 µmol kg -1 of DIC have been observed in the
upper 1000 m of the water column between the
1993 and 20 03 occupations of a track designated as A16N along 25°W in the North Atlantic
(Figure 3.13a). Although the magnitude of the changes
is not surprising, the patchiness of the changes was
not expected. More surprising is the fact that there
have been similar changes in the apparent oxygen
utilization (AOU) of the waters (a measure of the
decomposition of organic matter in the ocean),
indicating signifi cant changes in the organic matter cycling over the last decade that was previously
believed to be in steady state (Figure 3.13b). The
complicated patterns of these changes clearly show
that carbon is being influenced by more than simple
secular increases in atmospheric CO2. In some cases,
changes in circulation and organic matter cycling
may be masking anthropogenic changes, while in
other cases, these changes may enhance the apparent ocean carbon uptake.
Figure 3.13. Changes in (a) DIC (µmol kg-1) and (b) AOU
(µmol kg-1) in the upper 2000 m between the 2003 and
the 1993 occupations of A16N; positive values are an
increase in concentrations between 1993 and 2003
(modified from Feely et al. 2005).
Another intriguing preliminar y finding from a
comparison of a 2004 cruise in the North Pacifi c to
the aforementioned North Atlantic results is that
anthropogenic carbon inventories may be increasing
in the Pacifi c at about twice the rate of the Atlantic
over the last 10 years (Feely et al. 2005). This is in
contrast to the long-term anthropogenic CO2 inventory that shows larger column inventories in the
North Atlantic. The interpretation of these recent
fi ndings may lie in understanding the effects of
climate modes like the NAO or the Pacifi c Decadal
Oscillation (PDO) on the decadal-scale circulation.
These results also point to the need for improved
approaches for isolating the anthropogenic and
natural components of the observed variability.
25
4. THE TROPICS
A.
H. J. DIAMOND19 AND K. A. SHEIN,82 EDS.
OVERVIEW H. J. DIAMOND19
This tropics section consists of global input on two
primary topics: 1) ENSO and the tropical Pacific,
including ENSO seasonal variability, and 2) tropical
cyclone activity for the 2005 season in the following
seven basins: the Atlantic, northeast Pacific, northwest
Pacific, North and South Indian, southwest Pacific, and
Australia. The Pacific ITCZ also is discussed.
Persistent anomalous rainfall that characterizes mature
warm-phase ENSO conditions failed to develop near
the date line, except in February 2005. The relative
absence of such convection indicated that the ocean
and atmosphere were only weakly coupled during this
event. It should be noted that although conditions in
the Pacific met the definition for El Niño according to
the criterion recently formulated by the NOAA/National
Centers for Environmental Prediction (NCEP), namely,
five consecutive overlapping three-month seasons
with Niño-3.4 SST anomaly ≥ 0.5°C, this definition
is not universally agreed upon, either because it is
focused on a single region or defined only in terms
of SST. Regarding tropical cyclone activity, while the
2005 season was unprecedented in the Atlantic, setting
numerous records, other basins were characterized
by near- to below-normal levels of activity. Two sidebars expand upon the record Atlantic basin tropical
storm season.
B.
EL NIÑO–SOUTHERN OSCILLATION G. D. BELL,5
M. S. HALPERT,32 AND M. J. MCPHADEN53
I)
OVERVIEW
Conditions throughout the tropical Pacific during
2005 reflected a weak warm episode that ended in
March, followed by the development of below-average
SSTs during November and December (Figure 4.1a).
Intraseasonal variability often associated with the MJO
also was present. The weak Pacific warm episode early
in the year was accompanied by anomalously warm
waters at approximately 110–130 m depths across
most of the equatorial Pacific (Figure 4.1b). Subsurface
temperatures returned toward normal during April,
before cooling during the following months.
The weak warm episode did not show a strong relationship to the 850-hPa zonal wind anomalies, which
featured an oscillating pattern indicative of the MJO
(Figure 4.1c). However, a well-defined strengthening
of equatorial easterlies did accompany the transition to below-average SSTs during November and
December. These enhanced easterlies also contributed
to increased equatorial upwelling and a shoaling of
the oceanic thermocline.
There was generally minor reflection of the anomalously
warm SSTs in the pattern of deep convection over the
central and east-central equatorial Pacific early in the
year, as indicated by OLR anomalies (Figure 4.1d). An
El Niño signal was only evident during February, when
convection was enhanced over the central equatorial
Pacific (negative OLR anomalies) and suppressed across
Indonesia early in the year. This pattern impacted circulation over the Southern Hemisphere, probably resulting
in the observed drop in rainfall across southern Africa
during the heart of the rainy season.
Figure 4.1. Monthly time series of (a) SST (°C), (b) depth of the
20°C isotherm (m), (c) 850-hPa zonal wind speed (m s-1),
and (d) OLR (W m-2 ) over the central equatorial Pacific;
values were averaged over the region bounded by 5°N–5°S
and 180°–100°W. Five-day (solid line) and climatological
(1979–1995; dashed line) mean values are shown, as are
positive (orange) and negative (blue) anomalies (except for
OLR, where shading convention is reversed) relative to a
1979–1995 base period.
27
Figure 4.2. Time–longitude section
(5°N–5°S) of daily 200-hPa velocity
potential anomalies (m2 s-1) during 2005;
the shading interval is 3 × 10 6 m2 s-1,
and the thick solid contour is the zero
line. Anomalies are departures from
the 1971–2000 base period daily means,
and plotted using a five-day running
mean smoother.
Figure 4.3. Five-day average anomalies of (left) zonal wind (m s-1), (centre) SST (°C),
and (right) 20°C depth (m; an index for the depth of the thermocline) relative to the
mean seasonal cycle averaged over 2°N–2°S based on TAO/TRITON moored time
series data; the white line on the left panel indicates the 29°C isotherm, which
marks the eastern edge of the western Pacific warm pool. Ticks on the horizontal
axis indicate longitudes sampled at the start (top) and end (bottom) of record.
Convection was slightly suppressed across the central
equatorial Pacific for much of the year, and positive
OLR anomalies became increasingly pronounced
during November and December as the equatorial
easterlies strengthened and SSTs dropped.
II)
THE MADDEN–JULIAN OSCILLATION, KELVIN WAVE
ACTIVITY, AND ATMOSPHERIC CIRCULATION
Low-frequency variability in the tropics is strongly
influenced by the MJO (Madden and Julian 1971, 1972,
1994), a tropical disturbance that modulates tropical
convection and atmospheric circulation patterns with a
typical period of 30–60 days. The MJO tends to be most
active during ENSO-neutral years and can produce
ENSO-like anomalies (Mo and Kousky 1993; Kousky
and Kayano 1994). Low-level (850 hPa) and upper-level
(200 hPa) equatorial zonal winds and streamfunction,
200-hPa velocity potential and tropical convection,
and both sea surface and subsurface temperature
anomalies exhibited considerable intraseasonal variability during 2005 in association with the MJO.
The MJO is indicated in a time–longitude section
by continuous propagation of the 200-hPa velocity
28
potential anomalies around the globe (Figure 4.2).
Periods when the MJO was active include March–May
and July–September. This MJO activity contributed
to alternating periods of enhanced and suppressed
convection from the Indian Ocean to the date line,
and to alternating periods of low-level easterly and
westerly wind anomalies across the tropical Pacific
(Figure 4.1c).
The low-level wind anomalies associated with the
MJO can generate eastward-propagating oceanic
Kelvin waves, typically traveling 10° longitude week-1,
that feature downwelling in the mixed layer at their
leading edge and upwelling in their wake (Zhang
et al. 2001). Major Kelvin waves occurred during
March–May in response to very strong westerly wind
bursts over the western equatorial Pacific linked to
the MJO (Figure 4.3). As indicated by a time series of
area-averaged SST anomalies in the Niño-3 region, the
upwelling phase of one Kelvin wave produced a marked
drop in SSTs across the east-central equatorial Pacific
during February, followed by a return to near-normal
SSTs in March (Figure 4.4). The downwelling phase of
a second Kelvin wave reached the region in early May,
causing an abrupt increase in SSTs to 0.9°C above
average. These anomalies exceeded earlier values
associated with the weak warm episode. Several ENSO
forecast models interpreted this anomalous warming as
an indication of a possible return to El Niño conditions.
However, the upwelling phase of the MJO followed
behind, which dissipated the warmth soon thereafter.
An unusual feature of this El Niño was that excess
heat content along the equator, typically a precursor
to subsequent ENSO SST anomaly development (Jin
1997, Meinen and McPhaden 2000), did not precede but
rather developed in phase with Niño-3.4 SST anomalies
during 2004/2005 (Figure 4.5). The lack of a subsurface
heat content precursor may account for the relative
weakness of the 2004/2005 El Niño and the difficulty
in predicting its onset (Lyon and Barnston 2005). By
the end of 2005, the excess equatorial heat content
prevalent during most of 2004/2005 had disappeared
in association with the onset of cold La Niña conditions
(Figure 4.5).
Anomalous warming associated with the El Niño–like conditions during 2004/2005 was centred near the date line.
Conversely, near-normal SSTs prevailed in the eastern
Pacific and along the west coast of South America. The
failure of persistent warm SST anomalies to develop in
the eastern equatorial Pacific and along the west coasts
of the Americas limited the effects of this El Niño on
marine ecosystems and fisheries in those regions.
C.
TROPICAL CYCLONES
I)
SEASONAL ACTIVITY OVERVIEW H. J. DIAMOND19 AND
D. H. LEVINSON46
Averaged across all basins, the tropical storm seasons
of 2005 (2004/2005 in the Southern Hemisphere) saw
an above-normal number of named storms relative
to the 1981–2000 mean. Of these, fewer than normal
became hurricanes/typhoons/cyclones (HTCs), but
the number of major HTCs was slightly above average. Globally, 103 tropical storms (≥ 34 kt) were
recorded, with 53 becoming HTCs (≥ 64 kt) and 28
attaining major/intense (≥ 96 kt) status (compared
to an average of 97.25, 55, and 25.35 storms, respectively). The 2005 season was unprecedented in the
Atlantic, with numerous seasonal and individual
storm-related records. The highlights in the Atlantic
included the existence of three Saffir–Simpson category 5 storms (e.g., Wilma; Figure 4.6), which are
discussed in a sidebar on the Atlantic season. While
Figure 4.4. Time series of the Niño-3 region SST
anomaly index calculated over the area (5°N–5°S,
90°–150°W)
Figure 4.5. Monthly anomalies of Niño-3.4 SST (°C)
and warm water volume (WWV; × 1014 m3 ) from
January 1980 to December 2005; WWV, which is an
index of heat content along the equator, is based on a
blended thermal field analysis of TAO/TRITON moored
time series data and ship-of-opportunity expendable
bathythermograph (XBT) data integrated over the region
of 5°N–5°S, 80°W–120°E above the 20°C isotherm.
Niño-3.4 SST represents an average anomaly over
the region of 5°N–5°S, 120°–170°W. Time series have
been smoothed with a 5-month running mean filter.
Fig 4.6. Mexican Meteorological Service radar
image of category 5 Hurricane Wilma striking the
Yucatan Peninsula on 21 Oct 2005 [Source: Servicio
Meteorológico Nacional]
29
the Atlantic had its all-time busiest season ever with
27 total storms (see note in abstract), other basins
were characterized by near- to below-normal levels
of activity.
II)
Figure 4.7. June–October (a) 200–850-hPa vertical
wind shear magnitude total (m s-1) and (b) anomalies;
in (a) only values less than 8 m s-1 are shaded. In (b)
red shading indicates below-average strength of the
vertical shear; green box denotes the MDR. Anomalies
in (b) are departures from the 1971–2000 base period
monthly means; (c) a time series of August–October
area-averaged 200–850-hPa vertical shear of the zonal
wind (m s-1) across the MDR (inset). Blue curve shows
unsmoothed three-month values, and red curve shows
a five-point running mean applied to the time series.
ATLANTIC BASIN G. D. BELL,5 E. BLAKE,8 K. C. MO,58
C. W. LANDSEA,44 R. PASCH,65 M. CHELLIAH,15 AND
S. B. GOLDENBERG29
The tropical multidecadal signal incorporates the
leading modes of tropical convective rainfall variability occurring on multidecadal time scales. Three
important aspects of this signal responsible for the
increased hurricane activity since 1995 are 1) a stronger West African monsoon system; 2) suppressed
convection in the Amazon Basin; and 3) the warm
phase of the Atlantic multidecadal mode (Goldenberg
et al. 2001; Bell and Chelliah 2006). This tropical
multidecadal signal is very important to Atlantic
hurricanes because it affects an entire set of critical conditions across the main development region
(MDR) for decades at a time. The MDR consists of the
tropical Atlantic and Caribbean Sea south of 21.5°N
(Figure 4.7a, green box). During 2005, this signal again
set the backdrop for many of the observed atmospheric
and oceanic anomalies. The 2005 hurricane season
featured an extensive area of low vertical wind shear
(less than the 8 m s -1 threshold for tropical cyclone
formation) throughout the MDR and Gulf of Mexico
(Figure 4.7a), with the largest anomalies centred
over the central tropical Atlantic and Caribbean Sea
(Figure 4.7b). Since 1995, area-averaged vertical shear in
the heart of the low-shear area has been approximately
5–6 m s-1, with values during individual years such as
2005, 1999, and 1995 dropping to an incredibly low
2 m s -1.
Figure 4.8. June–October 2005 mean
(contours, interval is 10 × 10 6 m2 s-1)
and anomalous (shading) 200-hPa
streamfunction; anticyclonic anomalies
are indicated by positive values in the NH
and negative values in the SH. Cyclonic
anomalies are indicated by negative values
in the NH and positive values in the SH.
Anomalies are departures from the
1971–2000 base period monthly means.
Green box denotes the MDR.
30
This extensive area of low shear in 2005 resulted in
part from a stronger-than-average tropical easterly jet
and an expanded area of upper-level easterly winds
across the western MDR, both of which were related to
an enhanced subtropical ridge at 200 hPa. The 200-hPa
streamfunction field (Figure 4.8) shows that these conditions were part of a larger-scale pattern during 2005,
characterized by anticyclonic anomalies (Figure 4.8; red
in NH, blue in SH) and enhanced subtropical ridges in
both hemispheres from the eastern Pacific to Africa,
and by cyclonic anomalies in both hemispheres over the
central tropical Pacific. This tropics-wide pattern with
its pronounced interhemispheric symmetry is a classic
signature of very active Atlantic hurricane seasons.
Bell and Chelliah (2006) indicate that this pattern
signifies a response to anomalous tropical convection,
partly related to the ongoing tropical multidecadal
signal, and partly related to suppressed convection
over the central equatorial Pacific.
During 2005, long periods of anomalous upper-level
convergence were evident over the central tropical
Pacific, with shorter-period fluctuations sometimes
related to the MJO (Figure 4.9, solid line). Throughout
the season, these periods of suppressed convection
led to a strengthening of the 200-hPa subtropical ridge
over the western North Atlantic (Figure 4.9, orange
shading), which acted to focus periods of TC activity.
For example, 10 of the season’s 15 hurricanes and all
seven major hurricanes formed during these periods.
All seven early-season TCs and two mid-November
Figure 4.9. Five-day running mean time series showing areaaveraged anomalies of 200-hPa velocity potential (line) and
200-hPa streamfunction (shading) during 10 June–30 November
2005 (m2 s-1); velocity potential anomalies are calculated for
the central tropical Pacific region bounded by 10°–20°N,
160°E–170°W. Streamfunction anomalies are calculated for
the Gulf of Mexico and Caribbean Sea bounded by 10°–30°N,
60°–100°W. Small boxes below time series indicate when
tropical storms (blue), Category 1–2 hurricanes (green), and
major hurricanes (red) formed in the Atlantic basin. Anomalies
are departures from the 1971–2000 period daily means.
TCs also occurred when convection was suppressed
near the date line. This finding is consistent with the
study of Mo (2000). Also consistent with that study is
the break in activity during the first half of August in
association with MJO-related enhanced convection
near the date line.
The low vertical shear during 2005 was also associated with westerly wind anomalies in the lower
troposphere that reflected a markedly reduced
strength of the tropical easterly trade winds from
the eastern tropical Pacific to Africa. In combination
with the anomalous upper-level easterlies, this wind
ATLANTIC MONTHLY AIR–SEA FLUXES AND THE 2005 HURRICANE SEASON
M. A. BOURASSA,9 S. R. SMITH,84 P. HUGHES,33 AND J. ROLPH76
Latent heat flux refers to the rate at which
water vapour energy is transferred from the
ocean to the atmosphere. The release of this
energy occurs at higher altitudes in association with deep convective clouds and is a
critical energy source for developing tropical cyclones. Latent heat flux during 2005
was estimated using a newly completed
version of the Florida State University air–
sea flux fields (FSU3; Bourassa et al. 2005).
The objectively analyzed latent heat fluxes
across the tropical Atlantic were generally
100–140 W m -2 during June–September,
with larger values coinciding with tropical cyclone genesis regions (Figure 4.10).
Area-averaged anomalous latent heat flux
during the period was 10 W m -2 , which
is 20% greater than the largest value in
the 1978–2003 time series (not shown).
It is suggested that this increased latent
heat flux, combined with above average
SSTs and a strong cross-equatorial flow of
deep tropical moisture into the heart of the
MDR, led to a more unstable and deeper
boundary layer that favored increased tropical
cyclone activity.
Figure 4.10. Model-derived latent heat fluxes (W m-2 ) during June–September 2005; the first
letters of named storms are located where the storm was first named. The numbers indicate
similar positions for tropical depressions that did not become named storms.
31
distribution is consistent with the baroclinic response
of the atmospheric circulation to anomalous tropical
convection linked in part to the tropical multidecadal
signal (Bell and Chelliah 2006). Over the Caribbean
Sea, the weaker easterly trade winds corresponded
with exceptionally low SLP and 1000-hPa heights in
response to a weakening and northeastward shift of
the Bermuda high. These conditions were associated
with anomalous low-level cyclonic vorticity across the
northern half of the MDR, western North Atlantic, and
Gulf of Mexico.
These favourable conditions extended up to 700 hPa
along the equatorward flank of the African easterly jet
(AEJ) (Figure 4.11). The associated anomaly patterns at
this level are also consistent with the ongoing tropical
multidecadal signal. Conversely, during the preceding
period (1971–1994) of below-normal activity, higher
vertical shear (Figure 4.7b), combined with stronger
easterly trade winds and reduced cyclonic vorticity
south of the AEJ axis, were not conducive to Atlantic
hurricane formation in the MDR.
Favourable lower-level conditions during 2005 were
also associated with record warm SSTs across the
tropical Atlantic, Caribbean Sea, and Gulf of Mexico
(Figure 4.12, left). For the entire season SSTs across the
western tropical Atlantic and Gulf of Mexico reached
an 1870–2005 all-time high of 28.7°C (Figure 4.12,
right). The above conditions meant that African easterly
waves were embedded within an extended region of
anomalous cyclonic vorticity as they moved westward
over very warm SSTs into the low-shear, low-SLP,
high-vorticity environment of the central and western MDR. This combination is known to favour very
active hurricane seasons (Bell et al. 1999, 2000 2004;
Lawrimore et al. 2001).
While the combination of the supportive
multidecadal signal and suppressed convection
near the date line established an environment
conducive for a very active season, these factors
cannot account for periods when atmospheric
anomalies were exceptionally strong. Nor can they
account for the propensity of many TCs to develop
in and around, or traverse the Gulf of Mexico. Three
additional factors probably contributed to these
conditions. The first was a northeastward shift
and strengthening of the ITCZ over the eastern
North Pacific, with a corresponding increase in
convection over Central America and southern
Mexico. Associated with this pattern, an extensive
area of low SLP over the Gulf of Mexico contributed
to periods of actual low-level westerly winds across
the Caribbean Sea, which further strengthened the
low-level cyclonic circulation over the Gulf of Mexico.
Second was a persistent ridge of high pressure in the
middle and upper troposphere over the southeastern
United States and Gulf regions, which can be
linked to large-scale extratropical anomaly patterns
(Figure 4.9) and contributed to the low vertical shear
across the Gulf of Mexico and western MDR. This ridge
was particularly strong in July (Figure 4.8), when it
contributed to a sharp drop in vertical shear across the
southern and western MDR. Three hurricanes formed
in the MDR during this period, with two becoming major
hurricanes. Last, above-average SSTs in an already
very warm Gulf of Mexico (Figure 4.12, left) made
the environment over the western Atlantic and
Gulf of Mexico even more conduciveto tropical
cyclogenesis and major hurricane formation. In effect,
the environment in these regions during much of
the season typified the central MDR during an active
season. South of the anomalous upper-level ridge
over the eastern United States, enhanced easterly
Figure 4.11. June–October 2005 700-hPa
anomalous zonal winds (contour
interval is 1.0 m s-1) and relative vorticity
(shading, × 10 -6 s-1); solid (dashed) contours
indicate westerly (easterly) winds. Yellowred (blue) shading indicates anomalous
cyclonic (anticyclonic) relative vorticity.
Arrow shows mean position of the African
easterly jet axis during August–October.
Green box denotes the MDR. Anomalies are
departures from the 1971–2000 base period
monthly means.
32
Figure 4.12. Seasonal June–November (left) SST anomalies (°C) during 2005 and (right) time series of area-averaged SSTs (°C)
in the (left) green boxed region; (right) red line shows the corresponding 11-year running mean. Averaging region corresponds to
where seasonal activity was heavily focused during 2005.
flow also helped to steer many of the developing
TCs into the Gulf of Mexico, all of which eventually
made landfall.
Finally, the failure of many tropical storms to develop
until they reached the western part of the basin
is related to a pronounced eastward shift of the
mean upper-level trough to the extreme eastern
North Atlantic (Figure 4.8). This shift occurred in
association with the persistent upper-level ridge
farther west, and resulted in periods of anomalous
upper-level westerlies and increased vertical wind
shear that suppressed tropical wave development
in the eastern MDR.
III)
EAST PACIFIC BASIN D. H. LEVINSON46
(i)
Overview of the 2005 season
The hurricane season in the eastern North Pacific
(ENP) basin typically begins in mid-May and runs
through the end of November, with a climatological
peak in September. The 2005 hurricane season in
the ENP was below normal, with the majority of
the activity in the basin occurring during August
and September. A total of 15 named storms (NSs),
7 hurricanes (Hs), and 2 major hurricanes (MHs)
developed in the ENP basin in 2005, which was less than
the NOAA National Hurricane Center (NHC) 1971–2003
seasonal climatology of 15.5 NSs, 9 Hs, and 4.3 MHs.
A tropical depression (TD) formed in mid-October
(TD 16-E) and was relatively long lived, but did not
intensify to tropical storm strength. In addition to the
total number of tropical cyclones, the 2005 ENP
hurricane season was also below average for landfalling
systems along the Pacific coast of Mexico.
(ii)
Comparison of the 2005 season with climatology
The variability of ENP tropical cyclone activity is summarized with several widely used parameters and
indices (Figures 4.13a–d). In terms of the number of NSs,
which is the total number of tropical cyclones that reach
at least minimal tropical storm strength (sustained winds
≥ 34 kt), activity in the ENP basin has been near average
since the mid-1990s, fluctuating between slightly above
or below the long-term mean of 15.5 yr -1 (Figure 4.13a).
However, the annual number of hurricanes and major
hurricanes have been below average in most years
since 1995, except for the El Niño years of 1997/1998.
In addition, there were no landfalling named systems
during the 2005 ENP season.
Despite the obvious importance of the number of tropical cyclones in any particular year, several other indices
are useful in determining the historical significance of
an individual hurricane season. Figure 4.13b shows this
activity in terms of the annual number of days with
hurricanes as well as hurricane days, and both of these
statistics have been below average since 1995, again
with the exception of the 1997/1998 El Niño years.
In recent years, the Accumulated Cyclone Energy
(ACE) Index (Bell et al. 2000) has been used as a
diagnostic tool for understanding the intensity and
duration of tropical cyclone activity. The ACE Index
value for the ENP basin in 2005 was approximately
97 × 104 kt2, which was below both the long-term mean
and median and also within NOAA’s definition of a
“below normal” season. Historically, the ENP basin had
a well-defined peak in the ACE Index during the early
1990s, with the highest annual value occurring in 1992.
With the exception of 1997, there has been a marked
33
decrease in the ACE Index for the ENP basin beginning in
1995. The below-normal ENP activity since 1995 appears
to be inversely related to the increased North Atlantic
activity (Lander and Guard 1998; Landsea et al. 1998,
1999; Goldenberg et al. 2001; Bell and Chelliah 2006).
(iii) Impacts
None of the ENP tropical storms or hurricanes that
formed in 2005 made landfall along Mexico’s Pacific
coast, which on average has at least one tropical storm
and one hurricane landfall each year (Jauregui 2003).
However, remnants of several tropical storms and
hurricanes brought heavy precipitation to Mexico and
Central America, as well as the Hawaiian Islands.
The first tropical cyclone of the season, Adrian, formed
in mid-May from a tropical wave that crossed the Central
American isthmus and intensified into a tropical storm
on 18 May and to a hurricane the following day. Adrian
reached peak intensity on 19 May, about 139 km southwest
of El Salvador, with sustained winds of 70 kt. Poststorm
reports from ships and reanalysis of satellite-derived
intensities using the Dvorak (1984) technique indicate
that the storm had weakened to a tropical depression
just prior to landfall. Despite this rapid dissipation, heavy
rainfall associated with the remnants of Adrian-generated
flooding and mudslides resulting in one storm-related
death in Nicaragua (source: NOAA/NHC).
The other tropical cyclones that impacted Mexico were
TS Dora, whose centre of circulation did not cross
the coast but passed just 65 km off Zihuatanejo, and
H Otis, that dissipated about 148 km northwest of Cabo
San Lazaro. Otis generated tropical storm conditions
in some coastal locations and heavy rainfall in the
mountains of the southern Baja Peninsula.
The only tropical cyclone to impact the Hawaiian Islands
during the 2005 season was MH Kenneth, which had
the longest duration and reached the highest intensity of any tropical cyclone in the ENP basin in 2005.
Kenneth initially became a tropical depression on
14 September, intensifying to a tropical storm on the
15th, and to a hurricane the following day. It reached its
maximum intensity of 115 kt (minimal category 4 on the
Saffir–Simpson scale) on 18 September, moved slowly
east-northeastward and weakened to a tropical storm.
Despite moving into a region of cooler SSTs, Kenneth
reintensified into a hurricane and moved northwestward
toward the island of Hawaii, finally dissipating just east
of the island on 30 September. The remnants of Kenneth
produced locally heavy rainfall, but there were no official
reports of damages or injuries.
Figure 4.13. Seasonal tropical cyclone
statistics for the east North Pacific Ocean
over the period 1970–2005: (a) number of
NSs, Hs, and MHs, (b) days with hurricanes
(the number of days with winds ≥ 64 kt) and
hurricane days (days with winds ≥ 64 kt times
the number of storms with winds ≥ 64 kt),
(c) the Accumulated Cyclone Energy (ACE)
Index (× 104 kt2 ) with 2005 highlighted in red,
and (d) the maximum and mean maximum
wind speed (kt), and mean hurricane
duration (days); all time series in (a)–(d)
include the corresponding 1971–2003 base
period means.
34
(iv)
Environmental influences on the below-normal east
North Pacific hurricane season
Tropical cyclone activity (both frequency and intensity)
is influenced by several large-scale environmental
factors, including SSTs, vertical wind shear in the
mid- and upper-troposphere, the phase of the QBO
in the tropical lower stratosphere, and the phase of
the ENSO in the equatorial Pacific region (Whitney
and Hobgood 1997). In 2005, above-normal vertical
wind shear in the midtroposphere (850–200 hPa) was
observed in the ENP basin during the 3-month (July–
September) peak of the hurricane season (Figure 4.14,
top). Vertical wind shear anomalies exceeding 6 m s-1
occurred in the ENP basin’s MDR, which is defined here
between 10°–20°N and 90°–130°W (Figure 4.14, red box).
In addition to increased vertical wind shear, SSTs
were below normal in the MDR during the peak of the
hurricane season in 2005, with an area of -1° to -1.5°C
SST anomalies off the coast of the Baja Peninsula
(Figure 4.14, bottom). These cool SST anomalies were
located at the northern edge of the MDR, in a region
of normally cooler water due to coastal upwelling
and cold advection from the north associated with
the California current.
(JTWC). The 16 TYs in 2005 equaled the climatological
median, as did the four supertyphoons.
Other indices of tropical cyclone activity indicate a
slightly below- to near-normal level of activity in the
WNP in 2005. The 2005 ACE Index (Figure 4.16a) was
298 × 10 4 kt2 in 2005, which was slightly above the
climatological median of 289.8 × 10 4 kt2 and mainly
due to the occurrence of four supertyphoons in 2005,
which accounted for 75% of the total.
Monthly ACE values (Figure 4.16b) were slightly above
normal in the early season until September, and below
normal during October–December. The June ACE Index
Tropical cyclones in the ENP basin typically attain a
higher intensity when the QBO is in its westerly phase
at 30 hPa in the tropical lower stratosphere, but there
is also a corresponding decrease in the observed
seasonal frequency of storms (Whitney and Hobgood
1997). In 2005, the phase of the QBO anomaly was
westerly, but with relatively weak zonal wind anomalies
< 10 m s-1 at 30 hPa during the majority of the hurricane season (except in October when the phase of
the 30-hPa zonal wind became easterly). Therefore,
the activity in the ENP basin during 2005 was mixed
relative to the phase of the QBO.
IV)
WESTERN NORTH PACIFIC BASIN S. J. CAMARGO14
Tropical cyclone [typhoon (TY)] activity in the western
North Pacific (WNP) was slightly below average in
2005, with 25 tropical cyclones (TDs, TSs, and TYs),
which was below the 1971–2004 climatological median
of 31 (Figure 4.15a). Only one of the tropical cyclones
(TD 20W) failed to reach tropical storm intensity or
higher. There were 24 NSs in 2005 (8 TSs and 16 TYs),
which was below the climatological median of 26. The
eight TSs in 2005 were also slightly below the climatological median of nine, and one of these (25W) was not
officially named by the Joint Typhoon Warning Center
Figure 4.14. (top) July–September 2005 200–850-hPa vertical
wind shear anomaly (m s-1; relative to 1979–2004 means;
source: North American Regional Reanalysis data set); (bottom)
July–September 2005 sea surface temperature anomalies (°C)
from NOAA’s OI data set (Reynolds and Smith 1994; Reynolds
et al. 2002). The main development region for ENP hurricanes
is the area delineated by the red box in both panels.
35
the climatological mean (19.0°N, 134.2°E). This shift
to the northwest was due to several tropical cyclones
that formed at higher latitudes (at or north of 20°N)
and the absence of low-latitude (at or south of 15°N)
tropical cyclones east of 155°E.
Figure 4.15. (top) Number of tropical storms, typhoons
and supertyphoons per year in the WNP for the period
1945–2005; (bottom) cumulative number of named
storms per month in the WNP: 2005 (black squares and
line), and climatology (1971–2004) shown as box plots
(box = interquartile range (IQR), red line = median,
* = mean, + = values in top or bottom quartile,
and diamonds (circles) = high (low) records in the
1945–2005 period) [Source: JTWC]
value was the eighth highest in the historical record.
In contrast, November and December ACE Index values were the 8th and 13th lowest, respectively, in the
historical record. The high June ACE value was due to
typhoon Nesat, which was in the top 5% of the historical record of ACE per storm. The ACE index values of
supertyphoons Longwang and Nabi were both in the
top 10% of the historical record of ACE per storm.
The cumulative number of named storms per month
(Figure 4.15b) also shows an active early season
(March–May), a slightly above-normal mid-season
(August and September), and below-normal activity
in the late season (October–December). The 131 NS
days (days in which at least one tropical cyclone with
tropical storm intensity or higher occurred) in 2005
was below the climatological median of 163.1 days.
In contrast, the 29.25 days with intense typhoons
(number of days having at least one typhoon with
sustained winds ≥ 96 kt) was well above the climatological median of 20 days, and in the top 15% of
the historical record.
WNP tropical storms and typhoons in 2005 had an
average cyclogenesis position of 14.5°N, 138.4°E,
which is slightly northwest of the climatological mean
position (12.8°N, 143.5°E; 1.8° lat, 6.6° lon std dev).
The average track position of all named storms was
20.4°N, 132.3°E, which is also slightly northwest of
36
The 2005 WNP typhoon season was responsible for
many fatalities and economic losses. Eight named
tropical cyclones made landfall in China, resulting in around 300 deaths and losses estimated at
US$ 3 billion (see also section 6F). Four cyclones
affected Vietnam in 2005 (Damrey, Haitang, Longwang,
and Kai Tak), and three strong typhoons (Talim, Haitang,
and Longwang) made landfall in Taiwan. While Japan
was struck by 10 tropical cyclones in 2004, only two
(Nabi and Mawar) made landfall there in 2005. The
high number of landfalls in China, Vietnam, and Taiwan
(Figure 4.17a) and the relative absence thereof in
Japan were related to anomalous 500-hPa winds and
vertical wind shear (Figure 4.17b). Also present were
anomalous low-level (850 hPa) easterlies, probably
responsible for the absence of typhoons east of 150°E
at low latitudes.
V)
INDIAN OCEAN BASINS K. L. GLEASON28
(i)
North Indian Ocean
The North Indian Ocean (NIO) tropical cyclone season
extends from May to December, with two peaks in
activity during May–June and November when the
monsoon trough is climatologically positioned over
tropical waters in the basin. Tropical cyclones in the
NIO basin develop in the Bay of Bengal and the Arabian
Sea, typically between latitudes of 8° and 15°N, and are
usually short lived and weak, quickly moving into the
subcontinent. However, severe cyclonic storms with
winds > 130 kt can develop (Neumann et al. 1993).
Using reliable records from 1981 to 2000, a mean of
less than one major cyclone (MCYC: sustained winds
≥ 96 kt), 1.6 cyclones (CYC: sustained winds ≥ 64 kt),
and 4.75 NSs (sustained winds ≥ 34 kt) forms each
year in the NIO. The 2005 tropical cyclone season
was near normal with one CYC and six NSs forming from January to December (Figure 4.18, top).
The sole CYC lasted at that strength for only 1 day,
resulting in a below-average number of NIO cyclone
days for 2005.
The estimated 2005 ACE Index for the NIO basin was
7 × 10 4 kt 2, which is less than half of the 1981–2000
mean of 19 × 10 4 kt 2 (Figure 4.18, bottom). In fact,
NIO activity has been below normal for the past
6 years. This period follows four active tropical cyclone
seasons with above-average ACE Index values during the late 1990s.
Tropical Cyclone Fanoos was the only hurricane-strength
storm of the NIO season. Fanoos developed over
the Bay of Bengal west of the Andaman Islands
in early December and tracked west toward
the southeast Indian coastline. On 9 December,
CYC Fanoos briefl y became a category 1 tropical
cyclone before making landfall on the Tamil Nadu
coast near Vedaranyam with 55-kt sustained winds.
Two tropical storms also made landfall in 2005. Tropical
storm 03B formed in early October along the Indian
coastline and came ashore near the West Bengal
state with 35-kt sustained winds. Tropical storm 04B
formed near the Indian coast and made landfall in the
Andhra Pradesh state in late October with sustained
winds of 35 kt. Impacts from both of these storms
were minimal.
Figure 4.16. (a) ACE Index (x 104 kt2 ) per year in the western
North Pacific for the years 1945–2005; the solid green
line indicates the median for the 1971–2004 base period
(climatology), and the dashed green lines show the 25th and
75th percentiles. (b) ACE Index per month in 2005 (red line)
and the median in the years 1971–2004 (blue line), where the
green error bars indicate the 25th and 75th percentiles. In the
cases of no error bars, the upper and/or lower percentiles
coincide with the median. The blue plus signs (+) denote the
maximum and minimum values during the period 1945–2005
[Source: JTWC]
In mid-September, a tropical depression developed
over the South China Sea and moved inland over
Thailand. Cyclonic Storm Pyarr, identified by the
Indian Meteorological Department, moved into
the Bay of Bengal and intensified into a tropical
storm with estimated sustained winds of between
35 and 45 kt. Pyarr made landfall over the northeast Indian coast and quickly dissipated. This
storm was not included in the NIO analyses for
2005 due to a lack of available track data and wind
observations.
(ii)
South Indian Ocean
The tropical cyclone season in the South Indian Ocean
(SIO) is typically active from December through April
and officially extends from July to June, spanning
two calendar years. The SIO basin extends south
of the equator from the African coastline to 105°E,
although most cyclones develop south of 10°S
latitude. Cyclones in the SIO that remain east of 105°E
are included in the Australian summary (see next section). The vast majority of SIO landfalling cyclones
impact Madagascar, Mozambique, and the Mascarene
Islands, including Mauritius. Due to a sparse historical observational record and the lack of a centralized
monitoring agency, the SIO is probably the least
understood of all tropical cyclone basins (Atkinson
1971; Neumann et al. 1993). As a result, the SIO
Figure 4.17. (a) Observed tracks of tropical cyclones that made
landfall in China, Taiwan, and Vietnam in 2005; (b)
anomalous 500-hPa winds (vectors) and anomalous vertical
wind shear (m s-1, shading) during July–October 2005
37
Figure 4.18. Annual tropical cyclone statistics for the
NIO 1970–2005 (top). Number of tropical storms,
cyclones, and major cyclones (bottom). Estimated
annual ACE Index (x 10 4 kt 2) for all NIO tropical
cyclones during which they were at least tropical storm
or greater intensities (Bell et al. 2000).
The ACE Index is estimated due to a lack of consistent
6-hour sustained winds data for every storm.
Figure 4.19. Annual tropical cyclone statistics for the
SIO 1980–2005 (top). Number of tropical storms,
cyclones, and major cyclones (bottom). Estimated
annual ACE Index (x 10 4 kt2 ) for all SIO tropical
cyclones during which they were at least tropical storm
or greater intensities (Bell et al. 2000).
The ACE Index is estimated due to a lack of consistent
6-hour sustained winds data for every storm.
statistics presented are incomplete but are based
upon verifiable information.
proceeded southeast over the cooler waters of the
SIO and dissipated. Later in the month, Tropical Storm
Felapi affected the same region. The combined effects
from both systems caused widespread flooding across
southern Madagascar. Earlier in the season, tropical
cyclone 02S formed east-southeast of the Seychelles
Islands in late October and moved west toward the
African coastline. This storm briefly strengthened into
a tropical storm before weakening and made landfall
near Dar es Salaam, Tanzania, as a TD.
Using reliable data from 1980 to 2000, the SIO averages 2.7 MCYCs, 6.1 CYCs, and 11.95 NSs each year.
During the 2004/2005 season (from July 2004 to June
2005), the SIO tropical cyclone occurrences were
near average with 3 MCYCs, 6 CYCs, and 14 NSs
(Figure 4.19, top). However, the estimated 2004/2005
SIO ACE Index was 53 × 10 4 kt 2 , which was less
than half of the 1981–2000 average of 112 × 10 4 kt 2
(Figure 4.19, bottom). This suggests a decreased
intensity of tropical cyclones during this season,
because both occurrences and mean cyclone
duration (~4 days) were near average.
Three SIO tropical cyclones made landfall during the
2004/2005 season. The first of these formed northeast of Madagascar in mid-January and moved west
toward the African coastline. This TD became MCYC
Ernest after intensifying to 100-kt sustained winds
(category 3) while moving south over the warm waters
of the Madagascar Channel. Ernest weakened slightly
before brushing the southwestern coast of Madagascar
on 22 January with 70-kt sustained winds. It then
38
The most intense SIO tropical cyclone of the
2004/2005 season was MCYC Bento, with sustained
winds of 140 kt. Bento developed in late November
2004, east-southeast of Diego Garcia in the central
Indian Ocean, and quickly intensifi ed. Fortunately,
Bento remained well south of Diego Garcia and
there were no signifi cant impacts on the island.
Bento was the first tropical cyclone on record in
the SIO to reach category 5 intensity equatorward
of 10°S latitude.
CYC Adeline-Juliet formed near the Cocos Islands
in the southeast Indian Ocean during early April
2005 and began its westward movement toward the
central SIO. Adeline-Juliet reached peak intensity
of 120-kt sustained winds (category 4) on 8 April
while tracking west-southwest to within 213 km of
Rodrigues Island. It then quickly dissipated as it
moved southward into an unfavourable environment
of strong vertical shear, drier air and cooler SSTs.
VI)
SOUTH PACIFIC BASINS M. J. SALINGER,80
A. B. WATKINS,90 AND S. M. BURGESS12
(i)
Southwest Pacific
The 2004/2005 southwest Pacific tropical cyclone season had a climatological average nine tropical storms
east of 150°E (Figure 4.20), four of which reached MCYC
strength. All but one of the nine tropical cyclones
originated east of the date line (Figure 4.21), and all
occurred between December and April.
The most devastating tropical cyclones of the 2004/2005
season were Meena, Nancy, Olaf and Percy. These
systems all occurred in February in association with
an active phase of the Madden–Julian oscillation.
Judy was the first tropical cyclone of the season
(25 December) and brought torrential rainfall to
parts of the Tuamotu Islands, French Polynesia.
CYC Kerry developed northeast of Vanuatu on 6 January,
passing over the Pentecost and Malekula Islands the
next day, with pressures as low as 987 hPa and maximum sustained wind speeds of 90 kt. Lola affected the
region near Tonga from 31 January through 2 February,
with strong winds at Fua’amotu Airport.
Figure 4.20. The number of southwest Pacific tropical
cyclones for the 2004/2005 season (solid red bar) compared
to frequencies during the past 30 years; the horizontal green
line indicates the 30-year average (not including Ingrid,
which originated west of 150°E).
MCYC Meena formed east of Samoa on 3 February and
tracked toward the southern Cook Islands. Estimated
maximum sustained wind speeds reached 125 kt, with
gusts to 155 kt. Tropical storm-force winds occurred at
Mauke on 6 February with pressures as low as 986 hPa.
Gale-force winds and gusts to 62 kt occurred in Rarotonga
on the same day, preceded by about 100 mm of rain.
MCYC Nancy affected the northern Cook Islands from 13
to 15 February tracking south, forcing Aitutaki residents
and tourists into shelters on 15–16 February. Estimated
sustained maximum winds in Nancy reached 125 kt,
while storm-force winds gusted to 88 kt in Rarotonga
with reports of 100 kt elsewhere. Heavy rainfall, pressure
as low as 988 hPa, and high seas also accompanied the
storm. In Aitutaki, trees were uprooted, roofs damaged,
and low-lying areas flooded. Waves caused widespread
destruction along the northern and eastern coasts of
Rarotonga. The island of Mangaia was also badly hit.
Olaf was named on 13 February, and reached MCYC
strength with 145-kt maximum sustained winds.
Observed winds exceeded 65 k t in Samoa on
16 February, damaging numerous structures. Gales
buffeted the northern and southern Cook Islands on
17 February, with gusts to 51 kt in Rarotonga.
MCYC Percy, with maximum sustained winds reaching
140 kt came next. Gales, storm surge, and high tides
affected Tokelau on 26 February, where Percy was
reportedly the worst tropical cyclone in living memory.
Many homes were damaged and roads washed out,
with water up to 1 m deep in some areas.
Figure 4.21. Southwest Pacific tropical cyclone tracks for the
2004/2005 season (including Ingrid)
39
Rae followed (near the southern Cook Islands) on
6 March, but remained weak. Finally, Sheila formed
east of Niue on 22 April and tracked southeast,
with maximum sustained winds reaching 35 kt.
(ii)
Australian basin
As with the 2003 / 2004 season, the 2004 / 2005
Australian basin TC season saw only six TCs
between 105° and 160°E, compared with a longterm average of approximately 10. Three were
classified as severe tropical cyclones.
The most severe, TC Ingrid, occurred between 5 and
16 March ( Figure 4 . 22 ) . Ingrid formed in the
Coral Sea south of Papua New Guinea and
tracked first east- then west ward across Cape
York Peninsula and through of f shore islands
THE RECORD-BREAKING 2005 ATLANTIC HURRICANE SEASON
The 2005 Atlantic hurricane season was
unprecedented and broke many tropical
cyclone records. The season featured a
record 27 tropical storms (previous record
of 21 in 1933), a record 15 hurricanes (previously 12 in 1969), a record three category
5 hurricanes (previously two in both 1960
and 1961), a record estimated ACE Index
(Bell et al. 2000) of 285% of the median,
and a record ACE value of 131% of the
median from named storms forming outside the MDR. The MDR is the tropical
Atlantic and Caribbean Sea south of 21.5°N
(Figure 4.7, green box). The season also
featured a record 15 named storms making
landfall in the Atlantic basin. Seven of these
made landfall in the United States, including
a record four landfalling major hurricanes
(categories 3 – 5 on the Saffir–Simpson
scale; Simpson 1974). Also, a record seven
named storms occurred during June–July,
and a record 10 late-season storms formed
after 1 October. Seven storms became
major hurricanes during 2005, one shy of
the 1950 record.
Two tropical storms, two hurricanes and four
major hurricanes struck the United States during 2005 (Figure 4.23). These totals reflect
the effects of H Ophelia on the Outer Banks
of North Carolina.The devastating impacts
of H Katrina on the central Gulf Coast in late
August, one of the worst natural disasters ever
to strike the United States, made the 2005
season the costliest season in the country’s
history, conservatively estimated at over
US$ 100 billion. Elsewhere, three tropical
storms and three hurricanes (including
Wilma) struck Mexico, one tropical storm
40
made landfall in the Dominican Republic,
and one hurricane struck Nicaragua.
In Guatemala, Stan claimed over 1,000 lives
in mudslides and flooding. Wilma was the
costliest natural disaster in Mexican history
with estimated damage at US$ 1–3 billion,
eight deaths, and a reported 1637 mm
(64 in.) of precipitation recorded on Isla
Mujeres. The number of additional monthly
and individual records is too numerous to list
here, but several are of particular note:
• Since the inception of the current naming
system in 1953, this year was the first
time the Greek letter naming convention
was used.
• Dennis became the most intense hurricane
on record before August when a central
pressure of 930 hPa was recorded.
Figure 4.23. Satellite montage of US-landfalling hurricanes [courtesy: C. Velden,
University of Wisconsin—Madison, Cooperative Institute for Mesoscale
Meteorological Studies (CIMMS)]
of f the Northern Territor y coast, before making
landfall in the far nor th of Western Australia.
Ingrid is the only TC in Aus tralia’s re corded
history to impact the coastline of three dif ferent
states or territories as a severe tropical cyclone
( categor y 3 or above on the Australian tropical cyclone warning scale). For tunately, some
weakening before its coastal crossings resulted
in only modest damage.
Figure 4.22. Observed track of Tropical Cyclone Ingrid along
the northern coast of Australia. [Courtesy: M. Foley and
M. Lesley, Australian Bureau of Meteorology (BOM)]
G. D. BELL,5 E. BLAKE,8 K. C. MO,58 C. W. LANDSEA,44 R. PASCH,65 M. CHELLIAH,15 S. B. GOLDENBERG,29 AND H. J. DIAMOND19
• Emily eclipsed the record set by Dennis
for the lowest pressure recorded for a
hurricane before August when its central
pressure reached 929 hPa.
• Vince was the first tropical cyclone in
recorded history to strike the Iberian
Peninsula. While Alberto (1988 ) was
the most northerly and Ginger (1967)
the most easterly, Vince was the most
northeasterly, and was also the furthest
east when it became a hurricane.
• Wilma’s central pressure dropped to
882 hPa, which was the lowest pressure ever measured in the Atlantic basin,
eclipsing the old record of 888 hPa set by
H Gilbert in 1988.
ongoing tropical multidecadal signal; and
(4) exceptionally conducive upper-level
and lower-level wind and air pressure
patterns over the western Atlantic and
Gulf of Mexico.
• Zeta tied Alice (1954) for latest naming
(30 December), and was the longest lived
into January (6 January).
NOAA’s ACE Index is a measure of
seasonal activity that accounts for the
combined strength and duration of tropical
storms and hurricanes during the season
(Figure 4.24; Bell et al. 2000). ACE is
calculated by summing the squares of the
6-hour maximum sustained wind speed
in knots for all periods while the system
is a tropical storm, subtropical storm or
hurricane. The 2005 ACE value was a record
249 × 10 4 kt 2 (285% of the 1951–2000
median value). The 2005 season marks
a continuation of the current active
hurricane era that began in 1995. The
historical time series of the ACE Index
indicates large multidecadal fluctuations
in seasonal activity (Goldenberg et al.
2001; Bell and Chelliah 2006). During the
11-year period from 1995 to 2005,
seasons have averaged 14.7 TSs, 8.4 Hs, and
4.1 MHs, and every season has been
classified by NOAA as above normal except
for the two El Niño years of 1997 and
2002. In contrast, seasons during the
below-normal period 1971–1994 averaged
only 9 TSs, 5 Hs, and 1.5 MHs, with only
three seasons classified as above normal
(1980, 1988, 1989). These large differences
between the above-normal and belownormal eras result almost entirely from
differences in the number of tropical storms
that form in the MDR and eventually become
hurricanes and major hurricanes (Landsea
1993; Goldenberg et al. 2001).
The activity of the 2005 season is attributed
to four main factors: (1) long periods of
anomalous upper-level convergence and
suppressed convection over the central
tropical Pacific, reminiscent of La Niña
conditions; (2) record warm sea surface
temperatures across the MDR; (3) the
Figure 4.24. ACE Index expressed as percent of the 1951–2000
median value; season types are indicated by the background
shading, with pink, yellow, and blue approximating above-,
near-, and below- normal seasons, respectively.
41
D.
PACIFIC INTERTROPICAL CONVERGENCE
ZONE A. B. MULLAN59
Prominent rainfall and cloudiness maxima and OLR
minima are associated with the Pacific ITCZ and the
SPCZ. The Pacific ITCZ has its main branch in the
Northern Hemisphere between 5° and 10°N, where
it is strongest during June–August (Figure 4.25).
Between December and February, the ITCZ branches
south of the equator, near 5°S, and is associated with
the monsoon trough near Australia. Convection frequently extends northward from the monsoon trough
across the equator in the region of the Maritime
Continent. The southern ITCZ rarely extends eastward
of about 160°W, except during March and April when
it can reach 90°W. High-resolution Tropical Rainfall
Measuring Mission (TRMM) rainfall data show that
the southern ITCZ is particularly prominent east of
160°W in La Niña years, such as 1999/2001.
Near the date line, the southern ITCZ merges with the
SPCZ, which extends southeastward across the subtropical South Pacific. The SPCZ is most active in austral
summer (December–February) and is located in a region
of strong SST gradient south of the maximum SSTs.
In contrast, the Northern Hemisphere ITCZ is located
near the axis of maximum SSTs (Vincent 1998).
The northern ITCZ was continuous across the Pacific
from 140°E to 90°W in almost all months except
February (Figure 4.25, top). Also during February,
the SPCZ was extremely active from about 160°E to
160°W in association with more extensive equatorial
westerlies and several tropical cyclones (see previous
section). From May to August the northern ITCZ was
positioned slightly equatorward of its normal location, and during June and August it was particularly
active in the far eastern tropical Pacifi c between
120°W and 90°W (Figure 4.25, bottom).
For the year as a whole, high-resolution TRMM rainfall
data (0.25° latitude × 0.25° longitude grid) suggest
that precipitation was slightly above the 1998–2004
average in the northern ITCZ (Figure 4.26), and also
in the southern ITCZ and SPCZ to about 160°W.
Figure 4.25. Average rainfall rate (mm h-1) from TRMM
0.25° analysis for (top) February and (bottom) August 2005;
contours are unevenly spaced at 0.05, 0.10, 0.15, 0.20,
0.30, 0.40, 0.60, 0.80, and 1.00 mm h-1.
Figure 4.26. Annual average rainfall rate (mm h-1) from TRMM
0.25° analysis for 2005, as a percentage of the 1998–2004 mean
42
5. THE POLES
A.
A. M. WAPLE,89 ED.
OVERVIEW A. M. WAPLE89 AND J. RICHTER-MENGE74
The permanent presence of sea ice, ice sheets and
continuous permafrost are unique features of the
polar regions. The Arctic is further distinguished
because it sustains human and wildlife populations
in a harsh environment, as documented in the Arctic
Climate Impact Assessment published in November
2004 (online at www.acia.uaf.edu). These characteristics amplify the impact of climate change on
the region’s physical, ecological, and societal systems. Such impacts reach beyond the Arctic region.
For instance, studies are under way to determine
the extent to which the loss of sea ice cover and
the conversion of tundra to larger shrubs and wetlands, observed to have occurred over the last two
decades, have impacted multiyear persistence in
the surface temperature fi elds, especially in the
Pacifi c sector.
In this section, observations that indicate continuing
trends in the current state of physical components of
the Arctic system, including the atmosphere, ocean,
sea ice cover, and land, are discussed. The temporal
extent of the data provides a multidecadal perspective
and confirms the sensitivity of the Arctic to changes
in the global climate system. The destabilization of
several known relationships between climate indices
(e.g., Arctic Oscillation) and Arctic physical system
characteristics (e.g., continued reduced sea ice cover
Figure 5.1. Time series of the annually averaged
AO Index for the period 1950–2005, based on data
from the following Web site: www.cpc.ncep.noaa.
gov [Courtesy: I. Rigor]
and increased greenness of the tundra) presents an
intriguing and significant puzzle with respect to the
contemporary global climate system.
It was one of the warmest years on record for Greenland,
with an especially warm and foggy spring. Surface
temperature and ice melt are discussed for Greenland
in the context of a warming trend over the nation.
Also, trends and observations in temperature, sea ice,
and stratospheric ozone depletion are discussed with
respect to Antarctica. As the continent with more than
70% of Earth’s freshwater storage, trends in Antarctic
climate are of critical importance in determining the
long-term impact of climate change.
B.
ARCTIC RICHTER-MENGE,74 J. OVERLAND,62 A. PROSHUTINSKY,68
V. ROMANOVSKY,77 J. C. GASCARD,25 M. KARCHER,37 J. MASLANIK,52
D. PEROVICH,66 A. SHIKLOMANOV,83 AND D. WALKER87
I)
ATMOSPHERE
(i)
Circulation regime
The annually averaged AO index in 2005 was slightly
negative, continuing the trend of a relatively low
and fluctuating index that began in the mid-1990s
(Figure 5.1). This follows a strong positive pattern
from 1989 to 1995. Current characteristics of the AO
Figure 5.2. Arctic-wide (60°N–90°) annual average
surface air temperature anomalies (°C) over land for
the twentieth century based on the Climate Research
Unit (CRU) TEM2V monthly data set
43
are more consistent with the period from the 1950s to
the 1980s, when the AO switched frequently between
positive and negative phases.
(ii)
Surface temperatures
In 2005, annual average surface temperatures over land
areas north of 60°N remained above the mean value for
the twentieth century (Figure 5.2), as they have since the
early 1990s. Figure 5.2 also shows warm temperatures
in the 1930s and early 1940s, possibly suggesting a
longer-term oscillation in climate. However, a detailed
analysis shows different proximate causes and characteristics for the 1930s compared to the 1990s maxima.
The early warm anomalies appear to be characterized
by large region-to-region differences and are limited
to high latitudes (Johannessen et al. 2004; Overland et
al. 2004). The warm anomalies since the 1990s tend to
be Arctic-wide and reach into the midlatitudes.
Near-surface air temperatures in boreal winter and
spring 2005 continued to have the same general spatial pattern of warm anomalies as that of 2000–2004
(Figure 5.3a). The first major feature is positive (warm)
anomalies over the entire Arctic, consistent with
Figure 5.2. The second feature is the strong maxima
north of eastern Siberia and around the Davis Strait.
Spring anomalies (March–June) over the last 6 years
for these coastal areas are near 3°–4°C. The region
north of eastern Siberia is also a main location for loss
of sea ice cover over the last decade. While the Arcticwide pattern of positive anomalies was established in
the early 1990s, the locations of the 2000–2005 temperature anomaly maxima (consistent with the 2005
pattern shown in Figure 5.3a) contrast with those of
the early 1990s (Figure 5.3b).
The current pattern of near-surface temperature
anomalies (2000–2005) is distinctly different from the
near-surface temperature anomaly patterns associated
with the two major atmospheric circulation patterns
that characterized the second half of the twentieth
century (Quaddrelli and Wallace 2004). The positive
phases of these patterns were present during 1989–1995
(Figure 5.3b) and 1977–1988 (Figure 5.3c) and were
associated with the AO and Pacific–North American
(PNA) indices, respectively. A strong AO climate
pattern characterized the late 1980s and early 1990s
(Figure 5.1), with associated temperature anomaly
maxima in northern Europe and north-central Asia
in winter (Figure 5.3b), which expanded to northern
Alaska in spring. As part of the positive AO pattern,
44
Figure 5.3. 1000-hPa temperature anomalies
(°C) for (a) March–June 2005 (relative to
the 1968–1998 mean), (b) December–March
1989–1995, when the positive AO was
strong, and (c) December–March 1977–1988,
when the positive PNA was strong
[Source: NCEP/NCAR reanalysis; online at
www.ncdc.noaa.gov]
Figure 5.4. Summer (a), (b) heat (x 1010 J m-2 ) and (c), (d) freshwater (m) content; (a), (c) heat and freshwater content in the Arctic Ocean
based on 1980s averages (Arctic Climatology Project 1997); (b), (d) heat and freshwater content in the Beaufort Gyre (outlined in black in
(a) and (c)) in 2000–2005 based on hydrographic surveys (black dots depict locations of hydrographic stations); heat content is calculated
relative to the freezing point in the upper-1000-m ocean layer. Freshwater content is calculated relative to reference salinity of 34.8 psu.
west Greenland was anomalously cold. A positive
PNA pattern was dominant from 1977 to 1988, with
warm temperature anomalies over northern North
America (Figure 5.3c). The contrast of recent nearsurface temperature anomalies, with maxima in west
Greenland and northeast Siberia, to the temperature
patterns associated with the AO and PNA, suggests
that the current atmospheric circulation pattern is also
different from the main patterns of the second half of
the twentieth century (Overland and Wang 2005).
II)
ARCTIC OCEAN
(i)
Surface circulation regime
1970s and 1980s (Figure 5.4). From 2000 to 2005, the
western Arctic has been the focus of intensive investigation. This region includes the Beaufort Gyre, which is
the major reservoir of freshwater in the Arctic Ocean.
Although the total freshwater content in the Beaufort Gyre
has not changed dramatically, there is a significant
change in its distribution (Figures 5.4c,d). The centre
of freshwater maximum has shifted toward Canada
and intensified relative to the average. Significant
increases were also observed in the heat content of
the Beaufort Gyre (Figures 5.4a,b), primarily because
of an approximately twofold increase of the Atlantic
layer water temperature (Shimada et al. 2004).
The circulation of the sea ice cover and ocean surface
layer are closely coupled and are primarily wind driven.
Data from satellites and drifting buoys indicate that the
entire period of 2000–2005 has been characterized by
an anticyclonic circulation regime due to a higher sea
level atmospheric pressure over the Beaufort Gyre,
relative to the 1948–2005 mean. The dominance of the
anticyclonic regime is consistent with the AO index,
which has exhibited relatively low and fluctuating
values since 1996 (Figure 5.1).
(ii)
Heat and freshwater content
In recent years, the heat and freshwater content of
the Arctic Ocean have changed dramatically relative
to averages established by the Arctic Ocean Atlas of
the Arctic Climatology Project (1997), where water
temperature and salinity from observations were averaged and gridded for the decades of the 1950s, 1960s,
Figure 5.5. Annual mean relative sea level (cm) from nine
coastal tide gauge stations in the Siberian Seas (dotted line);
the blue line is the 5-year running mean sea level. The red
line is the 5-year running mean AO index (multiplied by 3 for
comparison).
45
The heat and freshwater content of the Arctic Ocean
depends on inputs along open boundaries. The most
important are fluxes in the Fram and Bering Straits. An
increase of the Atlantic water temperature in the Fram
Strait was observed in 2004 (Polyakov et al. 2005).
In the Bering Strait region, preliminary observations
from a mooring site, established and maintained since
1990, suggest that annual mean water temperatures
have been about a degree warmer since 2002, compared to 1990–2001. Since 2001, there has also been
an increase in the annual mean transport. Combined,
these changes have resulted in an increased northward
heat flux through the Bering Strait in recent years
(R. Woodgate, K. Aagaard, and T. Weingartner 2005,
personal communication).
(iii) Sea level
There is a positive sea level trend along Arctic coastlines (Figure 5.5). For 1954–1989, the rate of sea level
rise was estimated as +0.185 cm yr -1 (Proshutinsky
et al. 2004). The addition of 1990–2004 data increases
the estimated rate to 0.191 cm yr -1. Arctic sea level
also correlates relatively well with the AO index
(r = 0.83). Consistent with the influences of AO-driven
processes, Arctic sea level dropped significantly after
1990 and increased after the change from cyclonic to
anticyclonic circulation in 1997. In contrast, from 2000
to 2004 the rate of sea level rise has increased in spite
Figure 5.6. Sea ice extent in (left) March and (right) September 2005,
when the ice cover was at or near its maximum and minimum extent,
respectively; the magenta line indicates the median maximum
and minimum extent of the ice cover, for the period 1979–2000.
[Source: NOAA/National Snow and Ice Data Center (NSIDC)]
46
of a steady decrease in the AO index. At this point,
because of substantial interannual variability, it is difficult to evaluate the significance of this change.
III)
SEA ICE COVER
(i)
Extent and thickness
During 2005, every month except May showed a
record minimum sea ice extent in the NH relative
to the period 1979–2005 (J. Stroeve 2006, personal
communication). This time period is defined by the
availability of passive microwave images of sea ice
extent in the NH. The extent of the sea ice cover is
at or near its maximum in March and its minimum
in September (Figure 5.6). Ice extent in March 2005
was 14.8 million km2, decreasing to 5.6 million km2 in
September 2005, compared to the mean (1979–2005)
ice extent for March and September of 15.7 million km2
and 6.9 million km2, respectively. It is notable that in
March 2005, ice extent fell within the mean contour at
almost every location. In September 2005, the retreat
of the ice cover was particularly pronounced along the
Eurasian and North American coastlines.
To put the 2005 minimum and maximum ice extent in
perspective, the variability of ice extent in March and
September for the period of 1979–2005 is presented in
Figure 5.7. In both cases, a negative trend is apparent,
Figure 5.7. Time series of the variability of ice extent in March
(maximum) and September (minimum) for the period
1979–2005, normalized by the respective monthly mean ice
extent for the period 1979–2005; based on a least-squares
linear regression, the rate of decrease in March and
September was 2% decade-1 and 7% decade-1, respectively.
with a rate of 2% decade-1 for March and 7% decade-1
for September. Furthermore, the summers of 2002–2005
have experienced an unprecedented series of extreme
ice extent minima (Stroeve et al. 2005).
Ice thickness is intrinsically more difficult to monitor.
With satellite-based techniques only recently introduced, observations have been spatially and temporally
limited (Laxon et al. 2003; Kwok et al. 2004). Data from
submarine-based observations indicate that at the end
of the melt season the permanent ice cover thinned
by an average of 1.3 m between the period 1956–1978
and the 1990s, from 3.1 to 1.8 m (Rothrock et al. 1999).
On the other hand, measurements of the seasonal
ice cover do not indicate any statistically significant
change in thickness in recent decades (Melling et al.
2005; Haas 2004; Polyakov et al. 2003).
Trends in the extent and thickness of ice cover are
consistent with observations of a significant loss of
older, thicker ice out of the Arctic via the Fram Strait
(e.g., Rigor and Wallace 2004; Pfirman et al. 2004; Yu et
al. 2004) in the late 1980s and early 1990s. This event
coincides with the strong, positive AO period from
1989 to 1995 (Figure 5.1). When the AO is positive,
atmospheric and oceanic conditions favour a thinner
ice cover. A relatively younger, thinner ice cover, like
the one left behind from this event, is intrinsically more
susceptible to atmospheric or oceanic warming.
(ii)
Figure 5.8. APP-x (a), (b) surface albedo and (c), (d) skin
temperature for areas between 60°N and 90° and ice
concentrations of 15%–100% averaged over (a) June–
August, (b) April–September, (c) January–December, and (d)
March–May through 2004
Sea ice surface conditions
Data from 1982 to 2004, derived from AVHRR Polar
Pathfinder extended (APP-x) products (updated
from Wang and Key 2005a,b), indicate an overall
negative trend for boreal summer (June–August)
mean albedo of –0.4% yr -1 (Figure 5.8a). The trend
increases slightly to –0.5% yr -1 for the period from
April through September (Figure 5.8b), suggesting a
possible increase in the duration of the melt season.
In both cases, the surface albedo is relatively low
from 2001 to 2004 and is consistent with observations of an earlier, more spatially extensive onset of
melt and decrease in ice concentration (Belchansky
et al. 2004, Stroeve et al. 2005).
The time series of APP-x annual mean skin temperatures (Figure 5.8c) over the same period shows less
consistent change over time, with a general increase
in annual mean temperatures through the early 1990s
and a decrease from 1995 onward. When the time series
is limited to boreal spring (March–May) temperatures,
Figure 5.9. (a) Peak NDVI derived from 8-km-resolution
AVHRR data from 1981 to 2001 among the bioclimate
subzones and for the whole Arctic slope; (b) summer warmth
index (SWI; °C) over the past 22–50 years at meteorological
stations in each bioclimate subzone; dashed lines are linear
regressions. The shaded area in (b) highlights the period of
SWI covered by the NDVI data in (a). The arrows show years
of corresponding increases (red) and decreases (blue) in
NDVI and SWI (from Jia et al. 2003).
47
A more recent analysis covering the boreal forest and
tundra region of North America indicates that different patterns of greening have occurred in the boreal
forest and tundra areas (Goetz et al. 2005). The NDVI
has increased in tundra regions by an average of
about 10% for all of North America, whereas the NDVI
has declined in the boreal forest regions particularly
during the past 10 years.
(ii)
Water
Large regional variability, typical of Arctic conditions,
is observed in albedo, skin temperature, and ice concentration (Cavalieri et al. 1996). From 1996 to 2004,
the largest decreases in surface albedo correspond to
a reduction in ice extent in the Beaufort and Chukchi
Seas. Lower albedos over the central ice pack appear
consistent with the lower total ice concentrations over
this same period.
The R-ArticNet river discharge database (available
online at www.R-Arcticnet.sr.unh.edu) was extended
up to 2004 for 48 downstream river gauges. The last
five years were characterized by an increase of total
discharge to the Arctic Ocean mainly due to a contribution from Asian rivers. Mean 2000–2004 discharge from
Asia was 110 km3 (5%) higher than over the previous
20 years. Mean discharge to the ocean from North
America and Europe for 2000–2004 was practically
unchanged relative to 1980–1999. A consistent increase
in river discharge is observed from Eurasia for a longer
time interval as well. The maximum total discharge
of the six largest Eurasian rivers over 1936–2004 was
observed in 2002, at 2080 km3 yr -1 (Figure 5.10). Mean
discharge over 2000–2004 for the large Eurasian rivers
was 3%–9% higher than the discharge over 1936–2004.
Thus, the contemporary data further confirm the presence of a significant increasing trend in the freshwater
discharge to the Arctic Ocean from Eurasia documented
earlier by Peterson et al. 2002 (Figure 5.10).
IV)
LAND
(iii) Permafrost
(i)
Vegetation
Figure 5.10. Total annual discharge (km3 yr -1; red) to the
Arctic Ocean from the six largest rivers in the Eurasian
pan-Arctic for the observational period 1936–2004
(updated from Peterson et al. 2002); the least squares
linear increase (blue) was 2.3 (km3 yr -1) yr -1.
the 23-year linear trend is positive (0.14°C yr -1), with
greater interannual variability (Figure 5.8d), indicative
of the seasonal dependence of warming trends.
The most convincing evidence of widespread change
in Arctic vegetation comes from the historic trends
of tundra greenness as detected from satellites. The
NDVI is a measure of vegetation greenness derived
from surface reflectance in the red and near-infrared
wavelengths. Higher-latitude NDVI values might be
expected to increase under a warmed climate. Earlier
global studies of NDVI changes indicated a general
pattern of increased NDVI in the region between
40° and 70°N during the period of 1981–1999 (Myneni
et al. 1997, 1998; Zhou et al. 2001; Lucht et al. 2002).
Studies of the NDVI indicate an increase of 17% in NDVI
values in the tundra area of northern Alaska, where
summer warmth index (SWI) values at meteorological
stations increased by 0.16° to 0.34°C yr -1 (Figure 5.9)
during the same period (Jia et al. 2003).
48
Observations show a general increase in permafrost
temperatures during the last several decades in Alaska
(Osterkamp and Romanovsky 1999; Romanovsky et al.
2002; Osterkamp 2003), northwest Canada (Couture
et al. 2003; Smith et al. 2003), Siberia (Pavlov 1994;
Oberman and Mazhitova 2001; Romanovsky et al.
2002; Pavlov and Moskalenko 2002), and Northern
Europe (Isaksen et al. 2000; Harris and Haeberli
2003). Uninterrupted permafrost temperature records
over more than 20 years have been obtained by the
University of Alaska–Fairbanks along the International
Geosphere–Biosphere Programme (IGBP) Alaskan
transect, which spans the entire continuous permafrost zone in the Alaskan Arctic. All observatories
show a substantial warming during the last 20 years.
This warming varied by location, but ranged typically from 0.5° to 2°C at the depth of zero seasonal
permafrost temperature variations (Figure 5.11).
Figure 5.11. (left) Location of the long-term University of Alaska permafrost observatories in northern Alaska 1978–2005; (right) changes in
permafrost temperatures (°C) at 20-m depth during the last 20–25 years (updated from Osterkamp 2003)
These data also indicate that the increase in permafrost temperatures is not monotonic. During the
observational period, relative cooling has occurred
in the mid-1980s, in the early 1990s, and then again
in the early 2000s. As a result, permafrost temperatures at 20-m depth experienced stabilization and
even a slight cooling during these periods. An even
more significant cooling of permafrost was observed
during the very late 1990s and the early 2000s in
interior Alaska. A significant portion of this cooling is related to a shallower-than-normal winter
snow cover during this period. During the last three
years, there was a sign of recovery in mean annual
temperatures at shallow depths. Soil temperatures
in interior Alaska during 2005 reached the temperatures of the early to mid-1990s, which were
the warmest from during the last 70 years.
Data on changes in active layer thickness (ALT) in
the Arctic lowlands are less conclusive. In the North
American Arctic, ALT experiences substantial interannual variability, with no discernible trends; this is
probably due to the short length of historical data
records (Brown et al. 2000). A noticeable increase
in the active layer thickness was reported for the
Mackenzie Valley (Nixon et al. 2003). However, this
positive trend became negative at most of these sites
after 1998 (Tarnocai et al. 2004). An increase of more
than 20 cm in thickness between the mid-1950s and
1990, derived from the historical data collected at
Russian meteorological stations, was reported for the
continuous permafrost regions of the Russian Arctic
(Frauenfeld et al. 2004; Zhang et al. 2005). At the same
time, reports from several specialized permafrost
research sites in central Yakutia show no significant
changes in the ALT (Varlamov et al. 2001; Varlamov
2003). The active layer was especially deep in 2005
in interior Alaska. Around Fairbanks, the 2005 active
layer depth was the deepest observed in the past
10 years. Data from many of these sites show that
the active layer that developed during the summer
of 2004 (one of the warmest summers in Fairbanks
on record) did not completely freeze during the
2004/2005 winter. A thin layer just above the permafrost table remained unfrozen during the entire
winter.
V)
GREENLAND J. E. BOX10
(i)
Overview
Temperatures in 2005 around Greenland were anomalously warm, particularly during the boreal spring.
In the period of record that includes three stations that
began observations in the late 1800s, mean 2005 temperatures are rivaled only by recent years (e.g., 2003) and by
warm conditions during the 1930s/1940s. Unprecedented
spring coastal fog caused flight delays along western
Greenland that made European headlines. Widespread
positive sea surface temperature anomalies surrounded
Greenland for all seasons of 2005, according to the NOAA
SST OI version 2 (OIV2) data set (1982–2005; available
online at www.cdc.noaa.gov/cdc/data.noaa.oisst.v2.html).
Upper-air temperature soundings available from the
NCDC Integrated Global Radiosonde Archive (IGRA;
Durre et al. 2006) indicate a pattern of tropospheric
warm and stratospheric cool anomalies over Greenland,
particularly in the boreal winter.
49
Table 5.1. Greenland station temperature statistics: 2005 versus 1951–2005
Station
Location
Statistic
Winter
Spring
Summer
Autumn
Annual
Egedesminde
68.7 N
Rank
13
1
2
13
3
Central-west
52.8 W
Z score
-1.5
2.2
0.8
-0.1
1.9
Nuuk
64.2 N
Rank
N/A
1
10
19
N/A
Southwest
51.8 W
Z score
N/A
2.3
-1.4
0
N/A
Prins Christian Sund
60.0 N
Rank
7
1
1
17
1
South
43.2 W
Z score
-0.7
3.3
2.6
-0.3
2.4
Tasiilaq
65.6 N
Rank
7
2
2
34
3
Southeast
37.6 W
Z score
0.6
2.1
0.9
-1.1
3.2
Danmarkshavn
76.8 N
Rank
1
4
6
26
1
Northeast
18.7 W
Z score
2.8
-0.1
-0.8
0
2.6
Region
(ii)
Coastal temperature records
Although the coastal location of most meteorological
stations makes it difficult to construct a representative
climatology of temperature for Greenland, the existing
55-year (1951–2005) record allows the annual mean
temperatures of 2005 to be placed into historical context
being as among the warmest, if not the warmest, on
record (Table 5.1; see also Figure 6.30). In northeast
Greenland (Danmarkshavn), winter temperatures
were among the warmest on record (99th percentile).
The boreal spring was also unusually warm. Annual
mean temperatures at all stations fell within the
warmest decile (10%) of a normal distribution.
sublimation was slightly below normal due to decreased
snow entrainment by the wind as a result of increasing
melt duration (i.e., 10 days on average for the area below
equilibrium line altitude). Meltwater production was 40%
above normal owing to elevated summer temperatures
and low-albedo positive feedback, which was particularly
strong in August. Runoff rates were 45% above normal.
Because of an apparent expansion of the ablation zone
in this simulation, the accumulation area ratio was
7% below that of the most recent 17 years.
Previous observations have demonstrated accelerated
ice sheet dynamic flow during surface melt water production (Zwally et al. 2002). Remote sensing measurements
(iii) Greenland Ice Cap
Due to the paucity of regular observations of the
Greenland Ice Cap, analysis of its behavior must
necessarily be estimated by numerical modeling.
Polar fifth-generation Pennsylvania State University
(PSU) –National Center for Atmospheric Research
(NCAR) Mesoscale Model (MM5) calculations
(Box et al. 2006) indicate that increases in ice sheet melting exceeded increased snow accumulation in 2005, yielding a large (76%) negative surface mass balance anomaly
(Table 5.2). In addition to precipitation increases, more
of the precipitation was solid, despite a 1°C overall warm
anomaly, suggesting that the upper elevations of the ice
sheet received more snowfall than usual. Evaporation
rates were 30% greater than normal, driven by warmerthan-normal average air temperatures. Blowing snow
50
Figure 5.12. Monthly values of the Southern Hemisphere
annual mode index for 2005 (June value is missing)
Table 5.2. Greenland ice sheet surface mass balance parameters: 2005 departures from 1988 through 2004 average
% of 1988–2004 average
2005 minus 1988–2004 average [km3 y-1]*
Total precipitation
110%
73
Liquid precipitation
107%
2
Evaporation
130%
21
Blowing snow sublimation
98%
-1
Snow accumulation
108%
52
Meltwater production
140%
238
Meltwater runoff
145%
206
Surface mass balance
24%
-153
Mean temperature
—
1.0°C
Accumulation area ratio
93%
-0.06
* Unless otherwise indicated.
indicated widespread increases in glacier velocity south
of 70°N (Rignot and Kanagaratnam 2006), confirming that
flow rates did in fact increase in 2005. This suggests a
Greenland Ice Sheet contribution to sea level that explains
at least one-third of the observed recent accelerating
global sea level rise (Leuliette et al. 2004).
C.
ANTARCTIC
I)
ATMOSPHERIC CIRCULATION A. ARGUEZ3
rest of the continent have been mixed, with interior
stations such as Amundsen–Scott South Pole Station
and Vostok showing small, significant trends in surface temperature. Not surprisingly, many stations in
Antarctica are located near the more accessible coast
and, therefore, deriving a representative continental
temperature is virtually impossible. From the available temperature data, it is, perhaps, Antarctica’s
apparent continentwide temperature stability that is
remarkable in the context of global warming. Only
the Antarctic Peninsula (4% of the continental area)
shows a significant warming trend. The peninsula is
The Southern Hemisphere annular mode (SAM; also
known as the Antarctic Oscillation) is the Arctic Oscillation’s
(or Northern Hemisphere annular mode’s) counterpart
in the Southern Hemisphere. It is defined as the leading
mode of 700-hPa height anomalies south of 20°S. Since
early 2003, SAM positive and negative phase events have
been uncharacteristically weak and ephemeral. During
2005, SAM values were fairly small except for a value
of +1.2 in February and a considerably negative reading
(–2.0) in December. However, weak values prevailed
during the austral winter (Figure 5.12).
II)
TEMPERATURE A. M. WAPLE89
Surface air temperatures across the majority of stations on the Antarctic continent were near to above
average in 2005, with the largest warming trends over
the last several decades measured along the Antarctic
Peninsula, for example, at the Rothera meteorological station (Figure 5.13). Observed trends across the
Figure 5.13. Annual temperature anomalies from
Rothera Meteorological Station on the Antarctic
Peninsula; base period is 1947–2005 and years without
bars indicate unavailable data.
51
such as the Antarctic Peninsula and Bellingshausen
Sea (e.g., Vaughan and Doake 1996). The collapse of
the Larsen A and Larsen B ice shelves in 1995 and
2002, respectively, indicates that pronounced regional
warming can lead to ice shelf collapse and to glacial
acceleration (Waple and Rignot 2005).
Figure 5.14. Annual anomalies of Southern Hemisphere
sea ice; base period is 1979–2005. [Source: NOAA/
NSIDC]
also the warmest area of the continent on average,
rising above freezing for 2 months a year on its warmest coast (see information online at www.antarctica.
ac.uk/About_Antarctica/Above_Antarctica/Weather/
Temperature/index.php). Due to the paucity of stations
on Antarctica, the lack of long-term records and the high
interannual temperature variability, it is impossible to
state with certainty whether the Antarctic continent is
warming or cooling overall. However, despite the lack
of any clear warming signal in most coastal stations,
there are indications that ice sheet thinning is occurring in west Antarctica and is considered a likely result
of reduced buttressing by coastal ice shelves, which
are melting or disintegrating, perhaps in response to
coastal warming (e.g., Levinson 2005).
III)
SEA ICE A. M. WAPLE89
Southern Hemisphere sea ice extent has been increasing
since the late 1970s (Figure 5.14), with above-average
extent in 2005. However, seasonal and spatial variability
has been extreme in the Southern Hemisphere, with ice
shelf and sea ice retreat occurring in those areas that
have shown a warming trend over the last 50 years,
IV)
STRATOSPHERIC OZONE R. C. SCHNELL81
The year 2005 was the 20th consecutive year of NOAA
ozonesonde measurements at the Admunsen–Scott South
Pole Station, and the NOAA/ESRL/Global Monitoring
Division launched 68 balloon-borne ozonesondes
in 2005. Figure 5.15a shows the total column ozone
(blue line) in Dobson units and 20–24-km stratospheric
temperature (red line), illustrating the development of the
ozone hole over South Pole Station in 2005. The severity
of ozone depletion depends on wintertime stratospheric
temperatures, the stability of the polar vortex and active
chlorine levels. July and August profiles showed typical
cold temperatures throughout the lower stratosphere,
reaching a minimum of -92.3°C (180.9 K) on 10 July. The
cold temperatures provided favourable conditions for
the formation of polar stratospheric clouds that enable
the transformation of Cl and Br compounds into species
that destroy O3 when sunlight returns to the Antarctic
stratosphere. Total column O3 remained stable until
mid-August (Figure 5.15b). Stratospheric O3 declined to
a minimum of 110 DU on 29 September, and was nearly
completely destroyed in the 15–21-km layer (Figure 5.15c).
Compared to the years since 1986, 2005 was the 10th
lowest for O3 levels on record, with the record being 89 DU
on 6 October 1993. Ozone abruptly increased to 333 DU
on 13 November 2005 (Figure 5.15d), due to midlatitude
ozone-enriched air transported over the continent after
the polar vortex began to break up.
Figure 5.15. (a) Summary of South Pole total ozone in Dobson units and stratospheric temperatures measured by ozonesondes during 2005;
three selected profiles of altitude vs ozone partial pressure (mPa) are shown (b) prior to the 2005 ozone hole, (c) at minimum total ozone,
and (d) post-ozone hole. [Source: B. Johnson and S. Oltmans, NOAA/ESRL/GMD]
52
6. REGIONAL CLIMATES
A.
K. A. SHEIN,82 ED.
OVERVIEW K. A. SHEIN82
While the anomalous global warmth of 2005 is
generally reflected in regional temperatures, various
regions of the planet respond differently to climate
forcings at many scales, both spatial and temporal.
An analysis of globally averaged climate may mask
a number of important climatic conditions that have
impacted some areas more than others. This section
chronicles regional climatic conditions relative to their
historical context and highlights notable atmospheric
events of 2005. In fact, most regions experienced
some form of record-breaking weather or climate
conditions in 2005.
This section is distributed by continent or major
land region, and each regional subsection is further
divided into logical climatic divisions, either geographic or political. The use of national names in
no way implies political preference or precedence.
Also, it should be noted that while the large-scale
temperature and precipitation anomaly maps (i.e.,
Figures 6.1, 6.7, 6.16, 6.17, 6.24, 6.29, and 6.39) all
use a 1971–2000 base period for temperature and
a 1979–2000 base period for precipitation, discussions of anomalies in individual regions may refer
to alternate base periods.
B.
AFRICA
I)
EASTERN AFRICA C. OLUDHE,61 P. AMBENJE,2 AND
L. OGALLO60
The rainy seasons in the Greater Horn of Africa (GHA)
are influenced by the intra-annual north–south migration of the ITCZ. In the GHA region, rainfall exhibits
strong variability both in space and time. Much of the
variability is strongly accounted for by the existence
of complex topographic features, including the East
African lakes, and is also partly influenced by the
movement of the ITCZ. The subregion can, however,
be divided into three sectors (Southern, Equatorial, and
Northern) based on rainfall onsets and withdrawals.
The Southern sector (central and southern Tanzania)
experiences a unimodal precipitation regime, with
rain occurring between December and April. The
Equatorial sector (northern Tanzania, Kenya, southern
and extreme eastern Ethiopia, southern Sudan, and the
southern half of Somalia) generally exhibits a bimodal
rainfall regime, with the “Long Rains” season from
March to May and the “Short Rains” extending from
Figure 6.1. African 2005 annual (top) temperature
anomalies (°C; 1971–2000 base), and (bottom)
precipitation anomalies (mm; 1979–2000 base)
from the CAMS–OPI data set (Janowiak and
Xie 1999) [Source: NOAA/NCDC]
53
October to December. However, both the western and
coastal areas also receive substantial rainfall during
July and August. In the Northern sector (central and
northern Ethiopia, Eritrea, Djibouti, and the northern
half of Sudan), the major rainy season is between June
and September, but a few areas receive a secondary
peak from March to May.
The climate over the GHA is largely regulated by
sea surface temperatures in the Indian and Atlantic
Oceans, general atmospheric circulation and largescale anomalies (e.g., ENSO, Indian Ocean dipole),
Indian Ocean tropical cyclone activity, and the
variability of the monsoon.
(i)
Climate patterns in the GHA in 2005
Parts of the GHA were under persistent drought
throughout the year with most stations recording
rainfall much below their long-term mean (Figure 6.1).
In some of the arid and semiarid lands (ASALs), no
significant rainfall was recorded for the year. Erratic
rainfall and poor temporal distribution was common
during the rainy seasons, even in areas that recorded
normal to above-normal precipitation.
Most socioeconomic and subsistence activities in
the GHA depend directly or indirectly on rainfall.
Below-average rainfall during the year had farreaching socioeconomic impacts, including the loss
of life, livestock and property.
The Southern sector experienced abundant rainfall
between December 2004 and February 2005, providing
relief, especially in areas such as central and southern
Tanzania that had experienced extremely dry conditions during the 2003/2004 rainfall season.
heavy rainfall events, some exceeding 50 mm in
24 hours, significantly contributed to seasonal rainfall totals in some areas. Unfortunately, these events
generated flash flooding in some parts of the GHA,
displacing thousands of people.
Western and coastal areas of the Equatorial sector
recorded signifi cant rainfall from June to August,
although totals were slightly below the seasonal average in some areas (Figure 6.2).
From October to December (Short Rains), most parts
of the Equatorial sector received between 25% and 75%
of long-term seasonal mean precipitation (Figure 6.2).
Like the Long Rains, the Short Rains were also characterized by poor temporal distribution, and were
devoid of the heavy rainfall events common during
tropical rainy seasons. The rains ceased in the second
half of November instead of the usual mid-December.
Performance was extremely poor in the ASALs, enhancing the cumulative rainfall deficiencies these areas had
been experiencing for several consecutive seasons.
As a result, an estimated five million people in Kenya
and Tanzania were affected by famine.
The Northern sector had one major rainfall peak
concentrated in June–September, with few areas
receiving the usual secondary rainfall peak between
March and May. Near- to above-normal rainfall
was observed in parts of western and northern
areas of the subregion during June–August 2005
(Figure 6.2). Most locations recorded June–September
rainfall totals of 75% –125% of the long-term average. Much of the eastern and central parts of the
From March to May 2005 (Long Rains), most locations
over the Equatorial sector received near-normal to
below-normal rainfall amounts. There was a general
late onset and early withdrawal as well as poor temporal and spatial distribution of the seasonal rainfall
in most areas of the sector, especially the ASALs.
The month-by-month evolution of the Drought Severity
Index for the Long Rains season indicates that although
near-normal to wet conditions were observed at some
locations during March and May, long dry spells were
predominant, especially during the peak rainfall month
of April, which was relatively dry at many locations
within the Equatorial sector. Occasional short-lived
54
Fig 6.2. East African rainfall anomaly percentages for (left)
June to August and (right) October to December 2005
[Source: Kenya Meteorological Service]
Northern sector experienced dry conditions, with
most locations recording rainfall below 75% of the
long-term average.
II)
NORTHERN AFRICA M. A. BELL6 AND K. KABIDI36
(i)
Temperature
For 2005, mean temperature anomalies were generally
between 0.25° and 1.5°C above normal throughout
most of North Africa (Figure 6.1). The year started with
below-normal monthly mean temperatures in January,
and particularly in February. Chefchaouen, Morocco,
recorded 18 below-freezing days in January. Subfreezing
temperatures were recorded in many other areas and
broke numerous records. A low of -14°C was recorded
in a mountainous area of Morocco. By April, positive
temperature anomalies had begun to dominate, and
a heatwave in mid-July resulted in at least 13 deaths
in Algeria due to sunstroke, according to the British
Broadcasting Corporation (BBC). In Algeria, the heatwave pushed July temperatures as high as 50°C.
(ii)
Precipitation
The Mediterranean coast of North Africa receives the
majority of its rainfall during October–April, largely
from midlatitude cyclones and associated cold fronts.
In the Atlas Mountains of northern Morocco, Algeria,
and Tunisia, cold-frontal passages can bring subfreezing temperatures and heavy rain or snow, occasionally
causing floods and landslides.
Accumulated precipitation anomalies for the October
2004–April 2005 rainy season indicate below-normal
precipitation totals in most of Morocco (particularly
in the north), northwestern Algeria, the southern half
of Tunisia and much of northern Libya, as well as
above-normal precipitation in northeastern Algeria
and northern Tunisia (Figure 6.3).
The precipitation deficits in Morocco developed in
November 2004, in concert with the genesis of the
severe drought that would plague the Iberian Peninsula
and other sections of southwestern Europe for much
of the year. Although rain in mid- to late-February
provided some relief, dry conditions persisted
throughout most of the remainder of the 2004/2005
boreal winter rainy season in Morocco. October–
December 2005 began the winter rainy season
with near-normal precipitation throughout North
Africa.
As discussed in the CPC/Famine Early Warning System
Network (FEWS NET) Africa Weekly Weather Hazards
Assessments in 2005, although the cooler-than-normal
temperatures early in the year may have temporarily
reduced moisture demand and partially mitigated
precipitation deficits in the major wheat-growing
areas of North Africa, the persistence of below-normal
rainfall and the development of positive temperature
anomalies had a detrimental effect on the region’s
winter wheat crop, with grain production (including
wheat) in 2005 well below the previous year’s record
levels in Morocco, Algeria, and Tunisia, and also
below the average of the past five years in Morocco
and Algeria [Source: US Department of Agriculture
(USDA) and United Nations (UN) Food and Agriculture
Organization (FAO)].
(iii) Notable events
A winter storm in late January 2005 produced the
heaviest snowfall seen in Algiers in “more than
50 years,” according to the BBC, and was responsible for
at least 10 deaths, primarily due to traffic accidents.
During February, a synoptic low pressure system
brought heavy rainfall to the region, with a record of
193 mm falling in less than 24 hours at Tetuan, Morocco.
This system also brought high wind speeds exceeding
31 m s-1 in some places, and waves up to 10 m were recorded
along the northwestern Atlantic coast (Figure 6.4).
The most unusual climatic event recorded during
2005 was the landfall of Tropical Storm Delta. On
29 November, Delta passed to the north of the Canary
Figure 6.3. October 2004–April 2005 precipitation anomalies
(mm; 1979–2000 monthly means base) for northern Africa
from CAMS–OPI
55
Figure 6.4. Mean sea level pressure (hPa)
over North Africa on 28 February
[Source: NOAA/Cooperative Institute
for Research in Environmental Science
(CIRES)/Climate Diagnostics Center (CDC)]
Islands, where widespread damage and seven
fatalities were reported. Soon after, Delta crossed over
the southern coast of Morocco in the area of Tantan
and Layoune Ports, where it quickly dissipated, but
not before delivering much-needed rain to the area.
III)
SOUTHERN AFRICA W. M. THIAW,85 T. GILL,27 AND
W. A. LANDMAN43
(i)
Temperature
Annual mean temperatures across southern Africa for
2005 were generally 0.5°–1.5°C above the 1971–2000
mean (Figure 6.1). Most of Madagascar was 0.5°–1.0°C
above normal. Temperatures were up to 2°C above
normal in an area including northern Namibia, southern
Angola, northwestern Botswana, and western Zambia.
In South Africa, June–August mean temperatures were
2°–3°C above normal over the northeastern parts of
the country, 1°–2°C above normal over central regions,
and near normal in southern and western sections.
(ii)
Precipitation
The rainy season in southern Africa extends from October
to April, with the greatest amounts typically observed
between December and March. In general, ENSO conditions play an important role in the variability of southern
Africa rainfall, which tends to be drier than average during
El Niño and wetter than average during La Niña.
Overall, the 2004/2005 southern Africa rainy season
was characterized by near-average rainfall (Figure 6.1),
although delayed onset of the rains in October 2004
and inconsistent rainfall led to deficits in January and
February. Rainfall anomalies were quite variable in
the wet zone (east of 25°E; Figure 6.5 boxed region).
56
In this area, 200–400 mm of rain fell during the period
of November 2004–April 2005, ranking in the 10th–30th
percentile across northeastern Botswana, the eastern
half of Zimbabwe and northeastern South Africa. Rainfall
was also below normal in pockets along the east coast
of Madagascar. In contrast, central Mozambique and
southern Madagascar received from 700 to over 900 mm
of rainfall, which ranked in the 70th–90th percentile.
Average conditions prevailed in most of interior South
Africa, eastern Zimbabwe and southern Mozambique.
Climatologically dry areas of the region registered
near- to above-normal rainfall, with amounts in the
70th–90th percentile across southern Namibia.
The low-level atmospheric circulation for the 2004/2005
rainy season featured near normal easterly winds
(~4 m s-1) along the equatorward flank of the Mascarene
high. A significant reduction in low-level easterlies
associated with the presence of an anomalous
anticyclonic flow in the southwestern Indian Ocean
contributed to rainfall deficits in November 2004
and February 2005. In addition, an elongated ridge
extending from high latitudes into the continent
contributed to strong subsidence in southern Africa.
IV)
WESTERN AFRICA W. M. THIAW85 AND M. A. BELL6
(i)
West African Monsoon
West African rainfall can be divided into two quasihomogeneous regions: the Sahel and the Gulf of Guinea.
Rainfall in both areas is controlled by the annual progression of the ITCZ over the region. The African Sahel,
defined here as the region between 12°–20°N, 18°W–20°E
(Figure 6.6, boxed region), receives approximately 90%
of its mean annual rainfall during June–September. The
rainfall is monsoonal and its penetration into the region
Figure 6.5. November 2004–April 2005
(a) total rain and (b) precipitation anomaly
(mm; 1971–2000 base) for southern Africa;
boxed region is considered the “wet zone.”
is closely related to the position of the mid- and upperlevel jets, and to the ITCZ, which starts its northward
progression in March and reaches its northernmost
position in August. Seasonal precipitation exhibits a
strong meridional gradient, with average totals exceeding 600 mm in the south and 100–300 mm in the north.
These larger-scale circulation features are fairly sensitive
to changes in the global monsoon circulation on both
interannual and interdecadal time scales.
Further south, along the central Gulf of Guinea coast,
the rainy season is bimodal and runs from about
April to October, typically with a “little dry season” in
July–August. This configuration produces an extremely
marked north-to-south gradient in annual precipitation
totals across the region.
(ii)
Precipitation
The 2005 rainy season featured above-normal rainfall
across most of the Sahel (Figure 6.6). Rainfall totals
exceeded 100 mm above average across most of the
central and western areas of the Sahel. Overall, the 2005
rainy season was the second wettest since 1994. In particular, most of Senegal and southern areas of Mauritania,
northern Burkina Faso, eastern Mali, southern and western
Niger and southeastern Nigeria received above-normal
rainfall throughout most of the rainy season, with the
exception of July precipitation in southeastern Nigeria.
Rainfall anomalies were extremely strong over western
Senegal, which received around 700 mm of rainfall
between July and September (Figure 6.6). That is about
300 mm above the long-term mean, making 2005 the
rainiest season in this area since 1970. Seasonal totals
along most of the Gulf of Guinea coast, from Côte d’Ivoire
to western Nigeria, were below normal, particularly in
central Benin and Côte d’Ivoire. Rainfall deficits ranged
between 50 and 200 mm below the climatological mean
in the central areas of Ghana, Togo and Benin.
Heavy rainfall in June, near the start of the rainy season in Guinea and Guinea-Bissau, reportedly sparked
Figure 6.6. June–September 2005 (a) total rainfall and (b) anomalies (mm; 1971–2000 base) for western Africa; boxed region is the Sahel.
57
Figure 6.7. North American 2005 annual (left) temperature anomalies (°C; 1971–2000 base), and (right) precipitation anomalies
(mm; 1979–2000 base) from CAMS–OPI
a severe cholera epidemic that would eventually
spread to at least nine countries in West Africa over
the course of the last half of the year, according to
the World Health Organization (WHO). Initially, cases
were largely confined to the capital city of Bissau, but
it spread quickly. Heavy rainfall in Dakar, Senegal,
from mid-August through early September not only
flooded areas of the city’s outer suburbs and forced
the evacuation of approximately 60,000 people but
also triggered a sharp increase in the number of local
cholera cases. According to WHO statistics available
in late September 2005, at least 43,638 cases of the
disease and 759 deaths had been reported throughout West Africa. The end of the rainy season and a
lack of new reported cases by the end of December
allowed the Ministry of Health in Guinea-Bissau
to declare an end to the epidemic in that country,
according to the UN Integrated Regional Information
Networks (IRIN).
C.
NORTH AMERICA
I)
CANADA C. KOCOT,39 D. PHILLIPS,67 AND R. WHITEWOOD92
The climate of Canada in 2005 was characterized by
warmer and wetter conditions than normal (relative
to the 1951–1980 base period). Although Canada was
spared much of the extreme weather that impacted
other regions of Earth, it was not totally immune.
Anomalous winter warmth adversely impacted snowpack in British Columbia (BC). Several tornadoes
and heavy flooding in three provinces contributed to
58
2005 being the costliest year to date, weatherwise,
for insurers.
(i)
Temperature
Above-normal temperatures were observed throughout the country, with most areas at least 1°C above
normal (Figure 6.7). Departures of 3°C above normal
were experienced in the southwest corner of the Yukon
Territory. The 1.7°C above-normal (1951–1980 mean)
average national temperature experienced by Canada
in 2005 marked it as the ninth consecutive year of
above-normal temperatures (Figure 6.8). Overall, 2005
tied 2001 and 1999 as the third warmest year since
reliable nationwide records began in 1948. The year’s
national average is exceeded only by 1998 (+2.5°C)
and 1981 (+2.0°C).
Ten of the 11 Canadian climate regions had temperatures that ranked among the 10 warmest years
in their records. However, of those, only the north
BC mountains/Yukon region tied 1981 for its warmest year (+2.8°C). The remaining nine regions were
the following: Arctic mountains and fjords (second
warmest, +2.0°C); Pacific coast (fifth warmest, +1.2°C);
northwestern forest (sixth warmest, +2.0°C); Arctic
tundra (sixth warmest, +1.7°C); northeastern forest
(sixth warmest, +1.4°C); Mackenzie district (seventh
warmest, +2.1°C); Atlantic Canada (seventh warmest,
+0.9°C); south BC mountains (eighth warmest, +1.1°C);
and Great Lakes/St. Lawrence (ninth warmest, +1.1°C).
The lowest-ranked region, the prairies, experienced its
11th warmest year (+1.2°C). Over the 58-year period
of record (1948–2005), all 11 regions show a positive
annual temperature trend, with the greatest increase
(+2.2°C) in the north BC mountains/Yukon and the
smallest (+0.1°C) in Atlantic Canada.
(ii)
Precipitation
In 2005, Canada experienced its wettest year in the
58 years since reliable nationwide records commenced
(Figure 6.7). The 13.4% above the 1951–1980 mean
displaced the previous record of +9.1% (1996). Areas
with precipitation values over 20% above normal
in 2005 were most of Yukon, some of the southern
Northwest Territories, most of Nunavut, the southwest
coast of BC, southern Alberta, most of Saskatchewan
and Manitoba, the extreme north of Quebec and the
western part of Nova Scotia. Areas with precipitation
amounts at least 20% below normal were along the
west coast of BC, the eastern edge of BC, and the
western edge of Alberta. The remainder of the country
was close to normal.
Regionally, six climate regions experienced conditions in 2005 that would rank them among the
10 wettest years: Arctic tundra (second wettest,
+23.1%); northwestern forest (second wettest, +12.8%);
Arctic mountains and fjords (fifth wettest, +32.0%);
Mackenzie district (fifth wettest, +19.9%); north BC
mountains/ Yukon (fifth wettest; +19.1%); prairies
(seventh wettest, +18.2%). While no region ranked
2005 as the record wettest, a sufficient portion of
the country was enough above normal to collectively
produce Canada’s wettest year on record. The three
driest regions of the country were south BC mountains (-3.8%), Pacific coast (-5.5%), and Great Lakes/
St. Lawrence (-3.5%). These three regions recorded
only slightly drier-than-normal conditions.
(iii) Notable events
Although 2005 started out dry across the province,
a series of June storms drenched portions of southern Alberta, resulting in widespread flooding as
rain-swollen rivers escaped their banks. Extensive
damage was reported to dwellings and infrastructure in over 40 municipalities. At 247.6 mm, Calgary
recorded its wettest June on record (79.8 mm is normal). Outside the city, monthly rainfall approached
400 mm. Total losses are estimated at C$ 400 million
( US $ 3 4 4 million ) of which C $ 275 million
(US$ 232 million) was insured, making this weather
event one of the costliest in Alberta’s history.
Figure 6.8. Annual average air temperature anomalies
(°C; blue) and 1948–2004 trend (red) for Canada
[Source: Environment Canada]
Manitoba was treated to rare and record widespread
flooding as the result of frequent and intense June and
July thunderstorms. The Churchill River crested at its
highest levels ever recorded, and Manitoba Agriculture
estimated more than a quarter of the province’s farmland was inundated during the flooding.
A line of severe thunderstorms tracked across southern Ontario on 19 August, leaving record damage
estimated at over C$ 500 million (US$ 430 million).
The storms spawned two F2 tornadoes and a rare
tornado warning was issued for Toronto. In and
around Toronto, ~45 mm diameter hail, straight-line
winds with gusts reaching 72 km h -1, and rainfall
rates exceeding 100 mm h -1 generated the most
damage.
The 2005 sea ice extent in Canadian Arctic waters
dropped to its lowest level on record, 5.3 million km2;
down 20% from 1978 when satellite observations
began. The previous record minimum, of slightly
less than 6 million km2, was set in 2002. Since the
1970s, the geographical extent has been decreasing
by around 8% decade -1.
The Canadian International Forest Fire Centre (CIFFC)
reported a near average fire year in Canada in terms of
the number of fires (7438; -1.3% of normal), but with
significantly fewer hectares (ha) of forest consumed,
1.7 million ha, or ~68% of the 1995–2004 mean. The total
number of fires in 2005 was down in most provinces
and territories, with notable exceptions in Ontario
(1961 fires; +51% of normal) and Quebec (1,374 fires,
59
Figure 6.9. Statewide rankings of temperature as measured across the contiguous United States in 2005: (left) March–May
and (right) September–November. A rank of 111 (1) in the U.S. Historical Climatology Network (USHCN) record represents
the warmest (coldest) season since 1895.
+57% of normal). In contrast, and with respect to the
total area affected, all provinces and territories, with
the exception of Quebec, reported lower than average
area burned. The total area burned within the province
of Quebec in 2005 (831,022 ha), however, accounted
for ~49% of the national total.
II)
UNITED STATES OF AMERICA K. L. GLEASON28
(i)
Overview
Reliable weather records for the United States exist
from 1895 to the present, enabling the climate of
2005 to be placed in a 111-year context for the contiguous United States. The nationally averaged temperature in 2005 was the seventh (105 of 111 years)
warmest on record, with an annual mean of 12.3°C
(+0.8°C relative to the period of record). The linear
temperature trend for the 111-year record over the
contiguous United States is 0.056°C decade -1, with
an increase to 0.32°C decade -1 since 1976. Seven
of the ten warmest years on record for the United
States have occurred since 1986.
Precipitation in the United States in 2005 was variable throughout much of the country, with periods
of excessive rainfall in the Southwest and Northeast,
persistent drought in portions of the Northwest and
developing drought from the Southern Plains to
the Great Lakes. Nationally, it was the 43rd wettest
year on record, which is near the long-term mean.
Maine and New Hampshire had their wettest year
on record, surpassing 1909 and 1954, respectively.
Conversely, Arkansas had its second driest year
since 1895.
60
Temperature and precipitation anomalies for the
United States are based upon the 1895–2005 data
record, rather than on any particular 30-year normal
statistics (e.g., 1971–2000 mean). With respect to the
United States, temperature or precipitation is described
as “much above” or “much below” normal when the
value falls within the top or bottom 10% (decile) of the
historical record distribution. Temperatures are simply
“above” (“below”) normal if they fall within the upper
(lower) third, or tercile, of the distribution, but are not
in the top (bottom) decile. Values falling within the
middle tercile are considered near normal.
(ii)
Temperature
Temperatures were above average across most of
the contiguous United States from December 2004 to
February 2005, with no state ranking below average.
Colorado, Wyoming, and Utah were much above normal
for the season. From March to May (spring) was exceptionally cool from Texas to Florida and along the entire
eastern seaboard. Eleven states had below-average
seasonal temperatures. Three additional states reported
much-below-average temperatures for the season
(Figure 6.9, left). Extremely cool May temperatures
covered the Northeast and Mid-Atlantic coast, with
seven states (Connecticut, Rhode Island, Massachusetts,
Pennsylvania, Delaware, Maryland, and South Carolina)
experiencing one of their 10 coldest Mays on record.
In contrast, record to near-record summer (June–
August) heat occurred from the Great Lakes into
the Northeast with record seasonal heat in New
Hampshire and New Jersey. Much-above-average
temperatures in the southwestern United States
during July resulted from an upper-level ridge situated
over the region for most of the month. Temperatures
exceeded 38°C (100°F) and broke more than 200 daily
records in six western states. A new record of seven
consecutive days at or above 52°C (125°F) was observed
in July at Death Valley, California (previous record of
five days). A persistent upper-level cyclonic circulation
positioned over the central United States into August
contributed to above-average summer temperatures
across the eastern United States, where August temperature records were set in New Jersey and Rhode
Island. No state in the contiguous United States reported
below-average temperatures during the season.
An uncharacteristic blocking ridge over Alaska fostered
an exceptionally warm and dry summer throughout
the state. Statewide June–August 2005 temperatures
were third warmest since reliable records began in
1918. Overall, 2005 was Alaska’s sixth warmest year
on record, and was the sixth consecutive aboveaverage year for the state (Figure 6.10).
Autumn (September–November) temperatures were
much above average across large parts of the southern
and central United States and portions of the Northeast
and Mid-Atlantic, with warmer-than-average temperatures present throughout all but three states in the Pacific
Northwest (Figure 6.9, right). The contiguous United
States recorded its fourth warmest autumn in the last
111 years. This near-record heat resulted from a quasistationary 500-hPa ridge situated across eastern North
America. No state in the contiguous United States was
cooler than average during this season.
Figure 6.11. March–May 2005 statewide ranks of
precipitation for the contiguous United States;
a rank of 111 (1) in the USHCN record represents
the wettest (driest) year since 1895.
(iii) Precipitation and drought
Average precipitation for the contiguous United States
in 2005 was 755 mm, slightly above the long-term
(1895–2005) mean of 740 mm. Precipitation across the
United States in 2005 was characterized by persistent
moderate wetness in the Northeast and Southwest,
below-average precipitation in some parts of the
Northwest, and developing dryness from the Southern
Plains into the Great Lakes (Figure 6.9). An area from
Texas to parts of the Midwest and Ohio Valley was drier
than normal. Also, despite the significant rainfall associated with Hurricanes Katrina and Rita, Arkansas and
Louisiana reported much-below-normal precipitation
for the year. Conversely, Maine and New Hampshire
had their wettest year on record, and most states had
above- or much-above-normal precipitation.
December 2004 through February 2005 was very
wet from the California coast, through the Plains,
and into the Great Lakes and Northeast. There was
also much-above-normal precipitation around the
Great Lakes. A strong blocking high over the Gulf of
Alaska in conjunction with an amplified trough over
the southwestern United States generated an active
storm season along the West Coast. However, March
marked the beginning of a very dry period across the
central United States, extending from Texas to the Great
Lakes (Figure 6.11). An active storm track across the
western United States led to above- to much-abovenormal precipitation in the West, the Northern Plains,
the Southeast, and the far Northeast.
Figure 6.10. Alaska statewide average annual temperature
anomalies (°C), 1919–2005 [Source: NOAA/NCDC]
The boreal summer (June–August) brought muchabove-normal rainfall to the Southeast and parts of
61
the central and northern Great Plains. Only nine states
experienced below-normal precipitation, and just one
state (New Mexico) much-below-normal summer precipitation. Stormy conditions contributed to a record
wet October and autumn season across much of the
region. Six states (Maine, Vermont, New Hampshire,
Connecticut, Rhode Island, and Massachusetts)
reported their wettest autumn on record, and nine
states set a record for the wettest October, with two
additional states reaching their second wettest. Mt.
Washington, New Hampshire, set a record for the greatest October snowfall on record (200 cm), exceeding
the previous record set back in 2000 by 102 cm.
The year ended with a very dry December from the
Southwest, across the Southern Plains, to the Ohio
Valley and eastern Great Lakes. Several states from the
Southwest to the Lower Mississippi Delta had one of
their 10 driest Decembers; Arizona and Arkansas had
their driest December on record. December capped
a three-month period of much-drier-than-normal
weather in the Southern Plains, with Arkansas and the
Arklatex region (nexus of Arkansas, Texas, Louisiana,
and Oklahoma) all experiencing the driest October–
December on record.
At the beginning of the year, approximately 8% of
the contiguous United States was in moderate to
extreme drought, as defined by the Palmer Hydrological
Drought Index (PHDI; Palmer 1965; Heim 2002). The
areal extent of moderate to extreme drought grew to
reach a peak of 21% of the contiguous United States
in December 2005 (Figure 6.12). Precipitation deficiencies from March to June and again from September
to December resulted in the emergence and intensification of drought conditions from the Texas Gulf
Coast to the Great Lakes. Northeastern Illinois and
the Arklatex region were significantly impacted by
the emerging drought.
The development of drought conditions in parts of the
Midwest and Southern Plains can be attributed to a
pronounced shift in synoptic circulation in March–June
and October–December. During both periods, the
northern branch of the polar jet was active across the
western United States, while the southern branch was
active over the United States. Southeast and Atlantic
coast. The resulting absence of storms from the Deep
South to the Great Lakes created substantial precipitation deficits. The lack of precipitation from October to
December left Arkansas with its driest of such periods
on record and Louisiana with its third driest.
62
Figure 6.12. Change in the PHDI between
1 January–31 December shown by U.S. Climate Division
[Source: NOAA/NCDC]
The same synoptic conditions brought above to muchabove-normal rainfall to the northwestern United
States during the spring and the last two months of
2005, contributing to a significant reduction in total
drought area across parts of the Northwest by the
end of the year. The western United States drought
of 1999–2004 was one of the most severe droughts
in this region over the last 100 years. More than five
years of precipitation deficits lowered streamflows
and depleted reservoirs. Some reservoirs recovered
during 2005, but aggregated reservoir levels were still
below average at the end of the year.
(iv)
Snowpack and wildfires
(a)
Snow
The 2004/2005 snow season and snowpack was generally above average across the Southwest and much
below average across the northern Rockies and Pacific
Northwest. By the end of the 2004/2005 winter, the
Northwest snowpack was just 50% of normal. Snow
cover was slightly below average for the North American
continent as a whole over the winter and much below
average for the spring. This is consistent with a trend
toward reduced spring snow cover over North America
(Mote et al. 2005). Snow cover has been below average
in all but four years since the mid-1980s.
Notable snow storms in 2005 include a major winter storm, referred to as the “Blizzard of 2005,”
which deposited well over 30 cm of snow across
much of southern New England in January. Boston,
Massachusetts, had its snowiest January on record
partly as a result of that storm. NOAA’s operational
Northeast Snowfall Impact Scale (NESIS), developed
by Kocin and Uccellini (2004) to characterize and
rank high-impact northeastern US snowstorms,
ranked this January snow storm as the seventh
most intense on record for the region. In other
regions, a late-season (April) snow event produced
over 61 cm of accumulation in the mountains west
of Denver, Colorado, and a significant snow storm
on 27–28 November generated blizzard conditions
US CLIMATE EXTREMES INDEX (CEI)
How has the climate changed over the past
century? In what ways is it changing and by
how much? Many people, including climatologists, have been struggling with these
questions for some time now, not only for
scientific interest but also to aid in policy decisions (Houghton et al. 2001) and to inform the
general public. In order to answer these questions, it is important to obtain comprehensive
and intuitive information that allows interested
parties to understand the scientific basis for
confidence, or lack thereof, in the present
understanding of the climate system. One
tool, first developed as a framework for quantifying observed changes in climate within the
contiguous Unites States, is the United States
Climate Extremes Index (CEI).
across the nor thern High Plains, accumulating
up to 61 cm of snow in parts of Nebraska and the
Dakotas.
In contrast to the above-average snowfall season
in 2004/2005, the beginning of the 2005/2006 snow
season in the Southwest was nearly nonexistent. An
examination of USDA snowcourse/snow telemetry
(snotel) station data in Arizona revealed that 31 of
K. L. GLEASON28
The CEI was first introduced in early 1996
(Karl et al. 1996), with the goal of summarizing and presenting a complex set
of climate changes in the United States
so that the results could be easily understood and used to aid decision making by
policy makers. The CEI initially consisted
of a combination of five separate climate
change indicators. Recent revisions include
a sixth indicator related to extremes in
landfalling tropical system wind speed. The
CEI is now also evaluated for nine standard periods or seasons including: spring
(MAM), summer (JJA), autumn (SON),
winter [December–February (DJF)], warm
(April–September), cold (October–March),
hurricane (June–November), year to date,
Figure 6.13. Annual U.S. Climate Extremes Index values
(1910–2005) [Source: NOAA/NCDC]
and annual (January–December). The CEI
conveys the percentage area of the United
States that has been affected by climate
extremes as they relate to monthly maximum and minimum temperatures, daily precipitation, and the Palmer Drought Severity
Index (PDSI) within a given period.
The annual CEI for 2005 was about 41%,
which is much above the expected value
of 20% and is the second largest value
since reliable records began in 1910
(Figure 6.13). This high 2005 CEI was due
to the combined impacts from a record
active Atlantic hurricane season, extremes
in monthly maximum and minimum temperature, much-above-normal wet PDSI,
and extremes in daily precipitation. In
addition, the 20 05 hurricane and cold
seasons had record CEI percentages.
Approximately 44% of the United States
was affected by climate extremes during
the hurricane season and nearly 38%
during the cold season. All six indicators
were well above the expected percentage
during the hurricane season. Extremes
in much-above-average mean minimum
temperature were more than five times
the expected value. Much-above-average
mean maximum and minimum temperatures, a wet PDSI and the large number
of days with precipitation all contributed
significantly to the record extreme 2005
cold season. A more detailed explanation
of the CEI and graphs of the most current CEI and the individual indicators that
comprise the CEI may be viewed at the
NCDC CEI Web site at www.ncdc.noaa.
gov/oa/climate/research/cei/cei.html.
63
33 sites (94%) were snow free at the end of 2005—the
most snow-free locations in at least the past 40 years.
(b)
Wildfires
Preliminary estimates from the National Interagency
Fire Center suggest that 2005 will break the record set
in 2000 with over 3.45 million ha (8.53 million acres)
burned. During the 2000 fire season, roughly 3.41 million ha were consumed across the entire United States,
with over 2.83 million ha burned in the contiguous
United States. Despite the record area burned, the
total number of fires across the country continued to
decline in 2005, suggesting the average size of individual fires has increased over the past 20 years.
In Alaska, over 1.78 million ha burned in 2005, compared
to nearly 2.43 million ha consumed in 2004, which was the
worst fire season on record for the state. Above-average
temperatures coupled with below-normal precipitation
during the summer months contributed to the aboveaverage wildfire season across Alaska in 2005.
Atypical wildfire activity erupted across parts of the
central United States during December 2005. Numerous
large fires, enhanced by extreme drought conditions,
developed across parts of Oklahoma, Texas, and the
Southern Plains. Many of these fires continued to burn
into early January 2006. Over 162,000 ha had burned
across the Southern Plains during the first week of the
New Year, normally a time of very low fire activity.
(v)
Severe extratropical storms
Preliminary estimates indicate there were only
nine very strong to violent tornadoes (F3–F5 on the
Fujita scale) during the 2005 official tornado season
(March–August), all of F3 intensity. This was significantly below the 1971–2000 mean of 37, contributing
to a slight negative trend in very strong to violent
tornadoes observed since 1950. However, two lateyear (out of season) tornado outbreaks increased the
annual total.
A severe weather outbreak accompanied by over
30 reported tornadoes occurred across Mississippi and
Louisiana in April. In June, tornadoes ripped through
the town of Hammond, Wisconsin, causing over
US$ 3 million in damage. Severe thunderstorms in
August generated tornadoes that killed at least three
people in Wisconsin and Wyoming. Tornadoes also
touched down in September across parts of the central United States between Oklahoma and Wisconsin.
On 6 November, a deadly Midwestern tornado outbreak
claimed 24 lives in and around Evansville, Indiana. This
was the deadliest United States outbreak since 1998.
Additional severe weather impacted the same region
on 15 November, with over 30 tornadoes reported.
Among these was the strongest tornado of the year, an
F4 twister that reached a higher intensity than any of the
tornadoes that developed during the official season.
III)
MEXICO M. CORTEZ VÁZQUEZ18
(i)
Temperature
In 2005, the areally averaged annual mean temperature
for Mexico was 21.4°C, which is 0.7°C warmer than normal
(based on the period of record, 1980–2004). The year,
2005, was ranked as the second warmest, behind 1998,
since the start of the national temperature data set in
1980 (Figure 6.14). The warmth in 2005 continued the
trend of above normal temperatures in Mexico since the
mid-1980s. Nationally, the lowest minimum temperature
Figure 6.14. Annual temperature anomalies
(°C) over Mexico (1980–2005)
64
systems. However, an early withdrawal of the summer
monsoon in northern Mexico, along with a persistent
meteorological drought in the western part of the
country during the entire summer, resulted in limited
water storage at all dams and hydrological drought
declarations by year-end along the Lerma–Chapala
and Cutzamala Basins.
(iii) Notable events
Figure 6.15. Percent of normal precipitation across
Mexico during the 2005 rainy season (May–October)
relative to the 1941–2005 mean
for the year was -17°C, reported in the mountains of
Durango in northwest Mexico, the same area that holds
the long-term historical record minimum temperature of
-25°C reported in December 1997 (based on 1980–2005
data). In 2005, maximum temperatures of 49.5°C were
reported in Chihuahua and Michoacán, and these temperatures were only 0.5°C less than the national all-time
historical record temperature of 50°C.
(ii)
Precipitation
Nationwide, the areally averaged rainfall was 778 mm,
which was 14.5 mm (2%) above the long-term climatological mean defined by the period of 1941–2004. The
year ranks as the 25th wettest on record. Although
annual rainfall total was slightly above average, rainfall
distribution was very irregular throughout the year. The
rainy season (June–October) was characterized by short
events of heavy rainfall, which were mainly associated
with tropical cyclones that approached Mexico from
the Atlantic side of the continent. Wet conditions were
observed in February, followed by a dry trend from
March to May. The onset of the summer rainy season
started 3–4 weeks later than normal in southern Mexico,
and this delay influenced the northward progression of
the monsoon during the early summer. The 2005 rainy
season was finally established after mid-June, with
exceptionally wet conditions being recorded in July
and again in October. Large rainfall deficits developed
in September, which is normally the wettest month on
average for Mexico. Although the total amount of rainfall
was slightly above normal, precipitation was localized
over small areas in the south and southeast, around
the tracks of landfalling tropical cyclones. Portions of
northern Mexico also received significant amounts
of rain during February, associated with midlatitude
Climatologically, the annual rainfall distribution in
Mexico clearly reflects the influence of tropical cyclone
activity on both sides of the country (Figure 6.15).
The southwest coast and western Mexico typically
receive appreciable rainfall from Pacific tropical storms,
but during the 2005 season the storms developed and
tracked farther offshore than normal (see section 4C).
This helped to depress rainfall totals in western and
northwest Mexico, with only two systems (Dora and
Otis) approaching the Pacific Coast states. In contrast,
a very active season was observed in the Atlantic and
Caribbean basins, with seven systems making landfall
in Mexico: Hurricanes Emily, Stan and Wilma; Tropical
Storms Bret, Gert, and Jose; and Tropical Depression
Cindy. The number of landfalling tropical cyclones in
Mexico in 2005 represented a new record since the start
of the satellite era. In southeast Mexico, Stan produced
abundant rainfall across the Yucatan Peninsula on
2–3 October before moving into the states of Veracruz
and Oaxaca. Pentad and monthly rainfall totals exceeded
200% of normal along and to the right of Stan’s track
into mainland Mexico, with heavy flooding in portions
of northern Veracruz and Oaxaca. Stan developed a
large moisture tap across the Pacific slope of Chiapas,
promoting widespread flooding along the Pacific slope
of Chiapas, and sections of Central America. By far
the most destructive tropical cyclone during the 2005
season was Wilma, which moved slowly across the
Yucatan Peninsula on 20–23 October, causing severe
economic loss and several fatalities in the Cozumel
and Cancun areas. Based upon wind speeds and sea
level pressure readings, Wilma was the most powerful
hurricane on record to make landfall in Mexico.
D.
CENTRAL AMERICA AND THE CARIBBEAN
E. K. GROVER-KOPEC31
I)
TEMPERATURE
Annual mean surface temperatures were slightly above
average across Central America and the Caribbean
65
during 2005 (Figure 6.16). Temperatures were at
least 0.5°C above normal for the year over the entire
region except for western Cuba and the Pacific coastal
regions of Costa Rica and Panama. The warmest
conditions relative to climatology were observed in
Guatemala and Belize, where annual departures from
the 1971–2000 mean exceeded 1°C.
II)
bit as the eastern portion of the island received 25%
more precipitation than normal during the year. Much
of this excess precipitation came during May, June,
and October, which are among the wettest months of
the year in that area. The climatological precipitation
distribution across most of the Caribbean and Central
American region is bimodal, with relative maxima
occurring in May–June and September–October.
PRECIPITATION
Most of the Central American isthmus experienced
drier-than-normal conditions in 2005, though precipitation
deficits were not as severe and widespread as those
seen in recent years.The most significant standardized
12-month precipitation anomalies occurred across
Honduras, Nicaragua, central Costa Rica, and southern
Panama (Figure 6.16). The largest absolute annual
precipitation deficits compared to the 1979–2000 base
period were observed in this same region. Negative
anomalies exceeding 1000 mm were observed in
eastern Honduras and Nicaragua, accounting for
approximately half of the climatological mean annual
precipitation in these areas, typically among the
wettest regions in Central America.
The largest contrast between the precipitation regime
of 2005 and that of recent years was observed in the
Caribbean, particularly in Jamaica, eastern Cuba and
western Haiti. Drought conditions, which have had a
large impact on water resources and agriculture over
the past few years in Cuba (Levinson 2005), eased a
III)
NOTABLE EVENTS
The record-breaking 2005 Atlantic hurricane season
(see section 4 sidebar) caused devastating losses
across the region from July to November. Most damage came from Hurricanes Dennis, Emily, Stan and
Beta, and Tropical Storm Gamma, which primarily
affected the countries of Cuba, Grenada, Guatemala,
Nicaragua and Honduras, respectively.
While all of these storms had tremendous localized
impacts, Stan was arguably the most destructive
and certainly the deadliest in the region, affecting
eight countries in early October. The storm brought
150–400 mm of precipitation to western Guatemala.
One-third of the population of Guatemala was affected
by Stan and more than 1,000 deaths were reported.
Agence France-Presse and Reuters reported that most
of these deaths occurred when mudslides buried the
villages of Panabaj and Tzanchal in the southwestern
department of San Marcos, where some of the largest
precipitation accumulations were reported.
Figure 6.16. Central American and Caribbean 2005 annual (left) temperature anomalies (°C; 1971–2000 base), and (right) precipitation
anomalies (mm; 1979–2000 base) from CAMS–OPI
66
E.
SOUTH AMERICA
I)
OVERVIEW M. RUSTICUCCI78 AND J. L. CAMACHO13
South America experienced below-normal precipitation anomalies across a majority of the continent
in 2005, with some excesses in the northwest and
southwest (Figure 6.17). However, despite overall
deficits, extreme but inconsistent and widely scattered precipitation events over most of the continent
adversely impacted the population. Additionally,
anomalously frequent cold air advection from higher
latitudes was experienced by both the north and south
sides of the continent.
Most of northern and eastern South America experienced temperatures above the 1961–1990 normal, while
western parts were below average. This is generally
reflected in Figure 6.17.
One important recent improvement in the climatic
analysis of South America has been the addition of
annual or monthly averages from up to 516 individual
stations, as provided by the National Meteorological
and Hydrometeorological Services of Argentina,
Bolivia, Chile, Colombia, Ecuador, Uruguay, Paraguay,
Peru and Venezuela, and also from the Brazilian Centre
de Provisão de Tempo Estudos Climáticos (CPTEC).
Attention has been given to minimizing excessive
local-scale influence to better establish the regional
behavior of extreme climatic events. These data are
the primary source for subsequent precipitation analyses and have been blended with the NCDC/GHCN
database for temperatures, with 1961–1990 as the
reference period.
II)
NORTHERN SOUTH AMERICA AND THE SOUTHERN
CARIBBEAN J. D. PABÓN63
Typical responses to equatorial Pacific cold conditions
were observed in northern and northwestern South
America in 2005, with precipitation generally above
normal except in the Colombian–Venezuelan Amazonia
region, where precipitation deficits occurred. Intense
rainfall, particularly from March to May and during
the second half of the year, caused flash flooding and
landslides that resulted in death and considerable
damage. Around one million people were affected
during intense October–December rainfall events
in Colombia, which caused more than 500 deaths,
destroyed over 1,000 houses and heavily damaged
infrastructure.
Figure 6.17. South American 2005 annual (top)
temperature anomalies (°C; 1971–2000 base),
and (bottom) precipitation anomalies
(mm; 1979–2000 base) from CAMS–OPI
67
Although average temperatures for several months were
below normal over several regions of Colombia, northeastern parts of Columbia and Venezuela experienced annual
average air temperature up to 1.0°C above normal, while
western regions were slightly below normal. Northern
Hemisphere midlatitude synoptic activity penetrated further south than normal, generating perturbations in the
ITCZ and producing considerable precipitation over the
Caribbean and northern South America. The Caribbean
also was affected by the increased tropical cyclone activity
in the Atlantic basin in 2005 (see section 4C).
DROUGHT IN AMAZONIA
TROPICAL SOUTH AMERICA EAST OF THE ANDES
J. A. MARENGO50
Heavy rain in January caused flooding in Georgetown,
Guyana, and surrounding areas, affecting an estimated 290,000 people. Conversely, large negative
rainfall anomalies were measured east of the Andes
in the Amazon, northeast and southern Brazil, and in
the South American monsoon–Pantanal regions. The
rainy season in northeast Brazil during February–May
was below normal, reaching drought levels in some
J. MARENGO50
In 2005, large sections of the western part
of the Amazon Basin endured the worst
drought in 40 years and also one of the most
intense since the beginning of the twentieth century. While the Amazon normally
rises and falls in conjunction with seasonal
precipitation, 2005 rainfall was well below
normal (Figure 6.18), allowing rivers to drop
to record low levels. Levels of the Madeira
and Solimões Rivers, two of the Amazon’s
major tributaries, dropped to record and
38-year lows, respectively. In the Brazilian
states of Rio Branco, Rondonia, southern
Para and southern Amazonas, rainfall was
30%–50% below normal in January–April
2005, 33% below normal in June and
August, and 65% below normal in July.
According to the meteorological service of
Peru, the hydrological year of 2004/2005
exhibited rainfall well below normal in
Peruvian Amazonia, with mean rainfall for
the hydrological year September 2004–
August 2005 up to 39% below normal.
Rainfall on the basins of the Bolivian Beni
and Mamoré Rivers was about 20%–30%
below normal for January–April. Drought
conditions favored the occurrence of forest
fires, and in September the number of fires
was about 300% more than those detected
in September 2004.
Levels of the Amazonas River at Iquitos,
Peru, and Leticia, Colombia; the Solimões
River at Tabatinga and Fonte Boa, Brazil;
the Acre River at Rio Branco, Brazil ;
the Mamoré at Puerto Varador, Bolivia;
68
III)
and the Ibaré River at Puerto Almacén,
Bolivia all were well below normal during
most of 2005 until September, in some
cases as much as 2 m below normal
monthly means. At daily time scales, the
situation was even more dramatic. The
level of the Solimões River at Tabatinga
and Fonte Boa decreased from 11.5 and
21 m (respectively) in May to near 1 and
11 m (respectively) in September. Rainfall
star ted by the end of October 20 0 5,
reached a November mean of almost
107% above normal, and recharged the
Rio Amazonas in Iquitos to a normal level
by November. By January 2006, the Acre
and Madeiras Rivers achieved anomalously
high levels (11.08 and 12.34 m, respectively) due to the intense rains. In contrast
to the intense drought of the 1982/1983
and 1997/1998 El Niño years, the 2005
drought was concentrated in western and
southern Amazonia and was not related to
El Niño, which brings drought to central and
eastern Amazonia, but rather to a warming
of the tropical North Atlantic during most
of 2004 and 2005.
Figure 6.18. Rainfall anomalies (mm day -1) in central Amazonia
during the peak season (December–May) 1951–2005;
black arrows represent drought years 1983, 1998, and 2005.
semiarid interior regions and severely impacting over
two million inhabitants. Amazonia also experienced
intense drought during most of 2005, especially in
southern and western sections of the basin (see sidebar). More than 167,000 people have been affected by
the drought, both directly and indirectly. Low river
levels impacted the region’s main source of transport
and contributed to the deaths of large numbers of
already endangered manatees and river dolphins.
In west-central and southeastern Brazil, the rainy
season was from below to slightly below the normal.
Drought conditions were present in the Chaco region
of Bolivia and Paraguay during January and February
2005. Low water levels on the Paraguay River significantly reduced barge traffic in 2005.
Rainfall in southeastern Brazil was primarily in the
form of intense events that lasted several days.
Several of Brazil’s large cities were flooded by these
events, leaving much of the population without
power or shelter. In and around São Paulo and
Rio de Janeiro, dozens of people died due to landslides and flooding.
Annual air temperature anomalies reached almost
3°C above normal in eastern Brazil, with every month
above normal and the warmest months being April,
August and October. In October, typically the onset of
the rainy season in the southern areas, temperature
anomalies were up to 5°C above normal. From October
to December, temperatures were over 3°C above
normal in the far western Amazonia. Air temperatures
in Bolivia and northern Paraguay were 1°–4°C below
normal from September to November.
IV)
TROPICAL SOUTH AMERICA WEST OF THE ANDES
R. MARTÍNEZ51
As in northern South America, rainfall in Ecuador
and Peru was strongly influenced by SST over the
Niño-1+2 region during 2005. Despite weak warming
in the tropical Pacific, cold coastal SST anomalies led
to negative rainfall anomalies along the Ecuadorian
coast. Mean temperature was also below normal
during 2005. In November 2005, a strong frost caused
significant damage in the central and southern
highlands of Ecuador. In Peru, rainfall was below
normal along the central and southern highlands,
continuing several years of drought in this region.
Bolivia also experienced drier-than-normal conditions
in 2005, except for October and November when
intense rains generated flooding and damage. Mean
temperature in Bolivia was above normal across
most of the country.
V)
SOUTHERN SOUTH AMERICA M. BIDEGAIN7 AND
M. RUSTICUCCI78
Annual precipitation anomalies over southern
South America show light deficits over the east and
surpluses over central Chile and western and southern
Argentina. Above-normal precipitation for several
months contributed to the positive anomalies in these
regions. A series of intense summer (June–August)
precipitation events also contributed, with some
local anomalies exceeding 700% of the normal. On
26 June, 162.4 mm of rain fell over Concepción, Chile,
generating landslides that killed five and injured 4,800.
Between 26 and 28 August, 120 mm of rain fell in
48 hours in Santiago, Chile, resulting in 1,153 injured,
755 houses damaged, and an estimated economic
cost of US$ 10 million.
The regional core of negative precipitation anomalies
was in the Chaco region and southern Paraguay,
where intense drought prevailed until spring 2005.
Precipitation deficits produced livestock losses
and reduced water levels on the Uruguay River,
impacting hydroelectric generation. Strong negative
October–December rainfall anomalies dominated the
southern part of the region, affecting agriculture in
this productive region. In southern Brazil, seasonal
(December 2004–March 2005) rainfall 100–500 mm
below normal produced intense drought and heavy
agricultural losses. The southern state of Rio Grande
do Sul was the most affected, and while May rainfall
alleviated the drought, it produced flooding in some
cities. Damage attributed to the drought of 2005 in
southern Brazil was considerable: 2 million people
were affected by water shortages, 13 million tons
of agricultural products were lost, and economic
losses were on the order of US$ 3 billion.
Annual air temperature anomalies were generally
near normal, with eastern regions above normal and
central and western regions slightly below normal.
Uruguay experienced temperatures above normal (up
to +1.2°C), especially near the Brazilian border. From
January to August, most monthly temperatures were
above normal, with May–August having the largest
anomalies. June temperatures broke records (for the
1961–2004 period) over northeastern Argentina, and
winter was 2°C warmer than normal in Uruguay. In
69
contrast, cold air advection in September affected the
eastern part of the region. October–December temperature anomalies were up to 3°C below normal with
early December frosts, including a few intense frosts
in the Andes that killed thousands of sheep. Annual
air temperatures in Chile were slightly above normal
in the central region, and slightly below normal in the
south. April, May, and June temperatures were below
normal, especially in southern Chile. The week of
26 June, a severe cold air outbreak bet ween
34° and 36°S lef t 30,000 injured and af fected
12,000 homes.
On 23–24 August 2005, an exceptionally strong midlatitude cyclone occurred over Río de la Plata and
southern Uruguay. The gale was characterized by
unforced rapid deepening to a near–record (locally)
low mean sea level pressure, very high winds, and
anomalous cold surface temperatures. High winds
contributed to extensive damage and 10 deaths along
the Uruguayan riverside.
F.
ASIA
I)
RUSSIA O. N. BULYGINA,11 N. N. KORSHUNOVA,40 AND
V. N. RAZUVAEV72
(i)
Temperature
Russia experienced very warm conditions in 2005.
The mean annual air temperature anomaly relative to
the period of record (1936–2005) was +1.6°C, which is
the second highest value since 1936 (Figure 6.19).
The year began with January temperatures above
normal across all of Russia, although very cold weather
was observed in places. Northeast European Russia
experienced particularly warm conditions, with
mean monthly temperature anomalies exceeding
+8°C. Anomalies reached +7°C over central regions.
Moscow’s January 2005 temperature ranked third
highest on record, with record maximum daily air
temperatures observed on five days (e.g., 5.2°C on
the 9th). February temperatures in north European
Russia and western Siberia were up to 10°C above
normal.
Interestingly, January and February air temperatures in
the north of Asian Russia were often higher than those
to the south. At Turukhansk, the 12 January mean air
temperature was -5.5°C, which is 21.5°C above normal.
70
Figure 6.19. Departures of mean annual air temperatures (red)
over Russian territory for the period 1931–2005; a linear trend
of +0.14°C decade-1 (black) is also shown.
In contrast, at the end of January the Novosibirsk and
Kemerovo regions experienced temperatures as low
as -38°C, and temperatures in the Republic of Altai
reached -47°C. Particularly strong February frosts
occurred between the 15th and 18th in Altai (-40° to
-43°C), and from the 1st to the 10th in Trans-Baikal
(-38° to -44°C), while true “Siberian” frosts (-35° to
-40°C) were recorded during 14–19 February in the
Krasnoyarsk Territory and Khakasia (south-central
Siberia). The Republic of Tuva experienced its most
severe and persistent frosts in the past 20 years, as
temperatures fell to -48°C in the Tuva hollow.
March brought bitter cold across much of European
Russia, with record cold mean monthly temperature
anomalies (-5° to -6°C) in several northeastern areas
and colder-than-normal (-3° to -4°C anomaly) conditions in central and western regions. However, April
countered with positive temperature anomalies over
most of Russia. Western parts of the Sakha Republic
(northeast Siberia) were particularly warm, with mean
monthly anomalies from +7° to +8°C. Anomalous
warmth continued into May, with the mean May
temperature for Russia tying the record set in 1943.
May 2005 was the hottest May in the 105-year temperature record for the Ural Federal District.
During the first 20 days of June, central and southern
regions of European Russia recorded anomalously cold
air temperatures (0° to -2°C) as a result of frequent
cold air intrusion. Concurrently, temperatures ran
1° to 2°C above normal across most of the Russian
Far East. In early July, western and southern-central
regions of Siberia experienced a heatwave, with diurnal
temperatures climbing to 39°C in places.
Figure 6.20. Russian air temperature
anomalies (°C) in autumn 2005; insets show
November mean monthly air temperatures
at meteorological stations Ust’-Cil’ma
(1920–2005) and Uct’-Maja (1926–2005).
A strong anticyclone centred over European Russia
caused very dry and hot weather in August. The
Novosibirsk region and Altai in western Siberia experienced mid-August diurnal temperatures between
28° and 38°C. At 40°C, Zmeinogorsk exceeded the
previous August maximum temperature record by
2°C. For the whole of Russia, summer 2005 was one
of the warmest on record.
With a temperature anomaly of +2.7°C, autumn
2005 was the hottest autumn on record for Russia
(Figure 6.20). While September was warm, eastern
Siberia experienced its warmest October in the past
65 years, with October temperatures 2°–5°C above
normal. Record November temperatures were also
reported at several meteorological stations in northeastern European Russia and in the southeast of the
Sakha Republic, as two large heat domes formed
over those regions (Figure 6.20). Temperatures near
the centres of these areas of heat were 9° and 11°C
above normal, respectively.
The warm weather over European Russia and the
Sakha Republic persisted into December, with monthly
temperatures 1°– 4°C above normal. However,
December temperatures were 4°–5°C below normal in southern Siberia (Krasnoyarsk Territory
and Irkutsk region), as a cold pool formed in the
Siberian anticyclone zone following the warm
autumn.
(ii)
Precipitation
The warm winter temperatures also led to abovenormal January precipitation in places. Moscow
reported a new record January precipitation total
of 98 mm (232% of monthly average). Heavy March
precipitation was recorded in the eastern regions of
European Russia and in the Urals, in some regions
exceeding normal values threefold. Frequent March
snowstorms with heavy snow were observed across
European Russia (from the Nenets Autonomous
District to northern Caucasia). The Taimyr Peninsula
experienced strong winter blizzards with heavy snow
and winds exceeding 25 m s -1. In the east, TransBaikal, Sakhalin and Kamchatka were repeatedly
attacked by strong cyclones that brought heavy
snow and blizzard conditions. March precipitation in
these regions was more than double that of normal
amounts.
Heavy April precipitation (200% to > 300% of normal)
was recorded in central and southern regions of
Siberia (Krasnoyarsk Territory, Khakasia, Cis-Baikal
and Trans-Baikal). From 12 to 16 days of precipitation fell in Khabarovsk Territory and the Amur region,
more than double the normal frequency and totaling
over 200% of the normal precipitation for the month.
Wet conditions continued into May in Khabarovsk,
Maritime Territories and Sakhalin, resulting in high
river levels.
71
Summer precipitation across Russia was often accompanied by severe thunderstorms with hail and wind
squalls. Hail to 35 mm was recorded in the Krasnodar
Territory. In early June, the Arkhara River (Amur
region) flooded to a record June level of 4.1 m after
a 2-day, 100-mm rainfall. However, the hot June was
accompanied by precipitation deficits over western
Siberia (20%–30% of monthly normals). July precipitation in European Russia was inconsistent, with heavy
thunderstorms in places and precipitation deficits in
others. August precipitation was just 8%–30% of normal across European Russia, although the Kaliningrad
region in the far west received over 300% of normal
monthly precipitation.
Precipitation deficits continued into September.
Moscow experienced one of its driest Septembers
on record (12.2 mm, 18% of normal). With the high
temperatures, fire hazard increased over much of
European Russia and several peat bogs caught
fire. Near-normal precipitation returned to Russia
by December, with the exception of the south of
the Central Federal District, where around 200% of
monthly normal snow fell.
II)
CHINA F. REN73 AND G. GAO23
(i)
Temperature
The 2004/2005 winter (December–February) mean
temperature was near normal for China, but it ranked
the third lowest since the 1986/1987 winter. In midFebruary, rare icing events occurred in some provinces
in southern China (e.g., Hunan, Hubei, Guizhou). In
Hunan Province, the power grid was the most heavily
affected by icing since 1954.
As regards summer seasonal mean temperature,
2000 and 2001 tied as the highest ranked since 1951.
Heat waves occurred frequently in central-eastern
China and Xinjiang. Southeastern China experienced
5–15 days more than normal (20–40 days) with maximum temperatures at or above 35°C. Seasonal extreme
daily maximum temperatures were 38°–42°C in North
China, western Huanghuai Region and South China,
while records (1951–2005) of seasonal extreme daily
maximum temperature were broken in parts of Hebei,
Shanxi, Shandong, Zhejiang and Inner Mongolia.
China also experienced a warm autumn, and the seasonal mean temperature ranks second in the historical
record (1951–2005). In southeastern China, heatwaves
returned during the middle of September, when daily
maximum temperatures soared to 35°–39°C.
(ii)
In 2005, the annual mean temperature of China was
0.6°C above the 1971–2000 mean (Figure 6.21). It was
the ninth consecutive year of warmer-than-normal
temperature since 1997. Regionally, temperatures
were above or near normal across most of China, with
1°–2°C above normal in the middle Tibetan Plateau
and eastern Xinjiang.
Precipitation
In 2005, annual precipitation was 17.7 mm above the
1971–2000 mean across China (Figure 6.21). Regionally,
precipitation was 30%–100% above normal in the
Huanghuai region, southern and northern Xinjiang,
Qinghai, northwestern Tibet and the southeast coast,
and 30%–80% below normal in northern Heilongjiang,
the middle of Inner Mongolia and northern Ningxia
(Figure 6.22).
Figure 6.21. Mean annual (left) temperature (°C) and (right) precipitation (mm) averaged over China relative to the 1971–2000 mean
72
Basin. Between late September and early October,
heavy flooding occurred in the Hanjiang River and the
Weihe River as the result of frequent and widespread
rainfall in the southeast part of northwest China and
the Huanghuai region. About 5.52 million people were
affected, with 14 dead in Shanxi, Hubei and Gansu
Provinces and 2.5 billion RMB (US$ 311 million) of
direct economic loss.
Figure 6.22. Precipitation anomalies (%) across China
(1971–2000 base)
Regional and short-term drought was a major characteristic in 2005. In southern South China, precipitation
was only 300–600 mm from September 2004 to May
2005, or about 30%–80% below normal, resulting in
severe persistent drought. From April to May, rare
spring drought occurred in Yunnan Province as a
result of long-term rainfall deficiency. Early summer
drought occurred in middle and lower reaches of the
Yangtze River due to a delay in the onset and shortened duration of the plum rain season generating
below-normal precipitation. Summer–autumn drought
occurred in northeast part of northwest China and
Inner Mongolia, and autumn drought affected Hunan
and western South China.
(iii) Notable events
An above-normal eight tropical storms or typhoons
made landfall in China, of which six (Haitang, Matsa,
Talim, Khanun, Damrey, and Longwang) were severe,
with winds over 162 km h-1 (see section 4C). Heavy
rain and high winds generated mudflows and widespread flooding. About 92 million people were affected
(386 dead), and economic losses of over 82 billion
yuan renminbi RMB (US$ 10 billion) were exceeded
only by losses in the 1996 typhoon season.
From 17 to 25 June, consecutive heavy rainstorms
impacted South China, with 300–600 mm of rain falling
in parts of Fujian, Guangdong and Guangxi Provinces.
The Xijiang River in Guangxi and the Minjiang River in
Fujian exceeded flood stage. About 21 million people
were affected by the floods—171 people lost their
lives and direct economic loss was over 18 billion
RMB (US$ 2.2 billion).
During early and middle July, heavy rain and flooding occurred in the upper reaches of the Huaihe River
Although fewer dust storms affected China than in
2004 and 2005 had the lowest number since 1954,
13 storms occurred. The most widespread occurred
from 16 to 21 April, affecting 12 provinces in northern China, while the most intense storm occurred on
27–28 April, impacting nine northern provinces or
regions, including Beijing.
III)
SOUTHEAST ASIA F. REN73 AND G. GAO23
(i)
East Asian monsoon
The onset commenced over the South China Sea
(SCS) in the sixth pentad of May, about two pentads
later than normal. Stronger-than-normal southwesterly flow advanced to and persisted over South China
until the fourth pentad of June. In the last 10 days of
June, the monsoon advanced to the region between
the Yellow and Huaihe Rivers. In mid-August, rapid
retreat occurred to around 30°N, where it remained
until mid-September. In the sixth pentad of September
the warm and humid air had withdrawn from East Asia
and wind direction in the SCS shifted from the southwest to northeast, signifying a near-normal closing
date to the East Asian summer monsoon.
Figure 6.23. Time–latitude cross section of pentad
precipitation anomalies (%) for 110°–120°E [Source: National
Climate Center (NCC) China Meteorological Administration
(CMA)]
73
The SCS summer monsoon index (-1.42) was weaker
than normal. Intensity of the SCS monsoon was
also weaker than normal during summer except
for the periods from the sixth pentad of May to the
third pentad of June and from the second pentad of
August to the third pentad of August (Figure 6.23).
Precipitation was above normal in most of South
China in June and in the Upper Huaihe River from
July to September.
(ii)
Temperature
Annual air temperature anomalies were generally
0.5°–1°C above the 1971–2000 mean. However, annual
temperatures over the SCS were near to slightly below
normal (Figure 6.24). Seasonal mean surface air temperatures were above average in most of southeast
Asia during December 2004–February 2005, with
anomalies exceeding 1°C in the northern and southeastern Indo–China Peninsula. Generally, temperatures
were close to normal across southeast Asia through
the remainder of the year.
(iii) Precipitation
Precipitation was generally below normal across most
of continental Southeast Asia in 2005. Northern and
western Myanmar, southern Vietnam and portions
of Malaysia observed annual anomalies more than
400 mm below normal. The Philippines, western
Thailand and the northern Malay Peninsula received
above-normal precipitation, with northeastern Malaysia,
central Vietnam and Mindanao all receiving over
400 mm above normal for the year (Figure 6.24).
December 2004–February 2005 rainfall was below average over most of Southeast Asia, and more than 80%
below normal in the western Indo–China Peninsula.
March–May rainfall totals were well below normal in
the northern Indo–China Peninsula and close to normal over the remainder of Southeast Asia. In April,
Thailand experienced its worst drought in seven years.
June–August precipitation was close to normal, though
heavy rainfall caused flooding in northern Thailand.
In Myanmar, heavy monsoon-related rainfall affected
the southern coastal areas during the second week of
September. Otherwise, September–November rainfall
was near normal.
(iv)
Notable events
In Indonesia, heavy January rains hampered tsunami
(December 2004) relief efforts, and continued heavy
rain in February generated landsides that left 61 dead
and 90 missing. On 9 June, continuous heavy rain
brought mudslides with 12 deaths and 11 missing
in northern Vietnam. In West Sumatra, Indonesia,
heavy rainfall produced landslides near Padang on
2 September. There were 16 fatalities and at least
10 injuries. Heavy October rains across central Vietnam
produced flooding with at least 67 fatalities. The
most severely affected area was Binh Dinh Province,
where 3,200 houses were damaged and most of the
fatalities occurred.
Figure 6.24. Asian 2005 annual (left) temperature anomalies (°C; 1971–2000 base), and (right) precipitation anomalies
(mm; 1979–2000 base) from CAMS–OPI
74
IV)
INDIA AND SOUTHERN ASIA M. RAJEEVAN71 AND
K. R. KUMAR42
(i)
Temperature
This year was marked by extreme weather all across
South Asia, both in terms of temperature and precipitation. At the beginning of the year, parts of Afghanistan
and adjoining Pakistan experienced extreme cold
weather, with temperatures more than 5°C below
normal in February. Severe cold also prevailed over
northern India and adjoining regions in late February.
During the postmonsoon season and toward the end of
the year, heavy snow and extreme low temperatures
occurred over northern parts of South Asia, causing
several casualties and seriously affecting the rescue
and rehabilitation work in Pakistan following the
destructive earthquake of 8 October.
On the other extreme, May and June brought scorching heatwaves, with maximum temperatures around
45°–50°C in India, Pakistan, and Bangladesh. Delayed
southwest monsoon rain allowed the heat to persist
into June, claiming at least 400 lives in India. An anomalous anticyclone and northwesterly winds created a
severe heatwave over central and northeastern India,
with maximum temperatures 6°–8°C above normal.
(ii)
Precipitation
During the third week of February, sections of northern Pakistan and neighboring areas of northern India
received heavy snowfall, described as the worst in
two decades. Snowfall accumulations reached almost
2 m in some parts of Jammu and Kashmir in India.
In Pakistan, heavy rains in the south and snow in the
north triggered flooding and avalanches, causing the
extensive loss of life and property. Heavy rains in
March also caused flooding in parts of western Pakistan
and Afghanistan; Balochistan Province was the worst
affected. Over Pakistan, the 2005 January–March seasonal rainfall was 121% of its long-term average.
(a)
South Asian Summer monsoon
The summer monsoon this year was marked by
unprecedented heavy rains and extensive flooding
in parts of western and southern India, affecting
more than 20 million people and resulting in more
than 1,800 deaths. Rainfall activity in Nepal and
Bangladesh was, however, below normal. Central
and southern parts of India received excess rainfall
during the season (Figure 6.25). While northwest
India received normal rainfall, seasonal rainfall over
northeastern parts of India was below normal by
more than 20%.
The onset of the 2005 southwest monsoon was
delayed to 5 June, when it arrived over the south
peninsula and northeastern parts of India. Despite
unfavourable synoptic conditions, the monsoon
advanced more quickly over northwestern India
and Pakistan, covering the entire subcontinent by
30 June, 15 days ahead of the normal.
Figure 6.25. June–September
precipitation (mm) over
India in 2005: (left) actual;
(centre) normal; and (right)
percentage anomaly
75
Countering declining trends of recent years, 12 low
pressure systems formed (the most since 1998), of
which five developed into monsoon depressions
and one into a cyclonic storm (the first in September
since 1997). The storm tracked from the Bay of Bengal
across central India and the Gangetic Plains, resulting
in widespread flooding.
While the spatial distribution of rainfall this season was
normal, it was intermittent. There were prolonged dry
spells in June and August, though excess rainfall in July
and September ultimately helped the season end with
near-normal rainfall. Due to the late onset and sluggish
advance of the monsoon, rainfall during much of June
was limited. In August, monsoonal rainfall was 27%
below normal over India. Precipitation deficits lead to
moderate drought for 25% of India’s meteorological
districts (2% severe drought). For the season, average
rainfall over India was near normal (-1%).
(b)
(iii) Notable events
The most notable event of 2005 occurred on 27 July,
when Mumbai (Bombay) received its greatest-ever
recorded 24-hour rainfall of 944.2 mm (most of it in
just 6 hours, between 1430 and 2030 local time) at
Santacruz, an observatory at the airport (Figure 6.26).
The previous record of 575.6 mm was set at Colaba on
5 July 1974. Interestingly, in the 2005 event Colaba, just
20 km from Santacruz, recorded only 73.4 mm of
rain. This localized event was confined to a region of
a 20–30 km radius. Although a warning for regionally
heavy rainfall had been issued, torrential rain severely
disrupted life in the city, with numerous fatalities,
heavy damage and large economic losses.
Northeast monsoon
Heavy rainfall continued unabated in southeastern
parts of India and Sri Lanka during the northeast monsoon season of October–December. Five low pressure
systems (four depressions and one cyclonic storm)
affected southern parts of India and Sri Lanka. In India,
Tamil Nadu and Andhra Pradesh were the most affected
states. The northeast monsoon seasonal rainfall over
south India was 165% of normal, the highest since 1901.
Associated flooding affected more than three million
Figure 6.26. Rainfall totals (mm) on 27 July around
Mumbai, India
76
people, with at least 300 fatalities and considerable
socioeconomic impact. In Sri Lanka, approximately
29,000 families in 10 districts were affected and at
least six deaths were reported.
V)
SOUTHWESTERN ASIA F. RAHIMZADEH,70
M. KHOSHKAM,41 AND E. K. GROVER-KOPEC31
(i)
Temperature
All of southwest Asia experienced above-normal
temperature in 2005, with annual temperature anomalies of 0.5°–2°C. North and northeastern Iran were 3°C
warmer than normal for the year. Despite these positive
12-month departures, the northern half of the region
experienced cooler-than-normal conditions during
February, when mean temperatures ranged from 5°C
in western Turkmenistan to -18°C in the highlands of
Tajikistan and Kyrgyzstan. Temperatures were 1°–6°C
below normal in these areas. Northern Iran experienced winter temperatures 3°C below normal. Northern
Afghanistan, where hypothermia and other cold-related
illnesses claimed more than 100 deaths, recorded temperatures of 1°–2°C below normal during February.
In contrast, temperatures were above the 90th percentile across the central portion of Southwest Asia
during March (2°–7°C above the long-term average;
Figure 6.27). Afghanistan, Kyrgyzstan, Tajikistan, and
Pakistan experienced temperatures 1°–2°C above
normal in June. Spring temperatures in Iran were
split between cooler-than-average conditions in areas
of the east, centre and northeast of the country and
warmer-than-normal conditions elsewhere. Seasonal
mean temperatures in southern Iran were 10°–35°C.
Most of Iran saw positive summer temperature anomalies exceeding 2°C. A heatwave that affected Iran in
July produced monthly anomalies up to 4°C above
normal. Autumn remained on average 2°C above
normal across Iran, although cooler-than-normal
conditions were observed over the Persian Gulf.
(ii)
Precipitation
Southwest Asia generally receives most of its annual
precipitation from extratropical disturbances traveling eastward from the Mediterranean Sea between
November and April. From July through August, the
South Asian monsoon generally brings precipitation to
southeastern Afghanistan, but tends to suppress summer precipitation in areas farther north and west.
Annual precipitation accumulations were slightly below
normal across the majority of the region during 2005,
and most of these negative annual anomalies were
25–75 mm below normal (Figure 6.27). These departures
were modest, however, and generally accounted for
less than 25% of normal annual precipitation (i.e., 2005
annual totals were about 75%–100% of normal). A few
areas received above-average precipitation during 2005,
including portions of southern Afghanistan, western
and northern Pakistan, and south-central Kazakhstan.
However, these departures were also relatively small
(10–50 mm) compared to long-term mean accumulations.
Despite the relative precipitation deficits across
Southwest Asia, portions of the region experienced
record snow amounts during January and February.
As much as 200 cm of snow fell in just 2 weeks in parts
of Tajikistan, contributing to more than 475 avalanches
in the mountainous country. Similar impacts were
reported in northeastern Afghanistan, where avalanches claimed approximately 160 lives. The heavy
snowfall in January and February and above-average
precipitation in late 2004 (Levinson 2005) contributed to
a healthy snowpack in the highland areas of Southwest
Asia, which is responsible for providing most of the
region’s water supply later in the year.
The 2004/2005 winter was also wetter than normal
across most of Iran, averaging 154.1 mm, or 14%
above 2004 levels and 24% above the long-term
mean. The largest anomalies were in southeast Iran,
where some locations received up to 3.5 times the
normal seasonal amounts. Also, early winter snow
alleviated a 7-year drought in the region. Tehran
experienced record February snowfall, and in northern
Iran, heavy late-February snow damaged or destroyed
over 7,000 homes.
Unfortunately, abnormally warm conditions in March
hastened the melting of the highland snowpack and
swelled rivers across the region. Heavy rainfall in central
and western Afghanistan during March exacerbated
conditions and caused extensive flooding in those areas.
The June heatwave melted remaining snowpacks
in Afghanistan, Kyrgyzstan, Tajikistan and Pakistan.
Pakistan’s northern provinces were extremely hard
hit by the resulting flooding. More than 460,000
people were affected and nearly 1 million ha of crops
suffered damage.
Although spring is generally the rainy season for
Iran, and wetter-than-average (100%–200%) spring
conditions prevailed over northern Iran, it was much
drier than normal (0%–10%) across southern areas,
and also (50% of normal) in central regions. Overall,
Figure 6.27. Annual average (left)
temperature deciles and (right) precipitation
anomalies (mm; 1971–2000 base) for
southwest Asia
77
Figure 6.28. European average temperature anomalies (°C; relative
to 1961–1990 mean) 1850–2005; blue bars show the annual values
with uncertainties represented by the black bars. Red curves show
the annual anomalies and uncertainties after smoothing with a
21-term binomial filter. [Source: Brohan et al. 2006]
spring precipitation in Iran was just 42% of normal.
Summer precipitation in Iran was 20% below normal,
with especially dry conditions in the west half of the
country. However, heavy rains in the east did result in
flooding. Below-normal precipitation continued into
the autumn across much of Iran. Autumn precipitation
in Iran was 37% below the long-term mean, and some
parts in the east received just 0%–25% of the normal
seasonal precipitation due to the delayed onset of
late-season precipitation.
G.
EUROPE
I)
OVERVIEW J. J. KENNEDY38
The annual surface temperature anomaly (Brohan
et al. 2006) averaged over Europe in 2005 was 0.71
±0.07°C above the 1961–1990 average (Figure 6.28).
Only a small area extending north from Greece had
annual temperatures below average (Figure 6.29),
and that was only by around 0.1°C. Annual average
temperatures in the United Kingdom and northern
Norway and Finland were above the 90th percentile of
occurrence according to statistics based on the period
1961–1990 (all European temperature and precipitation
percentiles herein refer to this period).
Temperatures during the first three months of 2005
were significantly (meaning in the upper or lower
decile of the distribution) below average in southern
Europe, through Spain and the Mediterranean and into
Italy. In the same period, above-average temperatures
observed in the north and east exceeded the 90th
percentile only over Scotland. Between April and
June, temperatures were above average in all areas
and significantly above normal over much of Europe
west of 15°E and south of 55°N. Temperatures in Spain
and France exceeded the 98th percentile. From July
to September, temperatures once again were above
average in most areas, although temperatures were
close to average in southeastern Europe. Scandinavia
and Eastern Europe were significantly above normal.
78
Cooler conditions in southeastern and central Europe
coincided with the largest regional rainfall totals for the
season. October–December brought a north–south split,
with much of the Mediterranean and southern Europe
experiencing below-average temperatures, while in
the north temperatures were generally above average
with areas of the United Kingdom and Scandinavia
significantly above average.
Total precipitation (Rudolf et al. 1994, 2005; Rudolf and
Schneider 2005; Beck et al. 2005) between January
and November 2005 (Figure 6.29) was below average
in southwestern Europe, with parts of France and
the Iberian Peninsula receiving less than 40% of the
11-month 1961–1990 average. Precipitation in southeastern Europe was above the 1961–1990 average.
Romania and Bulgaria received signifi cant rainfall
excesses during the year, with August totals in some
areas approaching 500% of the monthly average.
II)
CENTRAL AND EASTERN EUROPE J. J. KENNEDY38
Annual average temperatures in the region ranged
from near average in Hungary and neighboring countries to over 2°C above average in eastern Ukraine
(Figure 6.29). January–November rainfall was significantly above average in Romania (Figure 6.29).
A warm January, with areas of eastern Ukraine more
than 5°C above average, gave way to colder conditions in February and March. Precipitation was
generally below average in the southwest, but exceeded
the average further north and east, with the largest
excesses occurring in January and February. March
rainfall was below average in most areas.
April–June temperatures in parts of Austria and the
Czech Republic were significantly above average, and in
Switzerland some western areas experienced temperatures above the 98th percentile. Further east, however,
temperatures were nearer the average. Rainfall anomalies were generally higher in the east than the west,
with the highest anomalies in Romania and Moldova.
In June, temperatures were significantly above normal
in Austria and Switzerland and, in the far west, were
high enough to exceed the 98th percentile. At the
same time, temperatures in Ukraine and Romania fell
below average. With only eastern Ukraine experiencing
below-average rainfall, April precipitation was above
average in most areas and the excess rainfall led to
flooding in Romania. May precipitation was close to
average in many areas, but Romania again experienced
above-average totals. Most areas were drier than
usual in June, with only eastern Romania and Ukraine
experiencing wetter-than-normal conditions.
Temperatures were above average between July and
September everywhere. The largest anomalies were in
the east where temperatures were significantly above
average. Smaller positive temperature anomalies
further to the west coincided with the largest rainfall
anomalies. Some areas of southern Romania received
more than three times the seasonal average, generating widespread flooding in July and August. July
temperatures were above average in all areas, but
above-average precipitation contributed to depressed
temperature anomalies in central Europe. Only Ukraine,
where precipitation levels were below normal, showed
significant warmth. August brought above-average
rainfall to all but the most easterly areas, and temperatures fell below average in the west. Ukraine showed
the highest temperature anomalies, with temperatures
significantly above normal in parts.
October and November were mainly drier than usual
in western areas, with wetter-than-average conditions confined to eastern Romania and Ukraine. Cold
anomalies spread from the southeast in October to
cover the central and western states in November.
In December, cold anomalies were confined to the
westernmost regions, with some parts of Austria significantly below normal.
III)
FENNOSCANDINAVIA, ICELAND, AND GREENLAND
C. ACHBERGER1 AND D. CHEN16
The climate of Fennoscandinavia (here Norway, Sweden,
Finland and Denmark) is to a large extent controlled by
the atmospheric circulation over the European and North
Atlantic region. Objective synoptic classification based
on the Lamb scheme (Lamb 1950) has been extensively
used in Sweden to quantify impacts of atmospheric circulation (e.g., Chen 2000). In 2005, occurrences of the four
dominant weather types—anticyclonic (A), cyclonic (C),
westerly (W), and southwesterly (SW)—were all above
the 1961–1990 mean. The generally warmer than normal
conditions over Fennoscandinavia in 2005 correspond
closely to increased frequencies of W and SW types.
(i)
Temperature
Annual mean temperatures for 2005 in the Nordic
countries (including Greenland and Iceland) were
around 0.5° to 4.5°C above the 1961–1990 mean
(Figure 6.30), depending on geographical location.
At Svalbard on Spitsbergen, annual mean temperatures
reached -3.0°C, which is 3.6°C above normal, ranking
2005 as the warmest year since 1912. Danmarkshavn,
in northeastern Greenland, reported an annual mean
temperature of -9.5°C, placing 2005 as the warmest
year since reliable measurements started in 1949.
Parts of Finland were also unusually warm, at around
2.3°C above the long-term mean. Sweden, Denmark,
and Norway, however, experienced more moderate
temperature deviations, ranging between 1.1° and 1.7°C
above normal, while the annual mean temperature for
Iceland was about 1°C warmer than the 1961–1990 mean.
Figure 6.29. European 2005 annual (left)
temperature anomalies (°C; 1971–2000 base),
and (right) precipitation anomalies
(mm; 1979–2000 base) from CAMS–OPI
79
The warmer-than-normal temperatures were most
pronounced at the highest latitudes.
Autumn (September–November) was extraordinarily warm at many Nordic locations, several of
which broke existing records. For example, on
11 October 2005, several Norwegian stations reported
daytime temperatures well above 20°C. Of these,
Molde Airport recorded the highest temperature
(25.6°C), which is a new Norwegian record for
October. Finland also experienced two unusually
warm spells in autumn, making November 2005 the
warmest November since 1900 in the central regions.
For Denmark, 2005 was the fourth sunniest year on
record and Reykjavik, Iceland, experienced 280 more
sunshine hours compared to the 1961–1990 mean.
In all, 2005 is ranked as the fi fth to eight warmest
year since the second half of the eighteenth century,
when regular measurements commenced in many
Nordic countries.
(ii)
Precipitation
Annual precipitation amounts across the region
were both above and below the 1961–1990 average
(Figure 6.31). Across much of Finland, northern
regions of Norway and Sweden, the Faeroese Islands,
and along the southern half of the Norwegian west
coast, annual precipitation was above average
(Figure 6.31, right), where positive departures reached
up to 40% of the long-term mean. Greenland also
experienced remarkable precipitation deviations from
the long-term mean (Figure 6.31, left). However, while
large regions of southernmost, western and northern
Greenland received from 50% to over 100% more
precipitation than the 1961–1990 mean, parts of eastern
and southern Greenland were considerably drier than
normal (Figure 6.31, left).
Southern Sweden, Denmark and Iceland received
annual precipitation amounts either somewhat below
or close to the long-term average. Many locations
across Fennoscandinavia, though, reported record or
near-record precipitation in 2005: Nuuk, Greenland,
received 1219 mm, ranking 2005 as the wettest year
there since 1958. For Norway, 2005 was among the
second to third wettest year since measurements
began. Parts of Finland experienced a very wet May,
with precipitation 200%–300% above normal. Severe
flooding occurred along several rivers in northern
Finland as a result of snowmelt. Also, Sweden experienced heavy rainfall in July and August.
80
(iii) Notable events
For much of the region, 2005 started with an extremely
severe storm. On 8–9 January, an intense low pressure
system formed west of the British Isles and intensified on its way toward Fennoscandinavia. Mean sea
level pressure dropped as low as 960 hPa when the
system, named Gudrun, reached Sweden, causing
major damage. Mean wind speeds and maximum
gusts reached 33 and 42 m s-1, respectively. In addition to widespread power failure and infrastructure
damage, Gudrun was responsible for several fatalities,
and became the worst tree-felling storm in Sweden
since 1930, when statistics on storm-related forest
damage started. In addition, record sea levels and
coastal flooding occurred along parts of the west coast
of Sweden, Norway and Finland. Swedish insurance
companies rank Gudrun as the costliest natural event
in their history.
Remnants of Hurricane Katrina reached Greenland as an
extratropical low pressure system in early September.
The low brought several thunderstorms and frequent
lightning to southwest Greenland.
On 14 November, an exceptionally strong low pressure
system named Loke reached Norway and brought over
200 mm of precipitation over 24 hours to several locations. In addition to one fatality, the storm generated
landslides across Norway, closing many roads and the
train line between Oslo and Bergen. Loke produced the
country’s second highest daily precipitation amount
(223 mm) on record, measured near Bergen.
IV)
CENTRAL NORTHERN EUROPE J. J. KENNEDY38
Annually averaged temperatures in 2005 were
above average throughout the region (Figure 6.29).
Precipitation totals for January–November were near
average in the west and east, but somewhat drier than
average in Poland (Figure 6.29).
January temperatures and precipitation were above
average everywhere, with the largest temperature
anomalies (more than 5.3°C) in the east. February
temperatures fell to below average everywhere except
the far northeast. These northeastern areas were also
drier than normal, but most areas experienced aboveaverage rainfall in February. March temperatures were
more than 3°C below average in Estonia and Latvia,
and only western Germany experienced warmer-thanaverage temperatures. Rainfall totals in March were
Figure 6.30. Annual temperature anomalies (°C) across
Fennoscandinavia, the North Atlantic, and Greenland in 2005
(1961–1990 base) [Source: NCAR–NCEP Reanalysis]
anomalously large in Belarus, Lithuania and eastern
Poland, but lower than average in other areas.
April–June temperatures in some areas of Poland
and Germany were significantly above average.
Temperatures were above average everywhere in April,
with significant warmth (> 2°C) in parts of Germany
and Poland. May anomalies were generally lower
than those in April, with below-average temperatures
in the east. After a dry April, May rainfall was above
average. Precipitation was particularly heavy in the
east with many areas receiving more than twice the
usual monthly amount. One area in Belarus received
nearly three times its average monthly total. In June,
temperature anomalies fell again in the east, with
temperatures dropping below average everywhere
east of central Poland. Temperature anomalies in
southwestern Germany, however, rose above 2.8°C
(98th percentile).
July–September was another warm period. Temperature
anomalies were highest in the east, exceeding the 90th
percentile in large parts of Poland, Belarus, and Estonia.
Precipitation was below average except in Germany.
July saw above-average temperatures in all areas,
although significant positive anomalies were found
in only a few areas of Poland, Belarus and Lithuania.
Above average rainfall in Germany and Poland may
have acted to reduce temperature anomalies in those
regions. Below-average August temperatures occurred
in Germany and much of Poland, and wetter-thanaverage conditions existed in the east. September
was drier and warmer than average in most areas,
with Belarus, northern Germany and western Poland
experiencing significant warmth. Only central parts of
Germany received above-average rainfall.
Figure 6.31. Annual precipitation departures from normal (mm yr -1)
across (left) Greenland and (right) Fennoscandinavia (1961–1990
base) [Source: Global Precipitation Climatology Center (GPCC)]
Temperatures between October and December were
above average and precipitation totals were below
average in all regions. October temperature anomalies
were highest in Germany, where temperatures were
significantly above normal over much of the country,
and even rose above the 98th percentile in the west
(+2.5°C anomaly). Some areas of Poland received
less than 20% of the average October precipitation. In November, southern Germany and Poland
had below-average temperatures, but anomalies
increased to the north. November was another dry
month, with no area experiencing above-average
rainfall. In December, temperatures were above
average in most areas, although southern Germany
had temperatures that were more than 1.6°C below
the average.
V)
NORTHWESTERN EUROPE J. J. KENNEDY38
Annually averaged temperatures in northwestern
Europe in 2005 were above average in all areas. The
United Kingdom and Netherlands experienced significantly above-normal temperatures, and anomalies in
Scotland exceeded the 98th percentile for the year.
France and the south of the United Kingdom were drier
than average between January and November 2005,
and some areas of France received less than 60% of
the average rainfall.
January saw significantly above-average temperatures in the United Kingdom. Most areas were dry,
but the north of the United Kingdom received aboveaverage rainfall, leading to flooding in a number of
areas. Temperatures fell below average in most areas
in February. In the south of France, both temperatures
and rainfall were significantly below average, with
81
some areas receiving less than 20% of normal precipitation. Drier-than-average conditions continued in
March across France and the United Kingdom. March
temperatures were above average in most areas, and
significantly so in Scotland and Ireland.
April–June average temperatures were significantly
above normal everywhere except Scotland and Ireland.
Temperatures in France exceeded the 98th percentile.
April was warmer than average in all regions, but
significant warmth was confined to the Low Countries,
southeast England and western Scotland. Precipitation
was above average in most areas. Temperatures in
May were below average only in Ireland and Scotland,
while the south of France experienced significantly high
temperatures. Rainfall was below average in England,
Wales and France, but above average in Scotland
and Ireland. Signifi cantly high June temperatures
covered all areas except central England, where subzero
temperatures were experienced near the start of the
month and the central England minimum temperature
index (Parker and Horton 2005) dropped below the 5th
percentile of occurrence (3.6°C on 7 June). Contrastingly,
temperatures in the southeast of France exceeded
the 98th percentile. Precipitation totals were below
average over continental regions.
July was warmer than average in all areas, with the
highest anomalies over continental areas. Rainfall
deficits were observed in the south of France and
Scotland, but other areas experienced above-average
rainfall. August was warmer and drier than usual in
the west, but cooler and wetter than normal in the
east. Anomalies increased in September with aboveaverage temperatures in all areas. Southern France had
the lowest temperature anomalies and highest rainfall
anomalies, but other areas were predominantly dry.
Figure 6.32. (left) Winter (DJF) and (right) summer
(JJA) anomalies of 500-hPa geopotential height
(contours; gpm) and corresponding 850-hPa
temperature anomaly field (colour; °C) over southwest
Europe and North Africa [Source: NCAR NCEP
reanalyses; 1961–1990 base]
Figure 6.33. Daily Tmax (black) and Tmin (dark blue)
temperature (°C) in Lisbon during 2005; red (green) and
orange (light blue) lines correspond to the 90th and
10th percentiles, respectively, of the Tmax (Tmin) for
each day of the year (10-day moving window),
and were computed using the period 1941–2000.
82
In October, temperatures in central areas exceeded
the 98th percentile. On 12 October, the highest ever
October minimum central England temperature
(15.2°C) was recorded. October precipitation was
above average in the United Kingdom and northwest
France but below average in other areas. November
temperatures in France and England were below
average, and it was drier than average over most of
the region. Temperatures in December were below
average in France, particularly in the south of the
country, where it was signifi cantly below normal
(< -2°C anomaly).
VI)
IBERIA R. TRIGO,86 R. GARCIA-HERRERA,24 AND
D. PAREDES64
(i)
Temperature
The average 850-hPa temperature across Iberia in
2005 was 0.3°C above normal (1961–1990 base period
mean). However, this relatively moderate annual
anomaly conceals considerably larger cold and warm
temperature anomalies observed in winter (-1.6°C)
and summer (+1.7°C).
Wintertime upper-level 500-hPa geopotential height
anomaly fields confirm the overall structure observed at
surface, with intense positive anomalies centred between
the Azores and Iceland (Figure 6.32, left). As a consequence
of this strong northwest–southeast geopotential height
gradient, Western Europe was under the influence of strong
cold air advection from higher latitudes. Several consecutive cold wave outbreaks were observed between the end
of January and early March affecting all of Iberia as well
as France and central Europe (Garcia-Herrera et al. 2006,
manuscript submitted to J. Hydrometeor). Daily maximum
and minimum temperatures for Lisbon, Portugal, reveal
that these cold waves reached the western coast of Iberia,
producing a number of days with temperatures in the
coldest decile (lowest 10%) of daily long-term records
(Figure 6.33). Interestingly, soon after the last cold wave,
a circulation shift to North African air penetration resulted
in late-March/early April daily temperatures within the
highest decile (hottest 10%) of long-term temperature
records for those days (Figure 6.33).
Spring and summer months were characterized by
warmer-than-normal temperature values, particularly
between late May and August, with most of Iberia
impacted by two intense heatwaves in June and August
(Figure 6.33). Positive SLP anomalies over Europe, due
to the extended Azores anticyclone, contributed to
the anomalous warmth in these months. The summer
THE EXTREME IBERIAN DROUGHT OF 2004/2005 R. TRIGO,86 R. GARCIA-HERRERA,24 AND D. PAREDES64
The year 2005 was characterized by one
of the worst droughts ever recorded in
the Iberian Peninsula, particularly in its
southern half. The hydrological year that
spans October 2004–September 2005 was
the driest on record for several locations
throughout Iberia, namely in the capital
cities of Lisbon, Portugal, and Madrid,
Spain, where reliable precipitation records
have been kept since 1865 and 1859,
respectively (Figure 6.34). In particular, in
Lisbon the 2004/2005 event surpassed the
previous record drought of the 1944/1945
hydrological year (Figure 6.35).
The precipitation regime over Iberia is
characterized by strong seasonal behavior,
with a unimodal rainy season concentrated
between October and March and relatively
arid conditions at other times. Therefore, all
major droughts in this region are characterized by lack of rainfall during several months
of the winter half of the year (Trigo et al.
2004). Using data from the GPCC (Rudolf
and Schneider 2005), spatially averaged precipitation over Iberia between October 2004
and September 2005 was roughly 45% less
than the 1961–1990 climatological average
(Garcia-Herrera et al. 2006, manuscript
submitted to J. Hydrometeor). However,
Figure 6.34. Percentages of normal (1961–1990 base)
precipitation accumulated between October 2004
and September 2005 [Source: GPCC] Lisbon (A)
and Madrid (B) are shown.
regionally, the drought was most intense
in the southern and southwestern sectors
of the Iberian Peninsula, where rainfall was
as little as 40% of normal during that period
(Figure 6.34).
The 2004/2005 drought had major socioeconomic impacts in both Iberian countries,
particularly for hydroelectricity and cereals
production, which decreased to 40% and
60%, respectively, of their long-term average values. Agricultural losses due to scarce
water resources and diminished pasture
land during this season have been estimated
at some US$ 2 billion.
Figure 6.35. Annual precipitation anomaly (mm) for Lisbon
from 1865 to 2005 (1865–2005 mean); anomalies correspond to
the hydrological year that spans between October of year n-1
and September of year n
83
500-hPa geopotential height anomaly field was dominated by a positive anomaly maximum centred
between Iberia and the United Kingdom (Figure 6.32,
right). This large-scale feature not only impeded
the natural eastward progression of low pressure
systems that frequently cross northern Europe in
the summer but also contributed to the advection of
warm air masses in the south as well as enhanced
adiabatic heating through subsidence (Trigo et al.
2004). Furthermore, extended periods of clear skies
associated with anticyclonic conditions contributed
to increased solar radiative heating over the region.
(ii)
Precipitation
The Iberian Peninsula experienced drier-than-normal
conditions during 2005 relative to the 1961–1990 base
period mean (Figure 6.29). The hydrological year
of 2004/2005 (October 2004–September 2005) was
among the driest since regular precipitation records
started in both Portugal and Spain (see sidebar).
In fact, based on the monthly Global Precipitation
Climatology Center (GPCC)-gridded data set (Rudolf
and Schneider 2005), drought conditions prevailed in
2005 over a large area of Western Europe, including
the southern United Kingdom, France, and Northern
Italy (Figure 6.29). During winter and spring, usually
the wettest seasons, dry conditions prevailed over
much of Iberia, with precipitation less than 50%
(and in places < 25%) of 1961–1990 mean values.
Winter months were characterized by the presence
of intense anticyclonic circulation, located northward
of its usual latitude (Azores). Drought conditions
continued through the spring months, although with
less intensity than in winter. However, unlike winter,
several storm tracks progressed from the Atlantic toward
Iberia and France, but were stopped from penetrating
the European continent by the development of extensive
blocking events.
The summer season of 2005 was also characterized
by reduced precipitation over Western Europe and
northern Africa, with the maximum amplitude over
Iberia. Precipitation totals rebounded across the northeastern sector of the Iberian Peninsula during autumn
2005, but normal precipitation fell in the remaining
areas, particularly in the southern sector.
Major climatic anomalies are of ten driven by
enhanced values of large-scale atmospheric circulation indices [e.g., ENSO, PNA, NAO, East Atlantic
(EA)]. Interestingly, this record-breaking drought
84
is only partially associated with extreme values of
teleconnection indices, particularly those that are
known to have a significant impact upon winter
Iberian precipitation (NAO and EA). Moderately
positive (negative) values of NAO (EA) recorded
between November and February are clearly associated with the scarce precipitation observed over
Iberia. However, the similarly dry conditions observed
in March are more diffi cult to associate with the
intense negative NAO (-1.8) and moderately positive
EA (+1.1) indices. Instead, an extremely intense blocking event positioned over unusually southern latitudes
contributed to the widespread March precipitation
anomaly over Europe.
VII) MEDITERRANEAN AND SOUTHERN EUROPE
J. J. KENNEDY38
Temperatures in the Mediterranean were above
average in most areas in 2005. Greece was one of the
few countries that experienced below-average
temperatures for the year as a whole. Precipitation
was generally below average in the west but above
average further east.
January–March temperatures were below average in
all regions and significantly below average in Italy.
In the western Mediterranean, temperatures fell
below the 2nd percentile of occurrence. The north
of Italy and the western Mediterranean were drier
than average, while southern Italy and Greece were
wetter. January temperatures were above average
only in Greece, and in the western Mediterranean
temperatures were below the 2nd percentile. Cold
conditions continued into February and the whole
of the Mediterranean area west of 25°E experienced
significant cold. The south and west were the coolest areas, with temperature anomalies of -3.7°C in
the Balearic Islands falling below the 2nd percentile. January was drier than average in the west,
but above-average January precipitation in Greece
and Italy spread westward in February, when only
northern Italy experienced below-average rainfall.
March was much drier, with most areas experiencing
below-average rainfall.
In April–June, average temperatures were signifi cantly above normal in all areas except Greece. West
of Sardinia, temperatures exceeded the 98th percentile. Precipitation was below average in most regions,
with the exception of southern Italy and Sicily. April
temperatures were near average in most regions.
May was warmer with large areas in the west showing signifi cant warmth. Heating continued through
June, with significantly above-normal temperatures
in all areas except Greece. Parts of northern Italy
exceeded the 98th percentile. Anomalous rainfall
occurred in southern Italy and Sicily, where totals
were above average in all three months.
Significant heat continued into July over large areas
of the region. Above-normal precipitation fell over
Greece, with eastern areas receiving more than twice
the monthly average. Temperature anomalies were
somewhat lower in August, dropping below average
in northeastern areas, but high temperatures in the
western Mediterranean brought large numbers of
jellyfish to Spanish beaches. Most areas saw aboveaverage rainfall, and northern Greece again received
signifi cant excesses of precipitation. Temperature
anomalies in the west fell in September, and most
areas experienced above-average rainfall.
Average October–December temperatures were below
average in most areas. October temperatures were
below average in the east and above average in the
west. In November, cold anomalies spread over much
of the Mediterranean, and by December, temperatures
had dropped significantly below normal in central
areas. In December, Greece was the only country in
the region to experience above-average temperatures.
October was mainly a dry month, with northern Italy
the only place to receive above-average precipitation. November in contrast was wetter than average
in most areas.
VIII) SOUTHEASTERN EUROPE J. J. KENNEDY38
Temperatures in southeastern Europe in 2005 were
within 0.2°C of the 1961–1990 average. Eastern areas,
particularly Bulgaria, received significantly aboveaverage precipitation totals for the year, causing
flooding in many areas.
The January–March period was colder than average
in all but eastern Bulgaria. January was generally
warmer than average but February saw significant
cold in most areas, with temperatures in Serbia more
2
than 3°C below average. Temperatures in March
were also below average, but not significantly so.
The greatest precipitation anomalies occurred in
Bulgaria in all three months, leading to fl ooding
there in February.
April saw the end of the anomalously cold weather,
with above-average temperatures in all areas. May
and June temperatures were above average in most
areas, significantly warmer than normal in Croatia and
Bosnia-Herzegovina, but below average for June in
Bulgaria. Rainfall totals in April and June were close
to average in most parts, but in Bulgaria significant
May rainfall once again brought flooding.
Although July temperature anomalies in southeastern Europe were positive, they were lower than in
most areas of Europe. Even so, temperatures along
the Adriatic coast were significantly above normal.
Temperatures in August were below average except in
eastern Bulgaria, which experienced significant warmth.
September was warmer than average. July, August, and
September rainfall totals were far above average, with
Bulgaria receiving more than twice the seasonal average, and large areas recording more than three times
the average. In August, 500% above-normal monthly
rainfall in Bulgaria produced flooding that continued
into September and left more than 30 dead.
Excessive rains continued in the east into October, but
November totals were close to average in all areas.
October and November temperatures were below
average in the south and east, but December saw the
colder weather shifting to the west. Temperatures in
the north were above average in October, but fell below
average for the last two months of the year.
H.
OCEANIA
I)
AUSTRALIA A. B. WATKINS90
Despite the notable absence of an active basin-wide
El Niño event, 2005 was the hottest year on meteorological record for Australia.2 Neutral to slightly warm
conditions in the equatorial Indian and Pacific Oceans at
For Australia-wide, as well as large-scale regional averages, high-quality monthly temperature data is available from 1950, with high-quality annual temperature
data starting 1910. For rainfall, high-quality area-averaged data commences in 1900. All records and percentile values are calculated with respect to these years.
Anomalies are calculated with respect to the 1961 to 1990 average, in accordance with World Meteorological Organization guidelines (WMO Publication No. 100:
www.wmo.ch/web/wcp/ccl/ GuideHome/html/wmo100.html).
85
Figure 6.36. Australian maximum April temperature (Tmax)
anomalies (°C) for 2005 (1961–1990 base)
the start of 2005 persisted until June, returning to near
normal in the latter half of the year. Correspondingly,
pressure over Australia was higher than normal during
the first half of the year, and from normal to below
normal during the remainder of 2005. The anomalously
high pressure over the country during early 2005
greatly reduced rainfall over the interior and inhibited
the northward penetration of frontal systems from the
Southern Ocean. High pressure also contributed to a
sporadic Australian monsoon, which failed to extend
far inland, resulting in anomalously warm and dry
conditions in the north.
(i)
Temperature
Due in part to the inconsistent Australian monsoon,
the tropical wet season (October 2004–April 2005)
was relatively warm. Northern Australia experienced
a +1.5°C maximum temperature (Tmax) anomaly
(0.6°C above the previous wet season record).
Additionally, April, ordinarily the start of the main winter
cropping season, was climatologically one of the most
remarkable months on record for Australia. Australiawide, April mean Tmax was 3.11°C above normal
(Figure 6.36). Not only was this 0.7°C above the previous
April record, but it was the largest anomaly recorded
for any month since Australia-wide temperature records
began in 1950, which is substantially higher than
the previous record (+2.68°C) set in October 1988.
Combined with record high minimum temperatures
(Tmin), the April mean temperature anomaly of
+2.58°C was 0.85°C above the previous April record
set in 2002 and 0.26°C above the previous record for
any month (June 1996). Notably, the April mean
temperature was the highest on record over
66% of the continent, with 86% of the continent
experiencing mean tempertures for the month in
the highest 10% the of recorded totals.
June–December Tmax and Tmin remained above
average across virtually the entire country (+0.75° and
+0.94°C anomalies, respectively). When combined
with the hot start to the year, temperatures for 2005
were exceptional. The Australia-wide Tmax anomaly
for 2005 was +1.21°C, equal to the record set in 2002,
while the Australia-wide Tmin anomaly of +0.91°C was
the third warmest on record behind 1998 (+1.12°C).
Overall, the mean temperature anomaly for 2005 of
+1.06°C was 0.23°C above the previous hottest year
(1998). Consequently, Australia experienced its hottest mean temperature since annual mean records
commenced in 1910. In total, 95% of the continent
Figure 6.37. Australian mean (left) temperature and (right) precipitation accumulation deciles for 2005 (relative to 1950–2005 for
temperature and 1900–2005 for precipitation) [Source: Australian BOM]
86
experienced above-average mean temperatures
during 2005 (Figure 6.37). With the absence of a decaying El Niño event, the record Australian heat of 2005
clearly highlights the impact of the long-term warming
trend of the global and Australian climate upon the
natural variability of year-to-year fluctuations.
(ii)
Precipitation
High pressure over the continent and the weak
Australian monsoon contributed to extremely dry
conditions during the first five months of the year
(Watkins 2005), with 44% of the country experiencing rainfall in the lowest 10% of recorded totals
(decile 1). Australia-wide, April was the eighth driest
April on record, with only 10.7 mm for the month (average of 31.1 mm). The Australia-wide average rainfall
of 168 mm was the second lowest January–May total
(after 1965) since Australia-wide monthly records
began in 1900. While short-term dry spells are not
unusual for Australia, it hindered the limited recovery from the devastating 2002/2003 El Niño–related
Australian drought (Coughlan et al. 2003), one of the
worst droughts in Australia’s recorded meteorological
history (Nicholls 2004).
In a remarkable turnaround, 80% of the country experienced above-average rainfall between June and
December, which corresponded to a change in ocean
conditions in the equatorial Pacific. June–December
precipitation over many previously drought-affected
parts of New South Wales was in the top 10% (decile
10) of recorded totals. Australia-wide, 23% of the
country experienced decile-10 precipitation between
June and December, compared to only 0.6% for the
January–May period.
Despite the average to above-average rainfall totals in
the second half of the year, the extremely dry conditions
during the first five months contributed to a generally
below-average rainfall year for Australia (Figure 6.37).
The 2005 Australia-wide average rainfall of 407.2 mm
was the 33rd driest such period since all-Australian
records commenced in 1910, which is 65 mm below the
1961–1990 mean of 472 mm. Overall, 63% of Australia
experienced below-average rainfall during 2005.
(iii) Notable events
In January, Nyang Station, in the Gascoyne region
of inland Western Australia, measured Australia’s
hottest month on record, when its average maximum
Figure 6.38. 2005 mean sea level pressure anomaly map
for the New Zealand region showing departures from
average (hPa); anticyclones were more frequent than
normal east of the South Island and in the Tasman Sea.
temperature of 44.8°C equaled the previous record,
also set at Nyang, from February 1998.
An intense low pressure system developed over Eastern
Bass Strait on 2 February, resulting in substantial
rainfall and low temperatures for Victoria, southern
New South Wales, South Australia and Tasmania. The
event made February 2005 Victoria’s wettest February
since 1973. Despite the 2-day event supplying 22%
of Melbourne’s annual mean rainfall (638.8 mm), the
city’s 2005 total precipitation of 589.8 mm was below
average for the ninth year in succession.
Tropical Cyclone Ingrid, which occurred between 5 and
16 March (see section 4c), reached category 5 (Australian
scale; information online at www.bom.gov.au/catalogue/
warnings/WarningsInformation_TC_Ed.shtml) on at least
two occasions, and is the only TC in Australia’s recorded
history to impact three different states or territories
(Queensland, Northern Territory, Western Australia) as
a severe tropical cyclone (category 3 or above).
Despite the generally mild winter, three major low-elevation snow events occurred during the season. These
affected the Northern Tablelands of New South Wales
and adjoining southern Queensland (22–23 June), the
Monaro district of New South Wales (8–9 July), and
southern Victoria and Tasmania (10 August). The August
event brought snow to sea level in Victoria for the first
time since 9 August 1951.
A notable heatwave affected large parts of central
and eastern Australia during late December 2005 and
early January 2006. The most abnormal conditions
87
Figure 6.39. South Pacific 2005 annual (left) temperature anomalies (°C; 1971–2000 base) and (right) precipitation anomalies
(mm; 1979–2000 base) from CAMS–OPI
occurred in the period from 30 December 2005 to
1 January 2006, when northwesterly winds brought
extreme heat to southeastern Australia. Arguably, the
most exceptional record occurred at Montague Island
(New South Wales), where a reading of 41.0°C broke
its previous all-time record by 3.8°C. Sydney reached
44.2°C on 1 January 2006, second only to the 45.3°C
reached there on 14 January 1939.
II)
NEW ZEALAND M. J. SALINGER80
New Zealand’s climate of 2005 was influenced by
more frequent anticyclonic activity in the Tasman Sea
and to the east of the South Island, resulting in less
wind, warmer temperatures and generally decreased
precipitation for much of New Zealand. However,
more cyclonic activity was present in February, March,
May, and December, and the northeast of the North
Island experienced more frequent easterlies at times
(Figure 6.38). Notable climate features in various parts
of the country included heatwaves, low soil moisture,
localized flooding, the Greymouth tornado, an unseasonable snowstorm and damaging hailstorms.
(i)
Temperature
The national average temperature in 2005 was 13.1°C,
0.5°C above the 1971–2000 normal. It was the fourth
warmest year nationally since reliable records commenced in the 1860s. Only 1971, 1998, and 1999 have
been warmer, with temperatures of 13.2°, 13.3°, and
13.3°C, respectively. For New Zealand as a whole,
there were seven warmer-than-normal months
88
(February, March, May, July through September,
and December), two cooler months (January and
April), and three months with mean temperatures
close to the climatological average (June, October,
and November).
A combination of anticyclones and northeasterlies
brought one of the warmest Februaries on record,
with maximum temperatures of 30°C or more in many
locations throughout New Zealand, and temperatures
of 35°C or more in sheltered inland areas of the South
Island during the first 10 days. The highest recorded
extreme air temperature for the year was 38.7°C at
Alexandra on 5 February (the highest temperature
there for any month, in records back to 1929). Overall,
February was the eighth warmest on record, with a
mean temperature of 18.6°C (+1.3°C anomaly).
Halfway through the year, more frequent anticyclones
over the North Island and northwesterlies over the
South Island produced the sixth warmest winter
(June–August) on record, even though June was
the coldest in a decade. With a mean temperature of
9.1°C (+1.2°C), July was the third warmest on record.
Record high August maximum temperatures were
recorded at Hanmer Forest (25.1°C on the 30th) and
Amberley (25.4°C on the 31st). August’s 9.8°C (+1.1°C)
made it the fourth warmest on record. Mild conditions accompanied continued anticyclonic activity
into spring (September–November). A changing
atmospheric pattern to warm northerlies produced
the third warmest December on record, with 17.5°C
(+1.9°C).
Mean temperatures in 2005 were at least 0.3°C above
average in most regions and 0.5°–0.9°C above average in parts of Auckland, Coromandel, western Bay
of Plenty and western North Island from Wanganui
to Wellington, as well as Wairarapa and much of
the South Island. Temperatures were near average
in coastal Wairarapa, along the Kaikoura Coast and
in coastal areas of south Canterbury. The warmest
locales were Cape Reinga and Whangarei Airport,
both with a mean temperature for the year of 16.1°C
(0.2° and 0.5 °C above normal, respectively). Several
locations observed record warmest annual average
temperature in 2005.
(ii)
Precipitation
New Zealand’s climate for 2005 was marked by
too little rain in some places and too much in
others. Annual rainfall during the year was less than
75% of normal over much of the South Island and
75%–90% of normal in the north and west of the
North Island (excluding Wanganui) and southern
Wairarapa (Figure 6.39). Clyde recorded the least
annual precipitation at 348 mm (76% of normal).
Near-record to record low precipitation was observed
at numerous locations. Conversely, well-aboveaverage precipitation (> 125%) fell in the western
Bay of Plenty, Hawke’s Bay and the far southwest
of the South Island. Precipitation was near normal
elsewhere. The wettest location was Cropp River in
Westland, with an annual total of 9290 mm.
to Nelson and the southwest of the North Island in
November, and developed in Hawke’s Bay, Auckland,
and parts of Northland in December. However, southeasterlies in both October and November produced
significant flooding in Gisborne.
(iii) Notable events
For the year, there were at least 26 heavy rainfall
events, half of which produced floods. There were
also 7 damaging hailstorms and 12 damaging tornadoes (or events attributed to tornadoes) in 2005.
The Greymouth tornado of 10 March was particularly destructive, leaving 30 people homeless and
resulting in damage worth at least NZ$ 10 million
(US$ 6.3 million). Wellington Airport was closed for
many more hours than usual in 2005. There were
52 hours with fog there, the highest for any year in
45 years of measurement.
The Bay of Plenty floods of 3–4 May and 17–18 May
were most disastrous, with the earlier of the two
causing widespread damage in parts of Tauranga,
and the later being phenomenal, with unprecedented
high rainfall for the district and a state of emergency
declaration from Tauranga to Matata. Hundreds of
people were evacuated. Homes were destroyed by
mudslides and flooding while rising waters threatened
hundreds of others, especially in Matata.
The extremely high temperatures during the first
10 days of February are notable because there are
Anticyclones in January commenced the trend of
low rainfall and severe or signifi cant soil moisture
deficits in the northern half of the North Island and
Canterbury, with these conditions persisting into
April. March was wetter in the North Island, but more
anticyclones in April kept conditions dry. However,
weather patterns changed abruptly in May, resulting
in widespread flooding in the Bay of Plenty.
Frequent winter anticyclones over the North Island
and northwesterlies over the South Island produced
extremely dry conditions in the east of the South
Island between June and September. Winter snowfall
was much less frequent than normal. However, an
early spring snow (19 September) down to sea level
in Canterbury was unusual for the month.
Below-average spring rainfall resulted in significant
soil moisture deficits developing much earlier than
usual from Southland to Marlborough. Deficits spread
Figure 6.40. South Pacific annual SST anomalies (relative
to 1971–2000 mean; °C); yellow or orange areas represent
above-average temperatures.
89
very few instances prior to this event where temperatures anywhere in New Zealand have exceeded
38°C (100°F).
In Canterbury, the Christchurch airport, schools, and
universities were closed due to an unusual early spring
snow (19 September) in Canterbury. Snow depths of
5–10 cm were recorded in the region.
III)
SOUTH PACIFIC ISLANDS M. J. SALINGER80 AND
S. M. BURGESS12
A high frequency of surface equatorial westerlies
occurred near the date line in February (the most
since the last El Niño in 2002). Trade winds generally were near normal in strength at other times of
the year. There was also some ENSO influence on
the location of the SPCZ during the year. West of
the date line, the SPCZ was further north than usual
from January through August, and further south than
usual from November through December. East of
the date line, the SPCZ was very weak from March
through August. It was further south than usual from
October through December (see section 4D). Aboveaverage equatorial SSTs occurred with the weakly
negative El Niño; however, the region of positive SST
anomalies drifted west as the El Niño faded. From
August through December OLR anomalies showed
enhanced convection over Papua New Guinea, and
suppressed convection over Western and Eastern
Kiribati, Tokelau, Tuvalu and the North Cook Islands.
For much of the year, mean sea level pressures were
above average west of the date line and below average in the east.
(i)
Temperature
Overall, 2005 was warmer than normal (1971–2000
mean) across much of the region (Figure 6.39). Aboveaverage SSTs occurred throughout much of the tropical
Southwest Pacific during 2005 (Figure 6.40). Notably,
SSTs were about +1.0°C above average around western
Kiribati and at least +0.5°C above average in many other
island nations, especially those north of 20°S. New
Caledonia, the Southern Cook Islands, the Austral Islands
and Pitcairn Island were surrounded by near-average
SSTs. Southwest Pacific island surface air temperature anomalies for 2005 were consistent with the SST
anomalies throughout the region. It was an extremely
warm year in Tahiti-Faa’a, central French Polynesia,
where the annual mean temperature was 27.0°C (+0.8°C
above the 1971–2000 normal) and equal to the highest
since measurements commenced in 1957.
Locally, a heatwave occurred from 4 to 7 January in
La Tontouta, New Caledonia, with maximum temperatures between 36° and 37°C. New Caledonia’s
mean temperatures were 1.3°C above normal for
the month. In general, southern locations such as
Vanuatu, Fiji, New Caledonia, and southern French
Polynesia experienced a cooler-than-normal July and
August. Tahiti-Faa’a recorded its hottest November
and December maximum temperatures in 2005
(33.9° and 28.1°C, respectively).
(ii)
Precipitation
Southwest Pacific 2005 OLR anomalies showed a
region of enhanced convection over Papua New
Figure 6.41. Annual South Pacific outgoing longwave radiation
anomalies (W m-2 ); high radiation levels (yellow or orange) are
typically associated with clearer skies and lower rainfall, while low
values (blue) often indicate cloudy conditions and more rain for the
region.
90
Guinea ex tending toward the Solomon Islands
(Figure 6.41). There was also an area of weakly
enhanced convection over Niue and the Southern
Cook Islands, as well as Pitcairn Island. Convection
was suppressed in 2005 over western and eastern
Kiribati, Tokelau, Tuvalu, Wallis and Futuna, the
North Cook Islands and the Marquesas Islands.
The year’s rainfall distribution shows similarities
to the OLR pattern. However, for rainfall there were
not many significant anomalies. Annual rainfall
was at least 110% of normal in an area affecting a
region extending from Niue to the north and east
of Fiji and also parts of Southern French Polynesia.
In contrast, 2005 rainfall was less than 90% of the
average throughout much of New Caledonia.
One location, Gambier, Rikitea, French Polynesia,
recorded an ex tremely high 2005 precipitation
anomaly of 127% of normal (2505 mm). Several
locations received record monthly rainfall amounts:
Vunisea, Fiji (786 mm, April), Viwa (205 mm, June),
Monasavu (640 mm, June), Navua (587 mm, June),
Nausori Airport (474 mm, September), Vatukoula
(353 mm, November), and at Lupepau’u, Tonga
(440 mm, June). On 4 March, Vunisea set a record
for daily rainfall of 251 mm, and on 18 November,
Vatukoula received a record 119 mm of precipitation.
Record low monthly precipitation was also recorded
at Udu Point, Fiji (84 mm), and Tuamotu, Takaroa,
French Polynesia (25 mm), in March, Nadi Airport
and Penang Mill, Fiji (1 and 7.8 mm, respectively),
in May, and Bora Bora, French Polynesia (23 mm)
in October.
of the date line, with most being triggered during
an active phase of the Madden–Julian oscillation.
On the Saffir–Simpson scale, Meena and Nancy
reached category 4 status, while Olaf and Percy
were category 5 at their height.
The first of the cyclones, Meena, passed close to
Rarotonga on 6 February, with wind gusts exceeding 115 km h -1. Cyclone Nancy tracked through the
Northern Cook Islands from 13 to 15 February, with
winds gusting to 163 km h -1 in Rarotonga and to
185 km h-1 elsewhere. In Aitutaki, trees were uprooted,
roofs damaged and low-lying areas flooded. Wind
and storm surge caused widespread damage along
the nor thern and eastern coasts of Rarotonga.
On 16 Februar y, Olaf ’s storm surge destroyed
numerous coastal structures, and high winds lifted
many roofs in both Samoa and American Samoa.
Rarotonga’s west coast suffered substantial damage.
Late in the month, Percy, with maximum sustained
winds reaching 260 km h-1, caused widespread damage and destruction on Pukapuka, Nassau, Swain,
and Tokelau Islands. In Tokelau, Percy destroyed
most of the island’s agriculture and was reportedly the worst tropical cyclone in living memory.
Fortunately, despite the intense nature of these four
cyclones, and reconstruction estimates exceeding
US$ 25 million, there were no confirmed fatalities
from these storms (two listed as missing at sea
following Olaf ).
Rainfall occurred almost ever y day from 6 to
20 April in Fiji’s Western Division. Ex tensive
flooding occurred in the Northern and Western
Divisions over 16–20 April, closing almost 50 roads.
There was one fatality. Vanuatu, Pekoa, recorded
950 mm for April, including 5 days exceeding
100 mm. Torrential rainfall occurred in Fiji’s Central
Division during the last week of September, leading
to flooding in parts of Suva, Nausori, and Tailevu.
A large number of villages were evacuated and there
was one fatality. Nausori Airport recorded 187 mm
of rain on the 28th.
(iii) Notable events
In February, a record four intense tropical cyclones
(Meena, Nancy, Olaf and Percy) impacted the South
Pacific Islands (see section 4C). All four occurred east
91
7. SEASONAL SUMMARIES
G. D. BELL5 AND M.S. HALPERT32
Figure 7.1. December 2004–February 2005 (top) surface
temperature anomalies (°C) and (bottom) precipitation
percentiles based on a gamma distribution fit to the
1979–2000 base period; temperature anomalies
(1971–2000 base period) are based on station data
over land and sea surface temperature over water.
Precipitation data were obtained from a combination
of rain gauge observations and satellite-derived
estimates (Janowiak and Xie 1999). Analysis was
omitted in data-sparse regions (white areas).
Figure 7.2. December 2004–February 2005
(top) Northern Hemisphere and (bottom)
Southern Hemisphere 500-hPa geopotential
heights (9-dam contour interval) and
anomalies (shading, m) from the 1979–2000
base period mean
93
Figure 7.3. March–May 2005 (top) surface temperature
anomalies (°C) and (bottom) precipitation percentiles
based on a gamma distribution fit to the 1979–2000
base period; temperature anomalies (1971–2000
base period) are based on station data over land and
sea surface temperature over water. Precipitation
data were obtained from a combination of rain
gauge observations and satellite-derived estimates
(Janowiak and Xie 1999). Analysis was omitted in
data-sparse regions (white areas).
Figure 7.4. March–May 2005 (top) Northern
Hemisphere and (bottom) Southern
Hemisphere 500-hPa geopotential heights
(9-dam contour interval) and anomalies
(shading, m) from the 1979–2000 base
period mean
94
Figure 7.5. June–August 2005 (top) surface
temperature anomalies (°C) and (bottom) precipitation
percentiles based on a gamma distribution fit to the
1979–2000 base period; temperature anomalies (1971–
2000 base period) are based on station data over land
and sea surface temperature over water. Precipitation
data were obtained from a combination of rain
gauge observations and satellite-derived estimates
(Janowiak and Xie 1999). Analysis was omitted in
data-sparse regions (white areas).
Figure 7.6. June–August 2005 (top) Northern
Hemisphere and (bottom) Southern
Hemisphere 500-hPa geopotential heights
(9-dam contour interval) and anomalies
(shading, m) from the 1979–2000 base
period mean
95
Figure 7.7. September–November 2005 (top) surface
temperature anomalies (°C) and (bottom) precipitation
percentiles based on a gamma distribution fit to the
1979–2000 base period; temperature anomalies
(1971–2000 base period) are based on station data
over land and sea surface temperature over water.
Precipitation data were obtained from a combination
of rain gauge observations and satellite-derived
estimates (Janowiak and Xie 1999). Analysis was
omitted in data-sparse regions (white areas).
Figure 7.8. September–November 2005
(top) Northern Hemisphere and (bottom)
Southern Hemisphere 500-hPa geopotential
heights (9-dam contour interval) and
anomalies (shading, m) from the 1979–2000
base period mean
96
ACKNOWLEDGEMENTS
This assessment would not have been possible without
the talents, assistance, and contribution of many scientists
and technical experts from around the world. In addition to
the cited authors, a number of additional contributors and
reviewers provided valuable information, guidance, figures,
and expertise. Where known, we gratefully acknowledge
these contributors in the appendix below.
The editors also wish to acknowledge Omar Baddour of the
World Meteorological Organization (WMO) for his help in
identifying and soliciting additional authors and contributors. J. Kennedy is supported by the UK Department of
the Environment, Food and Rural Affairs under Contract
Number PECD 7/12/37.
Special acknowledgement and thanks must be given to Glenn
Hyatt, Sara Veasey and Deborah Riddle at NOAA/NCDC for
the extensive time and graphical expertise they devoted to
preparing this document for publication. The editors also
wish to thank last year’s editor, David Levinson, for his advice
and guidance. This assessment was partially supported by
a grant from the NOAA Climate Program Office’s Climate
Change Data and Detection Program.
97
APPENDIX: contributors and reviewers
Alfred Wegener Institute, Bremerhaven, Germany
• V. Rachold
Danish Meteorological Institute, Copenhagen, Denmark
• J. Cappelen
Finnish Meteorological Institute, Helsinki, Finland
• R. Heino
• S. Saku
Hadley Centre for Climate Prediction and Research, Devon, United Kingdom
• D. Parker
Icelandic Meteorological Office, Reykjavik, Iceland
• T. Jonsson
International Arctic Research Center, Fairbanks, Alaska
• J. Walsh
Max-Planck Institute for Meteorology, Hamburg, Germany
• L. Bengtsson
Naval Postgraduate School, Monterey, California
• W. Maslowski
Nicolaus Copernicus University, Torun,
´ Poland
• R. Przybylak
Norwegian Meteorological Office, Oslo, Norway
• E. Førland
• K. A. Iden
NOAA National Climatic Data Center, Asheville, North Carolina
• P. Groisman
• H.-M. Zhang
• W.E. Angel
• N. Guttman
• T. Whitehurst
• W. Brown
• S. Stephens
• B. Gleason
• S. LeDuc
• D. Easterling
Swedish Meteorological and Hydrometeorological Institute, Norrköpking, Sweden
• H. Alexandersson
98
REFERENCES
Arctic Climatology Project, 1997: Environmental Working
Group joint US – Russian Atlas of the Arctic Ocean—
Winter Period. L. Timokhov and F. Tanis, Eds., Environmental
Research Institute of Michigan in association with the National Snow and Ice Data Center. Ann Arbor, MI, CD-ROM.
Atkinson, G. D., 1971: Forecasters’ guide to tropical meteorology. U.S. Air Force Tech. Rep. 240, 360 pp.
Beck, C., J. Grieser, and B. Rudolf, 2005: A New Monthly
Precipitation Climatology for the Global Land Areas for the
Period 1951 to 2000. DWD, Klimastatusbericht KSB 2004,
ISSN 1437-7691, ISSN 1616-5063 (Internet), ISBN 3-88148402-7, 181–190.
Bell, G. D., and M. Chelliah, 2006: Leading tropical modes associated with interannual and multidecadal fluctuations in North
Atlantic hurricane activity. J. Climate, 19, 590–612.
—, and Coauthors, 1999: Climate Assessment for 1998. Bull.
Amer. Meteor. Soc., 80, S1–S48.
—, and Coauthors, 2000: Climate Assessment for 1999. Bull.
Amer. Meteor. Soc., 81, S1–S50.
—, S. Goldenberg, C. Landsea, E. Blake, R. Pasch, M. Chelliah,
and K. Mo, 2004: Atlantic Hurricane Season. State of the
Climate in 2003. A. M. Waple and J. H. Lawrimore, Eds., Bull.
Amer. Meteor. Soc., 85, S1–S68.
Belchansky, G. I., D. C. Douglas, and N. G. Platonov, 2004:
Duration of the Arctic sea ice melt season: Regional and
interannual variability. J. Climate, 17, 67–80.
Bourassa, M. A., R. Romero, S. R. Smith, and J. J. O’Brien, 2005:
A new FSU winds climatology. J. Climate, 18, 3692–3704.
Box, J. E., and Coauthors, 2005: Greenland ice sheet surface
mass balance variability (1988–2004) from calibrated Polar
MM5 output. J. Climate, in press.
Brohan, P., J. J. Kennedy, I. Harris, S. F. B. Tett, and P.D.
Jones, 2006: Uncertainty estimates in regional and global
observed temperature changes: a new dataset from 1850.
J. Geophys. Res., in press.
Brown, J., K. M. Hinkel, and F. E. Nelson, 2000: The Circumpolar
Active Layer Monitoring (CALM) program: Research Designs
and Initial Results. Polar Geog., 24, 163–258.
Bryden, H. L., H. R. Longworth, and S. A. Cunningham, 2005:
Slowing of the Atlantic meridional overturning circulation at
25°N. Nature, 438, 655–657.
Cavalieri, D., C. Parkinson, P. Gloerson, and H. J. Zwally,
1996: Sea ice concentrations from Nimbus-7 SMMR and
DMSP SSM/I passive microwave data, June to September
2001. National Snow and Ice Data Center. Boulder, CO
[available online at http://arcss.colorado.edu/data/docs/
daac/nsidc0051_gsfc_seaice.gd.html; updated 2005.].
Cazenave, A., and R. S. Nerem, 2004: Present-day sea level
change: Observations and causes. Rev. Geophys., 42,
RG3001, doi:10.1029/2003RG000139.
Chen, D., 2000: A monthly circulation climatology for Sweden
and its application to a winter temperature case study.
Int. J. Climatol., 20, 1067–1076.
Christy, J. R., R. W. Spencer, W. B. Norris, W. D. Braswell, and D.
E. Parker, 2003: Error estimates of Version 5.0 of MSU/AMSU
bulk atmospheric temperatures. J.Atmos Oceanic Technol.,
20, 613–629.
Climate Prediction Center, 2005: Climate Diagnostics Bulletin:
January 2005. US Dept. of Commerce, NOAA/NWS/CPC
[available online at www.cpc.ncep.noaa.gov/products/analysis_monitoring/bulletin_0105.].
Conway, T. J., P. P. Tans, L. S. Waterman, K. W. Thoning,
D. R. Kitzis, K. A. Masarie, and N. Zhang, 1994: Evidence for
interannual variability of the carbon cycle from the NOAA
CMDL global air sampling network. J. Geophys. Res., 99,
22831–22855.
Coughlan, M., D. Jones, N. Plummer, A. Watkins, B. Trewin, and
S. Dawkins, 2003: Impacts of 2002–03 El Niño on Australian
climate. Extended Abstracts, DroughtCom: Improving the Communication of Climate Information, N. Plummer, M. Flannery,
C. Mullen, B. Trewin, A. Watkins, W. Wright, T. Powell,
S. Power, Eds., Vol. 2, Melbourne, Autralia, Bureau of
Meteorology, 7–12.
Couture, R., S. Smith, S. D. Robinson, M. M. Burgess, and
S. Solomon, 2003: On the hazards to infrastructure in the
Canadian North associated with thawing of permafrost.
Proc. Geohazards 2003, 3rd Canadian Conf. on Geotechnique
and Natural Hazards. Edmonton, AB, Canada, Can. Geotech.
Soc., 97–104.
Cummins, P. F., G. S. E. Lagerloef, and G. Mitchum, 2005:
A regional index of northeast Pacific variability based on
satellite altimeter data. Geophys. Res. Lett., 32, L17607,
doi:10.1029/2005GL023642.
Curry, R. G., and M. S. McCartney, 2001: Ocean gyre circulation changes associated with the North Atlantic oscillation.
J. Phys. Oceanog., 31, 3374–3400.
—, and C. Mauritzen, 2005: Dilution of the northern North
Atlantic in recent decades. Science, 308, 1772–1774.
Daniel, J. S., and S. Solomon, 1998: On the climate forcing of
carbon monoxide. J. Geophys. Res., 103, 13,249–13,260.
Dlugokencky, E. J., S. Houweling, L. Bruhwiler, K. A. Masarie,
P. M. Lang, J. B. Miller, and P. P. Tans, 2003: Atmospheric
methane levels off: Temporary pause or new steady state?
Geophys. Res. Lett., 30, 19, doi:10.1029/2003GL018126.
Durre, I., R. S. Vose, and D. B. Wuertz, 2006: Overview of the
Integrated Global Radiosonde Archive. J. Climate, 19,
53–68.
Dvorak, V. F., 1984: Tropical Cyclone Intensity Analysis Using
Satellite Data. NOAA Technical Rep. NESDIS 11, 47 pp.
99
Eisenman, I., L. Yu, and E. Tziperman, 2005: Westerly wind
bursts: ENSO’s tail rather than the dog? J. Climate, 18,
5224–5238.
Emanuel, K. A., 2005: Increasing destructiveness of tropical
cyclones over the past 30 years. Nature, 436, 686–688.
Feely, R. A., C. L. Sabine, T. Takahashi, and R. Wanninkhof,
2001: Uptake and storage of carbon dioxide in the oceans:
The global CO2 survey. Oceanography, 14, 18–32.
—, L. D. Talley, G. C. Johnson, C. L. Sabine, and R. Wanninkhof,
2005: Repeat hydrography cruises reveal chemical changes
in the North Atlantic. Eos Trans. Amer. Geophys. Union, 86,
399, 404–405.
Frauenfeld, O. W., T. Zhang, R. G. Barry, and D. Gilichinsky,
2004: Interdecadal changes in seasonal freeze and thaw
depths in Russia. J. Geophys. Res., 109, D05101, doi:10.1029/
2003JD004245.
Free, M., D. J. Seidel, J. K. Angell, J. Lanzante, I. Durre, and
T. C. Peterson, 2005: Radiosonde Atmospheric Temperature
Products for Assessing Climate (RATPAC): A new dataset
of large-area anomaly time series. J. Geophys. Res., 110,
D22101, doi: 10.1029/2005JD006169.
Ganachaud, A., and C. Wunsch, 2003: Large-Scale Ocean
Heat and Freshwater Transports during the World Ocean
Circulation Experiment. J. Climate, 16, 696–705.
Goetz, S. J., A. G. Bunn, G. J. Friske, and R. A. Houghton, 2005:
Satellite-observed photosynthetic trends across boreal
North America associated with climate and fire disturbance.
Proc. Nation. Acad. Sci., 102, 13,521–13,525.
Goldenberg, S. B., C. W. Landsea, A. M. Mestas-Nuñez, and
W. M. Gray, 2001: The recent increase in Atlantic hurricane
activity: Causes and implications. Science, 293, 474–479.
Gruber, N., J. L. Sarmiento, T. F. Stocker, 1996: An improved
method for detecting anthropogenic CO2 in the oceans.
Global Biogeochem. Cyc., 10, 809–837.
Haas, C., 2004: Late-summer sea ice thickness variability in the
Arctic Transpolar Drift 1991–2001 derived from ground-based
electromagnetic sounding. Geophys. Res. Let., 31, L09402,
doi:10.1029/2003GL019394.
Hansen, J., and Coauthors, 2005: Earth’s energy imbalance:
Confirmation and implications. Science, 308, 1431–1435.
Harris, C., and W. Haeberli, 2003: Warming permafrost in
European mountains. Global Planet. Change, 39, 215–225.
Heim, R. R., Jr., 2002: A review of twentieth-century drought
indices used in the United States. Bull. Amer. Meteor. Soc.,
83, 1149–1165.
Horton E. B., C. K. Folland, and D. E. Parker, 2001: The changing
incidence of extremes in worldwide and central England
temperatures to the end of the twentieth century. Climatic
Change, 50, 267–295.
100
IPCC, 2001: Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the Third Assessment Report
of the Intergovermental Panel on Climate Change. World
Meteor. Org. /United Nations Environment Programme,
J. T. Houghton, Y. Ding, D. J. Griggs, M. Noguer, P. J. van
der Linden, X. Dai, K. Maskell, and C. A. Johnson, Eds.,
Cambridge University Press, 881 pp.
Isaksen, K., D. Vonder Mühll, H. Gubler, T. Kohl, and J.L. Sollid,
2000: Ground surface temperature reconstruction based on
data from a deep borehole in permafrost at Janssonhaugen,
Svalbard. Ann. Glaciol., 31, 287–294.
Janowiak, J. E., and P. Xie, 1999: CAMS-OPI: A global satellite-rain gauge merged product for real-time precipitation
monitoring applications. J. Climate, 12, 3335–3342.
Jauregui, E., 2003: Climatology of landfalling hurricanes and
tropical storms in Mexico. Atmosfera, 16, 193–204.
Jia, G. J., H. E. Epstein, and D. A. Walker, 2003: Greening of
arctic Alaska, 1981–2001. Geophys. Res. Lett., 30, 2067,
doi:10.1029/2003GL018268.
Jin, F. F., 1997: An equatorial recharge paradigm for ENSO.
Part I: Conceptual model. J. Atmos. Sci., 54, 811–829.
Johannessen, O. M., and Coauthors, 2004: Arctic climate
change: Observed and modeled temperature and sea ice
variability. Tellus, 56A, 328–341.
Karl, T. R., R. W. Knight, D. R. Easterling, and R. G. Quayle. 1996:
Indices of Climate Change for the United States. Bull. Amer.
Meteor. Soc., 77, 2, 279–292.
Kasischke, E. S., and Coauthors, 2000: Contributions of 1998
fires in the boreal forest to atmospheric concentrations of
carbon monoxide and methane. Eos Trans. Amer. Geophys.
Union, 81, 260.
Kocin, P. J., and L. W. Uccellini, 2004: A snowfall impact scale
derived from northeast storm snowfall distributions. Bull.
Amer. Meteor. Soc., 85, 177–194.
Koltermann, K. P., A. V. Sokov, V. P. Tereschenkov, S. A. Dobroliubov, K. Lorbacher, and A. Sy, 1999: Decadal changes on
the thermohaline circulation of the North Atlantic. Deep-Sea
Res. II, 46, 109–138.
Kousky, V. E. and M. T. Kayano, 1994: Principal modes of outgoing longwave radiation and 250-mb circulation for the South
American sector. J. Climate, 7, 1131–1143.
Kwok, R., H. J. Zwally, and D. Yi, 2004: ICESat observations of
Arctic sea ice: A first look. Geophys. Res. Lett., 31, L16401,
doi:10.1029/2004GL020309.
Lamb, H.H., 1950: Types and spells of weather around the year in
the British Isles. Quart. J. Roy. Meteor. Soc., 76, 393–438.
Lander, M. A., and C. P. Guard, 1998: A look at global tropical cyclone activity during 1995: Contrasting high Atlantic
activity with low activity in other basins. Mon. Wea. Rev.,
126, 1163–1173.
Landsea, C. W., 1993: The climatology of intense (or major)
Atlantic hurricanes. Mon. Wea. Rev., 121, 1703–1713.
Mears, C. A., M. C. Schabel, and F. J. Wentz, 2003: A reanalysis
—, G. D. Bell, W. M. Gray, and S. B. Goldenberg, 1998:
The extremely active 1995 Atlantic hurricane season:
Environmental conditions and verification of seasonal forecasts. Mon. Wea. Rev., 126, 1174–1193.
J. Climate, 16, 3650–3664.
—, R. A. Pielke Jr., A. M. Mestas-Nuñez, and J. A. Knaff,
1999: Atlantic basin hurricanes: Indices of climatic changes.
Climatic Change, 42, 89–129.
of the MSU channel 2 tropospheric temperature record.
Meinen, C. S., and M. J. McPhaden, 2000: Observations of warm
water volume changes in the equatorial Pacific and their
relationship to El Niño and La Niña. J. Climate, 13, 3551–
3559.
Melling, H., D. A. Riedel, and Z. Gedalof, 2005: Trends in the draft
and extent of seasonal pack ice, Canadian Beaufort Sea. Geo-
Langenfelds, R. L., R. J. Francey, B. C. Pak, L. P. Steele, J. Lloyd,
C. M. Trudinger, and C. E. Allison, 2002: Interannual growth
rate variations of atmospheric CO2 and its δ13C, H2, CH4, and
CO between 1992 and 1999 linked to biomass burning. Global
Biogeochem. Cyc., 16, 1048, doi:10.1029/2001GB001466.
Mo, K. C., 2000: The association between intraseasonal oscil-
Lawrimore, J., and Coauthors, 2001: Climate Assessment for
2000. Bull. Amer. Meteor. Soc., 82(6), S1–S55.
—, and V. E. Kousky, 1993: Further analysis of the relationship
Laxon, S., N. Peacock, and D. Smith, 2003: High interannual
variability of sea ice thickness in the Arctic Region. Nature,
425, 947–950.
Leuliette, E. W, R. S. Nerem, and G. T. Mitchum, 2004: Calibration
of TOPEX/Poseidon and Jason altimeter data to construct a
continuous record of mean sea level change. Mar. Geodesy.,
27, 79–94.
phys. Res. Lett., 32, L24501, doi:10.1029/2005GL024483.
lations and tropical storms in the Atlantic Basin. Mon. Wea.
Rev., 128, 4097–4107.
between circulation anomaly patterns and tropical convection. J. Geophys. Res., 98, 5103–5113.
Montzka, S. A, J. H. Butler, B. D. Hall, J. W. Elkins, and D. J.
Mondeel, 2003a: A decline in tropospheric organic bromine.
Geophy. Res. Lett., 30, 1826, doi:10.1029/2003GL017745.
—, and Coauthors, 2003b: Controlled substances and other
source gases. Scientific Assessment of Ozone Depletion:
Levinson, D. H., Ed., 2005: State of the Climate in 2004. Bull.
Amer. Meteor. Soc., 86(6), S1–S86.
2002, Global Ozone Research and Monitoring Project—
Levitus, S., J. I. Antonov, and T. P. Boyer, 2005: Warming of the
World Ocean, 1955–2003. Geophys. Res. Lett., 32, L02604,
doi:10.1029/2004GL021592.
Mote, P. W., A. F. Hamlet, M. P. Clark, and D. P. Lettenmaier, 2005:
Lucht, W., and Coauthors, 2002: Climate control of the highlatitude vegetation greening trend and Pinatubo effect.
Science, 296, 1687–1689, doi:10.1126/science.1071828.
L ump k in, R ., a n d Z . G a r r a f f o, 2 0 0 5 : Eva lu a t in g t h e
decomposition of tropical Atlantic drifter observations.
J. Atmos. Oceanic Technol., 22, 1403—1415.
Lyon, B., and A. G. Barnston, 2005: The evolution of the weak
2004–2005 El Nino. U.S. CLIVAR Variations, Vol. 3, US CLIVAR
Office, Washington, DC, 1–4.
Report No. 47.
Declining mountain snowpack in western North America.
Bull. Amer. Meteor. Soc., 86, 39–49.
Moyer, K. A., and R. A. Weller, 1997: Observations of surface
forcing from the Subduction Experiment: a comparison
with global model products and climatological datasets.
J. Climate, 10, 2725–2742.
Munich Re, 2006: Annual Review: Natural catastrophes 2005.
Knowledge Series, Topics Geo [available online at www.
munichre.com/].
Madden, R. A., and P. R. Julian, 1971: Detection of a 40–50 day
oscillation in the zonal wind in the tropical Pacific. J. Atmos.
Sci., 28, 702–708.
Myneni, R. B., C. D. Keeling, C. J. Tucker, G. Asrar, and
—, and —, 1972: Description of global-scale circulation
cells in the Tropics with 40–50 day period. J. Atmos. Sci.,
29, 1109–1123.
—, C. J. Tucker, G. Asrar, and C. D. Keeling, 1998: Interannual
—, and —, 1994: Observations of the 40–50 day tropical
oscillation—A review. Mon. Wea. Rev., 122, 814–837.
Mantua, N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and
R. C. Francis, 1997: A Pacific interdecadal climate oscillation
with impacts on salmon production. Bull. Amer. Meteor.
Soc., 78, 1069–1079.
R. R. Menani, 1997: Increased plant growth in the northern
high latitudes from 1981 to 1991. Nature, 386, 698–702.
variations in satellite-sensed vegetation index data from 1981
to 1991. J. Geophys. Res., 103, 6145–6160.
Neumann, C. J., B. R. Jarvinen, C. J. McAdie, and J. D. Elms,
1993 : Tropical Cyclones of the North Atlantic Ocean,
1871–1992, NOAA National Climatic Data Center, in cooperation with the National Hurricane Center, Coral Gables, FL,
193 pp.
101
Nicholls, N., 2004: The changing nature of Australian droughts.
Climatic Change, 63, 323–336.
Nixon, F. M., C. Tarnocai, and L. Kutny, 2003: Long-term active
layer monitoring: Mackenzie Valley, northwest Canada. Proc.
Eighth Int. Conf. on Permafrost, Vol. 2, Zurich, Switzerland,
International Permafrost Association, 821–826.
NOAA OCO, 2006: The State of the Ocean Observing System
for Climate. FY 2005 Annual Report: Office of Climate
Observation, J. M. Levy, D. M. Stanitsky, and P. Arkin, eds.,
US Dept. of Commerce, National Oceanic and Atmospheric
Administration, Office of Climate Observation, Silver Spring,
MD, 337 pp.
Novelli, P. C., K. A. Masarie, P. M. Lang, B. D. Hall, R. C. Myers,
and J. C. Elkins, 2003: Reanalysis of tropospheric CO trends:
Effects of the 1997–1998 wildfires, J. Geophys. Res., 108, D15,
4464, doi:10.1029/2002JD003031.
Oberman, N. G., and G. G. Mazhitova, 2001: Permafrost dynamics in the northeast of European Russia at the end of the
twentieth century. Norweg. J. Geog., 55, 241–244.
Osterkamp, T. E., 2003: A thermal history of permafrost in
Alaska. Proc. Eighth Int. Conf. on Permafrost, M. Phillips,
S. M. Springman, and L. U. Arenson, Eds., Zurich, CH,
A. A. Balkema, Vol. 2, 863–868.
—, and V. E. Romanovsky, 1999: Evidence for warming and
thawing of discontinuous permafrost in Alaska. Permafrost
Periglacial Process., 10, 17–37.
Overland, J. E., and M. Wang, 2005: The third Arctic climate
pattern: 1930s and early 2000s. Geophys. Res. Lett., 32,
L23808, doi:10.1029/2005GL024254.
—, M. C. Spillane, D. B. Percival, M. Wang, and H. O. Mofjeld,
2004: Seasonal and regional variation of pan-Arctic surface
air temperature over the instrumental record. J. Climate,
17, 3263–3282.
Palmer, W. C., 1965: Meteorological Drought. U.S. Weather
Bureau Research Paper 45, 58 pp [available from NOAA
Library and Information Services Division, Washington, DC
20852].
Parker, D. E., and E. B. Horton, 2005: Uncertainties in the Central
England Temperature series 1878–2003 and some improvements to the maximum and minimum series. Int. J. Climatol.,
25, 1173–1188.
Pavlov, A. V., 1994: Current changes of climate and permafrost
in the Arctic and Sub-Arctic of Russia. Permafrost Periglacial
Process., 5, 101–110.
—, and N. G. Moskalenko, 2002: The thermal regime of
soils in the north of Western Siberia. Permafrost Periglacial
Process., 13, 43–51.
102
Peterson, B. J., R. M. Holmes, J. W. McClelland, C. J. Vörösmarty,
R. B. Lammers, A. I. Shiklomanov, I. A. Shiklomanov, and
S. Rahmstorf, 2002: Increasing river discharge to the Arctic
Ocean. Science, 298, 2171–2173.
Peterson, T. C., and R. S. Vose, 1997: An overview of the Global
Historical Climatology Network temperature database. Bull.
Amer. Meteor. Soc., 78, 2837–2849.
Pfirman, S., W. F. Haxby, R. Colony, and I. Rigor, 2004 :
Variability in Arctic sea ice drift. Geophys. Res. Lett., 31,
L16402, doi:10.1029/2004GL020063.
Polyakov, I., and Coauthors, 2003: Long-term ice variability in
arctic marginal seas. J. Climate, 16, 2078–2085.
—, and Coauthors, 2005: One more step toward a warmer
Arctic. Geophys. Res. Lett., 32, L17605, doi :10.1029 /
2005GL023740.
Prather, M. J., 1996: Natural modes and time scales in atmospheric chemistry: Theory, GWPs for CH4 and CO, and
runaway growth. Geophys. Res. Lett., 23, 2597–2600.
Prentice, I. C. and Coauthors, 2001: The carbon cycle and atmospheric carbon dioxide. Climate Change 2001: The Scientific
Basis, Cambridge University Press, 190 pp.
Proshutinsky, A., I. M. Ashik, E. N. Dvorkin, S. Häkkinen,
R. A. Krishfield, and W. R. Peltier, 2004: Secular sea level
change in the Russian sector of the Arctic Ocean. J. Geophys.
Res., 109, C03042, doi:10.1029/2003JC002007.
Quadrelli, R., and J. M. Wallace, 2004: A simplified linear framework for interpreting patterns of Northern Hemisphere wintertime climate variability. J. Climate, 17, 3728–3744.
Quayle, R. G., T. C. Peterson, A. N. Basist, and C. S. Godfrey,
1999: An operational near real-time global temperature
index. Geophys. Res. Lett., 26, 333–336.
Reynolds, R. W., and T. M. Smith, 1994: Improved global sea
surface temperature analyses using optimum interpolation.
J. Climate, 7, 929–948.
—, N. A. Rayner, T. M. Smith, D. C. Stokes, and W. Wang, 2002:
An improved in situ and satellite SST analysis for climate.
J. Climate, 15, 1609–1625.
Rignot, E., and P. Kanagaratnam, 2006 : Changes in the
velocity structure of the Greenland Ice Sheet. Science, 311,
986–990.
Rigor, I., and J. M. Wallace, 2004: Variations in the age of Arctic
sea-ice and summer sea-ice extent. Geophys. Res. Lett., 31,
L09401, doi:10,1029/2004GL019492.
Roemmich, D., S. Riser, R. Davis, and Y. Desaubies, 2004:
Autonomous profiling floats: Workhorse for broadscale
ocean observations. J. Mar. Technol. Soc., 38, 31–39.
Romanovsky, V. E., M. Burgess, S. Smith, K. Yoshikawa, and
J. Brown, 2002: Permafrost temperature records: Indicator
of climate change. Eos Trans. Amer. Geophys. Union, 83,
589, 593–594.
Ropelewski, C. F., and M. S. Halpert, 1987: Global and regional
Stroeve, J. C., M. C. Serreze, F. Fetterer, T. Arbetter, W. Meier,
scale precipitation patterns associated with the El Niño/
J. Maslanik, K. Knowles, 2005: Tracking the Arctic’s shrink-
Southern Oscillation. Mon. Wea. Rev., 115, 1606–1626.
ing ice cover: Another extreme September minimum
—, and —, 1989: Precipitation patterns associated with the
high index phase of the Southern Oscillation. J. Climate, 2,
268–284.
—, J. E. Janowiak, and M. S. Halpert, 1985: The analysis and
display of real-time surface climate data. Mon. Wea. Rev.,
113, 1101–1106.
Rothrock, D. A., Y. Yu, and G. A. Maykut, 1999: Thinning of the
Arctic sea-ice cover. Geophys. Res. Lett., 26, 3469–3472.
Rudolf, B., and U. Schneider, 2005: Calculation of gridded
precipitation data for the global land-surface using in-situ
gauge observations. Proc. Second Workshop of the Int.
Precipitation Working Group, Monterey, CA, EUMETSAT,
ISBN 92-9110-070-6, ISSN 1727-432X, 231–247.
—, H. Hauschild, W. Rueth, and U. Schneider, 1994: Terrestrial
in 2004. Geophys. Res. Lett., 32, L04501, doi:10.1029/
2004GL021810.
Talley, L. D., 2003: Shallow, intermediate, and deep overturning
components of the global heat budget. J. Phys. Oceanog.,
33, 530–560.
Tarnocai, C., F. M. Nixon, and L. Kutny, 2004: CircumpolarActive-Layer-Monitoring (CALM) sites in the Mackenzie Valley, Northwestern Canada. Permafrost Periglacial Process.,
15, 141–153.
Thorne, P. W., D. E. Parker, S. F. B. Tett, P. D. Jones, M. McCarthy, H. Coleman, and P. Brohan, 2005: Revisiting radiosonde
upper-air temperatures from 1958 to 2002. J. Geophys. Res.,
110, D18105, doi:10.1029/2004JD005753.
Trenberth, K., 2005: Uncertainty in hurricanes and global warming. Science, 308, 1753–1754.
Precipitation Analysis: Operational Method and required
Trigo, R. M., D. Pozo-Vazquez, T. J. Osborn, Y. Castro-Diez,
density of point measurements. Global Precipitations and
S. Gámis-Fortis, and M. J. Esteban-Parra, 2004: North
Climate Change, NATO ASI Series I, Vol. 26, M. Desbois, and
Atlantic Oscillation influence on precipitation, river flow and
F. Desalmond, Eds., Springer-Verlag, 173–186.
water resources in the Iberian Peninsula. Int. J. Climatol.,
— , C. Beck, J. Grieser, and U. Schneider, 2005: Global
24, 925–944.
precipitation analysis products Global Precipitation Clima-
Varlamov, S. P., 2003: Variations in the thermal state of the litho-
tology Centre (GPCC), DWD, 8 pp [available online at www.
genic base of landscapes in Central Yakutia. Proc. Second Int.
dwd.de/en/FundE/Klima/KLIS/int/GPCC/Reports_Publica-
Conf. “The Role of Permafrost Ecosystems in Global Climate
tions/QR/GPCC-intro-products-2005.pdf].
Sabine, C. L., and Coauthors, 2004: The oceanic sink for anthropogenic CO2. Science, 305, 367–371.
Change,” Yakutsk, Russia, Permafrost Institute for Biological
Problems of Cryolithozone Siberian Division of the Russian
Academy of Sciences, 52–56.
Schmittner, A., M. Latif, and B. Schneider, 2005: Model projec-
—, Y. B. Skachkov, P. N. Skryabin, and N. I. Shender, 2001:
tions of the North Atlantic thermohaline circulation for the
Thermal response of the lithogenic base of permafrost land-
21st century assessed by observations. Geophys. Res. Lett.,
scapes to recent climate change in Central Yakutia. Proc. Int.
32, L23710, doi:10,1029/2005GL024368.
Conf. “The Role of Permafrost Ecosystems in Global Climate
Shimada, K., F. McLaughlin, E. Carmack, A. Proshutinsky,
S. Nishino, and M. Itoh, 2004: Penetration of the 1990s
warm temperature anomaly of Atlantic water in the
Canada Basin. Geophys. Res. Lett., 31, L20301, doi:10.1029/
2004GL020860.
Simpson, R. H., 1974: The hurricane disaster potential scale.
Weatherwise, 27, 169–186.
Smith, S. L., M. M. Burgess, and A. E. Taylor, 2003: High Arctic
Change,” Yakutsk, Russia, Permafrost Institute for Biological
Problems of Cryolithozone Siberian Division of the Russian
Academy of Sciences, 44–45.
Vaughan, D. G., and C. S. M. Doake, 1996: Recent atmospheric warming and retreat of ice shelves on the Antarctic
Peninsula. Nature, 379, 328–331.
Vincent, D. G., 1998 : Pacific Ocean. Meteorology of the
Southern Hemisphere, Meteor. Monogr., No. 49, Amer.
Meteor. Soc., 410 p.
permafrost observatory at Alert, Nunavut—Analysis of 23
Vose, R. S., R. L. Schmoyer, P. M. Steurer, T. C. Peterson,
year data set. Proc. 8th Int. Conf. on Permafrost, M. Phil-
R. Heim, T. R. Karl, and J. Eischeid, 1992: The Global Histori-
lips, S. M. Springman, and L. U. Arenson, Eds., Zurich, CH,
cal Climatology Network: Long-term monthly temperature,
A. A. Balkema, Vol. 2, 1073–1078.
precipitation, sea level pressure, and station pressure data.
Smith, T. M., and R. W. Reynolds, 2005: A global merged land air
ORNL/CDIAC-53, NDP-041, Carbon Dioxide Information
and sea surface temperature reconstruction based on histori-
Analysis Center (CDIAC), Oak Ridge National Laboratory,
cal observations (1880–1997). J. Climate, 18, 2021–2036.
Oak Ridge, Tennessee, 324 pp.
103
Wallace, D. W. R., 2001: Storage and transport of excess CO2 in
the oceans: The JGOFS/WOCE Global CO2 Survey. Ocean
Circulation and Climate: Observing and Modelling the Global
Ocean, G. Siedler, J. Church, and J. Gould, Eds., Academic
Press, 489–521.
Wang, X., and J. R. Key, 2005a: Arctic surface, cloud, and radiation properties based on the AVHRR Polar Pathfinder dataset.
Part I: Spatial and temporal characteristics. J. Climate, 18,
2558–2574.
—, and —, 2005b: Arctic surface, cloud, and radiation properties based on the AVHRR Polar Pathfinder dataset. Part II:
Recent trends. J. Climate, 18, 2575–2593.
Waple, A. M., and E. Rignot, 2005: Antarctic. State of the Climate
in 2004, D. H. Levinson, Ed., Bull. Amer. Meteor. Soc., 86(6),
S42–S44.
Watkins, A. B., 2005: The Australian Drought of 2005. WMO
Bull., 54, 156–162.
Webster, P. J., G. J. Holland, J. A. Curry, and H.-R. Chang, 2005:
Changes in tropical cyclone number, duration, and intensity
in a warming environment. Science, 309, 1844–1846.
Whitney, L. D., and J. S. Hobgood, 1997: The relationship between sea surface temperatures and maximum intensities
of tropical cyclones in the Eastern North Pacific Ocean.
J. Climate, 10, 2921–2930.
Willis, J. K., D. Roemmich, and B. Cornuelle, 2004: Interannual
variability in upper ocean heat content, temperature, and
thermosteric expansion on global scales. J. Geophys. Res.,
109, C12036, doi: 10.1029/2003JC002260.
Xie, P., and P. A. Arkin, 1997: Global precipitation: A 17-year
monthly analysis based on gauge observations, satellite
estimates, and numerical model outputs. Bull. Amer. Meteor.
Soc., 78, 2539–2558.
—, and —, 1998: Global monthly precipitation estimates
from satellite-observed outgoing longwave radiation.
J. Climate, 11, 137–164.
— , J. E. Janowiak, P. A . Arkin, R. Adler, A . Gruber,
R. Ferraro, G. J. Huffman, and S. Curtis, 2003: GPCP pentad
precipitation analyses: An experimental dataset based on
gauge observations and satellite estimates. J. Climate, 16,
2197–2214.
Xue, Y., T. M. Smith, and R. W. Reynolds, 2003: A new SST
climatology for the 1971–2000 base period and interdecadal
changes of 30-year SST normal. J. Climate, 16, 1601–1612.
Yu, L., R. A. Weller, and B. Sun, 2004: Mean and variability of
the WHOI daily latent and sensible heat fluxes at in situ flux
measurement sites in the Atlantic Ocean. J. Climate, 17,
2096–2118.
Yu, Y., G. A. Maykut and D. A. Rothrock, 2004: Changes in the
thickness distribution of Arctic sea ice between 1958–1970
and 1993–1997. J. Geophys. Res., 109, C08004, doi:10.1029/
2003JC001982.
104
Zervas, C., S. Gill, and A. Stolz, 2005: Relative Sea Level Trends
from Tide Stations: How are they determined and what do
they tell us? Climate Change Science Program Workshop,
Climate Science in Support of Decision Making, Arlington,
VA, U.S. Climate Change Science Program, P-GC1.6 [available online at www.climatescience.gov/workshop2005/posters/P-GC1.6_Zervas.pdf].
Zhang, C., H. Hendon, W. Kessler, and A. Rosati, 2001: A workshop on the MJO and ENSO. Bull. Amer. Meteor. Soc., 82,
971–976.
Zhang, T., and Coauthors, 2005: Spatial and temporal variability of active layer thickness over the Russian Arctic
drainage basin. J. Geophys. Res., 110, D16101, doi:10.1029/
2004JD005642.
Zhou, L., C. J. Tucker, R. K. Kaufmann, and D. Slayback, 2001:
Variations in northern vegetation activity inferred from satellite data of vegetation index during 1981 to 1999. J. Geophys.
Res., 106, 20 069–20 083.
Zwally, H. J., W. Abdalati, T. Herring, K. Larson, J. Saba, and
K. Steffen, 2002: Surface melt-induced acceleration of Greenland ice-sheet flow. Science, 297, 218–222.
For information about WMO, please contact:
Communications and Public Affairs Office
World Meteorological Organization
7 bis, avenue de la Paix – P.O. Box 2300 – CH 1211 Geneva 2 – Switzerland
Tel.: +41 (0) 22 730 83 14 – Tel.: +41 (0) 22 730 83 15 – Fax: +41 (0) 22 730 80 27
E-mail:
[email protected] – Website: www.wmo.int
For information about the Global Climate System Review, please contact:
World Climate Programme Department
World Meteorological Organization
7 bis, avenue de la Paix – P.O. Box 2300 – CH 1211 Geneva 2 – Switzerland
Tel.: +41 (0) 22 730 82 68 – Fax: +41 (0) 22 730 80 42
E-mail:
[email protected] – Website: www.wmo.int/web/wcp/wcp_progr.htm