Download The greenhouse effect and climate change : a

Document related concepts

Effects of global warming on human health wikipedia , lookup

Fred Singer wikipedia , lookup

General circulation model wikipedia , lookup

Climate-friendly gardening wikipedia , lookup

Climate change adaptation wikipedia , lookup

Media coverage of global warming wikipedia , lookup

Climate change in Tuvalu wikipedia , lookup

Kyoto Protocol wikipedia , lookup

Attribution of recent climate change wikipedia , lookup

Climate engineering wikipedia , lookup

Climate change and agriculture wikipedia , lookup

Climate governance wikipedia , lookup

Scientific opinion on climate change wikipedia , lookup

Effects of global warming on humans wikipedia , lookup

Economics of global warming wikipedia , lookup

Surveys of scientists' views on climate change wikipedia , lookup

Global warming wikipedia , lookup

Climate change, industry and society wikipedia , lookup

Effects of global warming on Australia wikipedia , lookup

Citizens' Climate Lobby wikipedia , lookup

2009 United Nations Climate Change Conference wikipedia , lookup

Public opinion on global warming wikipedia , lookup

Carbon governance in England wikipedia , lookup

Climate change mitigation wikipedia , lookup

German Climate Action Plan 2050 wikipedia , lookup

Economics of climate change mitigation wikipedia , lookup

Solar radiation management wikipedia , lookup

Climate change feedback wikipedia , lookup

United Nations Climate Change conference wikipedia , lookup

Climate change and poverty wikipedia , lookup

Views on the Kyoto Protocol wikipedia , lookup

Climate change in the United States wikipedia , lookup

Climate change in New Zealand wikipedia , lookup

Low-carbon economy wikipedia , lookup

Carbon Pollution Reduction Scheme wikipedia , lookup

Politics of global warming wikipedia , lookup

Mitigation of global warming in Australia wikipedia , lookup

IPCC Fourth Assessment Report wikipedia , lookup

Business action on climate change wikipedia , lookup

Transcript
"It is neither technical nor economic constraints
that will prevent us from reaching our goals. We have the tools.
The real challenge is to find the political will. "
Ritt Bjerregaard
Environmental Commissioner
at First Conference of FCCC Parties, Berlin 1995 1
 Jim Hubbard2
"We can't ignore mounting scientific evidence on
important issues such as climate change.
The science may be provisional. All science is provisional.
But if you see a risk you have to take precautionary action
just as you would in any other aspect of business."
Sir John Browne
Chief Executive Officer, BP Amoco 3
1
Cited in Newell 1997, p.12.
Cartoon originally published in The Dominion 27/11/00, p. 10.
3
Cited by Pew Center on Global Climate Change, http://www.pewclimate.org/belc/bp_quote.cfm . BP Amoco is a member of the
Pew Center’s Business Environmental Leadership Council.
2
The greenhouse effect and climate change
Parliamentary Library, August 2001
Where to find key information
( s = section, ss = sections, ch = chapter)
observed changes & predictions:
temperature, rain, sea level, ice melt
for New Zealand ss 6.5, 6.6 for the world s 6.1
What is climate change?
projected impacts:
social, economic & ecological
for New Zealand s 7.2
for the world
What effect will it have?
Is this connected
with risk to the
Ozone Layer?
scientific understanding
greenhouse gases & sources
causes of warming & cooling
climate models & scenarios
natural ice ages & warm periods
the El Niño cycle in New Zealand
the carbon cycle & carbon sinks
s 4.5
How is New Zealand
contributing to the
problem?
New Zealand vs. other countries
Kyoto Protocol parties vs. others
s 7.1
ch 4
s 6.2
s 6.3
s 6.4
s 6.7
ch 5
ss 4.3, 4.4, 1.3
Figures 1.1-1.3
Which countries
contribute the most?
What are the
international
agreements?
How would
“carbon sinks” &
“emissions trading” work?
UNFCCC
s 1.1
Kyoto Protocol s 1.3
emissions trading
ss 5.2.3, 9.4, 9.8
carbon sinks
ch 5
New Zealand policy ss 2.2, 10.1
What can be done?
targets & principles
national & international
local government
individuals
What has New Zealand
done so far?
Government policy 1990-2001
s 2.2
more detail on Government action ch10
Select Committee inquiries
s 3.2
legislation
s 3.3
Table of Contents ⇒ page v
Executive Summary ⇒ page 1
ch 8
ch 9
ch 11
ch 12
The greenhouse effect and climate change
Parliamentary Library, August 2001
Any views in this report, express or implied, are the author’s and do not necessarily reflect those
of the Parliamentary Library or Parliamentary Service.
Members requiring further information, copies of any of the references, or an oral briefing on this
subject are welcome to contact Dana Peterson at (04) 471-9358.
text complete as of 30 August 2001
published 5 September 2001
Parliamentary Library
Wellington, New Zealand
Special thanks to all of the reviewers of the draft text,
and to Linda Chin for assistance with editing and printing.
Copyright  NZ Parliamentary Library
Except for educational purposes permitted under the Copyright Act 1994, no part of this document may be reproduced or transmitted in any
form or by any means, including information storage and retrieval systems, other than by Members of Parliament in the course of their official
duties, without the consent of the Parliamentary Librarian, Parliament Buildings, Wellington, New Zealand.
iv
The greenhouse effect and climate change
Parliamentary Library, August 2001
CONTENTS
Where to find key information (flow chart)
Lists of tables, figures and boxes
iii
viii
Executive summary
1
Part A: The policy context
1
International climate change agreements
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
2
15
Agencies and Ministers involved
15
A brief summary of Government climate change policy 1990-2001 16
Activities in the House of Representatives
3.1
3.2
3.3
3.4
Part B:
4
5
6
6
10
10
11
11
12
New Zealand climate change policy - an overview
2.1
2.2
3
The United Nations Framework Convention on Climate Change
COP1, COP2, and the Berlin Mandate
COP3 and the Kyoto Protocol
COP4 and COP5: developing the Kyoto Protocol mechanisms
COP6 part one: failure to agree
COP6 part two: compromise and agreement
New Zealand’s participation in UNFCCC negotiations
“Contraction and Convergence”: a possible way forward?
5
Report of the Controller and Auditor-General
Select Committee inquiries
3.2.1 Local Government and Environment Committee 2000:
role of local government in climate change initiatives
3.2.2 Transport and Environment Committee 1998:
environmental effects of road transport
Legislation
3.3.1 Energy Efficiency and Conservation Act 2000
3.3.2 International Treaties Bill (2000)
3.3.3 Road Traffic Reduction Bill (2001)
Energy efficiency in the Parliamentary Buildings
19
19
20
20
23
23
23
25
26
26
Greenhouse gases and sinks
The greenhouse gases
4.1
4.2
4.3
4.4
The “greenhouse effect” and contributing gases
Data uncertainties
New Zealand’s emissions: overview
4.3.1 Gross and net emissions
4.3.2 Mix of greenhouse gas emissions
4.3.3 Changes in total emissions 1990-1999
Emissions data on individual gases
4.4.1 Carbon dioxide (CO2)
4.4.2 Methane (CH4)
27
27
28
29
29
29
31
32
32
36
v
The greenhouse effect and climate change
4.5
5
5.2
5.3
5.4
5.5
Part C:
6.4
6.5
6.6
6.7
39
Overview: the role of forests in the global carbon cycle
5.1.1 The carbon cycle and carbon sequestration
5.1.2 The role of deforestation in climate change
The “Kyoto Forest”
5.2.1 Greenhouse gas reporting and accounting for the land-use
change and forestry sector
5.2.2 Rules governing carbon sinks and trading
Remaining land-use and forestry issues
5.3.1 Adequacy of forestry data for climate change monitoring
and reporting
5.3.2 Forest sinks may not provide permanent sequestration
5.3.3 Forests cannot absorb most of the anthropogenic
CO2 emissions
5.3.4 New Zealand’s “Kyoto Forest” is in private ownership
5.3.5 Climate change effects on forests
5.3.6 Protection of indigenous forests
5.3.7 “Polluter pays” principle not addressed
5.3.8 CH4 and N2O implications
Trends in New Zealand land use and afforestation
Other greenhouse gas sinks
5.5.1 Soil management
5.5.2 Ocean storage
5.5.3 Underground storage
39
39
40
41
42
43
46
46
47
47
48
48
49
51
51
52
54
54
55
55
The estimated risk and impacts of climate change
Summary of IPCC Third Assessment Report Working Group One
Natural vs. anthropogenic influences on the climate
Climate change models, emission scenarios, and climate change
projections
Paleoclimatic data: climate trends over millions of years
Climate changes in Australia and New Zealand over the last
140 years
Climate change projections for New Zealand
The El Niño-Southern Oscillation (ENSO) phenomenon
Summary of IPCC Third Assessment Report, Working Group Two
Likely impacts in New Zealand
7.2.1 Agricultural production
7.2.2 Indigenous species and ecosystems
7.2.3 Freshwater and marine ecosystems
7.2.4 Health
7.2.5 Impacts on Mäori communities
7.2.6 Hydro-electricity generation
7.2.7 Tourism
7.2.8 International links
57
57
59
63
64
65
66
68
Impacts, adaptation and vulnerability
7.1
7.2
vi
37
37
37
Climate Change: current scientific understanding and projections
6.1
6.2
6.3
7
4.4.3 Nitrous oxide (NO2)
4.4.4 HFCs, PFCs, and SF6
Links with depletion of the Ozone Layer
Carbon sequestration or “carbon sinks”
5.1
6
Parliamentary Library, August 2001
71
71
75
75
76
77
77
77
78
78
78
The greenhouse effect and climate change
Part D:
8
83
Addressing “market failure”
83
Removing barriers to energy efficiency and renewable energy 84
Tax policies
85
Emissions trading and quotas
88
Creating a market for “green energy”
89
Financial support through grants and loans
90
Energy efficiency standards and labelling
90
The Clean Development Mechanism and Joint Implementation 93
CDM and the nuclear energy issue
97
Economic instruments
Energy Efficiency and Conservation Authority (EECA)
Energy efficiency planning by government agencies
Public awareness and concern
99
99
100
104
106
Local authority initiatives
11.1
11.2
11.3
11.4
11.5
11.6
12
79
81
81
81
82
82
Additional detail: the New Zealand situation
10.1
10.2
10.3
10.4
11
What is the target?
Primary focus on energy
“Good practice” policies
The global commons
Leadership and assistance from developed countries
“Embeddeness” of issues and the “no regrets” approach
79
National and international initiatives
9.1
9.2
9.3
9.4
9.5
9.6
9.7
9.8
9.9
10
Options for action
Overview: targets and principles
8.1
8.2
8.3
8.4
8.5
8.6
9
Parliamentary Library, August 2001
Legal context
Local authority operations
Roading and transport
Building codes and energy conservation
Waste management
Resource Management Act consents
11.6.1 The Stratford Power Station case
Individual choices
109
109
109
110
111
111
112
113
117
Glossary
123
Appendix Summary of Cabinet papers released on climate change policy
125
References
127
vii
The greenhouse effect and climate change
Parliamentary Library, August 2001
List of Tables
Table 1.1
Table 2.1
Table 4.1
Table 4.2
Table 4.3
Table 4.4
Table 5.1
Table 5.2
Table 5.3
Table 5.4
Table 5.5
Table 5.6
Table 6.1
Table 6.2
Table 6.3
Table 6.4
Table 6.5
Table 7.1
Table 7.2
Table 7.3
Table 8.1
Table 8.2
Table 9.1
Table 9.2
Table 9.3
Table 9.4
Table 9.5
Table 10.1
Table 10.2
Table 10.3
viii
Summary of greenhouse gas emission reduction commitments under the Kyoto
Protocol
Summary of key agency responsibilities in the climate change area
Global Warming Potential (GWP) and lifetime of the greenhouse gases
Estimated uncertainty for greenhouse gas emissions and sinks data, New
Zealand's 1990 baseline inventory
New Zealand CO2 and total greenhouse gas emissions; by weight and percent of
world total
Percent change 1990-1999, New Zealand's greenhouse gas emissions,
population, and real GDP
New Zealand's anticipated assigned amount, projected net forestry carbon sinks,
and potential emissions for 2008-2012
Summary of carbon emission and sink accounting and reporting for the land-use
change and forestry sector
New Zealand's land-use, land-use change and forestry (LULUCF) accounting for
1999
Percent change in total greenhouse emissions data (net compared with gross) for
1998 with inclusion of estimated effects from land-use, land-use change, and
forestry (LULUCF)
The Bonn agreement: caps on volume of carbon sink trading in forest
management credits, by country 2008-2012, in Megatonnes of carbon per year
A possible New Zealand framework for trading sink credits as put forward for
discussion by the New Zealand Climate Change Programme, July 2001
Principal conclusions of IPCC Working Group I relating to climate change
Extreme weather and climate events: estimates of confidence in observed and
projected changes
Summary of the SRES emission scenario storyline groupings, IPCC 2001
Climate change effects observed over the last century in New Zealand and
Australia
Current climate change projections for New Zealand; prevailing winds,
temperature, rainfall, sea-level and heating energy demand
Summary of some projected negative and positive impacts of climate change.
Global scale projections of the IPCC relating to climate change impacts and
vulnerability
Regional IPCC summary of adaptive capacity, vulnerability and key concerns for
Australia, New Zealand, and small island states
Atmospheric CO2 stabilisation scenarios
Share of greenhouse gas emissions from the energy sector, Annex I countries,
1998
Brief summary of results from 2001 analysis of a low level carbon charge for New
Zealand, with and without revenue recycling
Overseas examples of tax policies that provide incentives to decrease
greenhouse gas emissions.
Overseas examples of subsidy, grant and loan programmes for encouraging
energy efficiency and use of alternative and renewable energy
Overseas examples of energy efficiency standards and labelling initiatives.
Examples of pilot Clean Development Mechanism and Joint Implementation type
projects and other pilot emissions trading
Summary of EECA cumulative benefits to 1999-2000
EECA’s Key Output objectives and budget for 2000/01
Summary of the initiatives proposed in the Draft National Energy Efficiency and
Conservation Strategy, 2001
The greenhouse effect and climate change
Table 10.4
Table 11.1
Parliamentary Library, August 2001
Aggregate quantitative results from the UMR telephone survey of public
awareness and concern about climate change
CO2 emissions from the Taranaki Combined Cycle Power Station, electricity
sector emissions 1998-2000, and mitigation measures required under Resource
Management Act consent
List of Figures
Figure 1.1
Figure 1.2
Figure 1.3
The “Top 10” Kyoto Protocol countries for 1990 emissions of CO2
Top contributors of CO2, accumulated contribution for all countries since 1950
Estimated global emissions of CO2 from fuel in 1998, Annex I and non-Annex 1
parties
Figure 4.1
Basic diagram of greenhouse and ozone layer effects
Figure 4.2
Total per capita greenhouse gas emissions, Annex I countries, gross and net of
LULUCF (Gg C per person)
Figure 4.2(a) Gross emissions (land-use, land-use change and forestry (LULUCF) carbon sinks
and emissions not included)
Figure 4.2(b) NET emissions (land-use, land-use change and forestry (LULUCF) carbon sinks
and emissions included)
Figure 4.3
Percentage of each greenhouse gas in New Zealand’s total emissions, 1999 (CO2
Figure 4.4
Figure 4.5
Figure 4.6
Figure 4.7
Figure 4.8
Figure 5.1
Figure 5.2
Figure 5.3
Figure 5.4
Figure 5.5
Figure 5.6
Figure 5.7
Figure 5.8
Figure 5.9
Figure 6.1
Figure 6.2
Figure 6.3
Figure 9.1
equivalent kilotonnes)
New Zealand’s emissions of greenhouse gases, 1990 and 1999 (CO2 equivalent
kilotonnes)
Percent change in greenhouse gas emissions 1990-1998. Original data in tonnes
of CO2 equivalents
CO2 emissions per capita 1998 from fossil fuel combustion: World average,
Annex I and non-Annex I Parties, regional groupings, and selected countries to
illustrate the range of values
New Zealand’s CO2 emissions by source, 1990 and 1999.
New Zealand’s methane emissions by source 1999, and percent change by
sector 1990-1999
Estimated annual carbon fluxes in the global system, Gigatonnes of carbon per
year
Estimated magnitude of natural carbon reservoirs in the global carbon cycle
CO2 emissions from deforestation compared with fossil fuel burning and cement
manufacture, cumulative 1850 to 1998
Atmospheric concentration of CO2 : anthropogenic contributions 1750
to 2100, compared with terrestrial biosphere carbon sink potential
Change in harvest of indigenous timber and clearance of scrub for establishment
of new plantation forests in New Zealand, 1989-2000
Time profile of carbon sequestration over 120 years, kauri compared to radiata
Pine
Trends in New Zealand livestock numbers and lands newly planted in production
forest, 1950 to 2000
Estimated percentage of New Zealand under forest cover, from before human
settlement to the present
Proportion of New Zealand land in forest and other land uses, 2001
Natural and anthropogenic “radiative forcing” factors known to affect climate
Projections based on natural, anthropogenic, and combined climate change
factors, compared to actual observed temperatures 1850 to 2000
Projected changes to temperature and rainfall for New Zealand, 1980s to 2080s
Change from 1991 to 1998 in the share of renewable and waste energy sources
in total primary energy supply (TPES) and total electricity: selected countries
ix
The greenhouse effect and climate change
Figure 10.1
Figure 10.2
Figure 10.3
Figure 12.1
Figure 12.2
Parliamentary Library, August 2001
Actual and projected funding for the Energy Efficiency and Conservation Authority
(EECA), 1993/94 to 2005/06 (GST inclusive, in $1,000, nominal (not adjusted for
CPI))
Total real funding for EECA 1993/94 to 2000/01 (adjusted for CPI, in current
dollar terms)
Ratings of government agency energy-efficiency policy, management, monitoring,
staff training, and funding, 1999
Greenhouse gas emissions from different forms of transport
Energy efficiency information: standby power, fluorescent lights, cooking
modes, and paper
List of Boxes
Box 1
Box 2
Box 3
Box 4
Box 5
Box 6
Box 7
Box 8
Box 9
Box 10
Box 11
Box 12
x
Share of 1990 CO2 emissions for Kyoto Protocol Annex B Parties, by percentage
of total and by UNFCCC negotiating groupings
Summary of agreements reached at COP6 part two, July 2001.
The New Zealand Delegation to the Kyoto Protocol negotiations at COP6, The
Hague, 13-24 November 2000
Conclusions and recommendations of the Controller and Auditor-General relating
to climate change agreements, April, 2001
Local Government and Environment Select Committee Inquiry into the
Role of Local Government in Meeting New Zealand’s Climate Change Target:
Terms of Reference
Recommendations to Government, Local Government and Environment Select
Committee, December 2000
Recommendations from Transport and Environment Select Committee to
Government, September 1998
Examples of local authority initiatives: reducing greenhouse gas emissions
Individual actions - Transport
How many trees do I have to plant to absorb the carbon emitted by my car?
Individual actions - In the home
Individual actions - At work
The greenhouse effect and climate change
Parliamentary Library, August 2001
EXECUTIVE SUMMARY
The greenhouse effect and observed climate changes
•
The “greenhouse effect” is a natural phenomenon in which certain gases in the lower
atmosphere prevent some of the heat energy radiated from the Earth from escaping. The
human-caused emissions of greenhouse gases (CO2, methane, nitrous oxide, and some
industrial gases) have over the last few centuries added to this effect, making global
temperatures warmer than they would otherwise be and affecting global weather patterns.
•
The hole in the ozone layer is a separate phenomenon, but there are a few linkages with the
greenhouse effect. For example, some gases which deplete ozone in the upper atmosphere
(CFCs) also act as greenhouse gases in the lower atmosphere, and the trapping of heat in
the lower atmosphere by the greenhouse effect leads to a cooler upper atmosphere and a
slower recovery time for the ozone layer. (section 4.5)
•
Average global surface temperature has already increased about 0.6°C since 1860.1 The
freeze-free season has lengthened in many regions over 1950-1993. In New Zealand and
Australia, temperatures have risen 0.5 to 0.9°C. (Tables 6.1, 6.4)
•
During the 20th century global sea level has already risen 0.1 to 0.2 metres and rainfall
patterns have changed in many areas. In New Zealand and Australia, sea level has risen on
average about 20 mm per decade over the last 50-100 years and rainfall trends have
followed the cyclical El Niño events. (Tables 6.1, 6.4)
•
The Intergovernmental Panel on Climate Change (IPCC)2 has reported new and stronger
evidence that most of the global warming observed over the last 150 years is attributable to
human activities. If only human or natural influences on the climate are separately
modelled they do not fully explain the historical changes, but there is a good match
for both human and natural influences combined. (sections 6.1, 6.3, Figure 6.2)
•
Before significant human influence, the climate of the Earth alternated between warm and
cold periods over cycles of tens of thousands of years (e.g. the Cambrian and Cretaceous
eras and a number of Ice Ages). However, since the Industrial Revolution human activity has
led to concentrations of CO2 and methane higher than at any time during the past 420,000
years, and CO2 the highest it has been for the last 20 million years. (section 6.4)
Predicted climate changes
•
The world is already committed to some climate change which cannot be avoided, due
to the long life in the atmosphere of the greenhouse gases already emitted over the last few
centuries and the inertia in aspects of the global climate system.
•
Over the next century, there is a 90-99% chance of higher maximum and minimum
temperatures, more hot days, fewer cold and frost days, and reduced daytime temperature
ranges over nearly all land areas.
•
If greenhouse gas emissions are not controlled, the result of 35 modelling scenarios predicts
that global average temperature will increase by 1.4°C to 5.8°C over the period 1990 to
2100, a rate of warming without precedent over the last 10,000 years. Sea-ice, glaciers,
snow cover and ice caps are predicted to decrease, contributing to a global mean sea level
1
As a global average, this includes higher and lower temperatures, including some areas (e.g. Antarctica) which have not warmed.
IPCC reports go through a detailed review process with hundreds of scientific and country representatives and are the best
international scientific consensus statements available on the issue of climate change. The 2001 reports use improved modelling and
careful consideration of the many uncertainties.
2
1
The greenhouse effect and climate change
Parliamentary Library, August 2001
rise of 0.09 to 0.88 metres over 1990-2100, and intense precipitation events (drought and
flood). Tropical cyclones are predicted to increase in some areas. (section 6.1)
•
The impacts are expected to fall disproportionately on the poorest people. Those with the
fewest resources have the least capacity to adapt and are the most vulnerable. (section 7.1)
•
Rainfall predictions for New Zealand arise from the expectation that cyclical El Niño events
will increase or be exacerbated by global climate change. During El Niño events in the
summer there are stronger and more frequent winds from the west, causing more rain in
western areas and more drought on the east coast. In the winter, the wind is more from the
south causing colder conditions. (sections 6.5, 6.6; Figure 6.3 p. 69)
•
Impact scenarios for New Zealand suggest that the resources most vulnerable to future
climate change are likely to include: coastal areas; lands in eastern areas already prone to
drought; crops grown near their current tolerances for temperature or moisture; some Mäori
lands; indigenous species; and ski fields. Possible beneficiaries include forest owners who
may be able to trade carbon sink credits and farmers of crops that may benefit from
increased warmth or CO2 levels. (chapter 7)
•
The present level of scientific knowledge does not allow precise predictions about the nature
and magnitude of human-induced climate change and its effects. Scientific uncertainty
cuts both ways: the actual results may be significantly less, or considerably worse,
than current estimates predict.
Greenhouse gas emissions
•
New Zealand produces only 0.4% of the total greenhouse gases emitted worldwide.
However, on a per capita basis New Zealand is the fourth largest emitter among the
developed countries, exceeded only by Canada, the USA, and Australia. When carbon
sink credits from land use change and forestry are included, New Zealand’s rate of
emissions per capita is 8th highest. (section 4.3.1)
•
For most developed countries, CO2 is the principal greenhouse gas they are contributing to
the atmosphere, but in New Zealand 60% of greenhouse gas emissions are methane and
nitrous oxide, primarily from agricultural activities. However this relationship may reverse
over the next decade as New Zealand’s CO2 emissions are increasing. (section 4.3.2)
•
Comparisons with undeveloped countries can only be done for CO2 emissions from fossil
fuel use and cement manufacture, as accurate global data by country is not available for the
other emissions. On a per capita basis, New Zealand’s CO2 emissions are more than twice
the world average. (section 4.4.1)
•
Over the last decade, New Zealand’s CO2 emissions have increased faster than population
and GDP. However, New Zealand’s total greenhouse gas emissions have increased more
slowly than GDP and population. (section 4.3.3)
•
New Zealand’s emissions of CO2 are below the OECD average if measured by tonnes per
person or by unit of total energy used, in part because a majority (although a decreasing
share) of New Zealand’s electricity is produced using hydro-electricity. However, when
measured against GDP as a measure of economic production, New Zealand’s emissions of
CO2 are above the OECD average, and in the company of Canada, the USA, Australia, and
Korea. (section 4.4.1)
•
New Zealand’s historical contribution to the greenhouse effect includes CO2 from major
clearance of forest over the last 150 years, and methane and nitrous oxide emissions from
the resulting agricultural production. (sections 4.4.2, 4.4.3, 5.2, and 5.3)
2
The greenhouse effect and climate change
Parliamentary Library, August 2001
Carbon sinks
•
A carbon sink is a process where CO2 is removed from the atmosphere, and cannot
contribute to climate change. The largest natural sink is the ocean. Fossil fuels are carbon
reservoirs from ancient warm periods. Carbon sinks that people can easily enhance are
vegetation and soils. Conversely, deforestation and poor soil management add more CO2 to
the atmosphere.
•
Under the Kyoto Protocol, specified land-use, land-use change and forestry (LULUCF)
activities are part of national greenhouse gas emission inventories and will be part of the
“compliance equation”. The net effect may be positive (net increase in emissions) or negative
(carbon sink credits against emissions: this is the case for New Zealand). (section 5.4)
•
On a global basis carbon sinks are not a complete solution: emissions must still be
reduced. Even if all the forests removed from 1850 to 1998 were reinstated, it is only
enough to absorb half of the CO2 emitted globally over that period from the use of fossil
fuels, and CO2 emissions continue to increase. For a few countries including New Zealand,
forestry carbon sinks can compensate for emissions over the medium term. (sections 5.2, 5.3)
•
Countries which have LULUCF credits will theoretically be able to trade these on a world
market. International rules have not yet been set up. Domestic carbon emissions trading
trials have commenced in Canada and Denmark, and options are being discussed for New
Zealand. Cabinet has agreed in principle that domestic LULUCF credits would be able to be
traded on the international market, some proportion of the benefits would go to those
undertaking the sink activities, and those emitting greenhouse gases would have to pay the
international market price for carbon sink credits if they required them to meet any future
domestic emission quotas. (sections 5.2 and 9.4)
International agreements to reduce greenhouse gases
•
New Zealand ratified the United Nations Framework Convention on Climate Change
(UNFCCC) in 1993. The objective of the UNFCCC is to achieve “stabilisation of greenhouse
gas concentrations in the atmosphere at a level that would prevent dangerous
anthropogenic3 interference in the climate system.” The UNFCCC parties agreed that in
pursuit of this goal they would initially pursue measures to return their emission of
greenhouse gases, individually or jointly, to 1990 levels. This target was not met. (section 1.1)
•
The Kyoto Protocol is an agreement under the UNFCCC which New Zealand signed in 1998.
When it becomes effective, it will require the developed countries to reduce their greenhouse
gas emissions in aggregate at least 5% below 1990 levels over the period 2008-2012.
However, each country has a different commitment, and for New Zealand it is to reduce
emissions back to 1990 levels. (section 1.3)
•
The Kyoto Protocol will not be legally binding until it is ratified by at least 55 Parties to the
UNFCCC, incorporating developed country parties to the UNFCCC which account for 55% of
total CO2 emissions for 1990. So far, 35 of the 84 signatories have ratified it, but only one
(Romania) can be counted toward the 55% of emissions ratification requirement. (section 1.3)
•
Because global greenhouse gas emissions have continued to increase since 1990, the
actual reduction to meet the Kyoto Protocol target will have to be more than 5% of 1990
levels. Compared to the expected global emission levels for 2000 with current trends, the
total reductions required at present would actually be about 10%. By 2010 the total reduction
required at the same rate of increase would be 29-30%. For each country, it will vary by their
individual target and rates of emissions increase.
3
anthropogenic = caused by people (as opposed to natural forces)
3
The greenhouse effect and climate change
•
Parliamentary Library, August 2001
The recent agreement in Bonn has clarified how some details of the Kyoto Protocol will work
(e.g. funding, use of carbon sinks, emissions trading, the “clean development mechanism”,
and enforcement), but technically has not changed the aggregate 5% target.
(section 1.6, also see i-brief 2001/6)
•
The Kyoto Protocol would not require greenhouse gas emission controls for developing
countries, which in the late 1990s contributed 39% of the global CO2 emissions from fuel
combustion. However, on a per capita basis the developing countries produce only 1.85
tonnes of CO2 per person, compared with 11 tonnes per person in developed countries.
(section 1.3, Figures 1.3 and 4.6)
•
The 5% below 1990 emissions target, even if met, may not be enough to meet the
UNFCCC goal of stabilising greenhouse gases at a level that will prevent dangerous
interference in the climate system. Emissions from developing countries will also need to
be addressed in due course, and the longer it takes the developed countries to stabilise at
the target level the more emissions will accumulate in the atmosphere and contribute to
climate change. (section 8.1)
•
Climate change scenarios with atmospheric CO2 concentrations of 550 ppmv, or about twice
the pre-industrial levels, would have about two-thirds the climate impact on new Zealand as
“business as usual”, or no attempt to control emissions. This would requite much more
stringent controls that currently required by the Kyoto Protocol. (section 8.1)
New Zealand Government action
•
The Auditor-General reported in April 2001 that New Zealand is meeting its UNFCCC
obligations except the first and most important one, to formulate and implement national
policies to mitigate climate change through limiting human-induced emissions of greenhouse
gases. A range of policy measures has been adopted, but the measures have been
ineffective. (section 3.1)
•
The national target announced in 1994 was to stabilise net CO2 emissions at 1990 levels by
the year 2000, 20% through emission reductions (voluntary agreements with industry, a
national strategy and deregulation of the energy sector and a more competitive wholesale
electricity market), and 80% from new carbon sinks. This target was not met. (section 2.2)
•
The 1994 policy also had provision to introduce a low level carbon charge (e.g. carbon tax or
similar) if by mid-1997 the policy measures were not on track to achieve the target. This
charge was postponed in 1997 and again in 1999. In 2001 it was referred to the Tax Review
which is scheduled to report by the end of September 2001. The Government has
announced that if the Tax Review recommends a carbon charge, it would not be introduced
until after the next election. (section 2.2)
•
In 1992 the Energy Efficiency and Conservation Authority (EECA) was established by
Cabinet, and subsequently established as a Crown entity under the Energy Efficiency and
Conservation Act 2000. This agency has initiated and encouraged many voluntary energy
efficiency and conservation activities in the industry, government, community, and household
sectors. The first national Energy Efficiency and Conservation Strategy under the Act is due
by 1 October 2001. (section 2.2)
•
The New Zealand Government has announced its intent to ratify the Kyoto Protocol in
September 2002. The Prime Minister has stated that although New Zealand’s contribution to
global climate change was relatively small, “we must lead by example and encourage other
countries to participate actively.” (section 2.2)
4
Part A: The policy context
1
1.1
International climate change agreements
The United Nations Framework Convention on Climate Change (UNFCCC)1
The first report of the Intergovernmental Panel on Climate Change (IPCC) in 1990 concluded
that human-induced climate change was a real threat.2 In response, the United Nations General
Assembly convened a series of meetings which culminated in the adoption of the United Nations
Framework Convention on Climate Change (UNFCCC) at the “Earth Summit” in Rio de Janeiro,
Brazil, in May 1992.
New Zealand ratified the UNFCCC on 16 September 1993. It came into force in March 1994,
and at 24 April 2001 it had been ratified by 186 countries.
The ultimate objective of the UNFCCC is to achieve:
… stabilisation of greenhouse gas concentrations in the atmosphere at a level that
would prevent dangerous anthropogenic interference in the climate system.
Such a level should be achieved within a time-frame sufficient to allow ecosystems
to adapt naturally to climate change, to ensure that food production is not
threatened and to enable economic development to proceed in a sustainable
manner.3
The principles of the UNFCCC recognise the need for developed countries to take the lead in
combating climate change (Article 3). Currently 29 developed countries (including New Zealand)
and 13 countries undergoing the process of transition to a market economy are listed in Annex I
of the UNFCCC. These countries agreed to take appropriate measures with the aim of:
…returning individually or jointly to their 1990 levels these anthropogenic
emissions of carbon dioxide and other greenhouse gases not controlled by the
Montreal Protocol.4
The parties also agreed to set up and regularly report on national inventories of emissions,
develop emission reduction programmes and report on progress; cooperate in scientific and
technical initiatives, education, information exchange, and adaptation strategies; and promote
sustainable management of sinks and reservoirs. The Convention recognises the need for
developed countries to assist developing countries with funding and technology to help reduce
emissions and cope with adverse climate change effects, and the 25 countries listed in Annex II
(including New Zealand) agreed to help fund this.5
Under this framework, an initial commitment was made to reduce emissions to 1990 levels by
the year 2000, and to review the adequacy of that commitment. Under UNFCCC regular
Conference of Parties (COP) meetings are held to develop policies and monitor progress.
1
Sources for sections 1.1 and 1.2 include the UNFCCC website http://www.unfccc.de ; Ministry for the Environment 1998 pp. 5-6,
Ministry for the Environment 1999 p. 22.
2
The subsequent Third Assessment Report (TAR) in 2001 provided substantial new evidence to further document this concern.
These findings are summarised in this report in chapters 6 and 7.
3
Article 2. Quotes from the full text of the Convention, on http://www.unfccc.de/resource/conv/conv_004.html
4
Article 4.2b. The Montreal Protocol controls gases known to deplete the ozone layer, some of which (e.g. CFCs and HFCs) also act
as greenhouse gases in the lower atmosphere. See also section 4.5 for more information.
5
New Zealand is in both Annex I and Annex II of the UNFCCC.
The greenhouse effect and climate change
1.2
Parliamentary Library, August 2001
COP1, COP2, and the Berlin Mandate
The first Conference of Parties (COP1) in April 1995 concluded that not only would the original
UNFCCC target not be met, but also that those commitments were not sufficient to prevent
dangerous human interference with the climate. The Berlin Mandate was agreed to, which set up
a process to develop additional commitments for developed countries beyond the year 2000.
At COP1 the Parties also agreed to establish a pilot phase for Activities Implemented Jointly
(AIJ), to operate under the Subsidiary Body for Scientific and Technical Advice (SBTA) in
coordination with the Subsidiary Body for Implementation (SBI). This is a precursor to Clean
Development Mechanism (CDM) and Joint Implementation (JI) activities referred to later in the
Kyoto Protocol.6
At COP2, in Geneva in July 1996, the Geneva Ministerial Declaration was accepted by most
ministers and heads of delegation. This endorsed the scientific advice of the IPCC that “there is
already a discernable human impact on global climate” and called for legally binding
commitments.
1.3
COP3 and the Kyoto Protocol
COP3 was held in Kyoto, and after intense negotiation the Kyoto Protocol was agreed to on 11
December 1997. In order to reach agreement, a number of matters were left to be worked out in
more detail later.
New Zealand signed the protocol in May 1998, but has not ratified it. Government intends to
ratify it in 2002 (section 2.2).
Under Article 3 and Annex B, the Protocol sets greenhouse gas emission targets for 33
countries are to achieve during the period 2008 to 2012. The aggregate reduction for these
countries is “at least 5%” below the 1990 emissions level (Article 3.1), but different countries
have different targets. In addition, signatory countries undergoing transition to a market economy
are allowed to have different base years than 1990 (Article 3.5). The emission reduction targets
set out in the Protocol are summarised in Table 1.1.
The “assigned amount” for emissions for each country for the first commitment period 2008-2012
is the base year gross emissions of all greenhouse gases in CO2 equivalents, multiplied by the
reduction target, and multiplied by five (for the five years 2008-2012). Emissions data continues
to be refined, and the base year level has not yet been “frozen”. If the current data is used, New
Zealand’s assigned amount would be about 365 m tonnes for 2008-2012.7
The assigned amounts are based on gross emissions, but accounting for those emissions in
2008-2012 will be based on net emissions, to allow for emissions and sinks from specified landuse (afforestation, reforestation and deforestation). The 1990 base year was selected as a
partial counter to a potential “gross-net loophole” (see section 5.2).
Although the aggregate global target is about 5% below 1990 emission levels for the developed
countries, the actual reduction required will be more than that. This is because for the Annex B
countries overall greenhouse gas emissions have continued to increase since 1990. Compared
to the expected emission levels for 2000 with current trends, the total reductions required will
actually be about 10%. By 2010 the total reduction required at the same rate of increase would
be 29-30%.8
6
Decision 5/CP.1, http://www.unfccc.de/program/aij/aij_back.html .
1990 emissions of 73,064.35 Gg x 5 = 365,321.75 Gg = 365 Mt (million tonnes). A recent Government publication has used 363 Mt
(New Zealand Climate Change Programme 2001a, p. 7). Emissions data is in chapter 4, and a comparison of emissions and forestry
carbon sinks in chapter 5 (section 5.2).
8
UNFCCC Secretariat 1997, UNEP 1998,
7
6
The greenhouse effect and climate change
Parliamentary Library, August 2001
The Protocol also provides for development of “flexibility mechanisms”, or ways to meet
emission reduction targets other than reducing domestic emissions at source. These are:
•
•
•
“joint implementation” (emission reduction from projects shared between Annex I countries Article 6); and the ability of countries to form a “bubble” to share targets (Article 4);
a “clean development mechanism” (CDM), or obtaining emissions credits through helping
developing countries move to cleaner technology (Article 12), and;
emissions trading (Article 17).
The practical details, particularly with regard to transparency, efficiency, and accountability, were
to be worked out in subsequent COP sessions.
The Kyoto Protocol will be legally binding once it has been ratified. It will not enter into force until
the 90th day after the date on which not less than 55 Parties to the UNFCCC, incorporating
sufficient Annex I Parties to the UNFCCC to account for at least 55% of total CO2 emissions for
1990, have deposited their instruments of ratification.
As of 11 June 2001, 84 countries had signed the Kyoto Protocol (originally, or by undertaking
accession) and 35 countries had ratified it. However, only one of the ratifying countries,
Romania, is required by the Protocol to reduce emissions and can be counted toward the 55% of
CO2 emissions required for the Protocol to take effect.9
A summary of the UNFCCC parties and their contribution to the 1990 CO2 emissions is shown in
Box 1 and Figure 1.1. The largest contributor to CO2 emissions is the USA (33.9%), but
ratification of the Protocol could theoretically be achieved without the USA (55% of emissions
must be represented).
When signing the Protocol the Cook Islands, Niue, and Kiribati, countries which stand to suffer
significant adverse impacts from rising sea levels, declared that their signing and ratifying:
“in no way constitute a renunciation of any rights under international law concerning
state responsibility for the adverse effects of climate change and that no provision in
the protocol can be interpreted as derogating from principles of general international
law”.
The Cook Islands also declared that: “in light of the best available scientific information and
assessment on climate change and its impacts, it [the Cook Islands] considers the emissions
reduction obligation in Article 3 of the Kyoto Protocol to be inadequate to prevent dangerous
anthropogenic interference with the climate system.”10
9
UNFCCC Secretariat http://www.unfccc.de/resource/convkp.html . The countries that have ratified or joined by accession are
(Annex I Parties in bold): Antigua and Barbuda, Azerbaijan, Bahamas, Barbados, Bolivia, Cyprus, Ecuador, El Salvador, Equatorial
Guinea, Fiji, Gambia, Georgia, Guatemala, Guinea, Honduras, Jamaica, Kiribati, Lesotho, Maldives, Mauritius, Mexico, Micronesia,
Mongolia, Nicaragua, Niue, Palau, Panama, Paraguay, Romania, Samoa, Trinidad and Tobago, Turkmenistan, Tuvalu, Uruguay and
Uzbekistan. When parties join the Protocol after its original signing by “accession”, it is equivalent to ratification.
10
UNFCCC Secretariat, Declarations made by Parties upon signature, web address as in previous footnote.
7
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 1.1:
Summary of greenhouse gas emission
reduction commitments under the
Kyoto Protocol.
Annex B countries
grouped by emission
target type
Emission
reduction
commitment
% change of
base year
emissions
less than 1990 levels
the European Community,
and most of the Eastern
European countries
USA
Canada, Japan, Hungary,
and Poland
Croatia
stabilised at 1990 levels
New Zealand, Russian
Federation, and Ukraine
greater than 1990 levels
Australia
Iceland
Norway
-8%
Box 1:
Share of 1990 CO2 emissions
for Kyoto Protocol Annex B Parties,
by percentage of total
and by UNFCCC affinity groupings
European Union
Belgium
Denmark
Finland
France
Germany
Greece
Ireland
-7%
-6%
Italy
Liechtenstein
Luxembourg
-5%
Monaco
Netherlands
0
Portugal
Spain
+ 8%
+10%
+ 1%
Source: Kyoto Protocol, Annex B (on http://www.unfccc.de )
Note: full list of Annex B countries is in Box 1.
Sweden
Switzerland
United Kingdom
Eastern Europe
Bulgaria
Croatia
Figure 1.1:
The “Top 10” Kyoto Protocol countries for
1990 emissions of CO2
Czech Republic
Estonia
Hungary (*)
Latvia
Lithuania
Poland
Canada
3.2%
UK
4.0%
Poland
3.3%
Romania
Italy
3.0%
Slovakia
France
2.7%
Slovenia
United
States
33.9%
Ukraine
Canada
Iceland (*)
Germany
7.0%
Japan
New Zealand
Norway
Japan
7.8%
Others
13.9%
8
Russian
Federation
16.4%
13.15%
0.72%
not reported
1.14%
0.26%
0.58%
0.17%
0.27%
3.29%
1.34%
0.43%
0.10%
4.86%
The "Umbrella Group"
Australia
Ukraine
4.9%
23.24%
0.43%
0.79%
0.36%
0.42%
2.67%
7.00%
0.59%
0.22%
2.98%
0.001%
0.09%
0.001%
1.11%
0.30%
1.56%
0.38%
0.31%
4.03%
Austria
63.60%
1.92%
3.21%
0.01%
7.76%
0.18%
0.24%
(*) denotes countries which are Annex I parties to the
UNFCCC, but did not sign the Kyoto Protocol. Belarus and
Turkey are Annex I Parties not in Annex B of the Protocol.
Source: UNFCCC emissions data tables, A.3, on
http://www.unfccc.de/resource/ghg/tempemis2.html
The greenhouse effect and climate change
Parliamentary Library, August 2001
Figure 1.2:
Top contributors of CO2,
accumulated contribution for all countries since 1950
Central
America and
Carribean
2%
South Oceania China
America
1%
8%
3%
USA
27%
India
2%
Japan
5%
Rest of Asia
6%
Germany
6%
Canada
2%
Sub-Saharan
Africa
2%
Middle East
& North
Africa
4%
Rest of
Europe
22%
Russian
Fedn.
10%
CO2 emissions from fossil fuel burning and cement manufacture,
original data in 1,000’s of metric tonnes.
Source: World Resources Institute et al. 2000, Data Table AC.2
Figure 1.3:
Estimated global emissions of CO2 from fuel in 1998,
Annex I and non-Annex I parties
nonAnnex I
39%
Annex I
61%
Total CO2 emissions from fuel combustion
Source: International Energy Agency 2000, Table 1.
Notes: This does not include the other greenhouse gases:
reliable global data for this is not available.
Annex I parties are the only countries with greenhouse gas
emission reduction commitments under the UNFCCC, and
all of the Annex I countries except Belarus and Turkey also
have commitments under Annex B of the Kyoto Protocol.
9
The greenhouse effect and climate change
1.4
Parliamentary Library, August 2001
COP4 and COP5: developing the Kyoto Protocol mechanisms
COP4 was held in Buenos Aires, Argentina, in November 1998. The Conference agreed to the
Buenos Aires Plan of Action, which established deadlines for finalising the outstanding details of
the Kyoto Protocol so that it could be fully operational once ratified and entered into force. The
Action Plan addressed the “flexibility mechanisms” as well as compliance issues, monitoring,
and the transfer of climate-friendly technologies to developing countries.11
COP5 was held in Bonn, Germany, in November 1999. Substantive issues on which agreements
were reached included improving the rigour of national reports and improving guidelines for
measuring greenhouse gas emissions. The process for negotiators leading up to COP6 was also
agreed on, in order to make it possible to finalise regimes for compliance, capacity-building,
international emissions trading, Joint Implementation (JI), and a Clean Development Mechanism
(CDM), and to progress resolution of accounting for greenhouse gas “sinks” and assess adverse
effects on developing countries.12
1.5
COP6 part one: failure to agree13
In November 2000, the Parties met at The Hague, in the Netherlands, hoping to agree on final
details on implementation of the Kyoto Protocol. The meeting was suspended without reaching
agreement, and rescheduled for 2001.
The meeting faced a difficult challenge given the scale and complexity of the issues involved.
There were some wide differences of view in developing countries, principally between the
“Umbrella Group” and the European Union (EU), as well as between developed and developing
countries. Late in the meeting a compromise emerged on several key issues but the full
membership of the EU could not agree. Sticking points included USA demands for forest sink
credits through sponsorship in other countries (Cleaner Development Mecanism), EU demand
for domestic action to to be a significant part of meeting targets (“supplementarity”) and some
aspects of compliance.
Other disagreements arose about whether developed country sponsorship of nuclear power
projects in developing countries should be granted greenhouse emission credits under the Clean
Development Mechanism (CDM). The parties agreed that under the CDM, “Annex I Parties will
declare that they will refrain from using nuclear facilities for generating certified emission
reductions under CDM”, and that priorities for CDM will be renewable energy and energy
efficiency improvements.14
New Zealand opposed the USA position (mildly in public, reportedly more strongly in closed
session), while continuing to support the principle of carbon sinks as a valid domestic
mechanism.
Also at COP6, representatives of 134 developing countries demanded a stronger voice in
negotiations and a detailed programme whereby industrialised nations would transfer clean
energy technologies to developing countries as well as extra funds to adapt to climate change
impacts.
11
UNFCCC Secretariat, press release 14/11/98, Climate Change meeting adopts Buenos Aires Plan of Action, http://www.unfccc.de
12
UNFCCC Secretariat, press release 5/11/99, Ministers pledge to finalise climate agreement by November 2000,
http://www.unfccc.de
13
Sources for this section include: CNN, 24 November 2000, EU rejects compromise climate deal; CNN, 21 November 2000,
Nations in standoff over issues at global warming conference; The Press (editorial), 29 November 2000, Shirking a commitment;
Environmental Defence Society 2000.
14
UNFCCC Secretariat 2001a, p. 9.
10
The greenhouse effect and climate change
1.6
Parliamentary Library, August 2001
COP6 part two: compromise and agreement
The continuation of the suspended COP6 took place on 16-27 July 2001 in Bonn, Germany.
Between the two COP6 meetings, there had been a reversal of USA policy, with President Bush
announcing in March 2001 that he did not support the Kyoto Protocol and would not require USA
power plants to cap CO2 emissions, contrary to previous campaign promises.15 The USA
produces more greenhouse gas emissions than any other Annex I Party, or indeed any country
in the world (see Figures 1.1 to 1.3). The USA stance was widely decried by almost all countries.
The Council of Europe’s Parliamentary Assembly unanimously adopted a resolution that said
this decision “casts doubt on the credibility of the United States as a reliable partner prepared to
shoulder its share of the responsibility.” 16
Part two of COP6 prominently featured UNFCCC Parties determined to move forward with the
Protocol despite the USA withdrawal.17 Broad political agreement was reached on the
“operational rulebook” for the Protocol, with further detail to await COP7 and possibly
subsequent meetings. A summary of the agreements that were reached is in Box 2.
The media has reported “environmental campaigner” sources as stating that the new agreement
will effectively reduce greenhouse gas emissions by 2% of 1990 levels by the year 2012, rather
than the 5% the Kyoto Protocol originally envisioned.18 This is not an explicit part of the Bonn
agreement, and is presumably based on extrapolations of the use of the allowed carbon sinks
and emissions trading rather than cutting greenhouse gas emissions per se. Generally
environmental groups praised the agreement as an essential (although small) step forward.19
The next Conference of the Parties, COP7, is scheduled to be held in Marrakech, Morocco, from
29 October to 9 November 2001.20
1.7
New Zealand’s participation in UNFCCC negotiations
New Zealand is part of the “JUSSCANNZ” group of non-European Union industrialised countries
which meet as a group to discuss various issues. The name of the group comes from the
countries involved: Japan, the USA, Switzerland, Canada, Australia, Norway, and New Zealand.
Iceland, Mexico, and the Republic of Korea may also attend meetings.21
Most, but not all, of these countries plus Russia comprise the “Umbrella Group”. which often but
not always have joined ranks to negotiate at UNFCCC meetings. The Umbrella Group at COP6
part I comprised Australia, Canada, Iceland, Japan, New Zealand, the Russian Federation,
Norway, and the USA.22 At COP6 part two, New Zealand, Norway and Iceland reportedly broke
away from Umbrella Group positions.23 While the USA, Russia, Canada, and Japan are among
the “top 10” emitters (Figure 1.1), New Zealand, Norway, and Iceland do not fall into this
category.
15
Letter of 13 March 2001 from President Bush to Senators Hagel, Helms, Craig and Roberts, on
http://usinfo.state.gov/topical/global/environ/climate/01031401.htm ; Ambassador Johnson 2001, Statement on Kyoto Protocol and
Climate Change, 5 April 2001, http://usinfo.state.gov/topical/global/environ/latest/01040601.htm .
16
Associated Press 26 April 2001, Euro Council Criticizes U.S. on Kyoto, on http://dailynews.yahoo.com
17
The USA was present at the negotiations, but was not party to the agreements reached.
18
E.g. BBC News 23/7/01, The Bonn deal: winners and losers at http://news.bbc.co.uk
19
E.g. WWF Climate Change Campaign (http://www.panda.org/climate/victory.htm ), Greenpeace (http://www.greenpeace.org),
Friends of the Earth (http://www.foei.org ). The agreement has been dubbed “Kyoto Lite” by those who hoped for more.
20
UNFCCC 2000, Press release: Morocco to host next climate change conference in 2001, http://www.unfccc.de Further information
on COP7 will be available on that web site, and the host country web site http://www.marrakech-web.net/cop7 .
21
UNEP 1998, COP4 Press Kit, Glossary Part I, the Players.
22
Environmental Defence Society Inc 2000, p. 4.
23
Press release 27/7/01, on http://www.greenpeace.org
11
The greenhouse effect and climate change
Parliamentary Library, August 2001
Another major negotiating block is the “Group of 77 plus China” which is comprised of
developing countries, including the Middle Eastern oil producing nations.
New Zealand’s position in international negotiations is not generally reported in detail.
Incomplete clues are available from Ministerial announcements and published comments of
outside observers.
New Zealand, together with Saudi Arabia, the United States, Australia, Japan, Kuwait, Nigeria,
and Turkey, was criticised by the Climate Action Network (a global grouping of climate- change
activists) for its obstructive stance at COP5. At COP6, New Zealand was criticised for wanting
to remove references to existing international environmental agreements from the eligibility rules
for land-use, land-use change, and forestry, together with Australia, Canada and the United
States. New Zealand has also been awarded the “Fossil of the Decade” award. 24
Such criticisms have been publicly countered by Government. In particular, it has been
categorically stated that:
•
“New Zealand amongst other nations has been accused of trying to ‘twist’ interpretation of
the Kyoto treaty to enable OECD countries to increase emissions 15 to 20 percent. New
Zealand intends no such thing and is in fact advancing proposals designed to limit windfall
gains.
•
A further bizarre accusation is that New Zealand, amongst others, favours ‘loopholes’ in rules
for forestry that would give incentives to chop down old-growth forests and replace them with
new plantations that would generate carbon sink credits. This is utterly untrue. New Zealand
has always opposed deforestation and our proposals in this area explicitly preclude such an
outcome.
•
New Zealand has been grossly misrepresented as supporting the use of nuclear power
under the provisions of the Kyoto Protocol’s Clean Development mechanism ... New Zealand
has a proud and well-known record of anti-nuclear advocacy and has never spoken in favour
of nuclear energy at climate change negotiations.” 25
1.8
“Contraction and Convergence”: a possible way forward? 26
The Kyoto Protocol does not require emission reductions from the developing countries, some of
which like China and India produce significant amounts of CO2 (Figures 1.2 and 1.3). This is one
of the arguments that has been used by the USA to support its withdrawal from the Protocol.
On a per capita basis, the developing countries produce only a fraction of the CO2 that
developed countries do (Figure 4.6), but this is predicted to change over time as they seek to
improve their standard of living and their populations increase.
The developing countries argue that the developed countries have grown rich exploiting fossil
fuels and creating the majority of greenhouse gas emissions, and thus should take the lead in
cutting emissions now without seeking to impose equal responsibilities on the rest of the world
without equal economic rights. Developed countries have taken the lead under the terms of the
Berlin mandate in 1995 and by accepting legally binding targets under the Kyoto Protocol.
The Clean Development Mechanism and the new Expert Group on Technology Transfer under
the Kyoto Protocol are designed to help transfer “clean technology” to developing countries so
24
Speech by the Rt Hon Helen Clark,NZ aims to ratify Kyoto protocol on climate change by mid-2002, 8/5/00, p. 1 para 8;
http://www.fossil-of-the-day.org . The concern of the Climate Action Network was that would mean land use activities given
greenhouse gas emission credits under the “flexibility mechanisms” of Article 3 of the Kyoto protocol would not be required to
conform with the Conventions on Biological Diversity, Desertification, Wetlands, and Forests, Agenda 21, or the ILO.
25
NZ committed to reducing greenhouse gas emissions, http://www.executive.govt.nz/speech?speechralph=31390&SR=0
26
Further details available at Global Commons Institute website http://www.gci.org.uk.
12
The greenhouse effect and climate change
Parliamentary Library, August 2001
that they can begin to disengage their economies from reliance on fossil fuels, but this is only on
a project-by-project basis.
13
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 2:
Summary of agreements reached at COP6 part two, July 2001
Flexibility mechanisms vs. domestic actions
Emissions trading and use of Clean Development Mechanism should be supplemental to
domestic action: i.e. domestic action shall constitute a significant element of the effort made by
each Party (VII.1.5). However, there are no quantitative restrictions on the use of flexibility mechanisms.
•
Land-use, land-use change, and forestry
There are caps on credits available from forestry management, different for each Party: see
Appendix Z (VII.6(c)). Russia, Japan and Canada negotiated special concessions in this area.
There are no caps on other eligible land-use activities (e.g. cropland management, grazing land
management and revegetation) (VII.4).
•
•
International emissions trading
Each Annex I Party should be required to maintain in its national registry (e.g.: not trade) a
commitment period reserve of at least 90% of the Party’s assigned amount or 100% of 5 times
its most recently reviewed inventory, whichever is lesser (recommendation only, VI.4.1).
Countries which fail to meet their 2008-2012 commitments will have their eligibility to participate
in emissions trading suspended (VIII.2(d)).
•
•
Rules for the Clean Development Mechanism (CDM)
Afforestation and reforestation are the only eligible LULUCF (land-use and land-use change and
forestry) activities for CDM credits in the first commitment period (VII.3.8).
There is to be a cap on CDM credits from eligible LULUCF activities for meeting a Party’s
emission reduction commitments (1% of a Party’s base year emissions x 5) (VII.8).
Whether CDM projects contribute to “sustainable development” is to be defined by the host
country (VI.2.1 and VI.3.1).
There is to be a 2% levy on CDM emissions credits, to support the Adaptation Fund for
developing countries under the Kyoto Protocol (section II and VI.1.10).
Parties “shall refrain from using certified emission reductions from nuclear facilities to meet their
commitments” (VI.2.2 and 3.2).
The new CDM Executive Board to recommend to the COP8 meeting simplified procedures for
small projects that involve renewable energy, energy efficiency and other anthropogenic
emission reduction (VI.6).
Technical advice on issues relating to forestry credits under CDM such as non-permanence,
additionality, leakage, uncertainties, and socio-economic and ecological impacts (including
biodiversity and natural ecosystems) to be provided for the first COP session after the Kyoto
Protocol comes into force (VII.9).
•
•
•
•
•
•
•
Compliance mechanisms (part VIII)
Consequences for failure to meet commitments under the Kyoto Protocol shall include:
owing 1.3 tonnes in the second commitment period (starting 2013) for each tonne of
commitment not met in the first commitment period (2008-2012);
- being required to prepare a compliance action plan;
- suspension of eligibility to participate in emissions trading; and
- compliance committee to have a facilitative branch (to assist compliance) and an enforcement
branch to deal with failures to comply).
Additional procedures and mechanisms to be developed after the Kyoto Protocol comes into
•
force.
•
-
New supervisory bodies
•
Three new groups to be established, an Expert Group on Technology Transfer, a CDM
Executive Board and a Compliance Committee. The membership to represent Annex I and nonAnnex I Parties to the Protocol, the five global regions, and island states. This potentially gives
greater voting rights to the developing and vulnerable countries (sections III.2, VI.3.5 and VIII.6).
Financial and technological support for developing countries
•
Three new (voluntary) funding initiatives to be established focusing on least developed
countries, adaptation to climate change and transfer of clean technology (sections I and II).
Sources: UNFCCC Secretariat 2001c, Decision 5/CP.6 and 2001d, Press release 23 July 2001.
The section references refer to Decision 5/CP.6.
14
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 3:
The New Zealand Delegation to the Kyoto Protocol negotiations at COP6,
The Hague, 13-24 November 2000
Members of Parliament
Minister of Energy and Forestry
Co-Leader, Green Party
Energy Spokesperson , National Party
Hon. Pete Hodgson
Jeanette Fitzsimons
Pansy Wong
Officials
Private Secretary, Minister of Energy
Ministry of Foreign Affairs and Trade
Environment Division (4)
Diplomatic staff (2)
Ministry for the Environment
Climate Change Group (4)
Maruwhenua (1)
Ministry of Economic Development
Resources and Networks (1)
Environmental Issues (1)
Ministry of Agriculture and Forestry
Sustainable Resource Use Policy (2)
The Treasury
Environment, Science & Technology (1)
Regulatory and Tax Policy (1)
Ministry of Research, Science and Technology
Chief Scientific Adviser (1
Ministry of Mäori Development
Economic Development section (1)
Others
National Institute of Water & Atmospheric Research (1)
Natural Resource Users’ Group (1)
Forestry Industry Council (1)
Environmental Defense Society (2)
Source: UNFCCC 2000, List of participants, http://cop6.unfccc.int/pdf/lopcop6.pdf
The “Contraction and Convergence” proposal from the Global Commons Institute could provide
a straightforward and equitable path out of this dilemma. It has reportedly been endorsed by the
UK Royal Commission on Environmental Pollution (June 2000 report), some of the European
leaders (e.g. Jan Pronk of the Netherlands and President Chirac of France), and representatives
of developing countries and environmental, business, and other lobby groups.
“Contraction” refers to a significant contraction in global emission levels over time, significantly
larger than those envisioned by the Kyoto Protocol in the first commitment period.
“Convergence” refers to allocation of emission rights initially proportional to country income and
population, which over time will coverage to a standard world value. The contraction targets,
levels of emission rights, and convergence year would be up for negotiation in order to forge a
global agreement.
Initial per capita emission rights would be based on population for a set year (e.g. 1990). This
would remove incentives for increasing population in order to increase emission rights.
International emissions trading would also be part of the scenario. Countries would have an
incentive to develop using greenhouse-friendly technologies in order to have surplus emission
entitlements to trade.
15
2
New Zealand climate change policy - an overview
2.1
Agencies and Ministers involved
Climate change policy is currently the responsibility of a group of Ministers convened by the
Minister of Energy, the Hon. Pete Hodgson. Advice is provided by working groups of officials
from relevant government departments, ministries and agencies, coordinated by the Department
of Prime Minister and Cabinet. The ministerial portfolios and departments/agencies involved
are:1
•
•
•
•
•
•
•
•
•
•
Prime Minister and Cabinet
Environment
Foreign Affairs and Trade
Treasury
Agriculture and Forestry
Economic Development
Te Puni Kokiri
Transport
Research, Science and Technology
Energy Efficiency and Conservation Authority
The responsibilities of the key agencies are shown in Table 2.1.
Table 2.1:
Summary of key agency responsibilities in the climate change area.
Ministry for
the
Environment
Ministry of
Foreign
Affairs and
Trade
•
•
Ministry of
Economic
Development
•
Ministry of
Agriculture
and Forestry
•
•
•
•
•
Taking a leading role in international negotiations on climate change.
Collating information on New Zealand’s greenhouse gas emissions and
sinks, and providing reports to the FCCC Secretariat.
Co-ordinating participation, and leading negotiations, in international
forums on climate change.
Co-ordinating papers to Cabinet for approval of negotiating positions.
Gathering and analysing information on the positions taken by other
countries in negotiations.
Providing advice to the Climate Change Steering committee, Ministers,
and others on matters such as:
- energy and resource markets;
- the impact of environmental and conservation policies on
business; and
- the use of economic instruments to achieve environmental
outcomes.
Conducting research on New Zealand’s climate change position in
regard to agriculture and forestry.
Supporting New Zealand’s climate change position on agriculture and
forestry at international climate change meetings.
Source: Controller and Auditor-General 2001, pp. 93-96; H. Plume (MFE) pers comm 8/2001.
1
Ministry for the Environment, Developing solutions, on http://www.mfe.govt.nz/issues/ccsolutions.htm .
The greenhouse effect and climate change
2.2
Parliamentary Library, August 2001
A brief summary of Government climate change policy 1990 - 2001
On 4 June 1992, New Zealand signed the UNFCCC. The same month, after criticism of
Government’s energy policy by the Parliamentary Commissioner for the Environment,2 the
Minister of Energy confirmed that the Government’s energy policy framework was:
“to ensure the continuing availability of energy services, at the lowest cost to the
economy as a whole consistent with sustainable development.”
Key initiatives included deregulation of the electricity and gas industries and controlling
environmental effects through the Resource Management Act.3
Also in 1992, low rainfall led to low storage levels in the South Island hydro lakes, a power crisis,
and renewed interest in energy efficiency. That year, the Energy Efficiency and Conservation
Authority (EECA) was established by Cabinet (later to be established as a Crown entity under
the Energy Efficiency and Conservation Act 2000).
In June 1993, the Government announced an interim climate change policy, intending to develop
a comprehensive long-term strategy. The interim policy featured:
•
energy efficiency measures and incentives, and investigation of renewable energy options
(via EECA);
•
increased CO2 absorption through afforestation as a temporary measure;
•
three policy principles: environmental effectiveness, economic efficiency, and equity.
New Zealand ratified the UNFCCC on 16 September 1993.4
In July 1994, Government announced that the national target would be stabilising net CO2
emissions at 1990 levels by the year 2000. The components of the strategy were:
•
achieving 20% of the emission reduction, regardless of GDP growth, through:
- voluntary agreements with industry to promote improved energy efficiency and
conservation and greater use of renewable energy;
- a ten-point energy efficiency strategy administered by EECA; and
- deregulation of the energy sector and the establishment of a more competitive
wholesale electricity market.
•
achieving 80% of the emission reduction through enhancing carbon sinks (essentially
new forest plantings).
•
providing for the option of Government introduction of a low-level carbon charge if by
mid-1997 the policy measures were not on track to achieve the CO2 net emission
target by 2000 and the ‘20%’ policy objective.5
The Resource Management Act 1991 (RMA) was expected to be able to contribute to meeting
this target through the power to control air discharge consents involving significant emissions of
CO2 and PFCs. In May 1994, the Electricity Corporation of New Zealand’s proposed Stratford
gas-fired combined cycle power station, projected to increase New Zealand’s CO2 by 5%, had
been “called in” by the Minister for the Environment under s 140 of the RMA. This resulted in a
condition on the consent that required mitigation of any net increases from the power station
above an electricity sector baseline.6 However, this remained the only example.
2
Parliamentary Commissioner for the Environment 1992
Parliamentary Commissioner for the Environment 2000, p. 23.
4
http://www.unfccc.int/text/resource/country/nz.html
5
Ministry for the Environment 1998, p. 23.
6
See section 11.6.1 for more detail.
3
16
The greenhouse effect and climate change
Parliamentary Library, August 2001
In 1994, an Energy Efficiency Strategy and the Energy-Wise Companies campaign were
launched. In 1995-1996, an Energy Saver Fund was established with a $18 m budget over five
years.
However, over 1994-1997 New Zealand’s greenhouse gas emissions increased rather than
decreased. This would have justified introduction of a carbon charge as per the 1994 policy, but
in March 1997 the decision on the carbon charge was deferred until Kyoto Protocol negotiations
had been finalised.
In December 1998, a national transport policy statement was released. The document did not
explicitly mention transport emissions or energy efficiency. However, it was expected “the costs
associated with the adverse environmental effects of the transport system [will be] faced by
providers and users of transport services.” 7
In January 1999, the Government released Climate Change: Domestic Policy Options
Statement. This proposed options for meeting New Zealand’s Kyoto Protocol target (stabilising
emissions at 1990 levels by 2008-2012). The options focused on “price signaling measures”
(transferable tradable emissions permits or a carbon tax), and “complementary measures” (e.g.
energy efficiency). In November 1999, policy decisions were deferred until after COP6. 8
In May 2000, the Prime Minister announced that the Government intended to ratify the Kyoto
Protocol by June 2002, when the Rio Plus Ten Earth Summit will meet (10 years after the
UNFCCC was originally signed). She said that wide consultation would contribute to the
necessary development of policy and appropriate legislation. The Prime Minister also noted that
although New Zealand’s contribution to global climate change was relatively small, “we must
lead by example and encourage other countries to participate actively.”9
In April 2000, the Government had announced its support for the Green Party’s Energy
Efficiency and Conservation Bill, after negotiating amendments. The Bill passed into law as the
Energy Efficiency and Conservation Act on 15 May 2000, nearly two years after the original Bill
had been introduced.10
In August 2000, the Minister of Energy announced that the Government’s domestic climate
change policy would focus initially on energy efficiency measures, with work continuing on more
complex economic and regulatory options. Some Cabinet papers were released, as they were
again in February 2001.11 These Cabinet decisions are summarised in Appendix A.
In November 2000, the Minister of Energy announced that projects under the Crown Energy
Efficiency Loan scheme administered by EECA had saved $4 m in central and local government
energy costs and reduced CO2 emissions equivalent to taking 8,000 cars off the road.12 The
loans scheme had not been originally continued into the 2000-2001 budget for EECA, but in
November 2000 EECA was granted additional funding from the “Greens Fund”, reversing a
previous decline in funding (more detail in section 10.2).
More voluntary agreements under the Government Energy Efficiency Leadership Programme
were to be pursued, with a target of 15% energy savings from the public sector by 2005.
Mandatory energy-performance labels for appliances were approved by Cabinet, and
consultation would take place before regulations were passed.13
7
Parliamentary Commissioner for the Environment 2000, p. 29-30; Ministry of Transport 1998, p. 5.
Parliamentary Commissioner for the Environment 2000, p. 31. See section 10.1 for more detail.
NZ aims to ratify Kyoto protocol on climate change by mid-2002, 8/5/2000, on http://www.executive.govt.nz
10
Govt to support Energy Efficiency Bill, 3/4/2000 on http://www.executive.govt.nz. More information on the Energy Efficiency and
Conservation Act is in section 3.3.1.
11
Climate change policy: early decisions and directions, 30/8/2000, on http://www.mfe.govt.nz/new/ccrelease.htm ; Climate change
Cabinet papers released, 27/2/01,on http://www.mfe.govt.nz/new/media_27_02_01.htm ; Cabinet papers available via
http://www.mfe.govt.nz/issues/cabdec_feb_01.htm
12
Govt saving taxpayers’ dollars on energy on http://www.mfe.govt.nz/new/media_15_11_00.htm
13
Energy efficiency loans saving $4m a year http://www.mfe.govt.nz/new/media_12_11_00.htm
8
9
17
The greenhouse effect and climate change
Parliamentary Library, August 2001
New Zealand devoted $17.1 m on climate change research in 1997-98, and the Government
announced in 2001 that annual expenditure in this area was about $24 m per year.14
New Zealand also provides money to help developing countries with climate change activities.
Over 1994-97 $10.4 million was contributed to the Global Environment Facility, and additional
funds were provided to Pacific countries of about $2 m per year in 1998, 1999 and 2000.15
In January 2001, Cabinet agreed that domestic emissions trading, implemented across a range
of sectors and supplemented with other measures where necessary, will be a central policy
measure for meeting New Zealand’s Kyoto Protocol target.16
On 29 March 2001, the Draft Energy Efficiency and Conservation Strategy was released for
public comment. By the deadline of 1 June 2001, more than 360 submissions had been
received. As required under the Energy Efficiency and Conservation Act, the final strategy will be
issued by 1 October 2001. A summary of the draft strategy is in Table 10.3.17
In June 2001 analysis of the economic impact of a low-level carbon charge was publicly
released, and the information referred to the Tax Review 2001 which is scheduled to report to
Ministers by the end of September 2001. A summary of the economic impact findings is in
section 9.3 (Table 9.1).
In July 2001, an information document on forest sinks and the Kyoto Protocol was released. A
summary of the possible rules for forest sinks and domestic carbon trading is in section 5.2.
During the winter of 2001, low hydro lake levels again focused attention to energy efficiency, and
the Government requested that citizens reduce energy consumption by 10% to reduce the risk of
power blackouts. Energy saving tips were provided by television ads and other means. The
emergency diesel generators at Parliament were activated to reduce the load on the national
grid, which also created 42 tonnes of greenhouse gas emissions per week (section 3.4).
In August 2001, the Minister of Energy announced that the Government intended working
towards having legislation passed by Parliament in 2002 that would enable New Zealand to ratify
the Kyoto Protocol in September 2002.18
The Prime Minister has stated:
“New Zealand is a good international citizen…we must lead by example and
encourage other countries to participate actively in the international effort on climate
change. Ratification of the Kyoto Protocol will position New Zealand to be up with
the leaders on climate change and play a small but worthy role in bequeathing future
generations a more sustainable world.”19
Further detail on New Zealand’s proposed economic instruments, programmes for energy
efficiency and conservation, and a recent survey of public opinion on climate change is in
chapter 10.
14
Controller and Auditor-General 2001, p. 101; Government press release as reported by The Press 23/7/01.
Controller and Auditor-General 2001, pp. 101-102.
CBC Min(01)1/7, item d, 23 January 2001, http://www.mfe.govt.nz
17
Energy Efficiency and Conservation Authority 2000; The Dominion 18/6/01.
18
Speech by the Hon Pete Hodgson, 9 August 2001, Climate change after Bonn, p. 5.
19
Speech by the Rt Hon Helen Clark, 8 May 2000, NZ aims to ratify Kyoto protocol on climate change by mid-2002.
15
16
18
3
Activities in the House of Representatives
3.1
Report of the Controller and Auditor-General
In April 2001, the Controller and Auditor-General released the report Meeting International
Environmental Obligations. Among the multilateral environmental agreements looked at in the
report were the UNFCCC and the Kyoto Protocol.
One of the conclusions of the Controller and Auditor-General was that information given to
Parliament on climate change is not adequate. Improved reporting was recommended (Box 4).
Box 4
Conclusions and recommendations of the Controller and Auditor-General
relating to climate change agreements, April 2001
CONCLUSIONS
New Zealand ratified the Framework Convention on Climate Change - FCCC without adequate information.
•
In 1992 New Zealand agreed to implement the FCCC; The ratification recommendation to
Cabinet appears to have met the criteria of the day, although by today’s standards it was
inadequate. It did not cover the costs of implementing the FCCC in New Zealand. For example,
no information was provided on the likely cost increases for fuel, building, waste disposal,
electricity generation, and industrial processes.
New Zealand is meeting the FCCC obligations except the first and most
important one.
•
•
•
•
•
New Zealand has not fulfilled the main FCCC obligation to formulate and implement national
policies to mitigate climate change through limiting human-induced emissions of greenhouse
gases. A range of policy measures has been adopted, but the measures have been
ineffective.
The lack of progress is despite intense policy debate on climate change since ratification in
1992. Views on FCCC have been polarised among government departments as there has
been a lack of incentive to reach a satisfactory accommodation that would allow progress.
However, over recent months there has been evidence of broader agreement with, for
example, unanimous recommendations appearing in the climate change recommendations to
Government in papers to Cabinet.
New Zealand agreed to aim at reducing human-induced greenhouse gas emissions to 1990
levels by the year 2000. However, gross emissions of carbon dioxide (the main greenhouse
gas) have so far increased by 19% over that period. If all greenhouse gas emissions and not
only CO2 are considered, then the increase is 4.8%.
New Zealand is meeting the FCCC obligations to provide detailed, annually updated,
inventories of greenhouse gas emissions and sinks information; promote climate change
research; provide money to help developing countries meet their obligations under the
Convention; and promote climate change education, training, and public awareness.
Parliament is not given a clear picture of climate change issues and progress.
•
Individual agencies responsible for climate change matters report separately to Parliament.
There is no single report or process that pulls together all this separately reported information
to provide Parliament with a clear picture of climate change issues and progress.
(Recommendations on next page ⇒)
The greenhouse effect and climate change
Parliamentary Library, August 2001
(continued from previous page)
RECOMMENDATIONS
•
Climate change is complex and wide-ranging, and requires an effective “whole of government”
approach to assist in resolving inter-agency differences on policy. We recommend that the
accountabilities of the main agencies concerned with climate change should be expanded to
encompass a requirement to collaborate with other agencies in achieving demonstrable
progress on climate change obligations.
•
We recommend that the national impact analysis supporting any decision to ratify the Kyoto
Protocol should, as far as possible, include an assessment of all direct and indirect costs and
benefits of ratification.
•
New Zealand has produced a wide-ranging consultation document on climate change policy in
its Climate Change Domestic Policy Options Statement. We recommend that Parliament is
provided with a similar single report on climate change issues and progress as part of the
preparation for New Zealand’s ratification of the Kyoto Protocol.
•
We also recommend that the main agencies concerned with climate change provide
Parliament with a regular joint report on how New Zealand is meeting its FCCC obligations.
The report should:
- provide a single source of information on agency performance;
- explain how New Zealand is meeting its international obligations; and,
- inform Parliament by outlining new policy development and issues.
Source: Controller and Auditor-General 2001, pp. 81-83.
3.2
Select Committee inquiries
3.2.1 Local Government and Environment Committee 2000:
role of local government in climate change initiatives
On 13 June 2000, the Local Government and Environment Committee agreed on the terms of
reference for an inquiry into the role of local government in meeting New Zealand’s climate
change target (Box 5).
After conducting preliminary inquiries and being briefed by a number of agencies and
organisations,1 the Committee issued an interim report in December 2000. As the Government
was currently developing climate change policy, the Committee issued the report in order to
highlight the importance of taking a pro-active, co-operative approach to the implementation of
climate change policy. The report conveyed 15 recommendations to Government (Box 6) and
submissions were invited, particularly on 29 key questions. The deadline for submissions was 15
March 2001.
1
The National Institute of Water and Atmospheric Research, the Ministry for the Environment, the Energy Efficiency and
Conservation Authority, Local Government New Zealand, and the Parliamentary Commissioner for the Environment.
20
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 5
Local Government and Environment Select Committee
Inquiry into the Role of Local Government in
Meeting New Zealand’s Climate Change Target
Terms of Reference
In conducting its inquiry, the committee will examine:
•
The contribution local government can make to reducing greenhouse gas emissions
through the exercise of planning and regulatory functions, and also through its own
actions, with regard to such matters as:
- land use and subdivision consents
- biodiversity conservation
- transport planning and traffic management
- operation of vehicles
- building consent processing
- management of buildings
- water and waste water
- landfill management and waste management generally.
•
Any obstacles to local government playing this role, including:
- legislative impediments in the above areas
- information co-ordination problems at local government level, for example the
appropriate roles of regional and territorial authorities in relation to land use and
transportation issues.
•
Any central government actions that could:
- assist local government in its role in reducing greenhouse gas emissions,
including whether there is a need for national policy statement or guidelines or
standards under the Resource Management Act 1991, or other legislative
change, or other policy initiatives
- improve co-ordination and synergies between central and local government
efforts to reduce greenhouse gas emissions.
The committee recognises that adaptation issues need to be addressed, to the extent
that world-wide mitigation activities do not succeed in reversing climate change, and may
later conduct a second part to this inquiry to examine the adaptation aspect of local
government responsibilities.
Local Government and Environment Committee, 2000, Appendix B
21
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 6
Recommendations to Government
Local Government and Environment Select Committee
December 2000
Leadership by local authority on energy use
That strategies be developed, in conjunction with Local Government New Zealand, for facilitating the
•
shift by local authorities towards more energy efficient operations.
Processes under the Resource Management Act 1991
That sustainable models for urban form, incorporating integrated transport and land use strategies, be
developed and promoted in the New Zealand context.
That (as already recommended by the Parliamentary Commissioner for the Environment) guidance be
•
provided to local authorities on the relative weight to be afforded to the protection of outstanding natural
features and landscapes under the Resource Management Act 1991 vis a vis the development of
renewable sources of energy such as wind power.
•
Transport planning and operations
That clear policies and frameworks be developed for co-ordinating and facilitating local government
initiatives for enhancing public transport or for improving the energy efficiency of transport across
different modes.
That the legislation governing land transport strategies be amended to clarify that they should deal with
•
greenhouse gas reductions.
That the objective of Transfund New Zealand, as set out in the Transit New Zealand Act 1989, be
•
amended from a ‘safe and efficient roading system’ to be a ‘safe, efficient and sustainable transport
system’.
That more funding be provided for public transport infrastructure and Transfund New Zealand
•
procedures be reformed to facilitate the switching of funds between roads and alternatives.
•
Waste minimisation
That appropriate resources be directed towards completing and implementing the waste minimisation
strategy.
•
Education and information
That easily accessible education resources be provided to local government for informing the public of
ways to reduce greenhouse gas emissions, including through building design, and of the benefits that
arise from doing so.
•
Obstacles to local authorities
That priority be given to identifying and addressing obstacles to local government responses to climate
change.
•
Local government involvement in policy development
That the Minister of Local Government be a member of the ministerial working group on climate
change.
That officials from the Local Government Policy group of the Department of Internal Affairs be
•
represented in the officials working groups providing advice to the Government about the domestic
climate change policy options.
That Local Government New Zealand be given a role in developing policy frameworks for achieving
•
New Zealand’s climate change targets, and, where appropriate, Local Government New Zealand be
involved in the officials working groups at the earliest opportunity.
That full account be taken of issues and recommendations set out in the Local Government New
•
Zealand report on its survey of local authorities on climate change issues.
•
Central government example: energy efficiency
That all departments, Crown entities and State enterprises be encouraged or required to implement
energy efficiency programmes in their premises and activities.
•
Local Government and Environment Committee, 2000
22
The greenhouse effect and climate change
Parliamentary Library, August 2001
3.2.2 Transport and Environment Committee 1998:
environmental effects of road transport
In September 1998, the Transport and Environment Select Committee tabled an interim report
entitled Inquiry into the Environmental Effects of Road Transport.
The terms of reference for the inquiry were to:
consider the nature and scale of the environmental effects of road transport;
review work currently undertaken by the Government to investigate these effects;
consider the management option recommended by the Roading Advisory Group, and
evaluate the official assessment of the environmental effects of that option or any variation
being proposed by officials; and,
identify possible mechanisms for minimising the environmental effects of road transport.
•
•
•
•
The Committee made 22 recommendations to Government (Box 7). Climate change was listed
in the text as one of the environmental impacts of transport. Although it was not specifically
mentioned in the recommendations, a likely outcome of implementing the recommendations
would be a reduction in CO2 emissions from land transport.
3.3
Legislation
3.3.1 Energy Efficiency and Conservation Act 2000
Efficiency of energy use and choice of energy source have a direct bearing on greenhouse gas
emissions, particularly CO2 .
The Energy Efficiency Bill was introduced as a Member’s Bill by Jeanette Fitzsimons on 20
August 1998. The Transport and Environment Committee received submissions on the Bill, and
reported it back to the House on 15 July 1999 in amended form, retitled the Energy Efficiency
and Conservation Bill. The Energy Efficiency and Conservation Act was subsequently given
assent on 15 May 2000, and came into effect on 1 July 2000.
The original Energy Efficiency Bill sought to set up an independent authority to develop national
energy efficiency policy. The Committee found that the intent of the Bill was generally supported
by submitters, but that the Government would not support the Bill without significant changes
being made. The Government’s main concern was that the Bill appeared to shift primary
responsibility for an area of policy development to a Crown entity, with the result that the
responsible Minister would be politically accountable for policy and initiatives over which they
had only limited control. There was also concern that the Bill pre-empted the outcome of a
Government review of the appropriate governance structure for the Energy Efficiency and
Conservation Authority (EECA).2
When reporting back to the House, the Committee recommended that the Bill be amended so
that the Minister rather than the Authority would be responsible for achieving the purpose of the
Act; that regulations rather than rules would be available for promoting policy; that the ability to
control domestic electricity prices be removed; and that the vehicle of “market development
plans” to address barriers to energy efficiency in specified sectors be deleted.3
2
3
Parliamentary Commissioner for the Environment 2000, pp. 118-119.
Parliamentary Library, Bills Digest No. 585.
23
The greenhouse effect and climate change
Recommendations from the Transport and Environment Select
Committee to Government, September 1998
Source: Transport and Environment
Select Committee 1998, pp. 2-3.
Box 7:
Parliamentary Library, August 2001
24
The greenhouse effect and climate change
Parliamentary Library, August 2001
The Energy Efficiency and Conservation Act 2000 has these key features:
•
The purpose of the Act is to promote, in New Zealand, energy efficiency, energy
conservation, and the use of renewable sources of energy.
•
The Energy Efficiency and Conservation Authority is given statutory status (s 20).
•
A National Energy Efficiency and Conservation Strategy must be established. The Act
requires the first Strategy to be released in draft for consultation by 1 April 2001, and
finalised by 1 October 2001. Thereafter, each Strategy lasts five years, and one must always
be in effect. Strategies must be consistent with any national policy statement in force (ss 819).
•
When preparing Strategies, the parties that must be consulted are industry and commerce,
environmental and community organisations, Mäori organisations, local authorities, and the
Parliamentary Commissioner for the Environment (s 13(2)).
•
The Authority must comply with Government policy and the Minister’s directions. Ministerial
directions to the Authority are to be published in the Gazette and presented to the House (s
23).
•
Regulations may be promulgated by the Governor-General on recommendation of the
Minister for: prescribing minimum energy performance standards and related compliance
documentation; prescribing energy efficiency labelling; requiring the provision of relevant
statistics; and establishing related offences and penalties for non-compliance (s 36).
3.3.2 International Treaties Bill (2000)
This Bill may have relevance to New Zealand ratifying the Kyoto Protocol or entering into other
international instruments on climate change. Officials consider that new legislation will be
needed before New Zealand can ratify the Kyoto Protocol.4
Currently, international treaties are entered into by the Crown without specific approval from
Parliament, except where new domestic legislation is required to give effect to the treaty.
However, each new treaty together with an national interest analysis is required to be presented
to the House under Standing Orders 384 and 385.
The International Treaties Bill provides that Government cannot enter into international treaties
without Parliamentary approval. The Bill would enact the provisions of Standing Orders 384 and
385 in similar form, but also require the analysis to include consistency with the Treaty of
Waitangi.5
The Bill was introduced as a Member’s Bill by Keith Locke on 21 September 2000, and was
referred to the Foreign Affairs, Defence and Trade Select Committee. The closing date for
submissions on the Bill was 31 March 2001, and the report to the House from the Committee is
due by 7 September 2001.
4
5
Controller and Auditor-General 2001, p. 87.
Parliamentary Library, Bills Digest no. 706.
25
The greenhouse effect and climate change
Parliamentary Library, August 2001
3.3.3 Road Traffic Reduction Bill (2001)
This Bill has relevance to New Zealand’s largest and fastest growing contributor to CO2
emissions, road transport. It was introduced as a Member’s Bill by Jeanette Fitzsimons on 3 May
2001.
The Bill has the aim of requiring the Minister of Transport and regional councils to develop
targets, timetables and measures for the reduction of motorised road traffic, and for these to be
completed within a year of the Bill coming into force. In addition the principal objective of Transit
New Zealand and Transfund New Zealand would be amended to focus on a “safe and
sustainable land transport system at reasonable cost”, rather than operating a “safe and efficient
roading system” as required under current law.6
As of 6 August 2001, the Bill had not yet had its first reading.
3.4
Energy efficiency in the Parliamentary Buildings
In 1999, evaluation was done by EECA of Government agencies’ commitment to on-site energy
efficiency. Parliamentary Service ranked 11 out of 32 (i.e. in the top half), and the Department of
Prime Minister and Cabinet ranked 31 out of 32 (i.e. second worst).7
In 2000, Parliamentary Service won the Energy-Wise Award in the Public Sector category for its
achievements in energy efficiency and conservation. Savings of 30% were obtained through
such measures as replacing incandescent lamps with fluorescent lamps; reducing hours and
intensity of corridor, carpark, and outdoor lighting; reducing variability in air conditioning
temperatures; and installing a building automation system which allows finer control of power
supply, lighting, and air conditioning. Shifting energy loads to times of the day when cheaper
electricity is available has also been used to save further on energy costs.8
The Parliamentary complex has a designated energy manager, and energy costs and usage are
tracked monthly as part of the business plan. A series of detailed audits are being conducted to
identify further savings; one has been completed for Bowen House, and others are scheduled for
later in the year. Currently a re-lamping exercise in Bowen House is expected to realise 30-40%
fewer lamps while maintaining effective lighting levels, and some 10% savings in building power
use.9
With the advent of the low hydrolake levels in winter 2001 and the potential for a power crisis,
the Speaker of the House requested staff to undertake energy efficiency measures and directed
Parliament’s diesel generators to be used to lighten the load on the national electricity grid. It is
anticipated that reminders about turning off unneeded lights and equipment will be re-issued to
staff over the long term to improve energy use behaviour.10
The software is being developed for the diesel generators at Parliament so they can be used for
“load lopping”, or lowering the peak demands to reduce monthly networking charges. Over the
three weeks prior to 22 August, power consumption from the national grid was reduced 24%.11
However, the climate change implication of using these generators is the emission of up to 42
tonnes of CO2 a week over the forecast 10 week hydrolake shortage period.12
6
Parliamentary Library, Bills Digest no. 779; Transit New Zealand Act 1989.
See section 10.3 for further detail.
Energy-Wise News, September 2000, pp. 26-27.
9
P. Ritchie, Parliamentary Service, pers comm 8/2001. Light meters are used to ensure 500 lux at desk areas and 350 lux in other
areas. The existing lighting was more than required for health and safety requirements.
10
Email to all staff from the Speaker 31/7/01; notice in InHouse 8/8/01.
11
InHouse no. 33, 22/8/01, p.2, Parliamentary Service.
12
There are four diesel generators at Parliament, two near the Library and two under the Beehive. At optimum operating capacity
under the current “load lopping” situation, each uses about 80 litres per hour and is used 50 hours per week, or 4,000 litres of diesel
a week. With 2.65 kg of CO2 per litre of diesel, this is 10,600 kg or 10.6 tonnes per week. At the outset two generators were used,
and the other two were brought on line later. (P. Ritchie, Parliamentary Service and T. Jamieson, EECA, pers comm 8/2001).
7
8
26
Part B: Greenhouse gases and sinks
4 The greenhouse gases
4.1
The “greenhouse effect” and contributing gases
The natural greenhouse effect acts to trap some of the sun’s warmth from escaping back into
space and makes life possible on Earth. It is caused by the natural compounds of water vapour,
carbon dioxide, methane, and nitrous oxide. Without the greenhouse effect, the Earth would be
about 30ºC colder.
In the last two centuries human activity has “enhanced”1 this natural effect by adding significantly
higher levels of carbon dioxide (31% increase), methane (up 151%), and nitrous oxide (up 17%),
as well as some artificial compounds which also act as greenhouse gases (perfluorocarbons,
hydrofluorocarbons, sulphur hexafluoride, and the ozone depleting gases).
When solar radiation enters the atmosphere, visible light passes through and ultraviolet light is
absorbed by the ozone layer. The visible light absorbed by the Earth is converted to heat
energy, which is radiated outwards. While most gases allow heat to pass through to the upper
atmosphere, the greenhouse gases absorb this heat energy and re-radiate it rather than let it
escape (Figure 4.1). The hole in the ozone layer is a separate phenomenon, but there are a few
linkages with the greenhouse effect. These are listed in section 4.5.
solar
radiation
UV
greenhouse effect
Solar radiation
absorbed by the
Earth is radiated as
infra-red (heat).
Green-house
gases in the
troposphere
absorb and reradiate some of
this heat back to
the Earth.
Figure 4.1
light
visible
light
hole in ozone layer
An intact ozone layer in
the stratosphere
prevents damaging UV
light from entering lower
atmosphere.
thinning of the ozone
layer lets UV light in
stratosphere
15 to 50 km
Earth
troposphere
0 to 15 km
Basic diagram of greenhouse and ozone layer effects
 Dana Rachelle Peterson 2001.
Adapted from Controller and Auditor-General 2001, Figure 11, and greenhouse effect and
ozone layer webpages published by the Ministry for the Environment, NIWA , and CDIAC.
1
The Ministry for the Environment refers to the “enhanced greenhouse effect” to describe the human contribution to climate change
(e.g. Ministry For The Environment 1998, p. 2; Ministry For The Environment 1999, p. 19).
The greenhouse effect and climate change
Parliamentary Library, August 2001
Each greenhouse gas has a different potential to contribute to the greenhouse effect. The other
gases have a more powerful per unit effect than CO2 in this respect, both as a function of their
chemical properties and their lifetime in the atmosphere. The Global Warming Potential (GWP)
is used to represent this, with CO2 assigned a GWP of one, and the other gases assigned
GWPs in relation to CO2 (Table 4.1). The GWP is a “best estimate” rather than a precise
measure. The GWP is used to calculate aggregate greenhouse gas emissions in “CO2
equivalents”.
Table 4.1: Global Warming Potential (GWP) and lifetime of the greenhouse gases.
Global Warming
Potential (GWP)
Greenhouse gas
CO2 (carbon dioxide)
CH4 (methane)
N2O (nitrous oxide)
HFCs (hydrofluorocarbons)
1
21
310
140 – 11,700
Lifetime in
atmosphere
(years)
50 - 200
14.5
120
1.5 - 250
Increase in
atmosphere
since 1750
31%
151%
17%
(HFC-134a = 1,300)
(gases with hydrogen, fluorine, and carbon, e.g.
CH2FCF, known as HFC-134a)
CFCs (chlorofluorocarbons)
SF6 (sulphur hexafluorine)
PFCs (perfluorocarbons)
4,800 – 8,100
23,900
6,500 – 9,200
(gases with fluorine and carbon, e.g. CF4 and C2 F4)
85 -102
3200
3,200 – 50,000
the gases did not
exist in 1750
(CF4 = 6,500)
Source: Ministry for the Environment 1998, Table 1.1; http://cdiac.esd.ornl.gov/pns/current_ghg.html ; for increases since 1750, IPCC
2001a, p. 4. Notes: The GWP were calculated using a 100-year period. They are the IPCC figures from 1996. Although the IPCC
reported slightly revised GWP in 2001, the 1996 figures are the ones used under the UNFCCC for country reports of emissions (H.
Plume (MFE) and M Manning (NIWA), pers comm 8/2001).
4.2
Data uncertainties
Data for the greenhouse gases is the most robust for CO2. Data uncertainties for the other
greenhouse gases are significant, and are very large for N2O (Table 4.2). The reasons for the
uncertainties include limited coverage of data collection sites and incomplete scientific
understanding of the processes relating to emission and sequestration of the greenhouse gases.
Table 4.2:
Estimated uncertainty for greenhouse gas emissions and sinks data,
New Zealand’s 1990 baseline inventory.
CO2
± 5%
CH4
± 22%
N2O
± 60%
CO2 sinks
± 25%
Source: Ministry for the Environment 1998, Table 2.6, p. 16; H. Plume, Ministry for the Environment, pers. com. 5/2001.
28
The greenhouse effect and climate change
4.3
Parliamentary Library, August 2001
New Zealand’s emissions: overview
4.3.1 Gross and net emissions
New Zealand produces a small minority of the total greenhouse gases emitted worldwide (Table
4.3). However, on a per capita basis New Zealand was the fourth largest emitter of greenhouse
gases among the Annex I (developed country) Parties reporting data for 1998, exceeded only by
Canada, the USA, and Australia (Figure 4.2a). When CO2 removals by land use change and
forestry are included, New Zealand’s rate of emissions per capita is 8th highest out of the 30
Annex I countries reporting 1998 data (Figure 4.2b).
Table 4.3: New Zealand CO2 and total greenhouse gas emissions; by weight and percent of
world total
Measure
CO2 only
all greenhouse gas emissions, in CO2
equivalents
all greenhouse gas emissions, in CO2
equivalents, including the estimated removal
by carbon sinks
1999
emissions
1998
emissions
kilotonnes
(= Gigagrams)
kilotonnes
(= Gigagrams)
30,523.13
76,831.07
28,824.31
75,010.14
54,712.68
54,050.94
percent of total
world
Annex I
(late
1990s)
countries
(1998)
0.15%
--
0.18%
0.4%
-0.3%
Sources: Box 1 (chapter 1); Ministry for the Environment 2001, Table 10; UNFCCC Secretariat 2001b, Table A.1; OECD 1999, Table 2.3C.
Comparisons with non-Annex I countries can only be done for CO2 emissions from fossil fuel use
and cement manufacture, as global data by country is not available for the other emissions. This
comparison is presented in more detail in Figures 1.1 to 1.3 (chapter 1) and section 4.4.1.
4.3.2 Mix of greenhouse gas emissions
CO2 is the principal greenhouse gas emitted by most Annex I countries. In contrast, New
Zealand has the highest ratio of non-CO2 greenhouse gas emissions of any of the Annex I
Parties to the UNFCCC.
In 1999, 60% of New Zealand’s greenhouse gas emissions were from CH4 and N2O (Figure 4.3).
These emissions are primarily from agricultural activities. Not surprisingly, the OECD countries
with the next highest non-CO2 emissions also have a large agricultural sector: Ireland and
Australia.2
However, New Zealand’s emissions of CO2 are increasing and methane emissions are
decreasing (Figure 4.4). CO2 has been projected to become New Zealand’s primary contribution
to climate change over the next decade, but recently revised data for methane emissions
suggests caution about this trend.3
2
3
Ministry for the Environment 1998, pp. vii, 8. Countries data for 1990 (Figure 1.3).
Ibid., pp. 31-32; M. Manning (NIWA) pers comm 8/2001.
29
The greenhouse effect and climate change
Figure 4.2:
Parliamentary Library, August 2001
Total per capita greenhouse gas emissions, Annex I countries, gross and net of LULUCF (Gg C per person)
22.6
24.5
26.2
19.7
Portugal
Hungary
Sweden
Ukraine
Spain
Italy
France
Slovakia
Austria
Bulgaria
Poland
14.4
14.4 14.8
15.1
15.2
Estonia
6.5
14.3
Netherlands
10.1
12.7
Czech Republic
9.9
12.4
Belgium
9.5
11.6 11.7
Norway
9.4
Switzerland
4.7
8.3
8.9
10.4
8.3
9.8
7.6
9.3
7.4
Lithuania
4.3
Latvia
15.0
Germany
17.3
Monaco
tonnes per person
(CO2 equivalents)
30.0
Australia
USA
Canada
NZ
Ireland
Finland
Denmark
Greece
UK
0.0
Figure 4.2 (a): Gross emissions (land-use, land-use change and forestry (LULUCF) carbon sinks and emissions not included)
Sources for both graphs: Emissions: UNFCCC Secretariat 2000, Table A.2. Population: http://www.popin.org/pop1998/2.htm1998.
1998 emissions data not reported for Iceland, Japan, Liechtenstein, Luxembourg, Slovenia, Romania and the Russian Federation.
28.0
Spain
Lithuania
Austria
Italy
Bulgaria
Slovakia
Poland
21.7
Canada
Norway
21.2
USA
France
14.3
Netherlands
Hungary
14.2
Belgium
Ukraine
14.2
Denmark
9.6
14.1
NZ
9.3
14.0
Czech Republic
8.7
12.9
Estonia
8.5
12.9
Finland
8.4
11.9
Germany
8.3
11.5
UK
8.2
11.4
Greece
7.7
Portugal
4.8
7.6
9.5
7.1
9.0
6.4
Switzerland
4.3
Sweden
15.0
Monaco
tonnes per person
(CO2 equivalents)
30.0
15.6
0.4
30
Figure 4.2(b): NET emissions (land-use, land-use change and forestry (LULUCF) carbon sinks and emissions included)
Note: LULUCF balances can be positive (LULUCF emissions exceed sinks) or negative (sinks exceed LULUCF emissions)
Australia
Ireland
Latvia
0.0
The greenhouse effect and climate change
Parliamentary Library, August 2001
Figure 4.3
Percentage of each greenhouse gas in
New Zealand’s total emissions, 1999
(CO2 equivalent kilotonnes).
CH4
43.7%
N2O
16.1%
CO2 equiv. kT
30000
CO2
39.7%
HFCs,
PFCs,
SF6
0.4%
Figure 4.4
New Zealand’s emissions of greenhouse
gases, 1990 and 1999 (CO2 equivalent
kilotonnes).
15000
1999
Source for both Figure 4.5 and 4.6: Ministry for the
Environment 2001, Table 10.
For comparison to other sources, kilotonnes (kT) are
equivalent to Gigagrams (Gg).
0
CH4
N2O
HFCs,
PFCs,
SF6
1990 25399
35211
11849
606
1999 30523
33594
12397
318
CO2
4.3.3 Changes in total emissions 1990-1999
Under the UNFCCC agreement, signatory countries agreed to aim to reduce their greenhouse
gas emissions to 1990 levels by the year 2000. No countries attained this goal by design. Many
have increased rather than reduced emissions, and others have achieved reductions but for
reasons other than preventing climate change.
The Eastern European countries, through major economic downturn, restructuring, and in some
cases military destruction, have significantly reduced their industrial and domestic fossil fuel
energy use, and therefore their greenhouse gas emissions (Figure 4.5).
Germany’s reduction in greenhouse gas emissions has been attributed to unification with East
Germany, which has lower per capita emissions (the “wall-fall effect”). Luxembourg’s ability to
meet its greenhouse gas emission stabilisation target has been attributed to the conversion (for
economic reasons) from coal to electric fired furnaces in the steel industry.4
The UK has achieved a reduction in emissions through the removal of subsidies from coal
mining (done for economic and political reasons, rather than environmental reasons), and
subsequent reduction of coal use. Similarly, New Zealand’s reduction in methane and nitrous
oxide emissions, due to the impact on livestock production from the removal of agricultural
subsidies and market contraction, has partially offset an increase in CO2 emissions.
Under the UNFCCC reporting guidelines, policies that lead to reduction of greenhouse gases do
not have to be introduced for climate change reasons. In terms of the real effect on climate
change, any action that leads to net reduction of greenhouse gases in the atmosphere could be
viewed as legitimate. However, there are restrictions on what will be accepted for compliance
with the Kyoto Protocol.
4
Newell 1997, p. 2,6.
31
The greenhouse effect and climate change
Parliamentary Library, August 2001
Over the last decade, New Zealand’s CO2 emissions have increased faster than population, and
CO2 emissions from transport and thermal energy generation have increased faster than
population and GDP. However, New Zealand’s total greenhouse gas emissions have increased
more slowly than GDP and population (Table 4.4).
Table 4.4:
Percent change 1990-1999,
New Zealand’s greenhouse
gas emissions, population,
and real GDP
Sources: Emissions: Ministry of
Economic Development 2000; Ministry
for the Environment 2001, Table 10.
Population and GDP: Parliamentary
Library databases ex Statistics New
Zealand. Energy use: International
Energy Agency 2000, p. 11.317 (note
the energy use data is for 1990-1998
rather than 1990-1999).
4.4
Measure
Emissions
CO2, gross
All greenhouse gases, gross
All greenhouse gases, net
(including land-use change and forestry)
CO2 emissions by major sector
Domestic transport
Thermal energy generation
Industry
Residential, commercial, agriculture
Energy use
TPES (Petajoules of energy) (1990-1998)
Population and GDP
Resident population
Real GDP (1995/96 $)
percent change
1990-1999
+ 20.0%
+ 5.0%
+ 6.0%
+ 58.0%
+ 38.6%
+ 20.7%
- 0.9%
+ 21.3%
+ 12.0%
+ 23.0%
Emissions data on individual gases
4.4.1 Carbon dioxide (CO2)
Comparisons with other countries
New Zealand’s emissions of CO2 are below the OECD average if measured by tonnes per
person or by unit of total energy used.5 This is in part because most (although a decreasing
share) of New Zealand’s electricity is produced using hydro-electricity,6 whereas other countries
burn more fossil fuels. France, which emits less CO2 per unit of energy than New Zealand,
produces 77% of its electricity using nuclear power plants7 (a power source relatively free of CO2
emissions, but of concern for other reasons: see section 9.9).
However, when measured against GDP8 New Zealand’s emissions of CO2 are above the OECD
average, and similar to those of Canada, the USA, Australia, and Korea.9 This could be
interpreted as indicating that New Zealand’s level of energy efficiency in economic production is
lagging behind many other OECD countries.
On a global basis, data is only available by country on CO2 emissions from fossil fuel
combustion. (Estimates for other sources and sinks have large uncertainties). Looking at this
data on a per capita basis, New Zealand is below the Annex I Parties average and less than half
that of the USA, but more than twice that of the world average (Figure 4.6).
5
1998 tonnes CO2 per person OECD average11.1 vs. New Zealand 8.51: tonnes CO2 per tonnes oil equivalent OECD average 2.4
Vs New Zealand 1.88. Source: http://www.sourceoecd.org/data
6
Data for 1998 was 76.1% of electricity and 15.4% of total energy was derived from hydropower (UNFCCC 1999, sections 5 and 6).
The hydro share of electricity dropped from 73% in 1990 to 64% in 1999 (Energy Efficiency and Conservation Authority 2001b, p.6.
7
Europe Feb 2001, p. 26.
8
GDP = gross domestic product, a measure of economic production.
9
OECD average 0.59 kg CO2 per dollar of GDP (USD 1990), Vs. New Zealand 0.62, Canada 0.72, USA 0.78, Australia 0.83, and
Korea 0.96. Source http://www.sourceoecd.org/data
32
The greenhouse effect and climate change
Parliamentary Library, August 2001
M onac o
Figure 4.5:
Percent change
in greenhouse
gas emissions
1990-1998.
Original data in
tonnes of CO2
equivalents.
30.6
S pain
21.0
Ireland
19.1
Greec e
18.1
P ortugal
17.2
A us tralia
14.5
Canada
Source:
IEA 2000, Table 1;
original data from
UNFCCC official data
FCCC/SBI/2000/11.
13.2
US A
11.2
Denm ark
9.5
Japan
9.4
Netherlands
8.4
Norway
7.7
B elgium
6.5
A us tria
6.5
S weden
6.4
Ic eland
4.7
Italy
4.4
New Zealand
Finland
1.5
S witz erland
1.3
Franc e
0.9
-8.3
-15.6
-17.7
2.5
UK
Germ any
Hungary
-22.2 Czec h Republic
-24.0
Rom ania
-28.5
-29.6
Rus s ian Federation
-29.7
P oland
S lovak ia
-30.8
-46.3
B ulgaria
-46.6
E stonia
Ukraine
-50.5
Lithuania
-53.7
Latvia
-67.8
-75
Luxem bourg
-50
-25
0
25
33
The greenhouse effect and climate change
Figure 4.6:
CO2 emissions per
capita 1998 from
fossil fuel
combustion: World
average, Annex I
and non-Annex I
Parties, regional
groupings, and
selected countries
to illustrate the
range of values.
Source: International
Energy Agency 2000, pp.
II.77-II.81.
Note: This does not
include CO2 from other
sources (e.g.
deforestation), the
other greenhouse
gases, or carbon sinks.
Comparable global
data by country is not
available.
Parliamentary Library, August 2001
45.19
Qatar
United Arab Emirates
23.98
Kuwait
20.12
USA
20.10
Singapore
19.88
16.57
Australia
15.75
Canada
Finland
11.59
Annex I Parties
11.00
Netherlands
10.92
Germany
10.45
Ireland
10.36
Russia
9.64
UK
9.28
Japan
8.92
South Africa
8.54
NEW ZEALAND
8.05
Korea
7.97
Norway
7.77
OECD Europe
7.71
Former USSR
7.56
Italy
7.48
France
6.38
Sweden
6.05
Middle East
5.78
4.62
non-OECD Europe
3.87
WORLD
China
2.30
Latin America
2.15
non-Annex I Parties
1.85
1.13
Asia
Africa
0.96
India
0.93
Bangladesh
0.19
Ethiopia
0.05
0
15
30
tonnes CO2 per person per year
34
45
The greenhouse effect and climate change
Parliamentary Library, August 2001
CO2 emissions by source 1990 - 1999
The main contributors to New Zealand’s growing CO2 emissions are domestic transport, thermal
electricity generation, and industry (Figure 4.7).
New Zealand has the highest proportion of CO2 emissions from transport in the OECD (45.5% vs.
the OECD average of 30%) and transport energy demand has grown, on average, 3.8% per year
over the period 1991 to 1998. Transport use rates over 1990-1998 increased faster than GDP
(freight up 30%, passenger transport up 36%, and real GDP up 23%).10 The increases in transport
have been attributed to the removal of import tariffs on imported cars and a steep increase in car
ownership 1991-1996, the relatively low taxes on transport fuels by OECD standards, and the lack
of fuel economy standards for vehicles.11
The increase in electricity generation emissions has been attributed to deregulation of the energy
sector, a higher than average demand year in 1997, and general growth in demand in the
residential and commercial sectors.12
1990
12000
5826
5411
4812
2866
`
2386
2800
2845
1220
615
2561
6000
3518
8748
1999
672
kilotonnes CO2
11594
Industrial process sources of CO2 include steel, cement and aluminium manufacture. Fugitive fuel
emissions include natural gas venting and emissions from geothermal energy use. Sources of
increases in the industrial sector include the switching from synthetic petrol to methanol production
by Methanex and the growth in dairy production.13
Figure 4.7:
New Zealand’s CO2 emissions
by source, 1990 and 1999.
Source: Ministry of Economic
Development 2000, H. Plume Ministry
for the Environment 2001, pers. comm.
Supercedes data in Ministry for the
Environment 1999, Table 3.2.
Domestic
transport
Industry
Thermal
electricity
generation
Industrial
processes
Residential,
commercial,
agriculture
Other
transformation
Fugitive fuels
0
percent change in CO2 emissions 1990-1999
% of that
% of the
sector
sector
total
Domestic transport
32.5%
58.0%
Thermal electricity
53.8%
38.6%
generation
Industry
21.1%
20.7%
Industrial processes
20.1%
9.8%
Fugitive fuels
9.3%
1.4%
Residential, commercial,
-1.6%
-0.9%
and agriculture
Other energy
-54.2%
-27.3%
transformation
100%
Total
+ 19.2%
10
EECA at http://www.eeca.govt.nz/Conetnt/Sustainable_Transport/transfacts.htm ; see also Table 4.4.
UNFCCC Secretariat 1999, paras. 44-46.
12
Ministry for the Environment 1999, p. 33; Ministry of Economic Development 2000.
13
H. Plume, Ministry for the Environment, pers. comm. 2001
11
35
The greenhouse effect and climate change
Parliamentary Library, August 2001
4.4.2 Methane (CH4)
Currently the largest part of New Zealand’s contribution to climate change is in the form of
methane emissions. This primarily comes from “enteric fermentation” or digestive processes,
particularly in ruminant livestock such as sheep and cattle (Figure 4.8). The current best estimate
for methane production in New Zealand conditions is 88 kg/head/year for cattle and 12
kg/head/year for sheep.14 In ruminant livestock, it is estimated that 90% of the methane emissions
come from the mouth (belching), and 10% from the other end.15
The reduction in methane emissions 1990 to 1999 (Figure 4.4) has been attributed to the
significant lowering of livestock numbers in response to the removal of agricultural subsidies and
unfavourable markets for livestock products.16
Figure 4.8:
New Zealand’s methane
emissions by source,
1999.
Source: Ministry for the
Environment 2001, Table 10.
field burning of land use
crop residues change &
forestry
0.01%
0.33%
industrial
processes
0.01%
manure
management
0.98%
energy
3.16%
enteric
fermentation
81.45%
Note: data uncertainties are
large, e.g. ± 20% for enteric
fermentation, ± 35% for landfills,
and ±25% for wastewater
(Ministry for the Environment
1998, pp. 12, 15).
landfills
7.23%
wastewater
6.84%
1999
Percent change in methane (CH4) emissions 1990-1999
sector
Agriculture
Waste (landfills & waste water)
Energy
Industrial processes
Land use change & forestry
Total
CH4 emissions
kilotonnes (=Gg)
1990
1,492.2
142.7
37.5
0.1
4.2
1999
1,415.5
124.2
54.2
0.1
5.7
% of that
sector
% of the
total
-5.1%
-12.9%
44.7%
0%
34.3%
-4.6%
-99.6%
-24.0%
21.7%
0%
1.9%
100%
New Zealand’s current approach to reducing methane emissions is to research livestock digestion,
pasture composition, and livestock breeding to find ways of improving the efficiency of digestion. It
is hoped that this will obtain the double benefit of reducing methane production and increasing the
efficiency of conversion of food to bodyweight, but results will not be known for quite a few years.17
Other significant human-influenced sources of methane overseas include rice cultivation, coal
mining, and natural gas venting and leaching.
14
Ministry for the Environment 2001, Section Four. Methane production varies with digestibility of feed and metabolic status of the
animals. New Zealand data for swine and poultry is not available, and this potential source is omitted from the New Zealand total.
15
The Press 23/7/01, Putting the squeeze on flatulent cows.
16
UNFCCC Secretariat 1999, para. 21.
17
Government press release 20/5/01, Research the answer to greenhouse gas.
36
The greenhouse effect and climate change
Parliamentary Library, August 2001
4.4.3 Nitrous oxide (NO2)
As with methane emissions, most of New Zealand’s nitrous oxide emissions come from agricultural
activity. These emissions are virtually all from soil management practices: synthetic fertilisers,
animals wastes and waste management systems, and nitrogen leaching and run-off. The largest
source of N2O emissions in New Zealand is thought to be the interaction of livestock urine with the
soil. A small portion also comes from human sewage, and the burning of crop and biomass
residues. Data uncertainties are estimated to be large but they are difficult to quantify; possibly in
the order of ± 60%.18
Other sources of NO2 overseas, besides the sources noted above, are industrial processes
including nylon manufacture.
4.4.3 HFCs, PFCs, and SF6
These compounds are artificial, and invented for a variety of industrial processes. Ironically, many
of the HFCs and PFCs (e.g. HFC-134a and HFC-152a) were created as substitutes for CFCs in
order to protect the ozone layer, but the new gases are also powerful contributors to the
greenhouse effect.
The principal sources of HFCs and PFCs in New Zealand and overseas are refrigeration and other
industrial applications, and aluminium smelting. The principal source of SF6 in New Zealand is
electrical switchgear, and other sources include fire suppression and magnesium production. Other
sources of SF6 overseas include aluminium smelting and industrial applications.
4.5
Links with depletion of the ozone layer
The processes of climate change and ozone layer depletion are quite different, but do affect each
other. These linkages are summarised below.
1. CFCs19, which are key ozone depleting gases in the upper levels of the atmosphere, also act
as greenhouse gases in the lower atmosphere.
2. HFCs and PFCs, which were devised as ozone-friendly substitutes for the CFCs in
refrigeration, are powerful greenhouse gases.
3. Ozone, which is essential to screen out ultraviolet light in the upper atmosphere, acts as a
greenhouse gas when generated as a pollutant to the lower atmosphere.
4. The warming of the lower atmosphere from climate change cools the upper atmosphere. This
is thought to slow down the formation of ozone. HFCs also slow the ozone recovery process.
The delay in ozone layer repair caused by these factors is currently estimated at 15 to 20
years.
The UNFCCC and Kyoto Protocol do not address control of the CFC gases, as they are already
controlled by the Montreal Protocol. This Protocol is working very well, and the ozone layer is
expected to repair itself in about 70 years.20
18
Ministry for the Environment 1998, p. 16; Ministry for the Environment 2001, section four. The New Zealand report to the UNFCCC
secretariat in April 2001 noted that a numerical estimate of uncertainty could not currently be provided.
19
Chlorofluorocarbons.
20
Ministry for the Environment, http://www.mfe.govt.nz/issues/ozone.htm and http://www.mfe.govt.nz/issues/ozone_climate.htm
37
The greenhouse effect and climate change
38
Parliamentary Library, August 2001
5
Carbon sequestration or “carbon sinks”
This chapter is divided into five parts. First to be explained is the global carbon cycle and
definition of carbon sinks (section 5.1). Next, the role of carbon sinks under the Kyoto Protocol is
explored in some detail (overview, including credits for carbon trading and issues; sections 5.2,
and 5.3). Last to be summarised are; the relevant historical context of deforestation and
afforestation in New Zealand, and human-influenced carbon sinks other than forests (sections
5.4. and 5.5).
Background information on the UNFCCC and the Kyoto Protocol is in chapter 1 and more
information on international use of carbon sinks and carbon trading in sections 9.3 and 9.8.
5.1
Overview: the role of forests in the global carbon cycle
5.1.1 The carbon cycle and carbon sequestration
Carbon is an essential building block of life on Earth, and cycles through a multitude of solid,
liquid and gaseous forms.
The natural annual fluxes in the global carbon cycle are massive compared to the carbon
emissions from human activities (Figure 5.1). However, the emissions from human activity
(anthropogenic emissions) are a relatively recent phenomenon and have altered the balance.
All of the released gases have to go somewhere, and without sufficient absorption capacity in
the ocean, soils, or terrestrial biomass, much of the anthropogenic greenhouse gases are
staying in the atmosphere, where they can contribute to climate change.
the atmosphere
90
100
8 GtC
one-way
GtC
GtC
two-way
two-way
people
Fossil fuels & cement
manufacture 6.2 GtC
Land use change 1.5 GtC
the surface ocean
the land
Atmosphere to vegetation 101.5 GtC
Vegetation to soils & detritus 50 GtC
Soils and detritus to atmosphere 50 GtC
Vegetation to atmosphere 50 GtC
Land erosion to ocean 0.8 GtC
Land erosion to ocean 0.8 GtC
Atmosphere to surface ocean 92.4 GtC
Surface ocean to atmosphere 90.8 GtC
Surface ocean to biota 50 GtC
Biota to surface ocean 40 GtC
0.8 GtC
100
GtC
twoway
the intermediate
& deep ocean
Biota to deep ocean 10 GtC
Surface ocean to deep ocean 92.1 GtC
Deep ocean to surface ocean 100 GtC
Deep ocean to ocean sediment 0.2 GtC
Figure 5.1:
Estimated annual fluxes in the global carbon cycle, Gigatonnes of carbon per year
Adapted from http://cdiac.esd.ornl.gov/pns/graphics/blobcarb.gif. Amounts have been rounded for totals in arrows.
GtC = Gigatonne = one billion tonnes = the mass of one cubic kilometre of water.
 Dana Rachelle Peterson 2001
The greenhouse effect and climate change
Parliamentary Library, August 2001
When CO2 is taken in and used by plants in their metabolism, part of the carbon remains as a
structural part of the plant. In trees, this is largely in the form of wood. Half the dry weight of
wood is elemental carbon.1 As long as the tree is alive, or the wood remains undecomposed
(whether standing as deadwood in the forest or in a human-made product), the carbon is held
out of circulation.
Other places where carbon may be held out of circulation for varying lengths of time include
fossil fuels (which hold concentrated carbon from ancient forests and swamps), soils, decaying
organic matter, living creatures, dissolved CO2 in water, and the water and sediments of the
deep ocean. The deep oceans are by far the largest natural global reservoir of carbon (Figure
5.2).
Figure 5.2:
Estimated magnitude of
natural carbon reservoirs in
the global carbon cycle
atmosphere
(1990)
1.75%
terrestrial
vegetation
1.42%
soils,detritus,
land animals
3.68%
surface ocean
2.38%
marine biota
0.01%
dissolved
organic
carbon
1.63%
Sources: Schloerer 1997, Soon et al
1999 p. 150; Hadley Centre for
Climate Prediction at http://www.metoffice.gov.uk/research/hadleycentre
deep ocean
surface
sediment
0.35%
intermediate
and deep
ocean
88.78%
carbon reservoirs
(in gigatonnes of carbon)
This process of carbon storage is often termed sequestration in the climate change literature.
The terms reservoir and sink are used to describe the location of the sequestered carbon,
whether in a forest, the ocean, or the soil. A further distinction is often made, referring to a sink
as a place where active sequestration of carbon is taking place (e.g. a stand of growing trees)
and a reservoir as a static storage place (e.g. a mature forest, where the intake of new carbon is
balanced by the decomposition of old trees).
5.1.2 The role of deforestation in climate change
In many countries humans have removed much of the original forest cover from the earth, and
New Zealand is no exception. Forest clearance has occurred at high and middle latitudes of the
earth over the last several centuries, and in the tropics in the latter part of the 20th century.2
Forest clearance, in addition to reducing biomass and biodiversity, reduces present and future
reservoirs for carbon. It also releases the carbon previously held in the forest through burning
(clearance or fuel wood), exposure of soil carbon to degradation, and wood degradation through
use. Conversion to agricultural land also allows activities which contribute to emissions of the
more powerful greenhouse gases; methane (mainly from ruminant livestock and rice cultivation)
and nitrous oxide (mainly from fertiliser use and livestock).
1
2
Ford-Robertson, Maclaren and Wakelin 2000, p. 189.
IPCC 2000, part 1, paragraph 2.5-2.7.
40
The greenhouse effect and climate change
Parliamentary Library, August 2001
Over the last 150 years land-use change, predominately in forest ecosystems, has contributed
about half the amount of CO2 compared with the burning of fossil fuels and manufacture of
cement, and comprises about a third of the total major sources of CO2 (Figure 5.3).
The data on land-use change (mainly loss of forest
ecosystems) is least certain: error margins are ± 55 GtC,
compared to ±30 GtC for fossil fuel burning and cement
manufacture.
A previous estimate, for 1850 to 1990, is 122 (±40) GtC/yr
for land-use change 230 GtC/yr for fossil fuel burning and
cement manufacture (cited in Ford-Robertson, Maclaren and
Wakelin 2000, p. 189).
300
Gigatonnes of carbon, 1850-1998
Figure 5.3: CO2 emissions from deforestation
compared with fossil fuel burning and
cement manufacture, cumulative 1850 to
1998.
270
200
136
100
0
fossil fuels
land use
and cement
change
manufacture (mainly forest
ecosystems)
5.2
The “Kyoto Forest”
Article 3.3 of the Kyoto Protocol makes provision for “direct human-induced land-use change
and forestry activities limited to afforestation, reforestation and deforestation” occurring since
1990 to be part of the calculations for Annex I Parties to meet their commitments under Article
3.1. The various new and replanted forests that could be eligible for CO2 reduction credits have
been dubbed the “Kyoto Forest”.
The presence of this clause and the “net emissions” approach (emissions minus sinks) was the
result of hard bargaining by New Zealand, Australia, the United States and others.3 New
Zealand’s policy since 1994 has been, in the short term, to achieve a majority of net greenhouse
gas emission reductions through private sector planting of new forests (section 2.2).
Theoretically New Zealand could meet its Kyoto Protocol target in 2008-2012 with the use of
forest sinks alone (Table 5.1). However, the forest sink potential is finite. It has been estimated
that forest sinks may buy New Zealand perhaps 20-50 years to pursue lasting emission
reductions at source. If emissions are not significantly reduced, meeting reduction targets in
subsequent commitment periods will become increasingly difficult.
3
E.g. Gillespie 2000, p. 169-173.
41
The greenhouse effect and climate change
Table 5.1:
Parliamentary Library, August 2001
New Zealand’s anticipated assigned amount, projected net forestry carbon
sinks, and potential emissions for 2008-2012.
total for 2008-2012, Mt CO2 equiv.
source of data
Assigned amount (= allowed emissions)
365
Projected net forestry carbon sinks
100
Emissions if trends 1990-1999 continue
409
balance
Alternative emission projections for
“business as usual”
56 in credit
34 to 50 over
the assigned
amount
current data for 1990 (73,064.35 Gg) times
5 for 5 year period 2008-2012
New Zealand Climate Change Programme
2001a, p.7.
Gross emissions for 1999 (76,831.07 Gg)
plus 4,917.19 from 6.4% increase (19901999 rate for 1999-2008), times 5 for
2008-2012.
Assigned amount plus carbon sinks less
emissions.
Ministry of Economic Development,
Greenhouse gas emissions trading:
Overview (see link below)
Notes for comparison with other sources: 1 Gg (Gigagram, or 1 billion grams) = 1 Kt (kilotonne, or 1,000 tonnes)
= 0.001 Mt (Megatonne, or 1,000,000 tonnes). 1 Gg CO2 equivalent = 0.2727 Gg C (carbon).
Ministry of Economic Development link http://www.med.govt.nz/ers/environment/climate/emissions/index.html
5.2.1
Greenhouse gas reporting and accounting for the land-use change and forestry
sector
Annex I Parties of the UNFCCC are required to report emissions data to the UNFCCC
Secretariat, including data on the effects of land-use change and forestry on greenhouse gas
emissions.4 The data is reported both as gross emissions, and as net emissions with domestic
land-use, land-use change and forestry (LULUCF) effects included.
Reporting vs accounting
It is important to make a distinction between reporting under Article 3.4 of the Protocol, and
accounting under Article 3.3. The reporting allows carbon balances for countries to be
understood and transparent to other Parties, is annual from 1990, and covers a wide range of
land-use activities. The accounting is for a limited range of activities (human-influenced
afforestation, reforestation and deforestation occurring since 1990), and is only for the
commitment periods (2008-2012 being the first one) (Table 5.3).
Table 5.2: Summary of carbon emission and sink accounting and reporting for the land-use
change and forestry sector
accounting 2008-2012
Article 3.3
Article 3.4
reporting
1990 forward
Article 3.4
Afforestation/ reforestation
forests
- existing before 1990
•
- newly planted after 1990
- replanted production forests
Deforestation
harvest of production forests
•
indigenous forests and scrubland
•
includes loss from fire
Other land use activities, post 1990
Bonn agreement: limited these to: forest
management, grazing land management,
cropland management, and revegetation
Other sinks
4
(ocean fertilisation, storage)
optional
cap for forest
management
not mentioned in Kyoto Protocol
Reporting obligations are under both the UNFCCC (Article 4) and the Kyoto Protocol (Article 3.4).
42
optional
The greenhouse effect and climate change
Parliamentary Library, August 2001
Gross vs net
Reporting is both gross and net. Assigned amounts (allowed emissions for the commitment
period) are based on gross emissions at a set year (1990 for most Parties). Accounting is
however based on net emissions (allowing for afforestation, reforestation and deforestation).
Unadjusted, this could have created a potential “gross-net emissions loophole” of some 10%,
reducing incentives to reduce emissions at source. The restrictions on land-use activities and the
1990 cut-off were inserted by negotiators as a pragmatic measure to offset this potential
loophole.
Some LULUCF activities such as deforestation and poor soil management can create emissions,
so net emissions may actually be higher than gross emissions in reporting for some countries
(Figure 4.2b, chapter 4).
New Zealand’s land-use change and forestry sector reporting includes CO2 sinks in new
plantation forests, and CO2 emissions from timber harvest, scrubland clearance for new
plantings, and wildfires. Sinks far outweigh emissions in this sector (Table 5.3).
Table 5.3:
New Zealand’s land-use,
land-use change and
forestry (LULUCF)
reporting for 1999
Source: Ministry for the Environment
2001, Appendix 5. Note: data in
tonnes of carbon rather than of CO2.
carbon
losses
scrub wildfires
harvest of radiata pine forests
harvest of native forests
forest wildfires
carbon uptake in radiata pine
forests before harvest
carbon
sinks
203,704
3,832,376
32,980
58,859
10,204,966
The reporting of greenhouse gas emissions and sinks in the Annual Greenhouse Gas Emissions
Inventory from Annex I countries requires these countries to also report their CO2 emissions
from soils, but New Zealand’s data is currently insufficient for reporting.5
New Zealand’s net emissions data for 1998 (which included the emissions and sinks from the
land-use and forestry sector) was 28% lower than the gross emissions data (without the
emissions and sinks from this sector). New Zealand had the fourth largest gross-net change of
all countries reporting 1998 data (Table 5.4).
5.2.2
Rules governing carbon sinks and trading
Under Article 3.3 of the Kyoto Protocol, countries can gain additional “assigned amount” (sink
credits) for CO2 absorbed during 2008-2012 by forests established by “afforestation and
reforestation” since 1990. This will be measured by the verifiable increase in carbon stock in
these forests over the first five-year commitment period (2008-2012). Any loss in this carbon
stock, which also includes deforestation of forests established prior to 1990, will also mean a
country having to give up equivalent assigned amount. To be in compliance with the Protocol at
the end of 2008-2012, countries need to have sufficient assigned amount to cover their
greenhouse gas emissions.6
At COP6 part two in July 2001, agreement was reached on some of the rules for international
trading of surplus carbon credits. Caps were set on the amount of credits from forestry
management that could be traded, and some countries negotiated much larger amounts than
others (Table 5.5).
5
6
Ministry for the Environment 2001, Documentation section 5. See also section 5.6.1 for a discussion on carbon sinks in soil.
New Zealand Climate Change Programme 2001, p. 8.
43
The greenhouse effect and climate change
Table 5.4:
Percent change in total
greenhouse emissions data
(net compared with gross) for
1998 with inclusion of
estimated effects from landuse, land-use change, and
forestry (LULUCF).
Latvia
Sweden
Norway
New Zealand
Estonia
Ukraine
Finland
France
USA
Switzerland
Ireland
Austria
Spain
Poland
Bulgaria
Portugal
Hungary
Italy
Germany
Slovakia
Canada
Czech Republic
Denmark
Netherlands
Belgium
Greece
UK
Australia
Lithuania
%
-91.3
-37.5
-31.3
-27.9
-15.4
-15.1
-12.7
-12.5
-11.5
-11.4
-10.1
-9.5
-7.9
-7.4
-7.4
-6.2
-5.3
-4.4
-3.3
-3.2
-3.2
-2.5
-1.3
-0.7
-0.7
0
2.2
7.3
32.3
data not reported for 1998
Japan
Iceland
Russian Federation
Parliamentary Library, August 2001
Table 5.5:
The Bonn agreement : caps on volume
of carbon sink trading in forest
management credits, by country 20082012, in Megatonnes of carbon per
year.
Additions to and subtractions from the
assigned amount of a Party undertaken
under Article 6 (carbon trading) shall not
exceed:
Australia
Austria
Belgium
Bulgaria
Canada
Czech Republic
Denmark
Estonia
Finland
France
Germany
Greece
Hungary
Iceland
Ireland
Italy
Japan
Latvia
Liechtenstein
Lithuania
Luxembourg
Monaco
Netherlands
New Zealand
Norway
Poland
Portugal
Romania
Russian Federation
Slovakia
Slovenia
Spain
Sweden
Switzerland
Ukraine
Mt
C/yr
0.00
0.63
0.03
0.37
12.00
0.32
0.05
0.10
0.16
0.88
1.24
0.09
0.29
0.00
0.05
0.18
13.00
0.34
0.01
0.28
0.01
0.00
0.01
0.20
0.40
0.82
0.22
1.10
17.63
0.50
0.36
0.67
0.58
0.50
1.11
Source: UNFCCC Secretariat 2001e
(FCCC/CP/2001/L.7), Appendix Z, p. 12.
Source: FCCC/SBI/2000/11, Table
A.2.
Note: The managed forests in Latvia, Sweden, Norway and New Zealand are largely
not Kyoto Forests under Article 3.3 of the Kyoto Protocol, as they were planted before
1990. Unless the increase in carbon sequestration in these forests can be shown to
be the result of improved forest management post-1990 under Article 3.4, they cannot
be accounted for as carbon credits during 2008-2012. . (P. Maclaren, pers comm, 8/2001).
The caps in Table 5.4 are to limit the use of such credits.
44
The greenhouse effect and climate change
Parliamentary Library, August 2001
Parties will also be required to retain in the national registry (i.e. not trade) at least five times
their most recently reviewed inventory or 90% of their Assigned Amount, whichever is lesser).
Parties that fail to meet their emission reduction targets will be barred from participating in
international carbon trading.7
International carbon trading rules have yet to be decided, but some countries are commencing
pilot domestic schemes (section 9.4). In New Zealand, a discussion document on possible rules
for a domestic forest sinks trading scheme has been released (Table 5.6). A second paper
providing more detail is to be developed later in the year.8
Table 5.6: A possible New Zealand framework for trading sink credits
as put forward for discussion by the New Zealand Climate Change Programme, July 2001
Defining and issuing
sink credits
Eligible activities are limited to afforestation, reforestation and
deforestation (Article 3.3 of the Kyoto Protocol).
•
Credits could be established by statute.
Those undertaking sink activities (“responsible parties”) could be issued
with sink credits in proportion to each unit of CO2 sequestered in a “Kyoto
Forest” (forests established by human action since 1990).
A carbon sink credit could be recognised as a right separated from trees or
land and able to be sold or borrowed against. The basis for claiming
ownership to the carbon would be defined.
•
Legal advice to Government suggests that there is presently no provision in New
Zealand law to define legal ownership of carbon credits.
•
“Responsible parties” will also be liable for carbon losses during
Obligations for
responsible parties
Sink credit and
emissions trading
interface
Measuring,
monitoring,
reporting, and
claiming sink credits
Enforcement and
compliance
Taxation
7
8
the defined period.
Carbon lost through deforestation will also need to be accounted for.
Sink credits could be surrendered or cancelled, or additional credits
purchased to cover emissions.
•
Receipt of credits could be annual, or for a set period.
Receipt of sink credits and incurring of obligation for debits could occur
each year, or at the end of the first commitment period (2008-2012).
•
A point of obligation could be placed with forest owners or other
responsible parties.
This would include all harvesting of Kyoto Forests included in the
accounting system and all deforestation of non-Kyoto forests.
The responsible party would be required to monitor and report emissions
from deforestation, and at the end of each reporting period to hold
“emission units” or have surrendered sufficient sink credits to cover their
emissions.
Sink credits and emission units would:
•
be interchangeable, with the same unit of measurement (e.g. 1 tonne
of CO2 equivalent);
•
be able to be bought and sold both domestically and
internationally; and
•
allow its holder to emit the specified quantity of CO2 equivalent,
and be able to be surrendered to the Government or a body responsible for
authorising emissions.
•
Costs could be borne by sink owners.
Cost-effective methods need to be developed so that the value of sink
credits is not uneconomic.
•
Verified demonstration of carbon sinks would be required.
Options include third party verifiers, self reporting against set standards,
and Government agents.
•
Regular reporting could be supplemented with field audits.
•
Penalties would be imposed for non-compliance.
The taxation implications will need to be considered in due course.
A more detailed summary of COP6 part two decisions is in Box 2, chapter 1.
New Zealand Climate Change Programme 2001, p.4. Other information from the paper in chapter 5 of this report.
45
The greenhouse effect and climate change
5.3
Parliamentary Library, August 2001
Remaining land-use and forestry issues
There are a number of unresolved issues remaining concerning the documentation and trading
of carbon sinks. Some of the key issues relating to Article 3.3 accounting and Article 3.4
LULUCF activities are discussed briefly below.
5.3.1
Adequacy of forestry data for climate change monitoring and reporting
Accurate calculation and verification of carbon sequestration is dependent on good quality data
on a global basis, which is not available. The data available on forests has largely been collected
for political, economic, scientific and cultural reasons unconnected with climate change.
Likewise, the data available on rates of carbon sequestration in forests has been done for only a
few species under limited circumstances, although there are pragmatic approaches available to
inventory biomass change.
All methods being proposed at the moment result in approximations rather than calculations of
actual real effects on atmospheric CO2 or climate change on a global basis, and each country
nominates its preferred carbon accounting method. There are also approximations and
uncertainties associated with emissions data for other gases and sectors (Table 4.2).
There can be a huge difference in the estimated carbon sequestration and thus potential carbon
credits, depending on the definition of activities and the accounting method chosen. For
example, for temperate region forests, a range of IPCC definitional scenarios yielded -126 to
+167 Mt C/yr-1 for average annual global carbon stock changes 2008-2012, and a range of eight
New Zealand carbon stock accounting options yielded -3.5 to +3.5 Mt C/yr (a negative number
means net removal of carbon).9
In submissions to the UNFCCC Secretariat, New Zealand has selected a land-based rather than
an activity-based accounting method. The various land and activity based accounting methods
have quite different implications for different countries for their initial assigned amounts for the
2008-2012 commitment period, and therefore obligations of reduction of emissions.10
A likely response of UNFCCC parties to the major uncertainties this creates will be to either set
an arbitrary standard (not necessarily related to real carbon sequestration effects) and/or to
award carbon sink credits in a very conservative way. The IPCC will develop good practice
guidelines to assist countries in reporting.11
In the past decades, the relevant New Zealand data was provided by a few major forest owners.
With the privatisation of the New Zealand Forest Service and the proliferation of small forest
growers, it has been unclear whether sufficiently reliable forestry statistics will be available, even
with the use of satellite imagery.12
Systems for improved reporting of changes to New Zealand’s carbon pools are being developed.
The new carbon monitoring system will involve five-yearly updates of New Zealand’s land cover,
derived primarily from satellite imagery with supplemental data from ground-based forest and
scrubland plots, soil sampling, and monitoring. Acquisition of a new remote-sensing based Land
Cover Database (LCDB) is scheduled for summer 2001/02.13
9
IPCC 2000, Table 3; Ford-Robertson, Maclaren and Wakelin 2000, p.198-206
e.g. FCCC/SBSTA/2000/9/Add.1, 25/8/00 p. 19 and FCCC/SBSTA/2000/INF.7/Add.1, 3/9/00.
11
IPCC 2000, Part 4, para. 38; J. Barton (MFE) pers comm 8/2001.
12
Ford-Robertson, Maclaren and Wakelin 2000, p. 195.
13
Ministry for the Environment 2001, Section five; J. Barton (MFE) pers comm 8/2001.
10
46
The greenhouse effect and climate change
5.3.2
Parliamentary Library, August 2001
Forest sinks may not provide permanent carbon reservoirs
Trees only continue to absorb additional CO2 while they are growing. The CO2 stored in forests
does not represent a permanent reservoir unless the forest is protected from unsustainable
harvest or destruction by fire, disease, or adverse climate change. Sustainable production
forests, through replanting and appropriate management of harvesting debris, can theoretically
maintain an equilibrium as a net carbon reservoir.
Once trees are harvested and enter trade as wood or paper products, tracking in perpetuity to
determine their permanence is virtually impossible. In recognition of this the IPCC protocols
ignore harvested wood in carbon sequestration calculations: emissions are counted at the time
of harvest rather than when the product decays.14 In future, methods may be agreed on to refine
this accounting.
5.3.3
Forests cannot absorb most of the anthropogenic CO2 emissions
On a global scale, the ability of vegetation to act as a carbon sink is limited compared to the
magnitude of greenhouse gas emissions. Even if all the forests that have been removed by
people were replaced, they would not be sufficient to absorb the CO2 from fossil fuel emissions
during the next century (Figures 5.3 and 5.4). Carbon sink capacity in vegetation is finite and
limited by suitable available land, whereas fossil fuel emissions can continue as long as there is
fossil fuel to burn.
Figure 5.4: Atmospheric
concentration of CO2 :
anthropogenic contributions
1750 to 2100, compared with
terrestrial biosphere carbon
sink potential.
parts per million CO2
750
690
500
260
250
40
70
0
Anthropogenic
CO2, low
estimate
Anthropogenic
CO2, high
estimate
Carbon sink
potential of
terrestrial
biosphere - low
estimate
Carbon sink
potential of
terrestrial
biosphere high estimate
Source: IPCC 2001a, p.7.
Forest sinks can however serve as a short-term and partial response to climate change while
dependence on fossil fuel is reduced. In addition, reductions in net greenhouse gas emissions
will occur if increased use of biomass fuels (e.g. wood and alcohol) are used to replace fossil
fuels, and wood is used to replace materials with high embodied fossil fuel energy such as steel
and concrete.
On a local rather than global scale, a few countries are anticipated to be able to meet a large
proportion of their emission reduction requirements from carbon sinks in the short term (e.g. 20
to 50 years), and New Zealand may well be in this position.15
New Zealand’s contribution to global climate change (even if small compared to the rest of the
world) is, as for most Annex I countries, the result of substantial land-use change and everincreasing emissions from fossil fuel burning. New Zealand’s present increase in forest cover,
14
IPCC 2000, part 4, para 39, on http://www.grida.no/climate/ipcc/land_use/006.htm
Noting however that this goes beyond the first commitment period (Articles 3.1 and 3.3), and the targets and rules for the second
and subsequent commitment periods are yet to be agreed on (Article 3.4).
15
47
The greenhouse effect and climate change
Parliamentary Library, August 2001
through replacing marginal farm lands with production timber plantings, has barely begun to
replace the forest cover that was present in the 1800s when European colonisation began
(section 5.4).
5.3.4
New Zealand’s “Kyoto Forest” is in private ownership
The choice of planting site and rotation cycles for commercial forests has a strong influence on
the effectiveness of forests as carbon sinks. In New Zealand these are individual private sector
decisions driven primarily by economic rather than climate change objectives. The behaviour of
future forest owners can be neither modelled nor controlled, but scenarios can be used to look at
a range of possibilities.
If, for example, timber crop rotation is changed from 28 to 25 years, the carbon density of the
stand is reduced by an estimated 8 tonnes of carbon per hectare. An increase from 28 to 35
years would increase the carbon density per hectare by 20 tonnes by harvest date.16 Multiplying
these figures by the 1.7 million hectares in the existing New Zealand timber estate suggests that
a change of average rotation from 28 to 25 years could remove 13.6 million tonnes of carbon
from New Zealand’s potential carbon sink credits, while an increase in rotation from 28 to 35
years could add 34 million tonnes. This would have a significant impact on New Zealand’s ability
to meet its Kyoto Protocol obligations using forest carbon sinks.
The choices of whether to plant and the timing of harvest will relate to the “discount rate” or time
value used by the owner. Usually commercial rates place a higher value on present or near
future benefits than on medium or distant future benefits. However, the appropriate time value to
support reduction of CO2 may be very different.
Currently new forest plantings continue to be established on previous pasture lands. Whether
and for how long private landholders will choose to continue this trend is unknown.
A further issue relates to ownership of the carbon potential of the sink. Should it belong to the
country or to the private owner of the land the trees are planted on, or a mix of both? Can a way
of compensating tree growers for the carbon sequestered be developed that has low monitoring
and compliance costs?
In July 2000, Cabinet decided that some proportion of benefits from sink credits would go to
those undertaking the sink activities in New Zealand, but details are as yet unclear.17 If domestic
emitters are held legally responsible for emission reductions they may choose to buy forest sink
credits in lieu of reducing emissions, suggesting potential wealth transfers and transaction costs.
5.3.5
Climate change effects on forests
Research has indicated that in laboratory and some field conditions, increased CO2 in the
atmosphere increases photosynthesis, the growth of plants, and therefore rates of carbon
sequestration. However, other factors could reduce this beneficial effect of climate change. For
example:
•
•
16
The ability of plants to increase their photosynthesis is limited by the availability of moisture
and nutrients (however, moisture stressed plants benefit the most from higher CO2 levels).
Different plants respond to increased CO2 differently.18
Ford-Robertson, Maclaren and Wakelin 2000, p. 194-195.
New Zealand Climate Change Programme 2001a, p.10.
18
For example plants with the “C3” metabolism pathway (including cotton, soybeans, sunflowers, oats, barley, wheat, rice, sugar
beets and groundnuts) respond well to increased CO2 , while plants with the “C4” metabolism (such as corn, sugarcane, sorghum,
and sudan grass) respond better to hotter temperatures (FAO 2000, p. 12).
17
48
The greenhouse effect and climate change
•
•
•
Parliamentary Library, August 2001
Heterotrophic respiration (e.g. CO2 from bacteria and fungi involved in decomposition) is
expected to increase as temperatures increase from climate change.
If, as predicted, climate change increases the risk of drought, fire, pests, and heat stress,
forest uptake of carbon may gradually diminish or forests even become a carbon source.19
If trees reach an acceptable sawlog size earlier, they may be felled or decompose earlier,
with a net result of no improvement in carbon sequestration. The uptake of carbon will also
diminish naturally as the age-class structure of global forests changes.20
The UK Hadley Centre for Climate Prediction has warned that as climate change progresses
and temperature rises, the CO2 released from soils through increased decomposition of organic
matter and increased vegetation die-off from heat stress and drought may well feed back into
and accelerate the process of climate change.21
The Bonn agreement excludes carbon sink credit for removals resulting from either elevated
CO2 levels above their pre-industrial level or indirect nitrogen deposition (Article Vii.1.h).
5.3.6
Protection of indigenous forests
The Kyoto Protocol focuses on carbon sinks resulting from direct human-induced LULUCF since
1990. Under Article 3.3 existing non-production indigenous forests are excluded for carbon sink
credit but are an emissions liability if deforestation takes place. Both standing indigenous forests
and their deforestation are reported by Annex I countries under Article 3.4.
The concern has been expressed by environmental groups and others that, as planting new
forests will gain carbon sink credits and the carbon reservoir in old growth indigenous forests is
not formally recognised, there will be a powerful incentive to log these forests in non-Annex I
countries and replace them with fast-rotation production forests. Theoretically in developing
countries, which are not subject to the Kyoto Protocol, the contribution of CO2 from destruction of
existing forests would not be accounted for, yet the developed countries could buy credit for
establishment of “new” forest sinks on those lands.22 In addition to loss of significant existing
carbon reservoirs, this would present a risk of irreversible loss of biodiversity and in many
countries displacement of indigenous people.23
At the Bonn negotiations there was a proposal to define afforestation eligible under the Kyoto
Protocol as human-influenced conversion of land “that has not been forested for a period of at
least 50 years” in order to protect indigenous forests, 24 but it was not formally agreed to. In the
meantime, there have been some pilot carbon sink credit trades involving protection of
indigenous forest in developing countries (Table 9.6).
In New Zealand, felling of indigenous forest is governed by the Resource Management Act 1991
and 1993 amendments to the Forests Act 1949. Export of indigenous timber is now prohibited
unless it comes from an area covered by a registered sustainable forest management plan or
permit.25 The nature of protection for indigenous forest stands under the Resource Management
Act depends on their location and the provisions of the relevant Regional Plan. Harvest of
indigenous forest timber has declined markedly over the last decade. However, clearance of
scrublands (which often contain native species and comprise regenerating native forest or scrub
cover) for new plantation forest planting has sharply increased and then declined (Figure 5.5).
19
IPCC 2000, part I, para. 9.
P. Maclaren, pers comm 7/2001.
21
http://www.met-office.gov.uk/research/hadleycentre/models/carbon_cycle/results.html
22
e.g. Gillespie 2000 pp. 173-177; CNN 13/11/00 Emission credits: case for trees isn’t clear-cut on http://europe.cnn.com ;
http://www.greenpeace.org
23
See reference to the Declaration of the First International Forum of Indigenous People on Climate Change in section 9.8.
24
UNFCCC 2001c, Annex A, clause 1(b), http://www.unfccc.int/resource/docs/cop6secpart/l07.pdf
25
Forests Act 1949 s 67C, inserted by the Forests Amendment Act 1993.
20
49
The greenhouse effect and climate change
Parliamentary Library, August 2001
400,000
14,000
350,000
12,000
300,000
10,000
250,000
8,000
200,000
6,000
`
150,000
100,000
4,000
50,000
2,000
0
LINE: scrubland cleared for
new production
forests (in ha, 3-yr. ave.)
BARS: indigenous forest
harvest (in m3
merchantable timber)
Figure 5.5: Change in harvest of indigenous timber and clearance of scrub for
establishment of new plantation forests in New Zealand, 1989-2000
0
198919901991199219931994199519961997199819992000
Calculations have been done comparing the carbon sequestration rates of production Pinus
radiata forests and kauri trees in New Zealand. While the pine trees have a much faster growth
rate up to harvest, and thus a faster carbon sequestration than kauri over the short term, over
the long term (120 years) the picture is different (Figure 5.6).
Figure 5.6:
Time profile of carbon sequestration over 120 years, kauri compared to radiata
pine. The time period represents one 120- year rotation of kauri and four 30- year
rotations of radiata pine.
1000
800
tonnes
carbon
per
hectare
600
400
200
0
0
20
40
60
80
100
120
Age
st
1 pine
rotation
nd
2 pine
rotation
3rd pine
rotation
Multiple pine
rotations
Kauri
Source: Horgan 1999, p. 78 and Horgan GP pers. comm. 2001. Originally from Hunt DG and Horgan GP (in press),
Some implications for commercial forestry of including a carbon sink value among the outputs, NZ J Forestry Science.
Note: Carbon sequestration rates will depend on the suitability of the site (soil fertility, moisture, temperature, etc.).
This graph will not represent all sites. High standing volumes of carbon may also be achievable on suitable sites with
redwoods, Douglas fir, and other pine species (P. Maclaren, pers comm 7/2001)
50
The greenhouse effect and climate change
Parliamentary Library, August 2001
Depending on the “discount rate” or time value used, the carbon sequestration value of kauri can
be viewed as equal to or greater than pine.26 The carbon sequestration potential of native forest
with mixed species and multiple canopy heights may also be different from a pine tree
monoculture with optimum spacing between trees.
5.3.7
“Polluter pays” principle not addressed
The “polluter pays” principle requires that those who contribute to an environmental problem
should contribute to solving it. In this context, planting carbon sink forests on previous
agricultural land in New Zealand may be viewed appropriate to a certain extent, in that it partially
compensates for the historical loss of carbon reservoir from land clearance and emission of
other greenhouse gases (CH4 and N2O) from agricultural activity. It may also partially
compensate for CO2 emitted by previous farm machinery and present forestry equipment.
However, it does not involve the people, communities, and sectors that have contributed, and
continue to contribute, most of New Zealand’s CO2 emissions (e.g. road transport and industry)
and CH4 and N2O emissions (from the remaining agricultural lands).
If parties which create emissions choose to purchase carbon sink credits to counteract their
emissions, then the polluter pays principle would be given effect. It can also be argued that to a
certain extent, the Kyoto Protocol framework itself addresses the “polluter pays” principle, insofar
as most of the countries causing the bulk of emissions have agreed to do something about it.
5.3.8
CH4 and N2O implications
To understand the effect of LULUCF activities on climate change, the CH4 and N2O implications
also need to be understood. There are CH4 and N2O implications for other land-use change
activities subject to LULUCF credits such as wetland restoration, biomass burning, and forest
fertilisation which are, as yet, poorly understood.
In New Zealand, there is some evidence that forest soils are net sinks for CH4, and of course
planting “Kyoto Forests” on agricultural land displaces a previous source of CH4 and N2O in the
form of livestock and fertiliser use. In New Zealand, commercial foresters tend not to use
fertilisers. The recent dairy boom may in due course have implications for drainage of previous
swampland, loss of soil carbon, and increases in CH4 and N2O.27
New Zealand’s LULUCF accounting includes estimated CH4 and N2O emissions from scrubland
clearance burning and wildfires.28
26
27
28
D. Hunt research and original analysis cited in Horgan 1999, p. 78.
Ford-Robertson, Maclaren and Wakelin 2000, p. 192, footnote 11; P. Maclaren pers comm 2001.
Ministry for the Environment 2001, Appendix 5.
51
The greenhouse effect and climate change
5.4
Parliamentary Library, August 2001
Trends in New Zealand land use and afforestation
Historical land-use and forestry activities prior to 1990 are technically irrelevant under the Kyoto
Protocol. However, they are very relevant to gaining an understanding of how New Zealand has
contributed to, and can reduce contribution to, climate change.
Ruminant livestock (sheep, cattle, deer and goats) are a major source of methane and fertiliser
and livestock are a major source of nitrous oxide. In New Zealand, livestock numbers rose from
38.8 million in 1950 to a peak of 78.36 million in 1982, and have since steadily declined to 57.1
million in 2000.
Source: Agricultural
Statistics data series,
Research and Analysis
section, Parliamentary
Library, on Parliamentary
Intranet (original data from
Ministry of Agriculture and
Forestry, Situation and
Outlook for New Zealand
Agriculture, September
2000 and Statistics New
Zealand INFOS database),
and MAF 2001a, Table 7.
90
2000
80
1800
70
60
50
1600
1400
1200
1000
40
30
20
10
600
400
200
0
19
50
19
55
19
60
19
65
19
70
19
75
19
80
19
85
19
90
19
95
20
00
0
800
LINE: expansion of planted forest are
(1,000 ha)
Figure 5.7:
Trends in New
Zealand livestock
numbers and lands
newly planted in
production forest,
1950 to 2000.
BARS: sheep and cattle (millions)
At the same time, land planted in production forest rose steadily from 0.4 million ha. in 1966 to
1.77 million ha. in 2000 (Figure 5.7). The proportion on previous farmland (about 30%)29
supplanted sources of methane and nitrous oxide. The remainder, on ex- native forest or scrub
lands, removed a carbon sink before starting new sequestration of CO2 . New-land planting of
production forest has averaged 55,000 ha per year since 1990, driven by economic rather than
climate change considerations. The Ministry of Agriculture and Forestry forecasts an average of
40,000 ha of new forest plantings to 2010.30
However, when set against the historic backdrop of deforestation in New Zealand, the new forest
plantings can be seen to so far have replaced only a small proportion of the forest carbon
reservoirs previously removed during the European settlement period to create farmland, or lost
during the Polynesian settlement period from agricultural clearances, fire, and natural climate
change (Figure 5.8).
29
30
P. Maclaren, pers comm 8/2001.
http://www.maf.govt.nz/statistics/primaryindustries/forestry/wsf2000/wsfcontents.htm; D. Lincoln (MAF), pers comm 8/2001.
52
Figure 5.8:
Estimated percentage of New
Zealand under forest cover,
from before human settlement
to the present. Forested lands
as a percent of total land area.
Parliamentary Library, August 2001
forested lands (%)
The greenhouse effect and climate change
100
78
75
53
50
28
29.9
after
European
settlement
at present
(11/99)
25
0
prePolynesian
at start of
European
settlement
Sources: Statistics New Zealand 2000, New Zealand Official Yearbook 2000, pp.
397, 425-426; OECD 1996, p.42.
Currently, of the 8.1 million hectares in forest, the majority is protected native forest in the
Department of Conservation estate. The remainder is on private lands (1.7 m ha. or 21 % in
production forest; 1.3 m ha. or 16% in native forest legally capable of being logged subject to
consents, and 0.2 m ha., or 2.5% of other native forest on private land).31 This data is presented
together with other land-use types in Figure 5.9.
Privately
owned natural
forest
4.0%
Figure 5.9:
Proportion of
New Zealand
land in forest
and other land
uses, 2001.
State-owned
natural forest
19.2%
Other nonforested land
8.8%
2001
Planted
production
forests
6.6%
Shrubland
10.0%
Tussock
grassland
7.5%
Pastoral,
horticulture,
and arable
44.0%
Source: Ministry of Agriculture and Forestry 2001b, Tables A2 and A3.
Notes: 83,000 ha on minor offshore islands has been excluded from the total.
The state-owned forest is all in protected areas managed by DOC, except for
12,000 ha in Southland (Waitutu Inc cutting rights).
31
Source as for Figure 5.8.
53
The greenhouse effect and climate change
5.5
Parliamentary Library, August 2001
Other greenhouse gas sinks
5.5.1 Soil management
Carbon can be held in the soil in such forms as organic matter, humus and roots, and can be
lost from the soil through erosion, compaction, mineralisation, decline of soil structure, and
humus oxidation.32 Agricultural practices which result in the decline of soil fertility and organic
carbon in the soil include:
•
ploughing;
•
burning of biomass;
•
deforestation and draining of wetlands for agricultural development;
•
over-grazing and grazing erosion-prone soils;
low productivity subsistence agriculture (“mining” of soil fertility).
•
Once lost from the soil reservoir, this carbon can enter the atmosphere and become available to
effect climate change.
The soil’s capacity to hold carbon can be enhanced through such activities as:
•
conservation tillage
•
liberal use of mulch, compost, cover crops, and other organic amendments;
•
elimination of bare fallow;
•
integrated nutrient management;
•
restoring eroded and salt affected lands;
•
agro-forestry, afforestation, and tree protection;
•
longer fallow periods in slash and burn agriculture;
•
improved pasture management;
•
crop rotation changes; and
•
preventing erosion and water conservation and management.33
“Conservation tillage” is defined as having at least 30% of crop residues covering the soil at
planting. It is being increasingly practised in some developing countries, and is an integral part of
organic agricultural practice. Reducing the amount of tillage (ploughing) protects soil organic
matter from decomposition by minimising the chances of soil erosion. A hectare of unploughed
field can absorb up to a tonne of carbon every year.34 Increasing numbers of developing country
farmers are finding that zero-tillage makes both practical and economic sense.35
The eventual global carbon sequestration potential if significant areas of degraded lands were to
be restored has been estimated at 125 Mt C/yr -1 for improved cropland management, 240 Mt
C/yr -1 for improved grazing land management, 26 Mt C/yr -1 for improved agroforestry, and 2 Mt
C/yr -1 for improved urban land management.36
In New Zealand, Landcare Research has commenced a two-year study to produce a national
soil erosion related carbon budget. New Zealand is considered to have a relatively high rate of
carbon loss from erosion compared many other Annex I countries. Preliminary results from the
Waipoua Basin north of Gisborne suggest that in some areas annual losses may almost equal
the carbon sequestered by New Zealand‘s plantation forests.37 Under the Kyoto Protocol, the
new data will in due course need to be compared against New Zealand’s 1990 baseline.
32
Humus is the organic constituent of soil, formed by the decomposition of dead plants and animals.
Tiwari 2000, p. 44 and Appendix 4.2.
This would not be indefinitely, and the flux will depend on the initial soil carbon level, cropping, and other management variables.
35
New Scientist, An ordinary miracle, 3 February 2001, pp. 16-17;
36
Lal 1997 and Dixon et al 1994, cited in Tiwari 2000, p. 47.
37
Landcare Research, press release 5/701, Eroding uncertainty in New Zealand’s greenhouse gas emissions,
http://www.newsroom.co.nz .
33
34
54
The greenhouse effect and climate change
Parliamentary Library, August 2001
5.5.2 Ocean storage
New Zealand scientists have been among those worldwide involved in experiments to enhance
the ocean’s absorption of CO2 by fertilising the open ocean and encouraging the growth of plants
to absorb CO2. Initial results using iron sulphate have resulted in massive blooms of
phytoplankton (green algae), but it is not yet known how much CO2 has been taken up, or
whether it is held in the sea or subsequently released back to the atmosphere. The implications
for ecosystem health, including oxygen depletion and the impact on fisheries, are also
unknown.38
Another research project proposes injection of liquid CO2 onto the seabed offshore from Kona,
Hawaii. Computer models from Norway suggest that the gas would dissolve and the heavier
gas-rich water would sink to the bottom. Questions have been raised about the effect of the
acidity of the CO2-rich water seawater on sealife and the seabed, and the interaction with ocean
circulation altered in future due to climate change.39
5.5.3 Underground storage40
Pilot projects and research are underway in Canada, Japan, Norway, the Netherlands, and the
USA on the capture and storage underground of CO2. The greatest potential is seen for large
point-sources of CO2 such as power stations, which are near to coal and oil mining areas where
underground injection may be feasible. CO2 is already used commercially in about 70 oil fields
worldwide to inject into oil reserves to reduce viscosity and maximise extraction. Potential is
seen for the use of injected CO2 to displace methane from unmineable coal seams, and capture
of the methane for power production. Technology is also being developed to remove CO2 prior to
emission from coal burning.
There remain significant technical and cost difficulties in the capture and transport of CO2, and
the permanence of this option has not yet been demonstrated.
The Sleipner Project in the North Sea off Norway involves removal of CO2 from natural gas,
which is compressed and injected into an aquifer 1,000 metres below the seabed. This is the
largest CO2 capture and storage project in the world, sequestering about 1 million tonnes of CO2
per year. Another project in the Weyburn oil field (Saskatchewan, Canada) proposes injection of
about 14 million tonnes of CO2 underground to recover an incremental 120 million barrels of oil
over the next 15-20 years.
38
CNN 23/1/01, Ocean fertilisation yields hope, uncertainty for global warming on http://europe.com ; INL 21/2/01, NIWA scientists
find way to decrease greenhouse gases on http://www.stuff.co.nz/inl
39
Carbon sunk, New Scientist 30/6/01, p.19.
40
The source for this section is the Natural Resources Canada 2000, pp. 1-13.
55
The greenhouse effect and climate change
56
Parliamentary Library, August 2001
Part C: The estimated risk and impacts of climate change
6
Climate change: current scientific understanding and projections
The Intergovernmental Panel on Climate Change (IPCC) was formed by the United Nations
Environment Programme (UNEP) and the World Meteorological Organisation (WMO) in 1988.
The IPCC reports go through a thorough scientific and political review process before being
approved by the consensus of government delegates at a plenary meeting, and are therefore the
best international scientific consensus statements available on the issue of climate change.1 In
order to reach an international consensus, extreme views are moderated and consequently the
published projections are conservative and carefully qualified.
The most recent reports, issued in early 2001, build on the earlier reports, using the latest
scientific data and improved understanding of climate processes. The earlier projections had
attracted some criticism because of the limitations of the data and therefore the uncertainties of
conclusions. The latest reports use careful and explicit language to describe the state of
scientific understanding and degrees of certainty. As well, distinctions are made for regional,
hemispheric, and global statements, based on the completeness of the data.
6.1
Summary of IPCC Third Assessment Report of Working Group One
The Third Assessment Report of Working Group I of the Intergovernmental Panel on Climate
Change (IPCC) was approved by member countries in Shanghai in January 2001. Many
hundreds of specialists contributed to it.2 The summary conclusions, supported by the scientific
data on climate change, are as follows.
•
An increasing body of observations gives a collective picture of a warming world and other
changes in the climate system.
•
Emission of greenhouse gases and aerosols due to human activities continue to alter the
atmosphere in ways that are expected to affect the climate.
•
Confidence in the ability of models to project future climate has increased.
•
There is new and stronger evidence that most of the warming observed over the last 50
years is attributable to human activities.
•
Human influences will continue to change atmospheric composition throughout the 21st
century.
•
Global average temperature and sea level are projected to rise under all 35 IPCC emission
scenarios.
•
Anthropogenic climate change will persist for many centuries.
•
Further action is required to address remaining gaps in information and understanding.3
The key details underlying these conclusions, and their degree of certainty, are summarised in
Table 6.1. A brief explanation of the “emission scenarios” and how they contribute to the climate
change projections is in section 6.2.
1
2
3
For more detail see the IPCC website http://www.ipcc.ch
122 co-ordinating lead authors, 516 contributing authors, 21 review editors, and 337 expert reviewers.
IPCC 2001a.
The greenhouse effect and climate change
Table 6.1:
Parliamentary Library, August 2001
Principal conclusions of IPCC Working Group I relating to climate change
Likelihood
IPCC conclusions and projections 2001
Observations
Temperature
The global average surface temperature has increased about 0.6°C since 1861.
(This is 0.15°C higher than the previous estimate up to 1994 due to higher temperatures 1995-2000, and takes
into account various adjustments including urban heat island effects.)
The 1990s was the warmest decade, and 1998 the warmest year, since 1861.
This is the largest increase for any century during the last 1,000 years.
Average night-time daily minimum temperatures over land increasing about twice the rate of daytime
temperatures. The rate of increase over the sea was about half that over land.
The lowest 8 km of atmosphere has also warmed since the 1950s. This zone is differently influenced
by cooling factors such as aerosols, ozone depletion, and atmospheric circulation patterns.
The freeze-free season has lengthened in many regions between 1950-1993.
Some areas have not warmed (parts of the Southern Hemisphere and Antarctica).
factual
90-99%
66-90%
factual
factual
factual
Interpretation
Most of the observed warming over the last 50 years is due to an increase in greenhouse gas
concentrations.
In the context of the climate over the last 1,000 years, the warming during the 20th century is unusual
and not of entirely natural origin.
Only those climate models which include both natural climate variability
and anthropogenic emissions successfully reproduce 20th century climate changes.
66-90%
66-90%
Projections based on 35 emission scenarios
Over the next century, there will be higher maximum and minimum temperatures, more hot
days, fewer cold and frost days, and reduced diurnal range over nearly all land areas.
Global average temperature will increase by 1.4°C to 5.8°C over the period 1990 to 2100. (Note:
this is more than the previous IPCC projection of 1.0°C to 3.5°C, due primarily to improved models and lower
projected sulphur dioxide emissions.)
This rate of warming will be without precedent during at least the last 10,000 years.
Observations
Snow &
ice
Spring and summer sea-ice has decreased about 10-15% in the Northern Hemisphere since the
1950s.
th
Mountain glaciers in non-polar areas have been in widespread retreat in the 20 century.
Snow cover has decreased about 10% since the late 1960s (satellite data).
There has been about two weeks’ reduction in annual duration of lake and river ice in the midth
to- high latitudes of the Northern Hemisphere over the 20 century (surface data).
Arctic sea-ice thickness has declined 40% during late summer/early autumn, and declined
considerably more slowly in the winter, in recent decades.
No significant trends in Antarctic sea-ice apparent since 1978.
90-99%
range from
scenarios
90-99%
factual
factual
90-99%
90-99%
66-90%
factual
Projections based on 35 emission scenarios
Further decrease in sea-ice and snow cover, retreat of glaciers and ice caps.
Local warming over Greenland will be 1 to 3 times the global average.
At 5.5°C over 1,000 years the partial melting of the Greenland ice sheet would contribute about 3
metres to sea level rise.
No major loss of ice from the West Antarctic Ice Sheet. Antarctica may gain ice mass from local
increases in precipitation.
Observations
Sealevel
th
Global average sea level has risen 0.1 to 0.2 metres during the 20 century (tide gauge data).
The 20th century warming has contributed significantly to observed sea level rise, through
thermal expansion of sea water and widespread loss of land ice.
Projections based on 35 emission scenarios
The global mean sea level will rise by 0.09 to 0.88 metres over the period 1990 to 2100.
(Note: this is less than the previous IPCC projection of 0.13 to 0.94 metres, due to improved models which give
smaller contributions from melting glaciers and ice sheets.)
66-90%
90-99%
factual
90-99%
range from
scenarios
(continued next page) ⇒⇒⇒
58
The greenhouse effect and climate change
Parliamentary Library, August 2001
(Table 6.1, continued)
Type of
effect
Rainfall
&
drought
IPCC conclusions and projections 2001
th
Likelihood
Observations
Precipitation has changed over the 20 century in many areas:
- Mid- and high latitudes of the Northern Hemisphere: increase of 0.5% to 1% per decade;
and since mid-century, 2% - 4% increased frequency of heavy precipitation events.
- Tropical latitudes (10°N to 10°S): 0.2% - 0.3% increase per decade, but not recently.
- Northern Hemisphere sub-tropics (10°N to 10°S): 0.3% decrease per decade.
No discernible hemisphere-wide changes in the Southern Hemisphere.
90-99%
66-90%
66-90%
66-90%
factual
90-99%
Projections based on 35 emission scenarios
Over the next century, there will be more intense precipitation events over many areas and
increased summer drying and associated risk of drought in most mid-latitude continental
interiors.
Observation
El Niño
Warm episodes of the El Niño-Southern Oscillation (affecting regional variation in precipitation and
temperature in New Zealand and many other areas) have been more frequent, persistent, and
intense since the mid-1970s, compared to the previous 100 years.
Projections based on 35 emission scenarios
There are some shortcomings in the current climate models that simulate El Niño. Projections from
these models show little change or a small increase in the amplitude of El Niño events over the next
100 years.
Even with little or no change in El Niño amplitude, global warming will lead to greater extremes of
drying and heavy rainfall (risk of droughts and floods) in many regions affected by El Niño.
Storms
Timing
of
change
Projections based on 35 emission scenarios
Over the next century tropical cyclone peak wind and rain intensity patterns will increase in
some areas.
Anthropogenic climate change will persist for many centuries.
Most of the greenhouse gases are long-lived in the atmosphere, and the deep ocean and ice-sheets
adjust to climate change over a very long timescale.
Even after greenhouse gases are stabilised, changes in temperature and sea level will continue for
centuries, but at much slower rates than if the gases weren’t stabilised.
see Table
6.2
factual
factual
66-90%
see Table
6.2
factual
Source: IPCC 2001a, at http://www.ipcc.ch
6.2
Natural vs. anthropogenic4 influences on the climate
Scientific understanding of the “greenhouse”, or heat retention, effect on the atmosphere from
CO2 and the other greenhouse gases is now well developed. New understanding is now
emerging of how many other factors, both natural and anthropogenic,4 can cause “radiative
forcing” or contribute to warming and cooling. These factors are summarised in Figure 6.1.
The model simulations that best agree with climate observations over the last 140 years include
both natural and anthropogenic radiative forcing effects (Figure 6.2). Climate change simulations
which use only the natural influences cannot explain warming over the last 50 years, but do
indicate that natural influence may have contributed to warming in the first half of the 20th
century.5
Data from many locations have indicated that surface temperatures in urban areas can be
warmer than in comparable rural areas. This “urban heat island effect” has featured in
arguments of climate change skeptics, and has been extensively studied and debated. The most
recent IPCC reports have allowed for this possible bias in analysing the available data.6 Another
new adjustment in the models is for sulphate aerosols (an emission from some fossil fuels,
especially coal), which act to cool rather than warm the climate (Figure 6.1).
4
5
6
Anthropogenic = created by human activity.
IPCC 2001a, p. 6 and Figure 4.
Soon et al. 1999, pp. 152-153; 2001a, p. 1; IPCC 2001d, Chapter 3.
59
The greenhouse effect and climate change
Parliamentary Library, August 2001
It is important to realise that improved scientific understanding of climate change will not
necessarily result in projections of lesser impact. Some of the recent discoveries are pointing to
possible feedback loops in the global climate and biological systems that may mean accelerated
climate change. Potential results from climate change which will have unknown and possibly
rapid feedback impacts include:
Increased rates of organic decomposition with increased temperatures, which could add
more greenhouse gas emissions to the atmosphere from natural sources;
Increased methane from melting permafrost;
Slowing of the ocean currents due to changes in sea ice and salinity (potential loss of carbon
absorption potential and loss of Gulf Stream warming to northern Europe);
Loss of the Amazon rainforest as a carbon sink due to heat and drought;
A lengthened fire season for northern hemisphere forests (potential for increased emissions
and loss of carbon sinks); and,
Greater absorption of heat from the sun to feed back into the greenhouse effect if large areas
of arctic tundra are afforested for carbon sinks, as is planned in Russia and Canada
(replacement of snow and ice covered reflective surfaces with conifer tree dark surfaces).7
•
•
•
•
•
•
Table 6.2:
Extreme weather and climate events: estimates of confidence in observed and
projected changes
Confidence in observed changes
(latter half of the 20th century)
Likely
Very likely
Very likely
Likely, over many areas
Likely, over many Northern
Hemisphere mid- to high latitude
land areas
Likely, in a few areas
Not observed in the few analyses
available
Insufficient data for assessment
Changes in Phenomenon
Higher maximum
temperatures and more
hot days over nearly all
land areas
Higher minimum
temperatures, fewer cold
days and frost days over
nearly all land areas
Reduced diurnal
temperature range over
most land areas
Increase of heat index over
land areas
More intense precipitation
a
events
Increased summer
continental drying and
associated risk of drought
Increase in tropical
cyclone peak wind
b
intensities
Increase in tropical
cyclone mean and peak
b
precipitation intensities
Confidence in projected changes
(during the 21st century)
Very likely
Very likely
Very likely
Very likely, over most areas
Very likely, over many areas
Likely, over most mid-latitude
continental interiors. (Lack of
consistent projections in other
areas)
Likely, over some areas
Likely, over some areas
Reproduced with permission of the Intergovernmental Panel on Climate Change
Source: IPCC 2001a, Table 1. at http://www.ipcc.ch
Confidence levels = chance of being true: virtually certain (greater than 99%); very likely (90-99%); likely (66-90%); medium
likelihood (33- 66%); unlikely (10-33%); very unlikely (1-10%); exceptionally unlikely (less than 1%).
Heat index” = a combination of temperature and humidity that measures effects on human comfort. “Diurnal” = daytime
a = For other areas, there are either insufficient data of conflicting analysis.
b = Past and future changes in tropical cyclone location and frequency are uncertain.
7
http://www.met-office.gov.uk/research/hadleycentre/models/carbon_cycle/results.html ; New Scientist 14/7/01, p.18 and 21/7/01
pp. 4-5.
60
The greenhouse effect and climate change
Parliamentary Library, August 2001
Figure 6.1 : Natural and anthropogenic “radiative forcing” factors known to affect climate.
Reproduced with permission of the Intergovernmental Panel on Climate Change
Source: IPCC 2001a, Figure 3. References to other figures and chapters relate to that report.
61
The greenhouse effect and climate change
Figure 6.2:
Parliamentary Library, August 2001
Projections based on natural, anthropogenic, and combined climate change factors,
compared to actual observed temperatures 1850 to 2000.
Figure 4: Simulating the Earth's temperature variations, and comparing the results to measured changes, can
provide insight into the underlying causes of the major changes.
A climate model can be used to simulate the temperature changes that occur both from natural and anthropogenic
causes. The simulations represented by the band in (a) were done with only natural forcings: solar variation and
volcanic activity. Those encompassed by the band in (b) were done with anthropogenic forcings: greenhouse gases
and an estimate of sulphate aerosols, and those encompassed by the band in (c) were done with both natural and
anthropogenic factors included. From (b), it can be seen that inclusion of anthropogenic forcings provides a plausible
explanation for a substantial part of the observed temperature changes over the past century, but the best match with
observations is obtained in (c) when both natural and anthropogenic factors are included. These results show that the
forcings included are sufficient to explain the observed changes, but do not exclude the possibility that other forcings
may also have contributed. The bands of model results presented here are for four runs from the same model. Similar
results to those in (b) are obtained with other models with anthropogenic forcing. (Based upon Chapter 12, Figure
12.7)
Reproduced with permission of the Intergovernmental Panel on Climate Change
Source: IPCC 2001a, Figure 4. References to other figures and chapters for that report.
62
The greenhouse effect and climate change
6.3
Parliamentary Library, August 2001
Climate change models, emission scenarios, and climate change
projections
There are three main bases for climate change projections: observed climate data, climate
models, and emission scenarios. The models are based on the best scientific understanding of
how climate works, currently including key “radiative forcing” elements summarised in Figure
6.1, and the known data on climate trends. When the models and the scenarios are combined, a
range of possible outcomes is obtained.
The climate change models used now are very sophisticated compared to those available for the
first climate change projections. In the early 1990s, the models included atmosphere, land
surface, ocean and sea-ice factors. The current models also include sulphate aerosols, nonsulphate aerosols, and carbon cycle dynamics. In future, vegetation dynamics and details of
atmospheric chemistry are likely to be added.8
The 35 emission scenarios currently used by the IPCC replace the previous IS92 scenarios, and
were approved in March 2001. They utilise variations in four narrative storylines that build on
current social, economic and fuel consumption trends (Table 6.3). All of the scenarios describe
futures generally more affluent than today, with lower populations than the IS92 scenarios. None
explicitly assume implementation of the UNFCCC or the Kyoto Protocol.
Table 6.3:
Summary of the SRES emission scenario storyline groupings, IPCC 2001.
A1 Storyline
Economic growth very rapid, rapid introduction of more efficient technologies, substantial
reduction in regional differences in per capita income, capacity building. Population peaks midcentury and declines thereafter. Further divided into three energy scenarios:
A1FI – fossil fuel intensive, A1T – non-fossil energy sources, and A1T – balance of fuel sources.
A2 Storyline
Emphasis on local identity, self-reliance, and regional economic development. Fertility and
economic differences between countries converge very slowly. Population continues to increase.
B1 Storyline
Population pattern as for the A1 storyline, but rapid development of service and information
economies and introduction of clean and resource-efficient technologies. Improved equity.
B2 Storyline
Population pattern as for the A2 storyline but increasing at a slower rate. Intermediate levels of
economic development and less rapid technological change than B1 and A1 storylines.
SRES = IPCC Special Report on Emissions Scenarios. Source: IPCC 2001c, pp. 4-5.
The greenhouse gas emission and climate change implications of these scenarios diverge
markedly over longer timeframes, which is why the projections for temperature and sea-level
change are expressed in a wide range.
A summary of the likely atmospheric concentrations of the greenhouse gases with a continuation
of current emission trends, and the likely associated climate changes, is presented in Table 8.1
with a discussion on emission reduction targets.
8
IPCC 2001d, pp. 48-49.
63
The greenhouse effect and climate change
6.4
Parliamentary Library, August 2001
Paleoclimatic data: climate trends over millions of years9
Scientists have collected various forms of data that allow reconstruction of an imperfect but
increasingly clear picture of what the Earth’s climate was like in the distant past. The data
sources include ice cores from Greenland, the Antarctic, and high mountain glaciers in Asia and
South America (which show both annual layers and some bubbles of preserved ancient air); the
fossil and pollen record (showing warm climate species in presently cold climate zones); and
some older historical records (e.g. 2,000-year-old Chinese records and 800-year-old records in
Europe).
The data shows earlier periods were as warm or warmer than the present, and that these
alternated with periods that were very much colder. Factors thought to have contributed to global
warming and cooling in the distant past are: changes in land and ocean-floor topography
(affecting patterns of land absorption, and air and ocean currents); changes in the tilt of the
earth’s orbit and axis; fluctuations in radiation output from the sun; and changes in circulation
and rising of deep ocean waters in the oceans. At least one occasion, a change in Atlantic
Ocean currents causing a marked ice age cooling of Europe, was relatively abrupt (over 5-10
years).10
•
Ice ages and interglacial periods
Over cycles of tens of thousands of years before human interference in the atmosphere the
climate of the Earth has alternated between warm and cold periods.
From about 23,000 to 15,000 years before present, huge ice sheets extended into areas in the
northern hemisphere that now have a temperate climate. In New Zealand, the last Glacial
Maximum was between 26,000 and 18,000 years ago, and temperatures were an estimated 4°
to 5°C lower than at present. New Zealand’s last warm period was between about 10,000 and
8,000 years ago and temperatures were about 1°C above modern values. The fossil record
indicates a mild climate, light winds, and lush forests at this time.11
In Europe about 1,000 years ago, the Medieval Climate Optimum occurred. While the climate in
Europe was colder than at present, it was mild enough in Greenland to allow colonists to settle
there until the “Little Ice Age” occurred. This latter period lasted from approximately 1450 to
1890, and the global mean temperature has been estimated at 0.5° to 1.0°C lower than today.
The medieval warm period was not synchronised around the planet: historical records show that
it had ended 900 years ago in China and Japan, but continued for some two to three more
centuries in North America and Europe.12
Air bubbles found in the Vostok ice core in Antarctica show that concentrations of the
greenhouse gases CO2, CH4, and N2O have varied systematically with the coming and going of
periodic ice ages and warm periods over the last 150,000 years. This demonstrates a
relationship between these gases and climate change which predates significant human
intervention in the atmosphere.
The human contribution to greenhouse gases since the Industrial Revolution has created higher
concentrations of CO2 and methane than at any other time in the past 420,000 years. The CO2
concentration in the atmosphere is higher than it has been for the last 20 million years.
9
Crowley 1996. Except where noted this is the source for the whole section.
Natural Climate Fluctuations on http://katipo.niwa.cri.nz/ClimateFuture .
11
Past Climate Variation over New Zealand on http://katipo.niwa.cri.nz/ClimateFuture/Past_Climate.htm
12
Soon et. al 1999, p. 151.
10
64
The greenhouse effect and climate change
Parliamentary Library, August 2001
The Cretaceous (Age of Dinosaurs) and Cambrian Periods
•
About 100 million years ago, the mean global temperature may have been as much as 6° to 8°C
warmer than it is today. The amount of CO2 in the atmosphere then is thought to have been
about three times what it is today, and roughly equivalent to the current “fossil fuel reservoir.“ In
other words, the “carbon sinks” in ancient forests and swamps converted over time into coal, oil
and natural gas are now being released by human intervention back into the atmosphere. It is
estimated that a doubling of current CO2 emission levels will result from the consumption of
about 20% of the theoretical fossil fuel reserves, and that at current emission rates we could
theoretically return to a Cretaceous type climate (i.e. 6° to 8°C warmer) in AD 2400-2700.13
About 500 to 550 million years ago, there was a peak in CO2 estimated to have been three times
that of the Cretaceous Period. After this, trees evolved, and one theory suggests that over time
their absorption of huge amounts of CO2 precipitated a major ice age.14
6.5
Climate changes in Australia and New Zealand over the last 140 years
Consistent with other parts of the world, temperatures and sea level in New Zealand and
Australia have increased over the last century. These and other related changes are
summarised in Table 6.4.
Table 6.4: Climate change effects observed over the last century in New Zealand and Australia.
Parameter
Temperature
- mean
- maximums
& minimums
- diurnal
range
- oceans
Rainfall
Sea level
Storms
Atmospheric
concentration of
greenhouse gases
Observed change over last century
Risen 0.5°C to 0.9°C since 1900. Large ranges between decades (presumed
natural origin), with average rise of 0.1°C per decade (New Zealand, 0.11°C).
Largest increases have been since about 1950.
Highest recorded temperatures ever in the last decade.
Frequency of very warm days increased.
Frequency of very cold nights decreased.
Night time temperatures have risen faster than day time temperatures, and the
difference between them has decreased by up to 1°C over the last 40 years.
This appears to be connected to increased cloud cover (e.g. 5% increase in
cloud cover since 1910 in Australia, with largest changes in spring).
Water temperature generally rising.
Rainfall is very variable from place to place, and changes correlate with cyclical
El Niño events. A regional trend is not clear.
Over large areas of Australia, increases in frequency of heavy rainfalls and
average rainfall recorded, with larger increases in the summer half-year.
Over the last 50-100 years, risen on average by about 20 mm per decade
(complicated by changes in land elevation due to geotectonic forces in some
areas, slow land uplift and redistribution of ocean currents since the last ice
age, and regional sea level variations).
Data not conclusive, especially for non-tropical (mid-latitude) storms.
More tropical cyclones reported over last few decades north of New Zealand,
but improved accuracy of data raises doubts about comparison with earlier
years.
In Australia between 1969/70 and 1995/96, the total number of cyclones
decreased but their intensity and duration increased.
CO2 concentration 30% higher than in pre-industrial times, and increasing by
about 0.4% each year.
Measurements of methane (since 1989 at Baring Head, New Zealand) and
nitrous oxide (since 1995 at Baring Head, and since late 1970s in Australia)
show a steady increase in concentration.
The diurnal temperature range is the difference between night time and day time temperatures.
Source: Basher and Pittock 1998, section 4.2.2; http://katipo.niwa.cri.nz/ClimateFuture/Gases.htm ; A. Reisinger, Ministry for
the Environment and B. Mullan NIWA, pers. comm. 7/2001.
13
Note, however, that the fossil record of the tropical climate and biota during this period does not relate to exactly the same
locations on Earth as at present. For example, during the Cretaceous period New Zealand did not exist and Antarctica was not at the
southern pole.
14
New Scientist 16/6/01, pp. 30-33, The Kingdoms of Gaia.
65
The greenhouse effect and climate change
6.6
Parliamentary Library, August 2001
Climate change projections for New Zealand15
The IPCC has concluded that the current global climate models provide useful projections at the
global and continental scale, but allow for little confidence at smaller scales. Reasons include:
•
•
•
the resolution of global climate change models and data sets is often not fine enough to
reflect local weather conditions;
the climate models cannot yet accurately predict the El Niño - Southern Oscillation
phenomenon, which has a major influence on the climate of the Australasian region; and,
knowledge about the sensitivities of natural and managed agricultural systems to climate
change is limited.
Projections for regions (e.g. Australia and New Zealand combined) are generally much less
certain than global projections, and different climate models tend to show greater differences in
their regional results than on global scales.
To create regional and local climate projections, scientists use global climate models and a
technique called “downscaling”. The historical correlation between local weather patterns and
large-scale regional climate patterns is applied to global models simulations to make projections
for the future. Agreement among scientists can be reached on the likelihood of some local
climate features but not others, so that it is better to use a range of plausible projections in
decision-making.
In 1992 (for Australia) and in 1994 (for New Zealand), scenarios were published which were
used for local projections in many subsequent publications. These models employed the best
global models and scenarios available at the time, but used an “equilibrium” level of greenhouse
gases at a future date and a static ocean model. The more recent models use “transient” levels
of the greenhouse gases (modelling changes as they occur) and a dynamic ocean model, and
produce projections with some different aspects than the earlier models (Table 6.5).
The latest projections for temperature and rainfall change in New Zealand are presented in
Figure 6.3. They are an average of downscaling of four global climate models for the New
Zealand situation.16 It is important to note in addition to these mean values that a significant
increase in the risk of extreme rainfall events (floods or droughts) is also predicted.
Projections for sea-level rise in New Zealand vary from international rates due to “glacial
isostatic rebound”: the land area is still rising about 4 cm a century as an after-effect of the
removal of glacial ice sheets from the last ice age. The most recent projections for New Zealand
are sea-level rises of 3 cm to 25 cm in 2050 (worst case scenario 30 cm) and 9 cm to 66 cm in
2100 (worst case scenario 84 cm).17
Likely local impacts are summarised in chapter 7. More detail is available in the recent report
Climate Change Impacts in New Zealand, published by the Ministry for the Environment in mid
July 2001.18
15
Unless otherwise noted, the main sources for this section are Basher and Pittock 1998, and the National Institute of Water and
Atmospheric Research website http://katipo.niwa.cri.nz/ClimateFuture .
The four models are from the Australian Commonwealth Scientific and Industrial Research Organisation (model CSIRO9), UK
Hadley Centre for Climate Prediction (model HadCM2), Canadian Centre for Climate Modelling (model CCC), and Japanese Centre
for Climate Research (model CCSR). The emissions scenario used was IS92a, roughly equivalent to SRES scenario A1. The
changes can also scale up or down, according to a more or less extreme emission scenario than that used by the four global models.
17
Ministry for the Environment 2001, p. 14 (see next footnote).
18
New Zealand Climate Change Programme 2001b. Available from http://www.climatechange.govt.nz , or
http://www.mfe.govt.nz/new/ImpactsReport.pdf (full version) and http://www.mfe.govt.nz/new/ImpactsReport-ExecutiveSummary.pdf
(Executive Summary). A published version is also available from the Ministry for the Environment.
16
66
The greenhouse effect and climate change
Table 6.5:
Climate
element
Current climate change projections for New Zealand; prevailing winds,
temperature, rainfall, sea-level, and heating energy demand
Projections for 2030 and 2070 to 2099
average of four “transient” models with interaction with deep ocean currents
Scenario 1:
Scenario 2:
no action taken to reduce emissions
emissions reduced
Prevailing
winds
Temperature
Rainfall/
precipitation
Parliamentary Library, August 2001
increased strength of mid-latitude westerly winds
For the year 2030
Approximately half that for the year 2070.
For the years 2070 to 2099, compared to 1970-1999
Northland & Auckland
1.0°C to +2.8°C
Western North Island, Waikato to Wellington
+0.8°C to +2.7°C
Eastern North Island, Bay of Plenty to Wairarapa
+0.9°C to +2.7°C
Nelson, Marlborough, to coastal Canterbury & Otago
+0.8°C to +2.5°C
West Coast and Canterbury foothills
+0.6°C to +2.5°C
Southland and inland Otago
-0.6°C to +2.2°C
For the year 2030
Approximately half that for the year 2070
For the years 2070 to 2099, compared to 1970-1999
Northland & Auckland
-10% to 0%
Western North Island, Waikato to Wellington
0% to +20%
Eastern North Island, Bay of Plenty to Wairarapa -20% to 0%
Nelson, Marlborough, to coastal Canterbury & Otago
-20% to +5%
West Coast and Canterbury foothills
+5% to +25%
Southland and inland Otago
0% to +30%
about two-thirds the
impact
about two-thirds the
impact
+ increased risk of extreme events (drought and flood)
Sea-level
Reduction in
energy
demand for
heating
2050
13 cm (range 3-25 cm)
2050
2100
34 cm (range 9-66 cm)
2100
Auckland
Wellington
Christchurch
Invercargill
2030
- 12-70%
- 25-33%
- 4-14%
- 12-19%
2070
- 69-79%
- 29-86%
- 9-62%
- 15-51%
overheating in some
areas would also
add to increased
air-conditioning
12 cm
(range 3-24 cm)
25 cm
(range 5-49 cm)
about two-thirds the
impact
Source: Basher and Pittock 1998, section 4.2.3; A. B. Mullan, NIWA, pers. comm.; Ministry for the Environment 2001b,
pp. 11, 16, 26. Scenario 1 assumes a target atmospheric concentration of CO2 of 700 ppm, and Scenario 2 assumes 500 ppm.
Further discussion of scenarios in sections 6.3 and 8.1
67
The greenhouse effect and climate change
6.7
Parliamentary Library, August 2001
The El Niño-Southern Oscillation (ENSO) phenomenon19
The IPCC projection for the Australasian region is for increased El Niño-like conditions and/or
exacerbation of El Niño conditions (greater rates of drying, risk of drought and heavy rainfall
leading to flooding) when they do occur. There is still scientific debate on whether climate
change would lead to long-term changes in El Niño frequency or intensity.
The El Niño-Southern Oscillation (ENSO) is a natural cyclical phenomenon in the Southern
Hemisphere which markedly affects global climate. The names La Niña and El Niño are
colloquial terms for its alternating mild and harsh effects seen along the west coast of South
America. Scientific understanding of the ENSO has developed over the last 30 years, and its
effects can be traced back through historical records over hundreds of years. The ENSO
phenomenon is now monitored by the difference between air pressure in Tahiti and Darwin,
termed the Southern Oscillation Index.
El Niños tend to occur every three to seven years, and last for about a year each time, although
there was a long-running El Niño over 1991-1995. There are also changes over multiple
decades: for example, the El Niño signal in global climate anomalies was weak between the two
World Wars, but has been strong since 1950, and there has been a higher frequency of El Niños
over the last two decades. This is now believed to be caused by another naturally recurring
climate pattern in the Pacific Ocean, known as the Interdecadal Pacific Oscillation, which has
alternating cold and warm periods in cycles of 20 to 30 years.
During an El Niño, the westerly trade winds in the Pacific Ocean weaken, leading to a rise in
sea-surface temperature, the reduction in nutrient-rich seawater upwelling off South America
and subsequent loss of fisheries, heavy rainfall and flooding in Peru, drought over Australia and
Indonesia, and more cyclones in areas such as the Cook Islands and French Polynesia.
The 1997-98 El Niño contributed to drought and uncontrolled forest fires in Indonesia,
Venezuela, French Guyana, Brazil, and New South Wales; severe drought and food shortages
in Papua New Guinea, and severe flooding in Ecuador, Peru and Chile. Effects were also felt in
the southern and western USA, Canada, and eastern Africa.
In New Zealand during an El Niño, stronger and more frequent winds come from the west in the
summer, resulting in more rain in western areas and more drought in the east coast. In winter,
winds are mostly from the south, resulting in colder conditions; and in the spring and autumn,
southwesterlies increase, resulting in a mix of summer and winter conditions.
Although the ENSO is estimated to account for only 25% of annual variance in rainfall and
temperature, and east coast droughts can also happen at other times, the probability of climate
variations is strong enough to warrant planning for them when an El Niño is predicted or in
progress.
19
El Nino and Forecasting Seasonal Climate and Global Climate Models on http://katipo.niwa.cri.nz/ClimateFuture ; Basher 1998
The 1997/98 El Nino Event: Impacts, Responses and Outlook for New Zealand on
http://www.morst.govt.nz/publications/elnino/index.htm
68
The greenhouse effect and climate change
Figure 6.3:
Parliamentary Library, August 2001
Projected changes to temperature and rainfall for New Zealand, 1980s to 2080s.
Averages from four AOGCM (Atmosphere-Ocean Global Climate Model) results for predicted mean temperature
increase (°C, top maps) and rainfall change (% - bottom maps) for New Zealand over the next 100 years (1980s to
2080s). On the left is summer (December, January, February) and on the right is winter (June, July, August).
 National Institute of Water and Atmospheric Research (NIWA)
http://katipo.niwa.cri.nz/ClimateFuture/Scenarios.htm
(“1980s” = the period 1970 -1999, “2080s” = the period 2070 - 2099)
69
7
Impacts, adaptation and vulnerability
7.1
Summary of IPCC Third Assessment Report of Working Group Two
The IPCC’s January 2001 report on the scientific basis for climate change was followed in
February by a report on implications for impacts, adaptation, and vulnerability of human
communities and natural ecosystems. The conclusions are summarised below.
•
Recent regional climate changes, particularly temperature increases, have already
affected many physical and biological systems.
Observed changes include shrinking of glaciers, thawing of permafrost, and earlier ice break-up on
water bodies; longer growing seasons; poleward and altitudinal shifts of plant and animal ranges and
decline in some plant and animal populations; and earlier flowering of trees, emergence of insects,
and egg-laying by birds.
•
There are preliminary indications that some human systems have been affected by
recent increases in floods and droughts.
However, as these systems are also affected by socioeconomic factors such as demographic shifts
and land-use changes, the effects of climate change are difficult to quantify.
•
Natural systems are vulnerable to climate change, and some will be irreversibly
damaged.
Vulnerable natural systems with limited adaptive capacity include coral reefs and atolls, mangroves,
boreal and tropical forests, polar and alpine ecosystems, prairie wetlands and remnant native
grasslands. Climate change is expected to increase risks for species already in danger of extinction.
•
Many human systems are sensitive to climate change, and some are vulnerable.
Sensitive human systems include water resources, agriculture, forestry, fisheries, human settlements,
energy supply, industry, insurance and financial services, and human health. A third of the world’s
population live in water-stressed areas, many of which are at risk of reduced rainfall with the
projected climate change.1
•
Projected changes in climate extremes could have major consequences.
Extreme events such as floods, heat waves, avalanches and windstorms are projected to increase in
frequency, while extreme low temperature events are expected to decrease.
•
The potential for large-scale and possibly irreversible impacts poses risks that have
yet to be reliably quantified.
Events of possibly low probability but potentially large consequences which are not yet adequately
understood include slowing of the North Atlantic ocean circulation systems (reduction in warming of
parts of Europe), major reduction of Greenland and West Antarctic ice sheets (greater sea-level rise),
and increased release of greenhouse gases from permafrost and coastal sediments (accelerated
warming).
•
Adaptation is a necessary strategy to complement climate change mitigation efforts.
Impediments to achieving the full measure of adaptation include decisions based on short-term
considerations, continued development of risk-prone areas, insufficient information, and over-reliance
on insurance mechanisms.
•
Those with the fewest resources have the least capacity to adapt and are the most
vulnerable.
The impacts are expected to fall disproportionately on the poorest people. Human systems can best
deal with the consequences of climate change if they have the necessary wealth, technology,
education, information, skills, infrastructure, and resources. Present disparities in well-being are
expected to increase with disproportionate impacts of climate change. Global aggregate estimates of
the costs and benefits of climate change misrepresent this effect by treating gains for some as
canceling out losses for others.
1
Approximately 1.7 billion people live in “water stressed” areas (indicator: where more than 20% of the renewable water supply is
used). With population growth this is projected to increase to 5 billion people by 2025 (IPCC 2001b, p.7).
The greenhouse effect and climate change
•
Parliamentary Library, August 2001
Adaptation, sustainable development, and enhancement of equity can be mutually
reinforcing.
Population growth, resource depletion, and poverty need to be addressed together with the likely
impacts from climate change.
The projected impacts of climate change will have positive and negative effects in different parts
of the world (Table 7.1).
However, summaries of such impacts can be misleading if they average out positives and
negatives on a global scale without acknowledging the extreme impacts that some local
communities could experience.
Table 7.1:
Summary of some projected negative and positive impacts of climate change.
area of
impact
water
supply
projected impacts
negative
positive
less in many water-scarce regions
more in some areas
(particularly subtropics), monsoon
(e.g. Southeast Asia)
areas greater extremes of dry & wet
and some floodplain aquifers recharged
food
supply
less in most mid-latitude, tropical and
subtropical regions (more disease,
heat shock, drought, flood)
more in some mid-latitude regions
(longer growing season)
human
health
greater risk of vector- and waterborne diseases, heat-stress mortality,
increased risk from flooding (rainfall
and sea-level rise), storm events
less winter mortality
(mid- and high- latitudes)
energy
more demand for cooling, reduced
reliability of energy supply, less
hydropower in drought areas
less demand for heating
forestry
increased risk of forest fire
more timber (if appropriately managed)
Source: IPCC 2001b, pp. 4, 16.
More detail of the IPCC conclusions and their confidence levels are provided in Table 7.2 for the
global scale. Keeping in mind the many remaining uncertainties, some conclusions about
possible and likely outcomes in various regions have also been made by the IPCC. Regional
projections for Australia, New Zealand and small island states (in the Pacific and elsewhere) are
summarised in Table 7.3, with the estimated level of confidence if reported.
Small island states are among the countries most vulnerable to climate change. In the Pacific
region, the countries of lowest elevation are expected to suffer “profound” impacts as the sea
level rises, including the Marshall Islands, Tuvalu and Kiribati. “Severe impacts” resulting in
major population displacement are projected for Micronesia, Nauru, and Tonga. “Moderate to
severe” impacts are projected for Fiji and the Solomon Islands, and “local severe to catastrophic”
effects are projected for Vanuatu and Western Samoa.2
Projections of key impacts for New Zealand are summarised in section 7.2.
2
IPCC, The Regional Impacts of Climate Change, on http://www.grida.no/climate/ipcc/regional/258.htm
72
The greenhouse effect and climate change
Table 7.2:
Parliamentary Library, August 2001
Global scale projections of the IPCC relating to climate change impacts and vulnerability.
IPCC projections
Ecosystems
Human
health
Water
resources
Plant
growth
and food
supply
Climate change will cause significant disruption to ecosystems.
Habitat for cold/cool water fishes will decrease, and habitat for warm water fishes will increase.
Future sea-surface warming will increase stress on coral reefs and increase the frequency of
marine diseases.
The ability of at-risk species to adapt to changing habitat boundaries will be limited by
human destruction and fragmentation of habitats. Some “critically endangered” species
will become extinct, and most of those “endangered or vulnerable” will become rarer and
closer to extinction.
The response of species and ecosystems to climate change will lag behind by a few to many
years.
The geographic range of potential transmission of malaria and dengue-two vector-borne
infections will increase. This currently affects 40-60% of the world population.
The impact of increased heat waves and heat-related death and illness, often exacerbated
by higher humidity and urban air pollution, will be greatest for urban populations, the
elderly, the sick, and those without access to air-conditioning.
Increases in flooding will increase the risk of drowning, diarrhoeal and respiratory
diseases; and in developing countries, hunger and malnutrition.
In developed temperate zone countries, net temperature-related deaths will decrease (more
summer deaths but fewer winter deaths) (Little published research on other countries).
confidence
level
67-95%
67-95%
67-95%
67-95%
67-95%
33-95%
67-95%
67-95%
33-67%
Annual mean streamflow will increase in high latitudes and in SE Asia, and will decrease in central
Asia, the Mediterranean, southern Africa, and Australia.
33-67%
An increased concentration of CO2 can stimulate crop growth and yield, but that may not
always overcome adverse effects from heat and drought. Field trials show smaller gains
than pot trials.
Increased heat stress to livestock in some areas.
In mid-latitude areas, temperature increases of less than a few °C will result in generally positive
crop effects, and increases over a few °C will result in generally negative crop effects. In tropical
areas, crop yields would generally decrease with even a few °C increase (some crops are already
33-67%
67-95%
6-67%
6-33%
near their maximum temperature tolerance, and dryland/rainfed agriculture predominates).
Climate change impacts will cause small percentage changes in global agricultural income, with
increases in developed regions and increases or declines in developing regions.
Climate change will worsen food security in Africa, mainly through increased extremes and
temporal/spatial shifts of climate.
Human
settlements
Economic
impacts
Insurance
& finance
established
(current
trends)
Communities on the coast and near rivers, and in urban areas with inadequate storm
drains, water supply and waste management, are at highest risk from increases in flooding.
Communities with little economic diversification and heavy reliance on climate-sensitive
primary production are more vulnerable to climate change.
Developed areas in the Arctic and where permafrost is ice-rich will need special works to mitigate
thawing damage to buildings and transport infrastructure.
Increased rainfall will increase floods, landslides, mudslides, and soil erosion, thus putting
pressure on insurance and disaster relief systems in some areas.
67-95%
Disproportionate impacts from climate change on the poor will increase income disparities.
Economic losses and GDP reductions will be greater for higher magnitudes of warming.
Global timber supply will increase, enhancing the rising market share in developing
countries.
For all warming magnitudes, there will be net economic losses for developing countries.
For increases up to a few °C, there will both economic gains and losses in developed countries.
Aggregated on a global scale, GDP would change ± a few percent for global mean temperature
increases up to a few °C.
Some tourist destinations will shift (losses and gains felt in different areas).
More people will be harmed than benefited by climate change, even for low levels of warming.
33-67%
33-67%
33-67%
Actuarial uncertainty in risk assessment will increase, placing upward pressure on
insurance premiums and/or lead to reclassification of some risks as uninsurable and
withdrawal of coverage.
67-95%
95% +
67-95%
5-33%
5-33%
5-33%
67-95%
5-33%
67-95%
Source: IPCC 2001b. Estimates of confidence used in the report were based on the collective judgement of the authors using observational
evidence, modeling results, and theory they have examined. The range was: very high (95% or greater), high (67-95%), medium (33-67%), low (533%) and very low (5% or less). Those termed medium to high in the original are noted here as 33-95%, and medium to low are noted as 5-67%.
73
The greenhouse effect and climate change
Table 7.3:
Parliamentary Library, August 2001
Regional IPCC summary of adaptive capacity, vulnerability, and key concerns for Australia,
New Zealand, and small island states (details in italics from IPCC background reports)
IPCC conclusions and projections
Adaptive
capacity
(human
systems)
Water
resources
Crops &
fisheries
Sea level
Storms
Ecosystems
Tourism
Australia and New Zealand
Generally high, except for indigenous groups in some regions with low adaptive capacity
and therefore high vulnerability.
Small island states
Generally low, and therefore high vulnerability. Likely to be among the countries most
seriously impacted by climate change.
Australia and New Zealand
Likely to be a key issue due to projected drying trends over much of the region and change
to a more El Niño-like average state.
Vulnerability is high with respect to hydrology. Of most concern are drought-prone areas,
flood-prone urban areas, low-lying islands, and alpine snowfields.
New Zealand’s glaciers are likely to shrink further.
Small island states
Islands which currently have limited water supplies will be highly vulnerable.
Australia and New Zealand
The net impact on temperate crops may initially beneficial, but the balance is expected to
become negative for some areas and crops with further climate change.
Small island states
Limited arable land and soil salinisation makes agriculture highly vulnerable to climate
change.
Impacts to coastal ecosystems will negatively impact reef fisheries and those who rely on
them.
Small island states
The next 100 years will see enhanced coastal erosion, loss of land and property,
dislocation of people, increased risk from storm surges, damage to coastal ecosystems,
saltwater intrusion into freshwater resources; costs of responding and adapting to these
changes will be high.
Australia and New Zealand
Intensity of heavy rains and tropical cyclones will increase, with consequent increased local
risk of flooding, storm surges, and wind damage.
Australia and New Zealand
Some species with restricted climatic niches and which are unable to migrate due to
fragmentation of the landscape, soil differences, or typography could become endangered
or extinct.
Vulnerable ecosystems include freshwater wetlands in coastal zones, areas vulnerable to
accelerated invasion of weeds, arid, semi-arid, and alpine systems, and coral reefs.
Knowledge of climate change impacts on aquatic and marine ecosystems is relatively
limited.
Aquatic systems will be affected by the disproportionately large responses in runoff, river
flow and associated nutrients, wastes and sediments that are likely from changes in rainfall.
Small island states
Coral reefs, mangroves, sea grass beds, and other coastal ecosystems will be negatively
affected, with implications for sustainability of the local economy.
Australia and New Zealand
Reduced snow amounts and a shorter snow season appear likely and would decrease the
amenity value of the mountains and the viability of the ski industry.
Small island states
Tourism (an important source of revenue and foreign exchange for many islands) would
face severe disruption from climate change, sea-level rise, and loss of coral reef resources.
confidence
level
(based on
present
situation)
67-95%
67-95%
33-67%
67-95%
33-67%
67-95%
33-67%
67-95%
33-67%
67-95%
Source for plain text summary and confidence level definitions as for Table 7.2 (pp. 2, 16-19 in IPCC 2001b). The predictions about small island
states includes those in the Pacific as well as all others around the world, and therefore of necessity are very generalised.
Additional key factors in italics from Basher and Pittock (eds) for Australia and New Zealand, and Nurse, McLean and Suarez (eds) for small
island states, in IPCC1998.
74
The greenhouse effect and climate change
7.2
Parliamentary Library, August 2001
Likely impacts in New Zealand
Recently, the Ministry for the Environment published Climate Change Impacts in New Zealand. 3
The following text summarises the key predicted impacts from this report and from the IPCC
background reports of 2001.
7.2.1 Agricultural production4
The impact of climate change on soil properties and plant growth are difficult to predict due to
limited understanding of the complex interactions involved, and uncertainties about the exact
nature of climate change at the local and regional scale.
Research indicates that increased CO2 concentration in the atmosphere will alter the carbonnitrogen ratios of biomass, soil nutrients, and soil carbon. Increased CO2 (“carbon fertilisation”)
can increase water-use efficiency in plants leading to higher productivity. However, this positive
response to CO2 enrichment:
•
is limited by available soil nutrients and moisture;
•
is stronger for legumes and weedy species than for most grasses;
•
does not have much effect on older trees;
•
seems to decrease over the long-term; and,
•
plants grown under elevated CO2 conditions have less protein.
In New Zealand, highest increases from carbon fertilisation are projected for cooler wetter areas,
such as the southern South Island, and areas that are already warm and dry are expected to
gain the least. By 2030, a 10-20% increase in annual pasture yield is projected for suitable sites.
Other changes relating to rainfall, temperature, and storm events are likely to include the
following.
•
•
•
•
•
•
•
•
•
•
•
•
Reduced or increased availability of moisture in soils, depending on region.
Increased risk of drought and flood, especially in areas already prone to such events.
Increased risk of soil erosion, of concern in New Zealand’s deforested hill country.
Increased risk of alkalisation and salinisation of soils where rainfall decreases and/or
evaporation increases (possibly offset in part by increased CO2 levels).
Increased rate of biochemical processes from elevated temperature and reduced frost
losses in temperate areas, but increased desiccation and sun scald in warmer areas.
Decreased winter stock losses in the New Zealand high country with warmer winters, but
also decreased suitability of areas for crops that require winter chilling.
Negative impacts for some crops and positive impacts for others, e.g. in eastern areas with
increases in temperature and dryness. 5
Increased need for irrigation in some areas, possibly competing with other sectors for water
supply.
Increased incidence in some areas of land degradation, weeds and pests, and diseases.
Increased risk of saltwater intrusion into groundwater aquifers in such areas as Hawke’s Bay
and parts of Canterbury.6
Reduced water availability for irrigation of pipfruit growing areas in Hawkes Bay.
Reduced suitability of the Bay of Plenty for growing kiwifruit due to loss of adequate winter
chilling, from 2050 if current greenhouse gas emission rates continue.7
3
New Zealand Climate Change Programme 2001b.
Basher and Pittock 1998, sections 4.3.1.2, 4.3.3.1; New Zealand Climate Change Programme 2001b, pp. 18-21.
5
For example during the 1997/98 El Niño, with increased dryness and warmth in eastern areas, unirrigated dryland farms grew 25%
less grain of poor malting quality due to its nitrogen content being too high, and apple growers reported more problems with sunburn,
overheating and lack of colour development in the fruit due to lack of cold nights. However, Hawkes Bay grape growers reported the
best vintage in 15 years, smaller fruit and higher sugar content, and less need for chemical sprays (Basher 1998, p. 11-12).
6
New Zealand Climate Change Programme 2001b, p. 16.
7
Competing kiwifruit growing areas in Italy and Chile may also become marginal with export market implications (New Zealand
Climate Change Programme 2001b, p. 20).
4
75
The greenhouse effect and climate change
•
•
•
•
Parliamentary Library, August 2001
Increased suitability of regions such as Marlborough, Canterbury and Central Otago for
growing other fruit crops such as apples and kiwifruit, subject to availability of irrigation.
Increased wheat productivity (10-15% by 2050) subject to adequate irrigation and nitrogen
fertilisation.
Further increase in subtropical “C4” pasture grasses, which are low quality and less
desirable for livestock diets, but may also provide feed during dry periods when traditional
forage plants die off.8
No increased pasture production from increased CO2 where the frequency of summer
droughts also increases (e.g. in low elevation New Zealand sheep farming).
Impacts will likely vary from district to district, from crop to crop, and from decade to decade. For
example, in the first few decades of global warming, grain crops may benefit from higher CO2,
but in later decades increased temperatures would reduce the grain-filling period.
In addition, impacts are likely (possibly both positive and/or negative in different parts of the
agricultural sector and in different decades) from significant changes to global food production
and commodity prices, and the ability of other countries to produce goods required locally or
purchase our exports.
Changes in the suitability of districts for particular crops and farming systems could have major
economic impacts unless anticipated and planned for early. The agricultural sector can
theoretically prepare for these risks by dedicated breeding programmes (e.g. drought resistant
forage species, high-quality subtropical grasses, pip and stone fruit that requires less winter
chilling, biological pest control), water management schemes, and wider use of current drought
resistant species. However, there is also a risk that adequate efforts may not be initiated until
traditional techniques increasingly fail.
7.2.2 Indigenous species and ecosystems
In the past indigenous species have adapted to natural climate changes, but the future rate of
change may exceed any that the present biota have previously experienced. Perhaps more
importantly, their ability to adapt will be restricted by considerable loss and fragmentation of
habitat from urban and agricultural development, and competition and predation by introduced
species. Slow-growing indigenous species (such as forest trees, high country tussock, and some
fish species) may also be less able to adapt.
The most vulnerable terrestrial ecosystems are considered to be fragmented native forests of
drier lowland environments in Northland, Waikato, Manawatu and in the east from East Cape to
Southland. Studies on individual species is limited. A study of the likely climate change impact
on kauri showed a decline in suitable habitat of 25% by 2050 and 65% by 2100 if the current
greenhouse gas emissions rate continues.9
During the droughts of the 1997-98 El Niño, exceptional die-back of native vegetation was
reported in eastern areas of both the North and South Islands. However, the warmer summer
temperature was also expected to result in heavy masting (flowering, fruiting and seeding) in
wetter areas, to the benefit of breeding of native birds (as well as increases in the predator
population) the following spring and summer. With climate change the natural fluctuation
between rodents, predator numbers, and native bird species may be intensified. 10
8
The occurrence of the subtropical grass Paspalum dilatatum has already spread southward by 1.5º latitude (spreading from midWaikato/East Cape to Wanganui/Cape Kidnappers) during a period of increasing temperatures 1976-1988. (New Zealand Climate
Change Programme 2001b, p. 18).
9
New Zealand Climate Change Programme 2001b, pp. 24-25.
10
Basher 1998, p. 15. After prey species populations increase, predator numbers naturally increase after a lag period. The decline
in predator numbers also lags after the decline in prey, and as the predators run out of rodents to prey on they are likely to increase
their predation on native birds.
76
The greenhouse effect and climate change
Parliamentary Library, August 2001
7.2.3 Freshwater and marine ecosystems
Knowledge of the response of freshwater and marine ecosystems to climate change is not exact
enough to allow confident prediction of effects. It is thought that increased flooding will affect
water quality, rising sea-level will affect estuaries, changes in temperature will affect spawning,
and changes in ocean currents will affect marine nutrient upwelling, food networks, reproductive
patterns and species ranges. Whether on balance these are positive or negative is not yet
known. Warmer seawater may encourage the spread of toxic algal blooms, but the influence of
other factors is not fully understood.
7.2.4 Health
Warmer weather is anticipated to create both positive and negative health effects. These include
the following.
•
Reduced incidence of cold-related illness and death in winter.
•
Increased summer mortality (a Christchurch study found 1.3% increase for each 1ºC).
•
Reduced use of open fires for heating, resulting in less winter air pollution.
•
Increased summer smog, particularly in Auckland.
•
Establishment risk for mosquito populations capable of transmitting infections such as Ross
River virus and dengue fever, but not malaria (by 2100, possibly including Northland,
Auckland, Waikato, Bay of Plenty, Gisborne, Hawkes Bay and coastal Manawatu).11
Increased variability of rainfall (droughts and floods) may also contribute to health effects.
•
Spread of diseases transmitted between animals and humans, such as cryptosporidiosis,
from heavy rainfall events washing animal wastes into water supplies.
•
Poorer water quality during drought periods in areas where water supplies struggle to meet
demand.
In addition, the impact of increasing levels of greenhouse gases on the recovery of the ozone
layer (section 4.5) is likely to mean 15 to 20 years more of elevated skin cancer risk due to
exposure to high levels of ultraviolet light.
As with current health problems, these risks are likely to be borne disproportionately by people
of lower socio-economic status with limited resources to prevent illness and seek treatment.
7.2.5 Impacts on Mäori communities12
The IPCC has concluded that communities with limited resources will have limited ability to
make the necessary adaptations to the impacts from climate change. Low income families and
communities in New Zealand, which are at the present time disproportionately Mäori, may
therefore suffer greater impacts.
The reliance of Mäori on the environment as both a spiritual and an economic resource also
makes them more vulnerable and less adaptable to climate change. Mäori-owned land in some
areas is of lower-than-average quality and may be more prone to erosion and invasion by
subtropical grasses.13 Land ownership structures and spiritual and cultural links to the land are
likely to make it harder for Mäori to consider relocating or making major changes to land use.
Climate change impacts on indigenous species of traditional and cultural importance to Mäori
would in turn impact on Mäori people and communities.
11
New Zealand Climate Change Programme 2001b, pp. 28-30.
Sources for this section include Ministry for the Environment (unpublished materials), New Zealand Climate Change Programme
2001b, p. 31 (which was compiled with the help of Te Puni Kökiri and others).
13
Particularly drier and less productive areas such as Northland and the East Cape (New Zealand Climate Change Programme
2001b, p. 31).
12
77
The greenhouse effect and climate change
Parliamentary Library, August 2001
Mäori see the world as a unified whole, where all elements including tängata whenua are
genealogically connected. Climate change affects the balance between living things (utu), and is
seen as a depletion of the Earth’s life force (mauri). Restoring the natural balance from a Mäori
perspective is not only a question of physical and economic benefit, but also of spiritual value.
The Ministry for the Environment has involved Mäori in climate change issues discussions since
1990, and in 2000 completed a round of 10 nationwide hui, the results of which will feed into
consultations planned for 2001-2002.14
7.2.6 Hydro-electricity generation
Reduced snowfall would reduce the risk of spring floods, and smooth seasonal differences in
hydro-electricity generation. However, with more frequent extreme rainfalls also predicted, dams
will need to be managed conservatively to avoid overtopping during floods and running out of
water during droughts. In addition, increased sediment load would accelerate the reduction of
hydro-dam storage capacity.15
With higher temperatures, there would be less demand for electricity for winter warming,
possibly about 6% with continuation of current greenhouse gas emission trends. However, there
could also be higher demand for electricity in summer for air-conditioning.16
7.2.7 Tourism
Global warming is likely to lead to reduced snowfall, higher snowlines, earlier spring snowmelt, a
shorter snow season in Australia and New Zealand, and further retreat of the glaciers in New
Zealand. It has been estimated that even with a 10% increase in precipitation, a 2°C warming
from global climate change would cause a reduction of 20% in snow cover on the Southern
Alps.17
The climate change impacts in the Australian skifield areas are predicted to occur earlier than in
New Zealand, so there may be a short-term gain for New Zealand if skiers visit here instead.
However, in the long term, there is likely to be a loss of amenity value of mountain landscapes
for locals and tourists threatening the viability of the ski industry, which has limited options for
relocation.
If skifield operators resort to artificial snowmaking, this may have implications for the use of local
water and its availability for other purposes.
7.2.8 International links
Pacific Island countries are particularly vulnerable to aspects of climate change such as sealevel rise and increased risk of tropical storm events. Measures under the UNFCCC and Kyoto
Protocol are aimed at limiting the rise in greenhouse gas emissions and at providing assistance
to developing countries, including small island states, in order to address the adverse effects of
climate change. If significant disruption occurs for Pacific Island and other communities, there
could be an increased need for development aid, disaster relief, and numbers of refugees forced
to permanently leave their homes. As a nearby country with better resources to cope with the
impacts of climate change, New Zealand may be subject to increased aid and immigration
pressure.
14
http://www.mfe.govt.nz/new/athague.htm
IPCC 1996, WG II, Section 14.3.3, cited in Basher and Pittock 1998, section. 4.3.2.3. It is also worth noting that during the 1992 El
Niño there was an electricity crisis, but during the 1997-98 El Niño the increased westerly rains spilled over to the eastern side of the
Divide, and the Tekapo, Pukaki, Hawea, Te Anau, and Manapouri hydro-electric storage dams were full to spilling (Basher 1998, p.
14).
16
New Zealand Climate Change Programme 2001b, p. 27.
17
IPCC 1996, WG II, Section 7.4.1, cited in Basher and Pittock 1998, section. 4.3.2.3
15
78
Part D: Options for action
8
Overview: targets and principles
8.1
What is the target?
The ultimate UNFCCC objective is “stabilisation of greenhouse gas concentrations in the
atmosphere at a level that would prevent dangerous anthropogenic interference in the climate
system.” In the absence of agreement on what that level might be, an arbitrary target of a return
to 1990 greenhouse gas emission levels by the year 2000 for the Annex I countries was agreed
to. It was not met.
The current Kyoto Protocol target is for at least 5% reduction on 1990 greenhouse gas emission
levels over the first commitment period of 2008-2012, for most of the Annex I countries. This is
also an arbitrary target, based on political reality rather than scientific data on the reduction
target that would best prevent dangerous anthropogenic interference in the climate system.
Two scenarios used by organisations such as the Hadley Centre for Climate Prediction (UK) and
the New Zealand Climate Change Programme compare a projection of current emission trends
without serious attempts to reduce emissions, and a future where the world explicitly aims to
reduce emissions much more than required by the Kyoto Protocol. Even this latter scenario
would result in some significant climate change (Table 8.1).
Unlike the Kyoto Protocol target, these scenarios are based on greenhouse gas concentration in
the atmosphere rather than estimated emissions. They do not include all of the greenhouse
gases, only CO2 (which has the best emissions database globally, for UNFCCC Annex I
countries as well as the rest of the world).
A 550 parts per million (ppmv) atmospheric CO2 concentration target requires about a 30% cut in
the current rate of emissions, and for some countries a cut of up to 70%. In contrast, the Kyoto
Protocol requires a cut in projected emissions in 2012 of about 10% overall and up to 30% for
some industrialised countries.1 Emission rates are projected to continue to rise and emissions in
2012 are expected to be higher than at present. The Kyoto Protocol, if ratified and actioned with
its present targets, will result in levels higher than 500-550 ppm.
For the “Contraction and Convergence” scenario (see section 1.8), the Global Commons
Institute proposes that to ensure the survival of current ecosystems and avoid accelerating
feedbacks in the climate system, a target of ideally 350 ppm but no more than 450 ppm is
required. The Institute has suggested a target year of 2050 or 2100.2
While agreement may be reached that “business as usual” is very likely to lead to serious
climate impacts over the long term, and therefore adjustments need to be made, there remains
no agreed “right answer” for greenhouse gas reduction targets. The scientists can provide best
estimates and projections, but the risk management decision will be political.
1
2
UNEP 1998; New Zealand Climate Change Programme 2001b, p. 8..
http://www.gci.org , The Detailed Ideas and Algorithms Behind Contraction and Convergence.
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 8.1: Atmospheric CO2 stabilisation scenarios
ppmv CO2
~280
368
450
550
about twice
preindustrial
levels of
CO2
scenario description and climate change projections
pre-industrial level
2000 level
stabilising below 1990 emission levels within a few decades
Assumes the world explicitly aims to reduce greenhouse gas emissions.
Current emission rates need to fall by 30% to stabilise at this level over the long term.
In 2100, global emissions would need to be lower than 1990 and decrease even further
into the future despite increased world population and energy demand.
Global projections
Temperature rise begins to slow and stabilise around 2100, 1.4 to 2.6°C higher.
The rise of 2°C predicted for 2050 with unmitigated emissions would be delayed by 100
years. Temperature by 2230s: 2°C higher than today.
The sea level rise of 40 cm predicted for the 2080s with unmitigated emissions would be
delayed by about 40 years. Sea level by 2100 to rise by 13 to 70 cm.
Dieback of tropical forests and loss of carbon sinks substantially reduced by the 2230s.
Number of people suffering from drought induced by climate change in the 2080s: about
one billion.
New Zealand projections
Temperature and rainfall changes about 2/3 of those described below. Sea level rises
about 12 cm by 2050, and 25 cm by 2100.
650
700 to 750
about three
times preindustrial
levels of
CO2
stabilising below 1990 emission levels within about 100 years
Assumes no explicit attempt to control emissions.
The future development scenario includes rapid economic growth, increased
technological exchange and capacity building, substantial reductions in regional per
capita income differences, and a mix of fossil fuels and alternative energy sources.
Global projections: based on 750 ppm
Temperature rise begins to slow and stabilise around 2200, 2.1 to 3.82°C higher.. The
rise of 2°C predicted by 2050 with unmitigated emissions would be delayed by 50 years.
Temperature by 2230s: 3°C higher than today.
The sea level rise of 40 cm predicted for the 2080s with unmitigated emissions would be
delayed by about 25 years.
Dieback of tropical forests and loss of carbon sinks delayed by 100 years but significant
losses still occur.
Number of people suffering from drought induced by climate change in the 2080s: about
three billion.
New Zealand projections, based on about 700 ppmv
By 2100 temperature increases 0.6 to 2.8°C depending on the region. Rainfall decreases
up to 20% in eastern regions and Nelson/Marlborough, and increases up to 20-30% in
western regions, Southland, and inland Otago,
Sea level rises about 13 cm by 2050 and 34 cm by 2100.
1000
stabilising below 1990 emission levels within a few centuries
ppmv = parts per million by volume: concentration in the atmosphere (also expressed as ppm)
Global projections: Hadley Centre for Climate Prediction 1999
New Zealand projections: New Zealand Climate Change Programme 2001b, pp. 7-8, 11, 14.
2000 level: http://cdiac.esd.ornl-gov/pns/current_ghg.html
450, 650 and 1000 ppmv levels: IPCC 2001a, p.12.
80
The greenhouse effect and climate change
8.2
Parliamentary Library, August 2001
Primary focus on energy
Carbon dioxide makes up the largest part of greenhouse gas emissions worldwide. The majority
of this comes from the energy sector, largely through the consumption of fossil fuels for
transport, electricity, and industrial processes (Table 8.2).
Table 8.2: Share of greenhouse gas emissions from the energy sector, Annex I countries, 1998.
A
Greenhouse gas
emissions
B Contribution from energy
sector
C Share of total greenhouse
gas emissions from the
energy sector (AxB)
CO2
82%
CH4
12%
N2O
4%
Others
2%
96%
35%
26%
data not
available
79%
4%
1%
Totals
100%
84%
Source: UNFCCC 1998, Second compilation and synthesis of second national communications, reported in International Energy
Agency 2000, p. 14.
Therefore, in order to reduce global greenhouse gas emissions, initiatives are needed to
effectively target the energy sector. Global energy use has increased nearly 70% over the last
three decades and is predicted to rise more than 2% per year over the next 15 years. If
unchecked, this will raise greenhouse gas emissions about 50% above current levels.3
While New Zealand’s greenhouse gas emissions are mostly in the form of CH4 and N2O rather
than CO2, the technology for reducing of CH4 and N2O emissions without reducing agricultural
production is not very well advanced. Therefore, in the foreseeable future New Zealand’s
greenhouse gas reduction efforts will still need to focus on CO2 emissions, primarily from fossil
fuel energy, and creation of CO2 sinks, primarily through new forest plantings.
8.3
“Good practice” policies
In order to work, policies must be accepted by communities and be seen to be fair and effective.
The International Energy Agency has suggested that “good practice” policies will meet the
following criteria:
ƒ
ƒ
ƒ
ƒ
8.4
maximise both economic efficiency and environmental protection (both in terms of climate
change and other environmental issues);
be politically feasible;
minimise red tape and overheads; and,
have positive feedback effects in such areas as competition, trade and social welfare (or at
least not conflict with policies in those areas).4
The global commons
Like most environmental problems, anthropogenic climate change has emerged from a
syndrome called “the tragedy of the commons”.
•
The global atmosphere belongs to everyone, and thus no one and no country has been
responsible for protecting it from abuse.
•
Each individual or industry freely uses the common resource, focusing only on their own
needs and the immediate effects which can be attributed to their own actions, although the
3
4
A. Reisinger, Ministry for the Environment, unpublished papers 2001.
IEA 2000, p.24.
81
The greenhouse effect and climate change
Parliamentary Library, August 2001
effect on the global atmosphere, which will affect everyone to some extent, is the result of
cumulative actions of everyone over the long term.
The atmosphere is considered to be “free”, and so “rational” economic decisions tend to
ignore climate change implications.
•
Effective solutions, both global and local, must instil individual and corporate incentives and
responsibility for the ensuring health of the global commons.
8.5
Leadership and assistance from developed countries
The majority of greenhouse gas emissions over the last 150 years have been from the
developed countries, and they continue to the primary source (see Figures 1.1 to 1.3, and 4.6).
These countries are also in a better position to afford changes in technology to reduce
dependence on fossil fuels. Even with continuing growth, per capita energy use and emissions
in developing countries are still likely to be much lower than in developed countries over the next
30 years.1
The latest Kyoto Protocol agreement in Bonn continues to reflect the understanding that
developed countries need to take the lead and financially assist the developed countries to both
adopt greenhouse-friendly technology and cope with climate changes (see section 1.6).
8.6
Embeddedness of issues and the “no regrets” approach
Many of the changes that could significantly reduce the emission of greenhouse gases have
other benefits. For example: more efficient energy systems; cleaner vehicles; reduced air
pollution and pollution related illness; better public transport; the development of renewable
technologies; more comfortable homes; reduced reliance on imported and non-renewable fuels;
improved efficiency in production, healthier soil and sustainable agriculture; protection of
biodiversity; new employment opportunities; and sustainable economies.5
Climate change is also embedded in issues of development. The people most vulnerable to
climate change are those who, for socio-economic and geo-political reasons, are reliant on
vulnerable ecosystems and do not have the ability to find or pay for alternatives (Chapter 7).
Empowering people to develop sustainable economies, through such means as assisting with
local industry development, transferring renewable energy technology, assisting with population
control, and addressing desertification, deforestation and overgrazing, can improve their quality
of life, avoid development along a path that contributes high levels of greenhouse gas
emissions, and help them more easily cope with the climate change effects that do occur.
When the social and economic impact of policies to reduce greenhouse gas emissions are
analysed, it is important to account for all of the costs and benefits, both direct and indirect, both
new and avoided.
A “no regrets” policy is one that has other benefits besides reducing the risk of climate change
and is therefore worth doing anyway, even if predictions of climate change prove to be
inaccurate. The classic example is energy efficiency, which can save money over the medium to
long term as well as reduce greenhouse gas emissions, and has been a key focus area of the
New Zealand policy response.
5
OECD 2000; IPCC 2001d, pp. 19-21; P. Bunyard 2001, Where now for the world’s climate?, The Ecologist 31(1):54.
82
9 National and international initiatives
9.1
Addressing “market failure”
“Market failure” is a syndrome common to most environmental issues. In the economic
decisions that people and industry make, use of the global atmosphere is “free”. Its good health
or otherwise is not valued in economic terms, and thus “rational” economic decisions tend to
ignore climate change implications. This market distortion is compounded by such things as
subsidies for fossil fuels and barriers such as lack of information on energy-efficient or
alternative sources of energy.
“Economic instruments”1 are considered by their proponents be the most efficient and low cost
way to address climate change. Rather than force people to take a particular action to reduce
emissions, economic instruments are designed to create a financial incentive and allow people
to choose the technology and method that best suits their circumstances.
Studies by the OECD (including the International Energy Agency) support the removal of fossil
fuel subsidies, targeted tax changes, and emissions trading as ways to encourage markets to
reduce CO2 emissions at lowest cost. Theoretical analysis suggests that emissions trading will
also be a sound method for cost-effective reduction of greenhouse emissions.2
If the Kyoto Protocol is successfully ratified and the first round of Annex I country commitments
come into force in 2008, a market value will be placed on greenhouse gas emissions. Countries
will need to keep emissions within their assigned amount, and will tend to choose the least
expensive method. This may be buying carbon sink credits, reducing emissions at source, or
paying for emission reductions in other countries where it is cheaper to do so. Some countries,
particularly in Eastern Europe where economies have contracted since 1990, will have surplus
assigned amount to sell.
Proposals to increase the value of the global atmosphere in the market include the following:
Tax policies: making emissions cost more
•
•
•
Emission taxes (at point of emission)
Product charges or taxes (e.g. a carbon charge)
Tax credits or reductions for climate-friendly alternatives
Trading: creating a market for reduced emissions
•
•
•
Trading in emission quotas and emission reduction credits
Creating markets for “green energy”
Clean Development Mechanism and Joint Implementation
Other fiscal measures
•
•
Removal of fossil fuel subsidies
Financial support for improving energy efficiency and developing renewable energy sources.
Information on these measures follows, after a discussion of market barriers.
1
I.e. techniques which try to harness the power of the market to make the desired changes. For environmental issues, this can
include changing the way the market values key resources and creating new markets for the sustainable use of resources.
2
IEA 2000b, pp. 26, 33.
The greenhouse effect and climate change
9.2
Parliamentary Library, August 2001
Removing barriers to energy efficiency and renewable energy
Reduction in energy use through energy efficiency and creation of energy from renewable
sources can replace demand for fossil fuel energy and thus reduce greenhouse gas emissions.
From an economic point of view, they will be worthwhile if their cost is less than the cost of the
thermal energy they replace. From the wider point of view, there are many ancillary benefits
(section 8.6).
In addition to the generic pricing issue discussed in section 9.1, other barriers to the optimal
uptake of energy efficiency and renewable energy have been identified. 3 The main types of
barriers and some of the measures which have been designed to help overcome them are
summarised below.
No value placed on the public benefits
Many of the environmental, health, and future generation benefits of renewable energy and
energy conservation are not reflected in market prices, thus eliminating much of the incentive for
consumers to switch to these technologies. Creating new price incentives, regulating renewables
market share, and regulating standards of product energy efficiency are some of the ways that
have been used elsewhere to address this barrier.
Lack of information
Many individual consumers are not aware of the potential energy savings or the environmental
implications of their choices; which of a range of products or behaviours are less polluting, or
where to get the best value for money. Labelling schemes and education programmes for
example are designed to help overcome this barrier.
Limited access to capital and rapid payback requirements
Energy efficiency and renewable energy measures often require up-front capital investment in
order to achieve long-term savings. Domestic and small business consumers in particular often
have limited capital and prefer investments with shorter payback periods. Targeted grants and
loans have been used to help overcome this barrier.
Emerging technology
Like all emerging technologies, renewables compete at a disadvantage against established
industries. Disadvantages include lack of economies of scale and lack of infrastructure.
Customers may find new products relatively difficult to obtain, or lack of familiarity may cause
customers to view them as more risky purchases. Assisted capacity building for the industry,
temporary financial support, publicity on products, and supplier information are some methods
that address this barrier.
Lack of responsibility/ spilt incentives
Consumers who rent homes or offices often have no incentive to make structural improvements
to improve energy efficiency and neither do landlords, as they do not usually pay the power bills
(termed the “landlord-tenant dilemma”). Housing developers have little incentive to provide
hidden benefits in a house such as insulation if a higher selling price does not result. Mandatory
energy efficiency building codes and financial assistance via grants or loans targeted at energy
efficiency retrofits for existing buildings have been used to help address this barrier.
3
Based on Parliamentary Commissioner for the Environment 2000, pp. 48-49; IPCC 2001d, pp. 33-37; Union of Concerned
Scientists 2000 http://www.ucsusa.org/energy/brief.barriers.html .
84
The greenhouse effect and climate change
Parliamentary Library, August 2001
Inappropriate price signals
Electricity and gas tariffs often do not reflect the full marginal cost of production, particularly
environmental costs not valued in the market such as climate change. Consumer prices do not
fully reflect variation in costs between peak and off-peak periods, nor do they rise over time
toward the marginal cost for constructing a new power station to reflect diminishing surplus
capacity. While electricity reforms in New Zealand have made the market more competitive,
companies continue to have a financial incentive to maximise energy consumption rather than
market energy efficiency to their customers. Regulations requiring energy companies to provide
energy efficiency and conservation services and tariff incentives for reduced use can be used to
address this problem.
Lack of rational decision-making
Energy bills are often a small part of consumers’ overall budgets, and may not receive the
attention required to maximise energy savings. Force of habit, the view of energy as a fixed
overhead, and the perceived difficulty of making changes to one’s behaviour are important
factors. Furthermore, rational choices cannot be made where the necessary information is not
available. Public education campaigns are used as one way of addressing this situation.
9.3
Tax policies
Taxes or charges are one way to partially levy emitters for the environmental cost of their
emissions, and give them an incentive to find less damaging ways to conduct their business.
Revenue can be retained by government to fund incentive programmes and offset the costs of
meeting emission reduction commitments, and/or recycled back into the economy to offset
economic impacts.
In OECD countries, revenues from environmentally related taxes averaged 2.5% of GDP and
around 7% of total tax revenues for 2000. Most of the OECD country “green taxes” apply to
motor vehicles, energy products, or waste management. It has been estimated that a 1%
increase in energy prices via taxes would in the long term reduce energy use by about 5%. 4
In 1999, 26 OECD countries were surveyed on their policy responses to climate change. A
majority (73%) had proposed or implemented tax instruments to encourage the reduction of
greenhouse gas emissions. The most prevalent type was indirect rather than direct emission
taxes. One-third of the tax initiatives related to transport and more than half addressed fossil
fuels. There were also 11 carbon or emission taxes. More than half were not yet enacted at the
time of the survey.5
New Zealand was one of the minority of OECD countries in this survey that did not seek to make
use of tax instruments to reduce their greenhouse gas emissions in 1999. The others were
Austria, Hungary, Spain, Sweden and Turkey.
Among the OECD countries, New Zealand also has one of the lowest levels of tax on automotive
fuels, resulting in relatively low fuel prices. Generally countries with higher petrol prices consume
less per capita and therefore emit less greenhouse gas per capita. 6
4
OECD Observer, Summer 2000, p.120. The OECD “green taxes” database includes 170 taxes, 160 fees or charges, and over 850
exemptions and refund mechanisms, that levy charges on substances or activities that can have a negative environmental effect,
regardless of the reasons behind implementation. It is considered that for example a tax on fossil fuels introduced for purely fiscal
reasons will have the same environmental impact as one introduced to reduce pollution. For New Zealand, the database includes
excise taxes on fuel, motor vehicle licence fees, and road user charges (http://www.oecd.org/env/policies/taxes/index.htm ).
5
IEA 2000b, pp. 26-32.
6
Parliamentary Library 2000, Petrol prices and taxes, pp. 2, 4. Available on the Parliamentary Intranet.
85
The greenhouse effect and climate change
Parliamentary Library, August 2001
One issue usually raised about “green taxes” is their impact on consumer prices, industry, and
economic competitiveness against other countries without similar taxes. Whether the taxes are
additional or merely a reallocation of funds in the economy is a central consideration.
A recent study predicting the effects of a “low level carbon charge” in New Zealand found
that using the revenue to repay debt would cause a reduction in all measures of macroeconomic
activity, but recycling the revenue through a reduction in GST or income tax would result in
increases in output, consumption, and employment. All options would reduce CO2 emissions.
Some industries would have reduced output (e.g. fossil fuels, cement, metals and electricity) and
others increased output (e.g. agriculture, forestry, and related industries).7
The detailed information has been provided to the Tax Review 2001, which is due to report at
the end of September 2001. Cabinet has decided that if as a result a decision were taken to
proceed with a carbon charge, that it would not be implemented until after the next general
election.8
A summary of the findings from the New Zealand carbon charge analysis is in Table 9.1.
Most overseas studies have shown that the distributional effects of a carbon tax can have
negative income effects on low-income groups unless the tax revenues are used directly or
indirectly to compensate for such effects.9
Table 9.1: Brief summary of results from a 2001 analysis of a low level carbon charge
for New Zealand, with and without revenue recycling
Percent changes in
variables
CO2 emissions
Options with a $30 per tonne carbon charge
Revenue used to
Revenue recycled
Revenue recycled
repay debt
through reduction in
through reduction in
(not recycled)
GST
personal income tax
- 3.2 %
-2.79 %
- 2.9 %
Real consumption
- 0.09 %
+ 0.45 %
+ 0.17%
Investment
- 0.04 %
- 0.14 %
+ 0.17 %
Exports
- 0.18 %
+ 0.14 %
+ 0.09 %
Imports
- 0.12 %
+ 0.07 %
+ 0.05 %
GDP
- 0.09%
+ 0.26%
+ 0.16 %
Employment
- 0.15 %
+ 0.38%
+ 0.29 %
CPI
- 0.02%
- 0.69 %
-0.06 %
Price of petrol
for all scenarios: + 1.9 cents/litre
Source: Harvey 2001, Tables 1.2 and 2.1. The GST reduction option assumes a fixed real wage for its calculations.
Fuel price increases vary by carbon content of the fuel: varies for types of gas, oil, and coal, and electricity made from fossil fuel.
Some examples of climate change related tax policies overseas are given in Table 9.2.
7
8
9
Infometrics Consulting 2001, Bertram 2001, and Harvey 2001.
CBC Min (01)3/4 refers (quoted in Harvey 2001, p.1).
IPCC 2001c, para. 15.
86
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 9.2:
Overseas examples of tax policies that provide incentives to decrease greenhouse
gas emissions.
Country
Tax programme relating to greenhouse gas emissions
(a) Taxes on carbon, CO2 or fossil fuel
Denmark
Finland
Germany
Ireland
Netherlands
Sweden
Switzerland
UK
CO2 taxes Introduced in 1992, revised 1993 and 1999. 100DKK (NZ$27.67) per tonne CO2.
Subsidy schemes in place to return the revenue to industry for work-related expenses and energy
efficiency. Companies which enter into voluntary agreements to reduce energy use can also obtain CO2
tax reductions. As of 2000, light industry paid 90% of the tax, and heavy industry 25%. Electricity is taxed
at the consumption rather than production level. All sectors but transport covered.
Energy tax on household and public sector, and heating used by industry and energy utilities.
Differentiated by energy type: per GJ, generally 41DKK (NZ$11.34), but unleaded gasoline 102DKK
(NZ$28.33), diesel 67DKK (NZ$18.61), and natural gas 31DKK (NZ$8.61).
Sulphur tax Per kilogram of sulphur or SO2 emitted. Coal and heating oil attract all three taxes.
Passenger car fuel efficiency tax Annual taxes by efficiency capacity and fuel type of vehicle.
48 different rates, from minimum of DKK440 (NZ$122) for petrol cars able to get 20+ km per litre, to
maximum of DKK22,020 (NZ$6,116) for diesel cars able to get only 4.5-4.8 km per litre.
National goal: reduce 1988 levels of CO2 emissions 20% by 2005.
Tax on CO2 emissions
In 1999 Parliament approved increase from FMK 82 (NZ$28.58) to FMK 102 (NZ$35.55) per ton of CO2.
Use of wood tax-exempt, but tax on peat increasing from FMK 4.9 (NZ$1.71) to FMK 9 (NZ$3.14) per
kWh of energy produced.
Energy taxes Effective April 1999, energy taxes increased by Pf6/litre for diesel and gasoline, Pf4/litre
for heating oil, Pf 2/kWh for natural gas, and Pf 0.32/kWh for electricity (NZ 6.4 to 0.3 cents). Exemptions
for manufacturing industry (pay only 20% of the tax), and oil and gas for power generation in industry (not
taxed). Further increases scheduled for 2000-2003.
Carbon tax Approved Dec. 1998, started 1999, and fully phased in by 2005. A progressive tax applying
to all energy products. The existing tax structure on other fuels will be retained.
Carbon tax
Introduced in 1999, the “BSB” tax discriminates between three fuel types, with the highest tax on coal.
Energy and CO2 tax A range of rates by fuel or energy type.
Unleaded petrol SEK4.5 (NZ$1.03) per litre, diesel SEK 2922 to 3446 (NZ$668 to $778) per m3, natural
3
gas SEK1033 (NZ$236) per m , electricity consumption SEK 0.162 (NZ 3.7 cents) per kWh.
Tax on non-renewable fuels
Approval granted by both Swiss Parliamentary chambers as of 1999. Tax on non-renewable fuels such as
petroleum, gas, oil, coal and uranium of 0.3 centimes (NZD 0.04 cents) per kWh produced.
450 million Swiss francs per year to be used to promote renewable energies
(e.g. solar and hydro), energy efficiency measures in buildings.
“Climate Change Levy” Effective as of 1 April 2001.
Taxes energy content rather than carbon content. Exemptions are allowed for the public transport sector,
“good quality” combined heat and power businesses, companies which generate power from “green”
sources, and businesses entering in energy saving programmes. After heavy lobbying, energy intensive
industries received a 30% concession on the Levy, and an additional 80% discount if they signed energy
efficiency agreements.
It is projected that this measure will produce half of the national emissions
reduction target (12.5% reduction in 1990 emissions by 2012) and GBP 1 billion in revenue.
(b) Tax reductions for alternative fuels and technologies
Australia
Canada
Japan
USA
Tax exemptions for trains Part of the “New Tax System”, from July 2000 (as reported in late 1999).
100% excise tax credit for rail transport to improve its competitive position.
Reducing wastage of flared fossil fuels As of 1999 Federal Budget.
Generating equipment fuelled by flare gas at oil fields eligible for higher capital cost tax allowance.
To help reduce emissions and displace coal-fired electricity generation.
Tele-work centres to reduce transport emissions
Tax system gives incentives for “tele-work” centres to discourage long-distance commuting are given tax,
and lower taxes on low fuel consumption and low-emission vehicles.
Biomass electricity tax credit Proposed for FY 2000 budget.
Extension for another five years for current tax credit of 1.5 cents (NZ 3.6 cents) per kWh for electricity
produced from biomass and 1 cent (NZ 2.4 cents) tax credit for co-fired biomass and coal plants. Type of
eligible biomass also extended to include certain forest and agriculture related resources.
Tax credit for renewable power sources
Tax credit for electricity from wind and other renewable sources. Expected to reduce the cost of wind
power between US 1.3 – 2 cents (NZ 3 to 4.7 cents) per kWh.
Sources: IEA 2000b; OECD 1999 (http://www.oecd.org/env/policies/taxes/index.htm); for Denmark, Kristofferson et al 1997; for Norway,
Offshore November 1999, p. 18; for UK, European Report 31/3/01, p. 342; Economic Review (UK) 17(2):5; Nitrogen & Methanol Journal,
January 2000, p.12. Abbreviations: GJ = Gigajoule and kWh = kiloWatt-hour (measures of energy).. Others are units of currency.
Conversions to NZ dollars as of May 2001 (source Pacific Exchange Rate Service at http://pacific.commerce.ubc.ca/xr/data.html )
87
The greenhouse effect and climate change
9.4
Parliamentary Library, August 2001
Emissions trading and quotas
The basic principle behind “emissions trading” is that an artificial element of scarcity is injected
into a market which regards the atmosphere (as a sink for emissions) as inexhaustible and
“free”. Once scarcity has been created, such as through legal quotas for emissions that reduce
over time, parties can take the cheapest option to reduce emissions, either doing it themselves
(and sell the surplus if they can do it relatively cheaply), or buying emission credits from another
party. Emissions trading is seen by proponents as a cheaper and more efficient alternative to
historic “command and control” regulatory approaches.
Theoretically a range of commodities could be tradable under the Kyoto Protocol, including:
•
carbon sink credits (e.g. from “Kyoto Forest”, either from a country or a private holder);
surplus emission rights (“assigned amount” not required by an Annex I country, or surplus
•
private emissions quota under a domestic scheme);
•
emission reductions that are cheaper to effect than one’s own emissions at home in nonAnnex I countries (Cleaner Development Mechanism or CDM), in Annex I countries (Joint
Implementation), or private transactions within a domestic quotas scheme.
The rules for international greenhouse gas emissions trading have not yet been set up, apart
from a few policy decisions made in Bonn in July 2001 (Box 2 in Chapter 1). It is hoped that
current negotiations and pilot schemes will lead to a system being in place by 2008. The
estimated value of “emissions units” for 2008-2012 is $13 to $50 per tonne of CO2 equivalent.10
An emissions cap and quotas for the electricity sector have been introduced in Denmark in
anticipation of an international emissions trading scheme being developed, the Greenhouse Gas
Emission Reduction Trading Pilot is in operation in Canada, and a domestic emissions trading
pilot is scheduled to start in the UK in 2002. Other countries investigating introduction of a
domestic emissions trading scheme include Australia, France, Germany, the Netherlands, New
Zealand, Norway, Sweden, and the USA.11
A discussion document on carbon sink credit trading for New Zealand was released in July
2001. The possible rules have been summarised in the chapter on carbon sinks (section 5.2.2,
Table 5.6) as a central issue is the rights and obligations of New Zealand’s forest owners.
With limited experience in emissions trading worldwide, the benefits for climate change at this
stage remain largely theoretical. There are two models with a longer track record: the US Acid
Rain Program and New Zealand’s fisheries management using Individual Transferable Quotas
(ITQs). The New Zealand fisheries quota management system has successfully created a
market (an estimated 77% of originally allocated quota having changed ownership) and the Acid
Rain Program has achieved a 25-30% reduction in SO2 emissions and acid deposition in the
most sensitive region since 1995.12
However, neither of these programmes has faced the exact suite of issues relevant to
greenhouse gas emission management. In particular, documenting “additionality” for CDM
(Cleaner Development Mechanism) project credits and ensuring accountability and transparency
could add significantly to transaction costs. Reconciliation of privately generated emission
credits with national emission reduction targets is another challenge.
Key design issues for either international or domestic schemes include;
•
defining the unit of trade (e.g. tonnes of CO2 equivalent);
•
administration (registry of emission unit holdings and transfers, emissions inventory, etc.);
•
assigning points of obligation (for reporting emissions and sinks);
•
initial allocation of emission units or ownership of carbon sinks.
10
Ministry of Economic Development, http://www.med.govt.nz/ers/environment/climate/emissions/index.html
IEA 2000b pp. 58, 69, 118; U.K. Department of the Environment http://www.environment.dtlr.gov.uk/consult/ggetrade , para. 1.15;
UK Emissions Trading Group http://www.uketg.com; Ministry for the Environment 1999.
12
United Nations Conference on Trade and Development 1998 pp. 1-12; U.S. Environmental Protection Agency 1999, pp. 5,9.
11
88
The greenhouse effect and climate change
Parliamentary Library, August 2001
One of the unresolved issues in international emissions trading is that of “hot air”, or credit for
emissions which are no longer being produced. Some countries, primarily in Eastern Europe,
have suffered economic collapse and a reduction in emissions of up to a third since the 1990
baseline under the Kyoto Protocol. If they do not significantly increase their current emission
levels, they will theoretically have carbon credits to trade with other countries or industries
wishing to offset their emissions. For Russia, the value of surplus assigned amount in 2008-2012
has been estimated at about $10 billion a year.13 Countries or private agents which purchase
such credits will be basing their “right to pollute” on assigned amount which is no longer being
used, and additional reduction in global emissions will not result.
A discussion of the Joint Implementation and Clean Development Mechanisms, which will
involve trading in some form, is in section 9.8.
9.5
Creating a market for “green energy”
Deregulation of the energy retailing market and separation of generation and delivery services
allows the creation of a special market for “green energy” for consumers to select if they choose,
as long as other market barriers do not remain and adequate incentives exist.
M-co, the company that runs New Zealand’s electricity market, developed an internet-based
Green Electricity Market (GEM) programme for Australia. The GEM will allow electricity retailers
to fulfil their new obligations under the Renewable Energy (Electricity) Act 2000 for an additional
2% of all electricity to come from renewable sources by 2010. The “green” electricity joins all
other electricity in the national grid, but generators receive Renewable Electricity Certificates as
to the amount generated, and it is these certificates which are traded. In June 2001 the average
price for the certificates in the 10-year forward trade market was A$25.14
The GEM was also tested in Europe in May 2001 at the invitation of the European Union.15
Other places with developing markets include Belgium, the European Union, Finland, France,
Germany, Italy, Japan, the Netherlands, the UK, and seven states in the USA.16
Unlike many other countries New Zealand already has a high percentage (73%) of electricity
from renewable sources, mainly hydroelectricity. However, this advantage is being eroded by
increasing use of fossil fuels. Like the USA, Canada, Japan and Norway, New Zealand has had
a declining share of renewables in both total energy and electricity (Figure 9.1).
In New Zealand, unlike in the countries where markets for “green energy” are developing, there
are no regulatory or market incentives to encourage development of renewable energy sources
or trade in green energy. For example, the developer of the Tararua Wind Farm has stated that it
is the only large wind farm in the world that does not have some form of financial assistance
from government. A recent campaign that sought to gain 16,100 subscribers willing to pay $2
more a week for green energy to subsidise stage two of the wind farm netted only 200.17
New Zealand’s draft National Energy Efficiency and Conservation Strategy contains voluntary
encouragement measures rather than legal and/or fiscal ones as used in other countries (Table
10.3).
13
Raab 2001, Major winners and losers at Bonn climate talks, http://www.earthtimes.org/bonn
The Independent 16 May 2001, M-co grows green electricity business, (p. 7); http://www.gemoz.com ; http://www.greenprices.nl ;
http://www.greenhouse.gov.au/markets/2percent_ren. Australia’s Mandatory Renewable Energy Target commenced on 1 April 2001
and requires the generation of 9,500 gigawatt hours of extra renewable electricity per year by 2010, enough power to meet the
residential electricity needs of four million people.
15
http://www.m-co.co.nz/Dnews/010516.htm
16
IEA 2000b, p. 69, 93; Energeia 16/1/01 cited on http://www.greenprices.nl/nl/newsitem.asp?nid=165;http://www.greenprices.com.
The USA states are Arizona, Connecticut, Maine, Massachusetts, Nevada, New Jersey and Texas (Union of Concerned Scientists
2000, http://www.ucsusa.org/energy/brief.rps.html).
17
Manawatu Evening Standard 20/7/01, Windfarm in the doldrums.
14
89
The greenhouse effect and climate change
Figure 9.1
Change from 1991 to 1998 in
the share of renewable and
waste energy sources in total
primary energy supply (TPES)
and total electricity. Ranked
by share of electricity.
Source: International Energy Agency
2000d, pp. II.326-11.354.
“Renewable” energy includes hydro,
solar, geothermal, wind, tide/wave/
ocean, heat pumps, biomass, methane
from biowaste,
and
bio-alcohols.
“Wastes”
includes
industrial
and
municipal wastes (combustion and
other).
Parliamentary Library, August 2001
Japan
Italy
USA
NZ
TPES
electricity
Australia
Canada
Norway
Sweden
Ireland
Germany
Finland
UK
Netherlands
Denmark
-100%
0%
100%
200%
300%
400%
percent change 1991-1998
9.6
Financial support through grants and loans
Historically fossil fuels have been subsidised in many countries. Among OECD countries these
subsidies have been slowly declining. Belgium, Portugal and the UK have eliminated their coal
subsidies since 1992, but 5% of the coal produced in OECD member countries remained
subsidised in 1999. Initiatives to remove the remaining subsidies on coal and other fossil fuels
are not apparent.18
Financial support to load additional market value on energy efficiency and renewable and other
alternative fuels are common measures used overseas. Examples are shown in Table 9.3. In
New Zealand, EECA operates the Energy Saver Fund and other grants to assist low income
groups to retrofit energy efficiency measures in homes, and loan schemes to assist government
agencies and the private sector to invest in improved energy efficiency.
9.7
Energy efficiency standards and labelling
Standards can set acceptable levels of energy efficiency in consumer products (either voluntary
or mandatory), and labelling can give consumers the information they need in the marketplace to
include energy efficiency among the factors they consider when choosing between products.
New Zealand is currently preparing to introduce a Minimum Energy Performance Standard
(MEPS) and an energy labelling regime for a range of products. There is currently a voluntary
system in place for refrigerators, modelled on the Australian standard.
Overseas, a variety of initiatives are in place (Table 9.4).
18
IEA 2000b, p. 26.
90
The greenhouse effect and climate change
Table 9.3:
Parliamentary Library, August 2001
Overseas examples of subsidy, grant and loan programmes for encouraging
energy efficiency and use of alternative and renewable energy.
Australia
Renewable Remote Power Generation
Up to 50% rebate for costs of installation conversion from diesel to renewable generation
in remote areas, through A$264 m (NZ$325 m) to states and territories.
Subsidy – photovoltaic systems (solar panels to produce electricity)
Rebate for up to 50% of the costs of installing household photovoltaic systems (max.
A$8,250 (NZ$10,157) per household), total funding A$31m (NZ$ 38.2m).
Alternative Fuel Conversion Programme
As of July 2000, grants for new alternative fuels vehicle (up to 50% of the difference in
price from conventional vehicle) and conversions to CNG or LPG (up to 50% of cost of
conversion).
Canada
Commercial Building Incentive Programme
Operating since 1997, expanded in 1999 to include multi-unit residential buildings. Onceonly grant of twice the estimated annual energy cost savings for approved designs up to a
maximum of $80,000 (NZ$122,984). Helps offset the cost of designing energy-efficient
buildings.
Natural Gas for Vehicles Initiative
Provides market, emission and safety studies, information, technology transfer, and direct
subsidies to encourage production and use of alternative fuel vehicles. C$2000
(NZ$3,075) is provided for each factory-built natural gas vehicle or C$500 (NZ$768) for
each road vehicle converted to natural gas. In 1999 the programme was renewed to 2001.
Germany
100,000 Roofs Solar Power Programme
A total of DM 1.1 billion (NZ$1.17 billion) over 1999-2005 for low-interest loans for
installation of photovoltaics (solar panels to produce electricity).
Renewable energy price subsidy
As of March 2000, guaranteed prices to producers for electricity from renewable sources.
Aim is to double the share of renewable energy by 2010. Fixed subsidy rates rather than
linkage to consumer prices, e.g. solar power, subsidised price DM.99 (NZ$1.05) per kWh,
up from market level of about DM.17 (NZ 18 cents). Geothermal, methane, and biomass
also benefit but at lower rates. The law was controversial, and calls for review of the solar
power subsidy once market share increases.
Renewable energy promotion
From 1999 to 2003, a total of DM 1 billion (NZ$1.06 billion) is allocated to support the
installation of solar thermal collectors (subsidy within certain limits), and energy
conservation measures in buildings (grants or low-interest loans).
Netherlands
Energy efficiency in industry
Subsidies are available from the Novem agency to firms in the asphalt, ceramics and steel
industries for improving energy efficiency. Tax breaks are also available if emission
reductions are agreed to.
Source: IEA 2000b.
Conversions to NZ dollars as of May 2001 (source: Pacific Exchange Rate Service at http://pacific.commerce.ubc.ca/xr/data.html )
91
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 9.4: Overseas examples of energy efficiency standards and labelling initiatives.
Australia
Minimum Energy Performance Standards (MEPS) introduced for refrigerators, freezers,
and electric water heaters.
New passenger cars will require mandatory model-specific fuel consumption labelling
under the Australian Design Rule.
Canada
An EnerGuide Labelling Programme is in place for home appliances and equipment. In
1999-2000 this was expanded to provide the public with information to enable them to
make energy-wise decisions in relation to home improvements and home buying.
Denmark
An energy label is required on all new passenger cars in salesrooms (from April 2000).
European
Union
Mandatory energy efficiency labelling in place for household appliances since 1992. All
household electric lamps will be required to carry a label showing energy efficiency
from the beginning of 2001.
New passenger cars sold in the EU will be required to carry a label on fuel economy
and CO2 emissions from the beginning of 2001.
Germany
Energy consumption labelling for domestic light bulbs and dishwashers is required (as
of 2000).
Japan
Environment and Energy Friendly Building Mark indicates energy conservation
performance above a certain standard, for structures other than houses (since 1999).
Standards for homes are also being developed.
USA
Federal energy efficiency standards are in place for equipment and appliances such as
heating and cooling equipment, water heaters, lighting, refrigerators, clothes washers
and dryers, and cooking equipment. It is estimated that this measure will prevent the
emission of 225 million tonnes of carbon (cumulative) by 2010.
Source: IEA 2000b; EECA 2000.
Energy efficiency in building codes
A particular form of energy efficiency standard is the building code. A significant portion of
households’ contribution to greenhouse gas emissions comes from heating and cooling.
Buildings can be built in such a way as to minimise the need for electricity and fossil fuels for
heating and cooling, through such measures as insulation and design for passive solar power.
Market barriers mean that this is often not done. Making the use of mandatory standards
ensures the nation's building stock is made progressively more energy efficient over time.
In New Zealand, minimum household insulation requirements were put into law in 1977.
However, the standard was based on the Auckland climate and not well suited to the majority of
the country, and it omitted consideration of other types of energy inefficiency in buildings.
Improvements were developed, but blocked for over 20 years primarily by Treasury.19
Enhanced building energy efficiency standards are now effective through amendments to clause
H1 of the Building Code (effective 31 December 2000). These improve the minimum insulation
requirements for colder areas, introduce heat loss and lighting energy level limits for commercial
buildings, and set maximum heat loss from hot water storage systems.20
19
20
Parliamentary Commissioner for the Environment 2000, pp. 60-65.
e.g. The Dominion 19 July 2000; EECA.
92
The greenhouse effect and climate change
9.8
Parliamentary Library, August 2001
The Clean Development Mechanism and Joint Implementation
The Kyoto Protocol allows for development of Clean Development Mechanism (CDM) and Joint
Implementation (JI) projects.21 These essentially involve Annex I Parties obtaining greenhouse
gas credits for reductions that occur in other countries through their sponsorship.
The theory behind CDM and JI is essentially that for climate change purposes it does not matter
where in the world reductions in greenhouse gas emissions occur, and it may often be cheaper
to reduce them in lesser developed countries where costs are lower. It is therefore a market
mechanism.
CDM projects would be between Annex I Parties and non-Annex I countries (i.e. between
developed and developing countries), and JI projects would be between Annex I Parties, such
as between the European Union and Eastern Europe.
CDM is seen as a way for wealthier countries to assist developing countries to adopt cleaner
energy technology sooner, and as a source of valuable revenue for developing countries,
potentially creating a “win-win” situation.
The rules for CDM and JI under the Kyoto Protocol have not been agreed upon, and were in fact
one of the points of contention on which the COP6 gathering foundered in November 2000.
However, many pilot projects have already taken place and their proponents are hoping that
they will qualify under the rules that finally do emerge.
Prior to the Kyoto Protocol, a number of projects that may subsequently qualify as CDM or JI
were initiated under the UNFCCC in a pilot phase for Activities Implemented Jointly (AIJ).
Currently on the AIJ database are 144 projects, the majority relating to energy efficiency,
renewable energy, and fuel switching. Other project types are forest preservation, afforestation,
agriculture, and fugitive gas capture.22 Examples of these and other trial projects are
summarised in Figure 9.5.
The World Bank has also established a Prototype Carbon Fund (PCF), a “learning by doing”
project for CDM and JI type projects. Private sector and government parties buy into the fund, in
exchange for a pro-rata share of the resulting greenhouse gas emission credits. In October 2000
five projects were nearing implementation. The PCF projects are expected to commence
development by the end of 2003 and be operational before January 2008.23
Before the international rules under the Kyoto Protocol for CDM and JI can be agreed to, some
major issues must be resolved. These issues include the following.
“Additionality”, “free riding”, and establishing baselines
In order to gain CDM or JI credits, there must be evidence that the project will create a genuine
reduction in the greenhouse gas emissions. In other words, the reduction must be additional to
any reduction that might have occurred anyway.
“Free riding” is a term used in the climate change literature to refer to the possibility that a
project will claim a reduction that is not truly additional, or generally that entities will benefit from
actions without contributing to their costs.
21
Kyoto Protocol, Articles 6 (Joint Implementation) and 12 (Clean Development Mechanism).
UNFCCC, Activities Implemented Jointly (AIJ), on http://www.unfccc.de/program/aij/aijproj.html
23
World Bank http://www.prototypecarbonfund.com The projects nearing implementation were in Latvia (methane capture from
solid waste), Costa Rica (renewable resources), Czech Republic (energy efficiency), Uganda (small hydro power), and Guyana
(biomass co-generation).
22
93
The greenhouse effect and climate change
Table 9.5:
Parliamentary Library, August 2001
Examples of pilot Clean Development Mechanism and Joint Implementation type
projects, and other pilot emissions trading
Countries
involved
Project summary
(a) Clean Development Mechanism (CDM) type projects
Annex I and non-Annex I Parties
Norway
Mexico
World Bank GEF
energy efficiency
Norway
Denmark
Netherlands
Burkina Faso
alternative fuels
Australia with
Solomon
Islands, Fiji,
Mauritius,
Indonesia, Chile
The “Ilumex” project involves subsidised replacement of 1.7 million household incandescent light
bulbs with energy-efficient fluorescent bulbs in Monterrey and Guadalajara, saving 940 Gwh of
electricity and preventing the associated CO2, CH4 , SO2 and NOx emissions. Norway has not
sought emission credits in exchange for its US$3m (NZ$1.3m) share. Commenced in 1995.
The overall objective is to meet growing urban demand for household fuels in Burkina Faso
without further loss of forest cover and carbon sequestration potential. Methods subsidised are
efficient carbonisation techniques to minimise fuelwood use, community-based forest
management, kerosene cooking stoves, and rural photovoltaic systems. Protection of 130,000
tonnes of timber and prevention of 1.5 million tonnes of CO2 over 5 to 6 years are anticipated.
The lifetime of these AIJ projects range from 1 to 20 years, and the estimated CO2 reduction
ranges from 13 to 5,200,000 tonnes. Projects involve energy efficiency (air conditioning in
Solomon Islands and power generation efficiency in Mauritius), renewable energy (solar in Fiji
and Mauritius and micro-hydro in Solomon Islands), fugitive gas capture (natural gas in Chile),
and renewable energy training and demonstration (Indonesia).
(b) Joint Implementation (JI) type projects
between Annex I Parties
Netherlands
Czech Republic
alternative fuels
Netherlands
Romania
energy efficiency
Norway
Slovakia
alternative fuels
Sweden
Balkan States
mostly energy
Project expected to be operational March 2000. The Netherlands helped fund a new biomass
heating grid and boiler plant in the town of Hostetin. The two countries will share equally the CO2
credits for the period 2008-2012.
Agreement signed in 1999. The Netherlands will provide 2 million guilders (NZ$1.88 m) for two
energy efficiency projects in the town of Targu Mures at a drinking water plant and a wastewater
treatment plant. The CO2 reductions created will be credited back to the Netherlands national
target after 2000.
Norway will contribute NOK 1.2 m (NZ$311,000) to convert two district heating schemes from
fossil fuels to biofuels. The net CO2 reduction is 50,000 tonnes. This is viewed as a basis for
more agreements in future.
As of 1999, Sweden financed 66 Activities Implemented Jointly (AIJ) with Balkan states, of which
51 have been approved under the AIJ programme.
The total CO2 reduction up to 1999 was 814,036 tonnes, and during 1999 was 201,954 tonnes.
(c) Other – involving private parties
Canada
USA
carbon sinks /
soil management
carbon credits/
energy production
carbon credits/
methane reduction
USA
Ecuador
forest protection
USA
Chile
renewable energy
•
A group of 10 Canadian energy companies will pay Iowa farmers for reducing CO2
emissions and improving CO2 sinks in soil, in exchange for obtaining the resulting CO2 credits.
The farmers must practise no-till or minimum-till soil management. The first deal is expected to
produce 1.3 million tons of carbon credits for 2000, and up to 6 million tons by 2012. The price
per carbon credit in 1999 was C$0.50 to $2.50 (NZ$0.77 to $3.84).
•
Canadian oil sands producer Suncor Energy purchased 100,000 tonnes of greenhouse gas
emission credits from the USA utility Niagra Mohawk Power Corporation which had significantly
reduced their emissions from 1990 levels. The agreement commits Niagra Mohawk to reinvest
at least 70% of the net proceeds in new projects to reduce emissions further, and includes
options of purchasing up to 10 million additional tonnes per year from 2001, depending on the
provision of credits by the respective governments.
•
In 1999, Ontario Power Generation Inc. bought 2.5 million tonnes of CO2 equivalent
emission reductions from Zahren Alternative Power Corp., a USA operator of landfill gas
collection and energy projects. PriceWaterhouseCoopers will independently verify the
reductions.
World Parks Endowment Inc. has sponsored the purchase of 2,000 ha. of lowland wet forest
threatened with logging to form a new Bilsa Biological Reserve. Protection of biodiversity was
the main objective, but 1,170,107 tonnes of CO2 will also not be released from sequestration by
logging. (AIJ project)
The Wind Energy Project will install 50 windmills with 37.5 MW generating capacity, saving
approximately 3 million tonnes of CO2 from coal-fired energy generation over a 20 year period.
The sponsor is the International Institute for Energy Conservation. (AIJ project)
Sources: UNFCCC http://www.unfccc.d/program/aij ; Work Bank http://www-esd.worldbank.org/aij/brief.htm ; IEA 2000b.
GEF = Global Environment Facility Trust Fund. Contributed to by Annex I Parties, and administered by the World Bank as Trustee and
Implementing Agency. About two-thirds of all project-related GEF resources are allocated to the World Bank GEF portfolio.
94
The greenhouse effect and climate change
Parliamentary Library, August 2001
The key to ensuring “additionality” is the setting of accurate baselines. The trouble is “what
would have happened anyway” is not known with accuracy and can only be estimated. Broad
industry or country averages and standardising have been proposed to reduce the costs of
estimating project-specific baselines and therefore encouraging more CDM or JI projects to take
place, but this may be at the expense of certainty that all of the claimed greenhouse gas
emission credits will actually help reverse climate change.
For example, a new gas-fired power station in India could generate almost US$300,000 of
credits per year (using a “low” carbon credit of US$5 per tonne CO2) if compared to recent
generating capacity construction in the electricity sector generally (all technologies), but
significantly less if compared only with recently constructed gas plants. In another example,
uncertainties in simple refurbishment-type efficiency improvement projects in the energy sector
can be as high as ± 80%.24
Accountability and transparency
CDM and JI reporting needs to be transparent enough to allow a third party to understand key
features of a CDM or JI project, the project baseline (“no project” scenario), and any greenhouse
gas emission reduction credits generated. In contrast, an analysis of 45 pilot projects found a
significant lack of documentation of the data and assumptions behind the project baselines.25
The Annex I Parties and industry players generally agree that accountability and transparency
are essential elements of any CM/ JI rules, but have not agreed on how this would work in
practice.
“Gaming”
The risk that parties will deliberately inflate individual project baselines to maximise greenhouse
gas emission-reduction credits is termed “gaming” in the climate change literature.
The International Energy Agency considers that the risk of gaming is a significant drawback of
project-specific baselines, and that the potential for gaming will be reduced if standardised
baseline assumptions are developed through a process with independent experts.26
“Leakage”
If upstream and downstream effects of a project on greenhouse gas emissions are not taken into
account, then the estimated actual effect of the CDM or JI project may be significantly under- or
over- estimated. “Leakage” is the term used in climate change literature to describe changes in
emissions that occur outside of the project boundary and evaluation of the project’s
performance.
If the project accounting net is thrown widely, the accuracy of the baseline should significantly
improve, but the cost of accounting may be prohibitive. “Double counting” may also occur for
indirect emissions over which the project has no control if another CDM or JI project accounts for
them as well.
“Fungibility”
This comes from a legal term meaning exactly equivalent for the relevant purposes. In the
climate change negotiations context, it means whether or not emissions reductions earned can
be readily directed, or redirected, at any time to meet any Annex I country’s obligations,
regardless of whether they are credits from surplus assigned amounts for countries’ emissions,
24
25
26
IEA 2000c, pp. 23, 34.
OECD 1999, cited in IEA 2000c, p. 23.
IEA 2000c, p. 27.
95
The greenhouse effect and climate change
Parliamentary Library, August 2001
carbon sinks, or credits from CDM or JI projects. In other words, will emissions credits be fully
exchangeable regardless of the country and method of origin?
Allowing Annex I countries to continue “business as usual”
When the USA proposed at COP6 to claim a large proportion of its emission reductions in CDM
projects, many Third World and European community nations objected. Essentially the concern
was that in order to reduce dangerous anthropogenic effects on the climate, greenhouse gas
emissions need to be substantially reduced at source in the industrialised nations as well as in
developing countries.
The Kyoto Protocol requires that JI projects be supplemental to domestic actions (Article 6(1)d),
and the text on CDM refers to the earned credits being eligible for meeting part of reduction
commitments (Article 12(3)b), but leaves open for debate the quantum actually allowed.
Impact on indigenous peoples and ecosystems
Indigenous peoples have voiced strong concern about the way CDM may operate in practice.
“Our intrinsic relation with Mother Earth obliges us to oppose the inclusion of sinks in
the Clean Development Mechanism (CDM) because it reduces our sacred land and
territories to mere carbon sequestration which is contrary to our cosmovision and
philosophy of life.
Sinks in the CDM would constitute a worldwide strategy for expropriating our lands
and territories and violating our fundamental rights that would culminate in a new
form of colonialism.
Sinks in the CDM would not help to reduce greenhouse gas emissions, rather it
would provide industrialised countries with a ploy to avoid reducing their emissions at
source.” 27
For example, evidence has been presented of Pygmy hunter-gatherers being driven from their
home to make way for a World Bank CDM-type forestry project.28
There are currently 13 “forest preservation” projects under the Activities Implemented Jointly
programme. One of these involves the introduction of “reduced impact logging” in East
Kalimantan (Indonesia) in areas where logging has already been scheduled. CO2 reduction
credits are calculated for logging of indigenous forests if techniques are used that cause less
damage to non-economic species and soil resources. While technically this approach may
release less carbon than “business as usual” logging, it may still pose a significant risk to
biodiversity and indigenous peoples.29
The latest Kyoto Protocol agreement from Bonn now excludes such “avoided deforestation” type
activities from CDM, as only afforestation and reforestation qualify. In addition, an agreed
principle guiding all LULUCF activities (CDM or otherwise) is that they contribute to “the
conservation of biodiversity and sustainable use of natural resources.”30 There is no specific
mention of protecting the interests of indigenous peoples.
27
Declaration of the First International Forum of Indigenous Peoples on Climate Change, 4-6 September 2000, available on
http://www.ienearth.org/climate_1-p2.html . Signatories represented alliances of native peoples of tropical forests including Asia, the
South Pacific, the Americas (including the Amazon), and Africa.
28
Climate Alliance, Indigenous peoples of the tropical rainforest and the CDM, quote from K. Zephryin of Rwanda, internet address
as for previous footnote.
29
http://www.unfccc.de/program/aij/aijact00/usaidn01-00.html , pp. 1, 17.
30
UNFCCC 2001c, section 3.8, and by reference section VII.1.e
96
The greenhouse effect and climate change
9.9
Parliamentary Library, August 2001
CDM and the nuclear energy issue
Overseas there is a strong lobby to get nuclear energy accepted as a “clean” fuel with regard to
climate change. At the last Conference of Parties (COP 6, The Hague), Australia, Canada and
Japan reportedly pushed for nuclear power to be officially accepted as suitable for Clean
Development Mechanism (CDM) projects under the Kyoto Protocol.31 Under this scenario,
developed countries could gain greenhouse gas emission “credits” by providing aid to
developing countries to build a nuclear power station instead of one powered by fossil fuels. The
proposal was not agreed to.
Greenhouse gas emissions from nuclear energy plants are low compared to generation of
electricity using fossil fuels.32 Nuclear energy generates about 35% of the electricity in
European Union (EU) countries, and the European Commission has estimated that EU nuclear
reactors prevent 312 million tonnes of CO2 from entering the atmosphere annually, equivalent to
7% of total EU emissions. The European Atomic Forum, a nuclear industry trade group, has
estimated that worldwide, nuclear energy plants help avoid the emission of 1.8 billion tonnes of
CO2.33
However, the risks of long-term radioactive contamination from power plant emissions,
accidents, or inadequate waste treatment are significant environmental liabilities from this
energy source. These were brought to vivid public attention following the 1986 Chernobyl
accident. In 2000, no reactors were under construction, on order, or planned in North America or
Europe; but in 14 other countries, primarily in Asia and Eastern Europe, 38 reactors were in the
planning and construction stages.34
The latest Kyoto Protocol agreement from Bonn requires Annex I Parties “to refrain” from using
certified emission reductions generated from nuclear power facilities to meet their greenhouse
gas emission reduction commitments.35
31
Tilting at Windmills: Nuclear Power and Climate Change, Habitat Australia 2001, 29(1):18.
Nuclear electricity production per se produces no greenhouse gas emissions, but emissions arise from the extraction, transport,
and processing of fuels, the development and maintenance of nuclear waste storage facilities, and the transport of nuclear waste.
33
Sains, A 2001, The Uncertain Future Of Nuclear Energy, Europe Feb.2001, p. 26; Laurent, C 2001, Beating Global Warming with
Nuclear Power?, UNESCO Courier Feb. 2001, p. 38
34
Laurent 2001, p. 39; reactor construction data from The International Atomic Energy Agency.
35
UNFCCC 2001c, section 3.2.
32
97
The greenhouse effect and climate change
98
Parliamentary Library, August 2001
10
Additional detail: the New Zealand situation
An historical overview of New Zealand’s climate change policy 1990 to 2001 is in chapter 2, and
summaries of Government action are noted throughout the report as appropriate. This chapter
provides more detail on some key elements.
10.1
Economic instruments
The three domestic climate change policy options proposed by the Government in 1999 for
public discussion primarily employed economic instruments. Measures to educate the public and
industry and to remove other barriers to energy efficiency were considered to be
“complementary measures”.1 The proposed domestic policy options were:
•
Forward trading
This option would focus on enhancing domestic awareness of the domestic and international
carbon emission trading systems that would operate during the first Kyoto Protocol
commitment period 2008-2012. This would allow New Zealand firms in a position to forward
trade (trade in anticipation of 2008-2012) to engage in trades with each other. There would
be no mandatory requirement to reduce domestic emissions prior to 2008-2012, but firms
could choose to do so if they decided it was cost effective.
•
Carbon charge with pilot trading
This option would focus on introducing a price for carbon into the market as soon as
possible, combined with a pilot carbon-trading programme for major point-source firms.
Participants in the pilot scheme would be asked to accept a cap on emissions below
“business as usual”, in exchange for being exempt from the carbon charge, and participation
would probably be voluntary. The carbon charge would be “low”, possibly $5-$10 per tonne
of CO2.
•
Carbon charge
This option would focus on introducing a price for carbon into the market as soon as possible
while work progresses on the design and implementation of a comprehensive domestic
emissions trading system. The charge would be applied at a uniform rate across all CO2
emitters, add to the price of the product, be subject to GST, and be a deductible business
expense in businesses’ liability for income tax. The carbon charge would be “low”, possibly
$5-$10 per tonne of CO2.
Submissions were received on these proposals, but decisions on any economic instrument were
deferred until further advancement of Kyoto Protocol negotiations. In 2001, an analysis of the
likely economic impacts of a low-level carbon charge was forwarded to Tax Review 2001. More
detail is in chapter 9, together with the discussion on tax policies (section 9.3). Government has
indicated that if Tax Review 2001 recommends a carbon charge, that it would not be
implemented until after the next election (section 2.2).
A discussion paper on forest sinks and carbon trading was released in 2001. The proposed
features of a trading regime are summarised in the chapter on forest sinks (section 5.2).
1
Ministry for the Environment 1999, pp. 67-71.
The greenhouse effect and climate change
10.2
Parliamentary Library, August 2001
Energy Efficiency and Conservation Authority (EECA)
The Energy Efficiency and Conservation Authority (EECA), originally established in 1992, is now
a Crown entity under the Energy Efficiency and Conservation Act 2000. The function of the
authority is now to encourage, promote, and support energy efficiency, energy conservation,
and the use of renewable sources of energy (s 21).
Over five years, Government funded EECA $30 million. Programmes with “hard quantifiable”
benefits have cost around $12.5m and achieved benefits of $59m (net present value $46.5m).
Programmes with “soft/indirect” benefits, such as the Energy-Wise Companies Campaign that
has served over 700 businesses but where energy savings are subject to commercial
confidentiality, are estimated to have produced benefits five to six times the programme costs. 2
A summary of cumulative benefits from EECA funding is presented in Table 10.1, EECA funding
trends in Figures 10.1 and 10.2, and the budget and output objectives in Table 10.2.
A draft of the National Energy Efficiency and Conservation Strategy, required by the Energy
Efficiency and Conservation Act 2000, was prepared by EECA and released by the Minister of
Energy in March 2001 for public comment. The draft proposed a wide range of initiatives
designed to reduce the market and other barriers to widespread adoption of energy efficiency
and conservation, which are summarised in Table 10.3. The Act requires the final strategy to be
published by 1 October 2001.
2
http://www.eeca.govt.nz/content/EECA_Corporate/overview.htm
100
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 10.1: Summary of EECA cumulative benefits 1999-2000.
Cumulative benefits
“soft/
indirect”
programmes
to 1999/00
“hard
quantifiable”
programmes
to 1999/00
TOTAL
$71 m
$135-$185 m
$206 to $256 m
Energy cost
savings
1.1 m tonnes 4.3 m tonnes
5.4 m tonnes
CO2 emission
reductions
“Hard quantifiable” programmes have fully documented savings. “Soft/indirect”
programmes involve a variety of actions by third parties with results that are
self-reported or estimates. If the total is averaged over the period 1992-2000, or
eight years, it equals 675,000 tonnes/year. If that amount is added to the 1999
national emissions data, one can estimate that without the EECA
programmes New Zealand’s total greenhouse gas emissions would have
been 0.9% higher, and total CO2 emissions 2.2% higher, in 1999.
Sources: EECA Annual Report for 1999/2000; Ministry for the Environment 2001, Table 10.
kT = kilotonnes = 1,000 tonnes
Figure 10.1:
Actual and projected funding for the Energy Efficiency and Conservation
Authority (EECA), 1993/94 to 2005/06 (GST inclusive, in $1,000, nominal (not
adjusted for CPI))
12500
10000
7500
5000
2500
0
1993/ 1994/ 1995/ 1996/ 1997/ 1998/ 1999/ 2000/ 2001/ 2002/ 2003/ 2004/ 2005/
94
95
96
97
98
99
00
01
02
03
04
05
06
Grants schemes
0
Crown Energy Efficiency loans
0
0
310
1850 4000 2500 2000 2000 2000 2000 2000 2000 2000
2900 2000
850
1000 1000 1000 1000 1000 1000 1000 1000 1000
758
498
539
Other revenue
469
Baseline appropriations
4366 5631 6984 6641 5644 4996 4640 7089 6565 6665 6665 6665 6665
691
567
563
356
389
391
391
391
391
Source: E. O’Connor, EECA, pers comm Note that the figure for 2000/01 will not match that in the Estimates of Appropriations 2001,
as additional funding was approved in November 2000. Projected funding is based on current Government policy as conveyed to EECA.
12000
11000
10000
9000
8000
7000
2000/01
1999/00
1998/99
1997/98
1996/97
1995/96
1994/95
6000
1993/94
Figure 10.2:
Total real funding for EECA
1993/94 to 2000/01 (adjusted for
CPI, in current dollar terms)
Total funding per year from Figure 10.1, adjusted using CPI to
2001 dollars, on a March year basis.
101
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 10.2: EECA’s Key Output objectives and budget for 2000/01
Description
Cross-sectoral
Budget
$’000
(Excl
GST)
2,392
Finalise the National Energy Efficiency and Conservation Strategy, including the report analysing
submissions on the draft and recommendations to the Minister of Energy
Establish and implement a methodology to measure New Zealand’s energy efficiency
Continue to gain a detailed understanding of New Zealand’s energy efficiency performance and energy
efficiency potential
Raise awareness of energy efficiency in general, and EECA, through a combination of activities
including publication of a ‘flag ship’ magazine
ƒ
ƒ
ƒ
ƒ
Energy supply
887
Develop and implement appropriate measures to promote greater uptake of renewable energy sources
Contribute to energy supply policy, including the gas sector review and Electricity Governance Board
process
ƒ
ƒ
Industry
865
Provide technical support, as appropriate, for Government in their development of Negotiated
Greenhouse Agreements
Undertake ‘business’ commitment programmes with three target audiences, business (Energy Wise
Companies), central (GEELP) and local government (Energy Wise Councils) supported by a range of
services including Crown loans and information services
Make energy management a mainstream practice through the development of a growth strategy
addressing both demand for energy efficiency service and fostering a market to meet the demand
ƒ
ƒ
ƒ
Buildings and Appliances
1,316
Implement mandatory, minimum energy performance standards (MEPS) and mandatory labelling
Administer the Energy Saver Fund (ESF) to provide a range of domestic retrofitting projects –paying
particular attention to particular target audiences and management of the funding to maintain continuity
across financial years for service providers
Initiate a review of the recently enacted H1 (energy efficiency) elements of the Building Code and work
with key elements of the new building industry to raise awareness of energy efficiency
ƒ
ƒ
ƒ
Continue research into current energy use patterns within homes and new research to start quantifying
the health effects of energy efficiency measures
ƒ
Transport
•
•
•
Implement and promote the fleet management guidelines
Undertake a review of current transport energy use patterns with a view to determining priorities for
future transport activities
Develop proposals for fuel efficiency information for new and imported light vehicle purchasers – which
may also include consideration of energy efficiency standards for light vehicles
Promote travel demand management proposals, including walking school buses and rideshare software
•
Begin to trial demand management initiatives being developed by others during the plan year
•
Total
929
6,389
This work programme includes EECA management of the Crown Energy Efficiency Loan Scheme. ($1,000,000 -GST not
applicable) which is available to publicly funded bodies, and Energy Efficiency Grants ($2,000,000 - GST not applicable) which
provide energy efficiency and renewable energy assistance to selected groups.
Source: EECA, 7/2001.
Note: “Walking school buses” (transport section) are when caregivers take turns walking with groups of children to school,
rather than each family driving separate cars.
102
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 10.3 : Summary of the initiatives proposed in the Draft National Energy Efficiency and
Conservation Strategy, 2001.
Sectors
Proposed objectives
Types of measures proposed
Government and
Local
Authority
1.
•
2.
Buildings
1.
2.
3.
Industry
1.
2.
Transport
1.
2.
3.
Leadership from central and local
government with a sector target
of 15% energy intensity reduction
from their own operations in five
years.
Integration of sustainable energy
outcomes with the goals,
objectives, statement of intent,
and planning processes of all
arms of central and local
government, including actions to
achieve the Strategy targets.
Leadership programmes Implement such programmes for
central and local government energy efficiency, develop
statements of intent, facilitate community projects.
•
Education Over the longer term, facilitate programmes for
schools, trade training establishments, and the wider public on
energy efficiency and renewable energy.
•
Pricing Develop pricing and taxes to support sustainable
energy (longer term - part of current tax review).
•
Planning and Resource Management Develop guidelines
under RMA for solar orientation and renewable energy
projects. Longer term: encourage prominence of energy issues
in plans under RMA, regional scale energy “accounting”, waste
as an energy resource.
Progressively upgrade the
energy performance across all
sectors of the existing building
stock.
Achieve “best practice” energy
performance in new residential
and commercial buildings.
In 15 years, existing and new
residential and commercial to be
retrofitted or constructed to
higher standards.
•
Progressive energy efficiency
improvement in all industry subsectors to meet international best
practice, industry by industry.
Greater utilisation of renewable
energy potential. In the short- to
medium-term this will be focused
on woody biomass from forest
waste.
•
Voluntary commitment Develop “Negotiated Greenhouse
Agreements” with major energy-intensive industries, and new
“Business Commitment” programmes for small & medium sized
industry.
•
Financial assistance Provide grants for energy audits,
investigate tax concessions & other measures. Longer term:
implement them.
•
Generic technologies standards, promotion Establish a
Minimum Energy Performance Standard for electric motors;
develop new “challenge” programmes. Longer term: further
standards and labelling.
•
Information & research Undertake sector studies. Longer
term: international benchmarking and research on woody
biomass collection & utilisation.
•
Industry training Investigate industry training needs &
opportunities. Longer term: implement industry training support.
•
Energy efficiency market promotion Longer term:
possibly promote ESCOs (energy service companies), efficient
shared and multi-site energy efficiency projects.
Reduce energy use through
travel-demand management.
Increase the use of more energyefficient and eco-efficient
vehicles and fuels.
Improve the provision and uptake
of low energy transport options.
•
Transport demand-reduction Encourage demandreduction trials (carpooling, tele-working); develop policies
consistent with links between transport, energy efficiency and
urban form.
•
Pricing Improve effectiveness of funding for alternatives to
roading; continuing policy development. Longer term: develop
road pricing and pro-efficiency trials.
•
Eco-efficient vehicles, other fuel options Facilitate ecoefficient vehicles in public fleet; monitor new technologies to
guide policy; investigate & develop vehicle efficiency standards.
Longer term: provide efficiency information (labelling), private
sector fund for eco-efficient vehicles, possibly implement
vehicle efficiency standards.
•
Energy efficient modes Provide information and
supportive policies (e.g. greater funding for public transport,
walking paths, and cycling facilities; more direction and advice
in national transport strategy). Longer term: provide more
explicit support policies & mechanisms to recognise energy
efficiency advantages of coastal shipping and rail.
Information Design guides for mass and glazing
optimisation; energy efficiency rating and labelling schemes;
building energy usage data; longer term, public information and
industry skills upgrading.
•
Standards Promulgate “better” and “best” energy efficiency
design practice for residential and commercial buildings,
upgrade Building Code Clause H1. Longer term: extend to
insulating standards, building design support.
•
Implementation support: Continue and redesign
residential assistance programmes, upgrade Housing NZ and
Government buildings. Longer term: develop commercial
building incentives.
continued next page
103⇒⇒
The greenhouse effect and climate change
Parliamentary Library, August 2001
Table 10.3, continued.
•
Energy efficient road networks & traffic management:
Reinforce the importance of better traffic-demand management
to improve energy efficiency, particularly in high-volume areas.
•
Education and information Run an energy-efficient fleet
management programme; publicise energy efficient driving
practices, vehicle choice, and vehicle maintenance.
Transport
continued
Energy
Supply
1.
2.
3.
Increase the amount (actual
supply and % market share
above business as usual) of
supply from renewable energy
source over time.
Improve whole-system
efficiencies of the energy supply
sector.
Improve the institutional
arrangements within the energy
supply sector so prices to energy
consumers consistently support
sustainability outcomes.
•
Electricity sector Establish appropriate incentives/ rules
through Electricity Governance Board; ensure policy reflects
good understanding of distributed generation & demand-side
management; investigate ways to reduce network energy
losses; ascertain support for energy-efficiency pricing;
investigate “competition by comparison”. Longer-term:
implement results of investigations.
•
Gas sector Ensure that Government’s review of the gas
sector includes issues of network expansion, market
development and pricing. Longer term: implement findings of
review.
•
Renewables Provide guidelines to local authorities; do
regional studies of renewable energy resources (including iwi
interests); facilitate use of wood waste as energy in forestry
processing; evaluate means to increase use of renewables in
electricity generation; review research funding. Longer term:
work with local authorities to get energy efficiency and
renewables into RMA Plans, implement findings & appropriate
mechanisms.
•
Industry development Support relevant industry
associations; develop an action agenda on industry
opportunities; develop support mechanisms for solar water
heating industry; investigate transport biofuels opportunities.
Longer term: implement findings & appropriate mechanisms.
•
Emerging technologies Locate and disseminate
information on fuel cell and hydrogen technologies; identify and
longer term implement alternative pathways for increasing use
of hydrogen energy technology.
Source: Energy Efficiency and Conservation Authority 2000, Draft Energy Efficiency and Conservation Strategy.
10.3
Energy efficiency planning by government agencies
The Government Energy Efficiency Leadership Programme (GEELP) was instigated by EECA in
October 1993. From 1992-93 to 1998-99, the programme contributed $550,000 to annual energy
cost savings, annual returns on investment of 40%, and cumulative cost savings to Government
of $1.7m. As at June 1999, there were 33 member agencies in the programme.
A comprehensive analysis of agency commitment to energy efficiency in 1999 showed a wide
range, with the National Library of New Zealand, Ministry of Education and Inland Revenue
being the three best, and Internal Affairs and the Department of Prime Minister and Cabinet
being the worst (Figure 10.3).
In 2000, a new programme evolved from this effort. Energy-Wise Government was set up with a
goal of 15% improvement in energy efficiency across the core public sector over 2000-2005.
Agencies that sign voluntary agreements to enter this scheme agree to appoint energy
managers, conduct detailed energy audits, implement all practical and cost-effective energyefficiency opportunities, and encourage building owners to improve energy efficiency where
agency accommodation is leased.3 As of August 2001, the newly launched programme has 21
agencies as members.
In 2000, Parliamentary Service won the public sector Energy-Wise Award. Energy efficiency
activities in the Parliament Buildings are summarised in section 3.4.
3
EECA, Energy Efficiency Agreement.
104
The greenhouse effect and climate change
Figure 10.3
Parliamentary Library, August 2001
Ratings of government agency energy-efficiency policy, management, monitoring,
staff training, and funding, 1999. Based on evaluation of 25 key performance indicators.
Most agencies are expected to achieve a minimum score of 30.
National Library of NZ
Ministry of Education
Inland Revenue Dept.
Dept. of Social Welfare
Ministry of Health
Ministry of Youth Affairs
Ministry of Commerce
Dept. of Conservation
Ministry of Agriculture
Ministry of Women's Affairs
Parliamentary Service
Reserve Bank of NZ
Audit New Zealand
Dept. for Courts
Ministry of Foreign Affairs and Trade
Crown Law Office
Ministry of Housing
NZ Customs Dept.
Te Puni Kokiri
Ministry of Transport
State Services Commission
Ministry of Research, Science and Technology
The Treasury
Ministry for the Environment
Statistics NZ
Office of the Auditor-General
Dept. of Corrections
Public Trust
Dept. of Labour
Land Information NZ
Dept. of Prime Minister and Cabinet
Dept. of Internal Affairs
15
30
45
energy efficiency policy performance indicator score
Data not reported for Education Review Office, Ministry of Fisheries and Serious Fraud Office.
Source: Energy Efficiency and Conservation Authority
105
The greenhouse effect and climate change
10.4
Parliamentary Library, August 2001
Public awareness and concern
In March 2001, UMR Research completed a study of public awareness and level of concern
about climate change commissioned by the New Zealand Climate Change Programme. The
study included a nationally representative telephone survey of 750 people aged 18 and over,
and four “focus groups” (number of people not specified). The margin of error for the telephone
survey was ± 3.5%.
Recommendations from analysis of the focus group results included:
•
•
•
•
using the term “global warming” rather than “climate change”;
clarifying the difference between the ozone layer and global warming;
providing viable options for people to make greenhouse-friendly choices (e.g. availability of
public transport); and
making the economic implications of policies clear.
As the press release and media reported only a very brief summary of the results, more detailed
summary is presented in Table 10.4.
106
The greenhouse effect and climate change
Table 10.4:
Parliamentary Library, August 2001
Aggregate quantitative results from the UMR telephone survey of public
awareness and concern about climate change
Question
“How much would you say you knew about the issues involved in
global warming?” (a lot, a fair amount, not that much, hardly anything)
“Have you heard of the Kyoto Protocol?”
“Natural weather cycles which have made the world hotter and
colder for tens of thousands of years are more important in
determining climate than anything people do.“
(Scale of 1 strongly disagree to 7 strongly agree)
“How much do you know about the New Zealand Government’s
response to global warming?”
(a lot, a fair amount, not that much, hardly anything)
“There is nothing a small country like New Zealand can do about
global warming.”
(Scale of 1 strongly disagree to 7 strongly agree)
“New Zealand should take an international lead on reducing global
warming.”
(Scale of 1 strongly disagree to 7 strongly agree)
“ I am prepared to pay a little more and put up with some
inconvenience to help the environment.”
(Scale of 1 strongly disagree to 7 strongly agree)
“How interested are you in finding out more about global
warming?”
(very interested, fairly interested, not very interested, not interested at all)
“How would you like to get this information?”
Grouping with
largest response
Percent
“a lot” + “a fair
amount”
63%
“no”
7 “strongly agree”
+ 6 agree
1 “strongly
disagree” + 2
disagree
60%
“not that much” +
“hardly anything”
22%
19%
84%
7 “strongly agree”
+ 6 agree
1 “strongly
disagree” + 2
disagree
15%
7 “strongly agree”
+ 6 agree
47%
7 “strongly agree”
+ 6 agree
46%
“very interested” +
“fairly interested”
76%
“television”
42.6%
54%
“Thinking about environmental issues facing New Zealand, how serious do you think the following
issues are?” (scale of 1 to 7 where 1 means not serious al all and 7 means they are extremely serious)
New diseases being established in New Zealand
Establishment of foreign pests such as spiders, ants, and
mosquitoes in New Zealand
The hole in the ozone layer
61%
7 “extremely
serious” +
6 quite serious
60%
58%
Pollution of lakes and rivers
55%
Global warming
52%
Waste disposal
48%
Radiation from cell phones and cell phone sites
18%
Source: UMR 2001.
107
The greenhouse effect and climate change
108
Parliamentary Library, August 2001
11
Local authority initiatives
Local authorities can actively encourage reduction of local greenhouse gas emissions, increase
in carbon sinks, and adaptation to the climate and sea level changes through their leadership,
service provision and consent-granting roles. Regional and district councils can influence
energy consumption patterns in such areas as urban design (and therefore transport demand)
and activities requiring resource consents under the Resource Management Act. Reduction in
the energy intensity of council services will reduce greenhouse gas emissions and can also save
ratepayers money.
The role of local government in meeting New Zealand’s climate change target is currently under
consideration by Parliament’s Local Government and Environment Select Committee, which
issued an interim report in December 2000 and called for submissions by 15 March 2001. Its
interim recommendations are presented in Chapter 3 (section 3.2.1), and its final report was in
preparation as of the end of August 2001.
The Energy Efficiency and Conservation Authority (EECA) runs an Energy-Wise Councils
Partnership Programme, which involves councils pledging energy efficiency targets and provides
access to the Crown Energy Efficiency Loan Scheme and a wealth of supportive information.1
Local authority members of the programme in 2000 were the Auckland, Christchurch, Hamilton,
Nelson, Waitakere, and Wellington City Councils, Auckland Regional Council, and Environment
Canterbury.
An international coalition of municipal authorities, Cities for Climate Protection Campaign, is
actively pursuing greenhouse gas emission reduction targets. Its website can provide access to
a model strategy, greenhouse gas auditing software, and other information to support councils
wishing to take strong local action. New Zealand members of this coalition are the Hamilton,
Waitakere, and Wellington City councils and the Waikato Regional Council. The Climate Alliance
of European Cities is a collective of 900 cities with more ambitious reduction targets than the
Kyoto Protocol or their national governments, and reductions of 25% in energy use over a
decade or less have been common. The Energie Cités network includes sustainable initiatives in
over 150 European municipalities. 2
11.1
Legal context
Four principal Acts pertain to the powers of local authorities to encourage reduction in
greenhouse gas emissions in New Zealand.
•
Local Government Act 1974 (e.g. energy efficiency of council services, waste
management, management of parks and reserves)
•
Building Act 1991 (e.g. Building Code and supplementary guidelines)
•
Resource Management Act 1991 (e.g. plans and consents)
•
Transit New Zealand Act 1989 (e.g. transport plans and funding)
11.2
Local authority operations
Central and local government authorities together consume about 2% of New Zealand’s energy
in their operations. In 2000 it was estimated that councils use about $75 million per year of
energy to provide or contract for services to ratepayers. Councils can have a direct role to play in
improving the energy efficiency of such facilities as buildings, street lighting, recreation facilities,
waste treatment plants, and vehicle fleets.3 The Local Government Act 1974 lists the “efficient
and effective exercise of the functions, duties and powers of the components of local
government” as a purpose of local government (s 37K(h), emphasis added).
1
http://www.eeca.govt.nz/default.asp
International Council for Local Environmental Initiatives on http://www.iclei.org/co2/co2.htm ; Local Government and Environment
Select Committee 2000, p.24.
3
Energy Efficiency and Conservation Authority 2001, p. 12; and government section of http://www.eeca.govt.nz/default.asp
2
The greenhouse effect and climate change
Parliamentary Library, August 2001
Central Government has adopted the goal of a 15% reduction in energy use by the end of 2005,
and invited local authorities to join them. This goal is not difficult to achieve if there is a will to
succeed and an energy manager is appointed, as shown by Christchurch City Council, which
has reduced its energy intensity by 25% (Box 8). This council also reduced the cost to
ratepayers of council energy use from $73 to $62 per household over five years, despite rising
energy prices.4
As managers of their own parks and reserves, district and regional councils have a significant
role in protecting and enhancing carbon sinks in the form of vegetation. Councils can also
enhance biomass on other lands through partnerships with the community and agencies such as
the Department of Conservation and the Queen Elizabeth II National Trust.
11.3
Roading and transport
Domestic transport is the largest and fastest growing contributor to New Zealand’s CO2
emissions (Figure 4.9). District and regional councils have a strong role to play in the design
and provision of local transport facilities, as well as the design of urban form which can reduce
the demand for car transport.5
A well-utilised public transport system requires significantly less fossil fuel for passenger
transport than car-based alternatives. Local authorities can assist in the necessary co-ordination
between central government (funding), regional councils (funding and planning), private
transport operators (service provision) and road controlling authorities (priority road space and
ancillary services) to find cost-effective alternatives to more roading. However, the adequacy
and flexibility of Transfund New Zealand funding for fuel efficient alternatives is a matter of
debate.6
In 1998 the Energy Efficiency and Conservation Authority (EECA) sponsored two seminars on
the role of local authority policies in changing urban form as a way to improve transport energy
efficiency, reduce CO2 emissions, and address other transport problems in New Zealand.7 EECA
also compiles a two-monthly e-mail newsletter to support the national Sustainable Transport
Network.
Use of fossil fuels can be reduced through improving public transport, cycling and walking
facilities, controlling car parking to create a disincentive for car commuting to the inner city, and
redeveloping residential uses in the inner city. In Wellington, the recent growth in this type of
accommodation has reduced the number of car trips to the CBD by about 290,000 per year.
Land Transport Strategies in Auckland and Canterbury regions actively promote more energy
efficient travel modes such as public transport, walking and cycling.4
A partnership between EECA and Lincoln University has produced successful internet-based
rideshare software to support carpooling. The Lincoln rideshare programme has reduced CO2
emissions by 136 tonnes per year, petrol use by over 68,000 litres per year, and increased
carpooling from 25% to 36% of journeys to and from the university. The software has now been
obtained by the Wellington Regional Council and Auckland City Council, as well as two
educational establishments and a major private sector employer.8
4
http://www.eeca.govt.nz/content/ew_government/councils/members.htm
For example, services and housing spread far apart without adequate public transport infrastructure requires greater use of private
vehicles and more fossil fuel consumption.
6
Local Government and Environment Select Committee 2000, p. 12.
7
Copies of the proceedings are available from EECA.
8
http://www.lincoln.ac.nz/rideshare ; EECA Annual Report 2000, p. 13 and http://www.eeca.govt.nz . Other bodies that had signed
the licence agreement for the software as of 2000 were the University of Canterbury, Fisher & Paykel Industries Ltd. in Mosgiel, and
the Eastern Institute of Technology in Napier.
5
110
The greenhouse effect and climate change
11.4
Parliamentary Library, August 2001
Building codes and energy conservation
For many years there have been attempts to improve the energy efficiency requirements in New
Zealand’s National Building Code, which is enforced by district councils. New changes to clause
H1 of the Code finally came into effect in 2001 (see section 9.7, p. 92).
In addition to enforcing national standards, district councils can help to disseminate information
about “eco building”, and thus encourage optimal energy efficiency of local buildings. Such
buildings will continue to exert a positive influence on energy efficiency for decades to come.
Waitakere City Council have developed the Sustainable Home Guidelines and a Better Building
Code for clauses in commercial building specifications.9 The Auckland Regional Council and
Hamilton Regional Council have collaborated with the Building Research Association of New
Zealand (BRANZ) to produce the Easy Guide to Eco-Building.10
Other guidelines available for councils to direct clients to include Design for the sun: residential
design guidelines for New Zealand available from EECA and Designing comfortable homes:
guidelines on the use of glass, mass and insulation for energy efficiency available from the
Cement and Concrete Association of New Zealand. BRANZ have certified architects in the
Auckland, Hamilton, and greater Wellington regions as Greenhome Accredited Assessors as
sources of expertise in this area.11
11.5
Waste management
Organic materials going to landfills and wastewater plants are a significant source of methane. In
1999, landfills in New Zealand emitted an estimated 117,430 tonnes and wastewater handling
added another 6,760 tonnes of methane, which is 21 times more powerful as a greenhouse gas
than CO2.12 The majority of these facilities would have been under the control of a district
council.
The two main ways that a council can influence these methane emissions are reducing the
amount of organic material going to waste and capturing the methane for use.
The amount of organic materials going to landfill can be reduced by such methods as: municipal
composting; regulations banning greenwaste in the landfill and thus encouraging private
composting operations; and encouragement of home composting (such as through education
and availability of at-cost bins).
Ways to reduce the volume of sewage requiring treatment include encouragement of
composting toilets and retention of septic tank systems (primarily in rural and semi-rural areas).
Methane can be captured by designing or retrofitting landfills and wastewater plants to collect
methane, and the gas can be made available for direct heat, electricity production, or running
vehicles.
Recycling of other waste materials (e.g. paper, glass, metal, plastic) reduces the energy required
to produce consumer goods by reintroducing already extracted materials into the production
cycle. Even for products where the fossil fuel emissions from extraction and processing of
materials occur largely overseas, the impact on global climate change is still relevant. The
benefits of recycling paper are further explored in the next chapter (Figure 12.2)
District councils that have undertaken such initiatives in New Zealand include Auckland,
Christchurch and Wellington (Box 8).
9
http://www.waitakere.govt.nz/ecocity/frameset.htm
Available via BRANZ at http://www.branz.org.nz/branz/resources/ecobook.pdf
11
http://www.cca.org.nz ; http://www.branz.org.nz/branz/resources/greenhomeassessors.htm
12
Ministry for the Environment 2001, Table 10 sheet 2. Greenhouse warming potentials (CO2 vs. CH4); see Table 4.1.
10
111
The greenhouse effect and climate change
11.6
Parliamentary Library, August 2001
Resource Management Act consents
As grantors of land use and land disturbance consents under the Resource Management Act
(RMA), councils have scope to minimise soil disturbance and loss of soil cover. The less
disturbance to soil cover, the greater the retention of organic carbon in the soil and therefore the
fewer CO2 emissions from local land-use. Likewise, district and regional plans may provide
protections for forests and greenbelt areas which can serve as carbon reservoirs as well as
community and ecological amenities. Encouragement of composting and mulching as part of
waste reduction policy (see previous section) also enhances the carbon sink capacity of the soil.
Theoretically, as grantors of air discharge consents under the RMA, regional councils should
have the ability to help control local emissions of greenhouse gases. Under s 15 of the RMA, a
discharge of any contaminant to air in contravention of a rule in a regional plan without a consent
is illegal, except where a discharge can be shown to be a legal activity existing prior to the plan
(s 20). This includes both discharges from industrial and trade premises (s 15(1)(c)) and from
any place or source, movable or not (s. 15(2)). As the definition of “contaminant” includes any
gas that “when discharged to air changes or is likely to change the physical, chemical, or
biological condition of the air into which it is discharged” (s 2), it would apply to greenhouse gas
emissions.13
However, the appropriateness of local authorities to actively control greenhouse gas emissions
through air discharge consents under the RMA has been questioned. The International Energy
Agency (OECD) review of New Zealand’s energy policy in 1997 stated:
“The decentralisation of most of the responsibility for implementing the Act to the
regional authorities makes it very difficult to ensure a consistent approach to
addressing greenhouse gas emissions throughout the country. Inconsistent
decisions might, for example, lead to the site for a proposed thermal power station
being moved from one region to another to take advantage of regulatory differences.
“Tackling small and mobile sources might be possible under the Act but would
involve high transaction costs.
“If regional authorities were empowered to use economic instruments (which is
uncertain), an array of regional instruments would be less effective and more costly
to administrate than a single national instrument.
“For these reasons, the resource consent process under the Resource Management
Act is not the most appropriate mechanism for addressing carbon emissions. It
follows that the requirement for resource consents to take account of carbon
emissions, which could lead to the consent being denied or conditions being
attached to it, should be lifted.”14
There are a few examples of regional councils specifying consent conditions which mention CO2
and PFC emissions, and require consent holders to use “best technology”, limit emissions, and
report annually (e.g. Huntly and Comalco).15 The Stratford Power Station is the only example of
specific constraints on quantity of CO2 emissions and is discussed below (section 11.6.1).
13
The causal link between a person’s activity and a discharge, and whether they could control the discharge given reasonable
precautions, have been addressed in RMA case law (Brookers Resource Management, s A15.04A). RMA case law has not yet
addressed whether greenhouse gases are “contaminants” under the Act (Brookers Resource Management, ss A2.24.04-.06).
14
OECD 1996, pp. 83-85, 90-91.
15
H. Plume (MFE) pers comm 8/2001.
112
The greenhouse effect and climate change
11.6.1
Parliamentary Library, August 2001
The Stratford Power Station case16
Currently the only known example of regional council specifically controlling greenhouse gas
emissions under RMA consents is the Taranaki Combined Cycle Power Station in Stratford,
whose air discharge consents are administered by the Taranaki Regional Council. The
conditions relating to the discharge of CO2 were the result of a Ministerial “call in” under s 140 of
the RMA, a Board of Enquiry, and Ministerial decision. The call-in was initiated because of
concern that the projected CO2 emissions would significantly increase New Zealand’s national
greenhouse gas emissions (some 5% of the total).
Stratford Power Ltd are required by their air discharge permit to “avoid, remedy, or mitigate the
effects of the additional amount of CO2 being discharged as a result of this consent” up to a
maximum of 1.5 MT of CO2 per year and report to the Council and the Minister for the
Environment on how they are fulfilling these obligations. The Minister required “additional” to be
defined in relation to the net effect on national electricity sector emissions normalised for the
average hydrological year, rather than emissions from the plant only. This occasions a
significant delay in reporting as national data is collected and analysed.
In contrast, the consent conditions recommended by the Board of Enquiry would have required
ECNZ to establish some 4,000 ha. of forest per year for the 34 year life of the plant. This would
provide a carbon sink to mitigate CO2 emissions. It was also proposed that ECNZ maintain the
resulting 136,000 ha. forest estate in perpetuity with harvesting and replanting on a regular
schedule.17
The plant commenced operation in February 1998, and has emitted over 1,548,410 tonnes of
CO2 since that time. The base year against which the net electricity sector emissions are
measured was higher than the subsequent two years of plant operation, so that no mitigation
measures were required by the consent for those years (Table 11.1). One reason for this may
be that the Taranaki Combined Cycle Power Station has significantly displaced use of the less
efficient and coal-burning Huntly Power Station.
Table 11.1:
CO2 emissions from the Taranaki Combined Cycle Power Station, electricity
sector emissions 1998-2000, and mitigation measures required under Resource
Management Act consent
Base year (2/97- 2/98)
1998-99
1999-00
CO2 emissions from
the Taranaki
Combined Cycle
Power Station
tonnes
-704,605
843,805
CO2 emissions from
the electricity sector
(normalised for average
hydrological flow)
tonnes
8,489,446
3,929,351
4,165,592
mitigation of CO2
emissions required
by the RMA consent
-none
none
Sources: Taranaki Regional Council Stratford Power Ltd Combined Cycle Power Station Monitoring Programme Annual Report
for 1998/99 and 1999/00; Stratford Power Ltd. Consent Compliance Report for 1998, 1998/99, and 1999/00.
16
Sources for this section: McSoriley 1995; Taranaki Regional Council Stratford Power Ltd Combined Cycle Power Station
Monitoring Programme Annual Report for 1998/99 and 1999/00; Stratford Power Ltd. Consent Compliance Report for 1998,
1998/99, and 1999/00; Hamilton 2000.
17
McSoriley 1995, p. 2.
113
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 8
Examples of local authority initiatives:
reducing greenhouse gas emissions
NEW ZEALAND
Auckland Regional Council
Active promotion of increased use of more energy efficient travel modes such as public
•
transport, walking and cycling through their Regional Land Transport Strategy.
The Buses First programme speeds up public transport and attracts passengers: includes
•
provision of bus lanes, traffic signal pre-emptions.
EECA rideshare software purchased to facilitate local carpooling.
•
Co-sponsorship of Easy Guide to Eco-Building.
•
City-wide recycling programme.
•
Encouragement of composting plants at landfills and home composting.
•
Hamilton City Council
Initiation of EnviroSchools, a model Council – schools partnership for environmental education.
A focus is whole-of-school life, with energy use one of the key elements.
Co-sponsorship of Easy Guide to Eco-Building.
•
•
Christchurch City Council
Appointed energy manager and increased the energy efficiency of its own operations (over five
•
years energy use has fallen 25% and energy bills reduced $2 million per year).
Major greenwaste composting plant, active encouragement of household composting .
•
•
Use of methane from sewage treatment plant for running council vehicles.
City-wide household recycling collection programme.
•
Environment Canterbury
Encourages low energy modes of transport (public transport, walking and cycling) and
•
advocacy of environmentally friendly vehicles through the Regional Land Transport Strategy.
Waitakere City Council
Publication of Sustainable Home Guidelines for the public and building professionals.
Preparation of Better Building Code guidelines for public buildings, including energy efficiency.
Preparation of the Green Print Purchasing Guidelines to encourage the reduction of
environmental impacts of printing (paper, energy and other resources).
•
•
•
•
Wellington City Council
Initiation of bus-only lanes to speed inner-city bus transport and improved pedestrian facilities
•
to encourage walking.
Shredding of all greenwaste from city street trees to create mulch, co-sponsor of major
•
greenwaste/biosolids composting plant, composting at both landfills.
City-wide household recycling collection programme, subsidised by user-pays rubbish
•
collection.
Collection of methane from the landfill.
•
Wellington Regional Council
Support of electric trolley buses and electric trains for public transport (vs. diesel).
Purchase of EECA rideshare software to facilitate local carpooling.
Promoting the use of renewable energy sources and recovery of landfill gas.
•
•
•
continued, next page
114
The greenhouse effect and climate change
Parliamentary Library, August 2001
(Box 8, continued from previous page)
OVERSEAS
Barcelona, Spain – solar energy
A housing regulation was adopted in 1999 to require provision for installation of solar thermal collectors on new
constructions and retrofitted buildings.
Copenhagen, Denmark – urban design
Over a number of decades urban growth has been managed along public transport corridors on development
and transport nodes with green open spaces in between, in conjunction with a strong emphasis on cycleways
and heavy taxes on cars.
Denver (Colorado), USA – vehicle emission reductions
Alternative Fuels Ordinance requires anyone owning a fleet of more than 30 vehicles to convert 10% of their
fleet to clean-burning fuels. In 1997, 141 fleets complied. Ordinances on smoking and idling vehicles reduce
emissions from individual cars through a citation system. City’s own Green Fleets programme has reduced CO2
emissions by 13% over 5 years.
Dortmund, Germany – wind energy
The municipal utility successfully sold bonds to the local community to finance a new 500kW wind turbine within
the city limits.
Portland (Oregon) and Davis (California), USA
Creation of an energy-efficient city through planning controls: more compact development, guiding growth to
more energy-efficient locations, active encouragement of public and non- motorised transport, and spatial layout
to achieve energy efficiency.
San Diego (California), USA – telecommuting
Telecommuting programme for 200 city employees, who telecommute (work from home with telephone and
electronic links) on average one day a week. Vehicle emissions for these workers have been reduced by 6373%. Direct benefits outweigh costs 5:1.
Stockholm, Sweden – methane use in vehicles
A partnership of public and private parties has resulted in methane from the local wastewater treatment plant
being used to power 200 dual petrol/biogas cars in the city.
Sources: Energy Efficiency and Conservation Authority 2001, p. 13; http://www.waitakere.govt.nz ;
http://www.iclei.org/co2/co2.htm ; Energie Cités http://www.agores.org/Publications/CityRES/fichesgblior.pdf ; Database of
Municipal Success Stories http://www.pembina.org ; Controller and Auditor-General 2001, p. 101.
115
The greenhouse effect and climate change
116
Parliamentary Library, August 2001
12
Individual choices
“What can I do, as just one person?”
Climate change is a global long-term problem, and actions of individuals can seem irrelevantly
small.
However, it was also actions of individuals, all over the world over the last 150 years, that
created the problem. It will be the cumulative actions of individuals around the world that will
either make things worse, or reduce greenhouse gas emissions to an appropriate level.
Every little action adds up, especially if you encourage your friends, family and neighbours to
join you.
Everything has an energy component
Every consumer and transport choice has energy implications, especially in developed countries
like New Zealand.
Energy is “embodied” in all consumer goods and services, before you even buy and use them,
through the energy it takes to extract and process raw materials and transport the finished
product to you. Products designed to be thrown away have higher embodied energy that those
designed to be reusable. It has been estimated that two-thirds of the “embodied” energy comes
from choices that the industrial sector makes, and a third can be controlled directly by consumer
choices.1 Some industrial choices can also be indirectly influenced by consumer demand.
The fuel consumption by appliances and transport are related both to their design (how
efficiently they use energy) and how you use them.
Any energy that comes from fossil fuels involved in these equations contributes to climate
change. In New Zealand virtually all of our transport and 27% of our electricity comes from fossil
fuels. For imported consumer goods, the majority of embodied energy is likely to be from fossil
fuels.
Soil, vegetation, and compost are carbon sinks
Many of us in New Zealand have control over a patch of land, and all of us have the option to get
involved in resource management issues in the wider community.
The greenery and soils in parks, forests, farms and home gardens can act as sinks for CO2.
Maintaining greenery, using composts and mulch, and choosing to compost kitchen and garden
scraps, either at home or through a local composting plant, not only enhances the local carbon
sink potential, but recycling organic waste also reduces emissions of methane from the local
landfill.
Choosing to buy organic and sustainably farmed food supports people who are building the
carbon reservoir potential of farmed soils. It also reduces the demand for pesticides and other
agrochemicals which have a high fossil-fuel embodied energy.
The following boxes summarise some of the actions recommended by agencies and experts to
help reduce your personal contribution to climate change.
1
US Environmental Protection Agency, http://www.epa.gov/globalwarming/emissions/individual/index.html
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 9: Individual actions - TRANSPORT
3 The average family car produces about 2.2 tonnes a year of CO2 3
3 Every litre of petrol saved reduces greenhouse gas emissions by 2.5 kg 3
your choices count!
“The Seven Habits of Highly Efficient Drivers”
Energy Efficiency and Conservation Authority
1. Avoid unnecessary driving
One-third of car trips in New Zealand are under 2 km and two-thirds are under 6 km.
To reduce the number of car trips, you can use public transport, walk or cycle, car pool, and plan ahead.
2. Drive with a smile
Less aggressive driving can improve fuel economy by up to 30%.
Accelerate smoothly, look ahead and avoid heavy braking and acceleration, keep to the speed limit, time
your trip to avoid road congestion.
3. Maintain your vehicle
Better vehicle maintenance can improve fuel consumption by 10-20%.
Keep tyres inflated to correct pressure, have wheel alignment checked regularly, ensure that engine
timing, spark plugs, and air filter are checked and maintained regularly.
4. Keep vehicle loads to a minimum
An extra 50 kg increases fuel consumption by 2%, and wind drag from roof racks, windows, and sunroofs
can use 5-10% more fuel. Remove unnecessary loads and roof racks not in use.
5. Turn off the extras
Air conditioners can add 10% to fuel use and rear screen demisters 3-5%. Use air vents instead when
possible.
6. Switch off the engine if idling for more than 30 seconds
Allow time to restart your engine, rather than leave it idling unnecessarily.
When starting from cold, drive off immediately, but be light on the accelerator for the first few minutes.
7. Longer term – choose the right vehicle
Choose a vehicle that suits your needs. Find out the fuel economy of models you are considering.
Generally the smaller the engine capacity, the more fuel efficient the vehicle. Large 4WDs do not have
good fuel economy and are not good commuter or inner city vehicles.
Plus
Consider dual-fuel, electric, bio-fuel and fuel cell vehicles as they become available on the market.
Source: brochure from Energy Efficiency and Conservation Authority (EECA).
Average family car emission estimate from EECA Annual Report 1998, p. 9. The Australian estimate is 6 tonnes of greenhouse gas
per year for the average family’s travel (may include other greenhouse gases and non-family car travel) (Australian Greenhouse
Office, Global Warming – Cool it ! , p. 20).
118
The greenhouse effect and climate change
Parliamentary Library, August 2001
Bike or walking
Figure 12.1:
Greenhouse gas
emissions from
different forms of
transport
0
Extra person on public transport
0.033
1.6 litre car, 4 people
0.05
0.08
4 litre car, 4 people
1.6 litre car, driver only
Source: Australian
Greenhouse Office, p. 21.
0.2
4 litre car, driver only
0.32
0
0.1
0.2
0.3
kg per person per km
Box 10:
How many trees do I have to plant
to absorb the carbon emitted by my car?
Assuming:
you drive 40 years and 16,000 km per year;
•
your vehicle has a fuel efficiency of 11 km per litre of petrol; and
•
the petrol produces 0.86 kg carbon per litre; then
•
your lifetime carbon “footprint” for car transport would be 50 tonnes of carbon,
or 1.25 tonnes/year.
3 Half of a hectare of production radiata pine forest
A “steady state” forest of radiata pine in New Zealand on a 30-year harvest cycle contains
about 112 tonnes of carbon/ha. You would have to establish the trees on non-forested land
(e.g. pasture) and trees would have to be replanted after every harvest in perpetuity.
3 One third of a hectare of protected native forest
Mature native forest contains at least 150 tonnes of carbon per hectare. This forest would
need to be newly established or allowed to regenerate, and protected in perpetuity.
Protecting already existing native forest will prevent loss of that existing carbon reservoir, but
will not compensate for the new CO2 emissions from your car.
Planting a tree does not remove carbon permanently, as the tree will eventually die or be
harvested. Planting a forest can be permanent if it is kept in sustainable harvest or
protected. However, it only provides a one-time benefit, whereas emissions are ongoing.
Source: Maclaren 2000, p. 56.
119
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 11 : Individual actions - IN THE HOME
3 The average home creates more CO2 than the average car 3
3 70% of New Zealand houses were built before insulation became mandatory 3
your choices count!
3 hot water (the largest user of energy in the home)
‰
‰
‰
‰
‰
‰
‰
Turn the thermostat on your water heater down to 60 C.
Use cold water cycles for washing clothes.
Use short showers rather than long showers or a bath when possible.
Install a low-flow shower head (yours is inefficient if it takes less than a minute to fill a 10 litre bucket).
If you do not have a Grade A hot water cylinder (look for the "watermark" label), install a cylinder wrap
(on average 42% and up to 70% of the energy used to heat hot water can be lost without insulation).
Buy a gas, solar or heat pump type hot water heater.
Repair dripping hot water taps (one dripping tap can waste as much as 5,000 litres a year).
3 lighting
‰
‰
‰
‰
Turn off lights that won’t be needed for 5 minutes or more.
Install compact fluorescent lights in high-use areas (see lighting box, next page).
Use natural light whenever practical.
Don't light unused rooms.
3 heating and insulation
‰
‰
‰
‰
Only heat rooms that are being used.
Weatherproof your house to minimise air leaks around doors and windows.
Insulate the ceiling (up to 40% of house heating can escape through the ceiling).
Block off any chimneys not in use.
3 appliances
‰
‰
‰
‰
‰
Look for an Energy Rating label when buying appliances.
Minimise use of “standby power” (see box on next page): turn off at the wall if possible.
Avoid unnecessary appliance use (e.g. run washers and dishwashers only when fully loaded).
Dry clothes on a clothes line whenever possible.
Buy an exhaust fan with automatic shutter doors for the kitchen or bathroom (reduce heat loss).
3 food and cooking
‰
‰
‰
‰
‰
Use a microwave oven when possible (see graph next page), but for boiling water use an electric jug.
Put lids on pots, match element to pot size, simmer rather than rapidly boil, use pressure cookers.
Thaw out frozen foods before starting to cook.
If using an oven, cook more than one dish at a time; also, oven fans reduce energy use.
Buy locally and seasonally grown food (others require more energy for production and transport).
3 recycling, composting, reducing waste
‰
‰
‰
Recycle paper, glass, metal, plastics, textiles and household goods when possible.
Compost kitchen scraps and garden trimmings, and give the compost back to the soil.
Buy food and other products in packaging that is reduced, reusable, recyclable, or has recycled content.
3 house design
If designing a new home or doing alterations:
‰ Make maximum use of the sun's heat and light, and install solar water heating.
‰ Install insulation in ceiling, walls, and floor than exceeds the Building Code minimum requirements.
‰ Ensure hot water systems are efficient (minimise pipe length and install lagging for insulation).
‰ Locate homes near to work places or good public transport, to minimise transport energy.
Sources: EECA (http://www.eeca.govt.nz ), Australian Greenhouse Office 2000, Consumer 2001, US Environmental Protection
Agency (http://www.epa.gov/globalwarming/emissions/individual/index.htm), World Resources Institute (http://www.safeclimate.net ).
120
The greenhouse effect and climate change
Parliamentary Library, August 2001
Figure 12.2: Energy efficiency information: standby power, fluorescent lights, cooking modes,
and paper
“Standby power”
all those little standby lights, digital clocks, and keeping the power on for appliances, stereo systems, TVs,
VCRs, computer screens, games machines… when added together consume:
2.9% of all electricity in New Zealand
The OECD average is 10% of residential energy = a 60-watt light burning continuously in each household.
q More than 40% of microwaves in New Zealand consume more power in standby mode than in cooking food.
q More electricity is consumed when VCRs are in the standby mode than when actively recording or playing.
Energy efficiency labels don’t give standby power information.
Turn off at the wall when not required.
Source: International Energy Agency 2001, Consumer 2001.
Greenhouse gas emissions from
different forms of cooking
benefits of using
Ñ fluorescent lights Ò
Ñ
(data for same
amount of light)
regular
incandescent
75 W
750 hours
compact
fluorescent
18 W
10,000 hours
Watts used
Rated lamp life
power per
10,000 hours
750 kWh
180 kWh
electricity cost
(@0.83/kWh)
(@0.83/kWh)
for 10,000 hrs.
$62.25
$14.94
bulb costs for
(13 @ $0.75)
(1 @ $20)
10,000 hours
$9.75
$20.00
Total life-cycle
cost
$72.00
$34.94
90% of the electricity used to run a regular
incandescent light is lost as heat
Microwave
0.03
Electric
cooktop
0.07
Gas
cooktop
0.10
0.00
0.15
kg greenhouse gas from cooking
(estimated average to cook vegetables)
Source: Australian Greenhouse Office, p. 12.
Adjusted for New Zealand mix of electricity
sources (90.2% fossil fuels in Australia,
27.1% in New Zealand).
Source: Rocky Mountain Institute, http://www.rmi.org. These are USA
prices. The same calculations can be done with local prices.
Paper
Papermaking uses 4% of the world’s energy.
Every tonne of paper recycled saves about 17 trees, 4,100 kW of electricity,
26,500 litres of water, and 27m3 of landfill space.
q Producing a tonne of paper takes as much energy as producing a tonne of iron or steel.
q Making a tonne of virgin paper takes 2 to 3.5 tonnes of trees, but making
100% recycled paper takes just over 1 tonne of paper.
Conserving and recycling paper saves energy and carbon sinks.
Source: Worldwatch Institute, http://www.safeclimate.net/action (US measurements converted to metric)
121
The greenhouse effect and climate change
Parliamentary Library, August 2001
Box 12: Individual actions - AT WORK
3 Four PCs left running continuously cause as much CO2 to be emitted per
year (from generation of electricity to run them) as the average family car. 3
3 Lighting consumes 30 – 50% of energy used by buildings. 3
your choices count!
3
lighting
Turn off lights that won’t be needed for 5 minutes or more.
Install compact fluorescent or triphosphor lights and reflectors, especially in high use areas.
Use natural light whenever practical.
Clean light fittings as a regular part of maintenance (dirt can lower output by over 20%).
‰
‰
‰
‰
3
‰
computers
Switch off your computer screen if away from your desk for more than 25 minutes.
Switch off your computer overnight and over weekends.
A computer is not damaged by turning on & off, uses no extra energy to start up. Screen savers do not save energy.
Activate the Energy Star provisions in your computer.1
‰
3 printers & other equipment
Turn off when not required, especially overnight and weekends (check with IT manager if networked).
Use two-sided copying and printing. When buying new equipment, select equipment with this option.
‰
‰
3
‰
‰
‰
3
paper
Establish paper recycling programme and use recycled content paper.
Use emails and electronic filing instead of hard copy where possible.
Use two-sided printing for internal and external publications.
transport
Follow the efficient driving tips in Box 9, for commuting and fleet management.
Consider telecommuting and flexible work hours where possible.
Use the stairs rather than the lift when possible.
For vehicle fleets, have an energy management programme (information and advice from EECA).
Encourage carpools and vanpools for larger businesses (rideshare programme available from EECA).
‰
‰
‰
‰
‰
3
heating & cooling
‰
‰
Make sure the air conditioning system is turned off at night.
Don’t use personal heaters in centrally heated offices.
3
‰
‰
‰
1
management
Set energy management policy and objectives, conduct a detailed energy audit, and implement and
monitor a plan of action. Join the Energy-Wise Business programme.
Information and advice available from EECA.
Designate and adequately support an energy manager.
Return energy savings to new initiatives to help motivate staff (e.g. new facilities, donations to charity).
For Windows: Screen Saver tab in Display under Control Panel; In “Energy saving features” set low power standby to 15 minutes and
monitor at 25 minutes. For more details contact EECA or Australian EnergyStar site (http://www.energystar.gov.au )
Source: EECA publications e.g. Energy-wise tip sheets; Energy-wise Office Equipment; Improving office lighting and
reducing energy costs; Energy-wise tips for efficient building operation (http://www.eeca.govt.nz )
122
The greenhouse effect and climate change
Parliamentary Library, August 2001
Glossary
albedo
The ability to reflect light. In the climate change context, the albedo of the Earth
is higher with snow, ice or cloud cover, reflecting solar radiation instead of
allowing it to be absorbed as heat which can then contribute to the greenhouse
effect.
Annex I Parties
Countries that are listed in Annex I of the UNFCCC: this includes New Zealand
and 41 other developed countries.
anthropogenic
Caused by human activity.
assigned
amount
The amount of greenhouse gas emissions allowed for each Party to the Kyoto
Protocol during the first commitment period 2008-2012. For most countries it
equates to 1990 gross emissions. Additional assigned amount can be gained
through LULUCF carbon sinks, or lost through emissions from LULUCF, since
1990.
biomass
Living matter. In the carbon cycle context, it is the plants and animals which
contain carbon as part of their body structure. In the renewable energy context,
it is wood, alcohol, and other energy sources from plant materials.
carbon sink
A process which removes CO2 from the atmosphere where it cannot contribute
to the greenhouse effect. Sinks include growing forests, soils, and the ocean. A
carbon reservoir is a store of carbon, such as a mature forest (see Chapter 5).
carbon sink
credit
Additional assigned amount (allowed emissions) due to documented humaninfluenced carbon sequestration since 1990 in agreed land-use, land-use
change and forestry (LULUCF) activities. Subject to agreed rules, it is
anticipated that surplus credits could be sold in domestic or international
carbon trading markets.
CDM (Clean
Development
Mechanism)
A mechanism allowed under the Kyoto Protocol for developed countries to earn
credit for greenhouse gas emission reductions which occur in undeveloped
(non- Annex I) countries as a result of their sponsorship (see section 9.8).
CH4
Methane, a key greenhouse gas primarily from agricultural activities (mostly
livestock and rice paddies), landfills, and some fossil fuel combustion, as well
as natural sources such as swamps.
climate change
Any change in climate or key variable (e.g. temperature, rainfall, wind, sea
levels and sea temperatures) over time spans greater than a decade. Shorter
term changes are termed “climate variability”.
CO2
Carbon dioxide, a key greenhouse gas primarily from fossil fuel burning and
land use change. Natural sources include respiration and decomposition.
COP
Conference of Parties, under the UNFCCC.
EECA
Energy Efficiency and Conservation Authority, the principal agency in New
Zealand involved in promoting energy efficiency and renewable energy.
GDP
Gross Domestic Product: a measure of the total monetary value of goods and
services produced by an economy, and a partial indication of living standards.
“Nominal” GDP is the value for that year in that year’s dollars. “Real” GDP is used for
comparisons between years and is adjusted for inflation and cost of living changes.
123
The greenhouse effect and climate change
global warming
Parliamentary Library, August 2001
The predicted warming of the global climate due to the greenhouse effect.
The term “climate change” is preferred, as it also includes other predicted
climate effects (e.g. droughts, floods, storms, local temperature variability
including cooling, shorter winters, melting ice, and rising sea levels).
greenhouse
effect
The trapping of heat near the Earth’s surface by gases that absorb and re-emit
infrared (heat) radiation. These gases are from both human and natural
sources, and without the greenhouse effect the Earth’s temperature would be
about 30°C colder than it is today.
Gg, Gt
Gg = Gigagrams (one billion grams), Gt = Gigatonnes (one billion tonnes).
One Gg is equivalent to one kT (kilotonne = 1,000 tonnes).
IEA
International Energy Agency, a subsidiary body of OECD.
IPCC
Intergovernmental Panel on Climate Change, established by UNEP and WHO.
It is comprised of hundreds of scientists and government representatives. Their
Summary for Policymakers reports are approved line-by-line in plenary
sessions with usually over 100 government representatives. The reports are
therefore considered conservative.
Kyoto Protocol
An agreement under the UNFCCC, signed in 1997 but not yet in effect as the
required full ratification has not taken place (see section 1.3).
J I (Joint
Implementation)
A mechanism allowed under the Kyoto Protocol for Annex I Party countries to
share costs, credits, and projects for greenhouse gas emission reduction.
LULUCF
Land-Use, Land-Use Change, and Forestry activities which may contribute
greenhouse gas emissions or act as carbon sinks or reservoirs, and are
subject to reporting under the UNFCCC and Kyoto Protocol. Certain LULUCF
activities will also be available for emissions accounting under the Protocol
(e.g. afforestation, reforestation, and deforestation under Article 3.3).
N2O
Nitrous oxide, a greenhouse gas primarily from agricultural activity. Natural
sources include organic wastes.
OECD
Organization for Economic Cooperation and Development. New Zealand is a
member.
ppmv
Parts per million by volume: a measure of gas concentration in the
atmosphere. Ten parts per million means that there are a million molecules of
air for ten molecules of the gas measured. Also expressed as ppm.
renewable
energy
Energy that is from naturally renewable sources within the human planning
timeframe. This includes solar, wind, biomass, and tidal energy. Fossil fuels
are considered to be non-renewable: while they are from natural sources, they
take millions of years to form in the Earth from organic matter deposits. Nuclear
fuel is not from natural sources.
sequestration
In the climate change context, lodging of carbon from CO2 into carbon sinks.
UNFCCC
United Nations Framework Convention on Climate Change.
124
The greenhouse effect and climate change
Parliamentary Library, August 2001
Appendix A:
Summary of key points from publicly released
Cabinet papers on climate change policy
as at 23 January 2001
•
The Government intends to ratify the Kyoto Protocol in mid - 2002. Demonstrable steps toward
meeting the Kyoto protocol commitments for 2008-2012 will need to be made by 2005. (CBC Min
(01) 1/1, a)
•
The Government intends to consult widely with industry, scientific and environmental groups,
local government, and Mäori in the process of developing climate change policy. (CAB (00) M 25/4A ;
CBC Min (01)1/2, a(ii) and h)
•
A communications strategy will be developed to promote public awareness of climate change
and the Government’s policy response. The strategy will inform businesses and individuals how
they can be involved in both policy development and action to reduce greenhouse gases.
Promotion of pre-2008 behaviour change is being considered. (CAB (00) M 25/4A; CBC Min (01)1/2,d;
CBC Min (01) 3/4, h and i).
•
A 1990 report on the likely climate change impacts to New Zealand will be updated and
released as a short summary report for public release [originally due April 2001: published July
2001]. Stage II (socio-economic analysis) will be funded during 2001/02. (CBC Min (01) 1/3)
•
The Government seeks to meet the 2008-2012 obligations under the Kyoto Protocol in a manner
that demonstrates environmental integrity and leadership while keeping as low as
practicable the social and economic costs of measures to achieve those obligations.
Domestic emissions trading is seen as providing the greatest assurance of achieving this, and
will be a central policy measure. Pilot emissions trading will be further considered after the
decision on a carbon charge. (CBC Min (01) 1/7, a, c, and d; CBC Min (01) 3/4, r)
•
In principle some proportion of New Zealand’s forest sink credits, expected to be internationally
tradable by 2008-2012, should accrue to those undertaking the activities. (CBC Min (01) 1/8)
•
Decisions on a carbon charge will only be taken as part of the current overall tax review process,
to be reported back to Ministers in September 2001. If there is a decision to proceed, it would not
be implemented until after the next election. If industries enter into Negotiated Greenhouse
Agreements before the charge is levied and reduce their emissions, this would be recognised in
the design and application of the charge. (CBC Min (01) 3/4, j - p)
•
As road transport generates 39% of New Zealand’s CO2 emissions and this is likely to increase
without policy intervention, Cabinet has requested a series of reports on themes including land
transport pricing, sustainable alternatives to road transport, and vehicle efficiency standards.
(CBC Min (01) 1/4)
•
As agricultural emissions of non-CO2 gases generate about 55% of New Zealand’s
greenhouse gas emissions and there are virtually no practical reduction options available, officials
have been directed to report on strategies to promote research.
(CBC Min (01) 1/5)
•
Vote: Research, Science and Technology will be re-focussed to better achieve climate
change policy objectives, to be reflected in the Foundation for Research Science and
Technology’s annual purchase agreement from 1 July 2001. Improvements are required in coordination, private sector investment, technology transfer, maintaining the research skill base, and
implementing social policy objectives.
(POL (00) M 35/5)
•
Officials will report back by 30 June 2001 on a national waste minimisation strategy, which has
potential to have energy efficiency and greenhouse gas emission benefits. (CBC Min (01) 1/1, e)
•
The Government is committed to leading the task of identifying, supporting and capitalising on the
economic opportunities presented by New Zealand’s climate change programme, and considers
that taking action now will improve New Zealand’s economic, environmental, and social position in
(CBC Min (01)1/2, e(v) and e(viii))
respect of the rest of the world.
125
The greenhouse effect and climate change
126
Parliamentary Library, August 2001
The greenhouse effect and climate change
Parliamentary Library, August 2001
Selected references
Note: other references in footnotes throughout the text
Australian Greenhouse Office, 2000, Global Warming - Cool it! A home guide to reducing energy costs
and greenhouse gases
Basher, Reid, 1998, The 1997/98 El Nino Event: Impacts, Responses and Outlook for New Zealand,
http://www.morst.govt.nz/publications/elnino/index.htm
Basher, Reid E. and Pittock, A. Barrie (eds), 1998, Australasia chapter, IPCC Special Report on the
Regional Impacts of Climate Change An Assessment of Vulnerability,
http://www.grida.no/climate/ipcc/regional/058.htm
Bertram, Geoff, 2001, Comments on "The Economic Effects of Low-Level Carbon Charges",
http://www.climatechange.govt.nz
Consumer, 2001, Save Money, Warm Up, Feel Better, Consumer magazine July 2001, pp. 20-23.
Controller and Auditor-General, 2001, Meeting International Environmental Obligations,
http://www.oag.govt.nz/homepagefolders/auditofficereports/mieo/mieo.htm
Crowley, Thomas J., 1996, Remembrance of Things Past: Greenhouse Lessons from the Geologic
Record, http://www.gcrio.org/CONSEQUENCES/winter96
Depledge, Joanna, 1999, Coming of age at Buenos Aires: the climate change regime after Kyoto,
Environment, September 1999
Energy Efficiency and Conservation Authority, 2001a, Draft National Energy Efficiency and Conservation
Strategy
Environmental Defence Society Inc, 2000, Report on United Nations Framework Convention on Climate
Change Sixth Conference of the Parties, The Hague, November 2000,
http://www.eds.org.nz/haguereport.htm
Ford-Robertson J.B., Maclaren J.P. and Wakelin S.J., 2000, The role of carbon sequestration as a
response strategy to global warming, with a particular focus on New Zealand, In Gillespie and Burns 2000,
Climate Change in the South Pacific: Impacts and Resources in Australia, New Zealand, and Small Island
States
Gillespie, Alexander and Burns, William C.J., 2000, Climate Change in the South Pacific: Impacts and
Responses in Australia, New Zealand, and Small Island States, Kluwer Academia Publishers, London.
Gillespie, Alexander, 2000, New Zealand and the Climate Change Debate: 1995-1998, In Gillespie &
Burns 2000
Hadley Centre for Climate Prediction and Research, 1999, Climate change and its impacts: stabilisation of
CO2 in the atmosphere,
http://www.metoffice.gov.uk/research/hadleycentre/pubs/brochures/B1999/contents.html
Hamilton, Kirsty, 2000, New Zealand Climate Policy Between 1990 and 1996: A Greenpeace Perspective,
In Gillespie & Burns 2000.
Harvey, Martin, 2001, Analysis of the rate and corresponding sectoral, distributional and competitiveness
impacts of a carbon charge, and revenue recycling options, http://www.climatechange.govt.nz
Horgan, G.P., 1999, Economic Issues in the Planting of New Zealand Native Trees, In conference
proceedings Native Trees for the Future, University of Waikato 8-10 October 1999, pp. 76-79
127
The greenhouse effect and climate change
Parliamentary Library, August 2001
Infometrics Consulting, 2001, The Economic Effects of Low-Level Carbon Charges,
http://www.climatechange.govt.nz
Intergovernmental Panel on Climate Change (IPCC), 2000, Land Use, Land-Use Change and Forestry:
Summary for Policymakers, Cambridge University Press, or
http://www.grida.no/climate/ipcc/land_use/001.htm
Intergovernmental Panel on Climate Change (IPCC), 2001a, Climate Change 2001: the Scientific Basis,
IPCC Working Group I Third Assessment Report, Summary for Policymakers (final draft),
http://www.ipcc.ch
Intergovernmental Panel on Climate Change (IPCC), 2001a, Summary for Policymakers: A Report of
Working Group I of the Intergovernmental Panel on Climate Change , http://www.ipcc.ch
Intergovernmental Panel on Climate Change (IPCC), 2001b, Climate Change 2001: Impacts, Adaptation,
and Vulnerability; IPCC Working Group II Third Assessment Report, Summary for Policymakers, Draft 1902-2001, approved in Geneva 13-16 February 2001., http://www.ipcc.ch
Intergovernmental Panel on Climate Change (IPCC), 2001c, Summary for Policymakers: Special Report
on Emission Scenarios, A Special Report of Working Group III of the Intergovernmental Panel on Climate
Change, http://www.ipcc.ch
Intergovernmental Panel on Climate Change (IPCC), 2001d, Technical Summary: A report accepted by
Working Group I of the IPCC but not approved in detail, http://www.ipcc.ch
Intergovernmental Panel on Climate Change (IPCC), 2001d, Technical Summary, Climate Change 2001:
Mitigation, A Report of Working Group III of the Intergovernmental Panel on Climate Change,
http://www.ipcc.ch
International Energy Agency (OECD), 2000a, CO2 Emissions from fuel combustion 1971-1998
International Energy Agency (OECD), 2000b, Dealing With Climate Change; Policies and Measures in IEA
member Countries, IEA/OECD, Paris
International Energy Agency (OECD), 2000c, Emission Baselines: Estimating the Unknown
International Energy Agency, 2001, Things That Go Blip in the Night: Standby Power and How to Limit It,
http://www.iea.org/public/studies/blip.htm
Local Government and Environment Select Committee, 2000, Inquiry into the Role of Local Government in
Meeting New Zealand's Climate Change Target: interim report of the Local Government and Environment
Select Committee, New Zealand House of Representatives, December 2000, Shoulder no. I.9A
Maclaren, J. Piers, 2000, Trees in the Greenhouse: the Role of Forestry in Mitigating the Enhanced
Greenhouse Effect, Forest Research Bulletin no. 219, Forest Research Institute, Rotorua
McSoriley, John, 1995, Climate Change, Forestry and the New Zealand Resource Management Act 1991,
Unpublished conference paper
Ministry of Agriculture and Forestry 2001, A National Exotic Forest Description as at 1 April 2000,
http://www.maf.govt.nz/statistics/primaryindustries/forestry/nefd2000/ntitle.htm
Ministry of Agriculture and Forestry 2001, New Zealand Forestry Statistics 2000,
http://www.maf.govt.nz/statistics/primaryindustries/forestry/nzstats2000/title.htm
Ministry for the Environment, 1998, Climate Change: More than just carbon dioxide. Significance, sources
and solutions for non-CO2 greenhouse gases in New Zealand, Ministry for the Enivronment, March 1998
Ministry for the Environment, 1999, Climate Change Domestic Policy Options Statement, a Consultation
Document, Ministry for the Environment, January 1999
128
The greenhouse effect and climate change
Parliamentary Library, August 2001
Ministry for the Environment 2000, Implementation of Article 3(3) of the Kyoto Protocol: Submission by
New Zealand, 1 August 2000, http://www.mfe.govt.nz
Ministry for the Environment, 2001, National Inventory Report, New Zealand Greenhouse Gas Inventory
1990-1999, including Common Reporting Format for 1999, as reported April 2001
Ministry of Economic Development, 2000, New Zealand Energy Greenhouse Gas Emissions 1990-1999
Ministry of Transport, 1998, National Transport Statement, Ministry of Transport, Wellington
Natural Resources Canada, 2000, The Capture and Storage of Carbon Dioxide Emissions: a significant
opportunity to help Canada meet its Kyoto targets, Office of Energy Research and Development, Ottawa,
Ontario, October 2000
New Zealand Climate Change Programme, 2001a, Forest Sinks and the Kyoto Protocol - An Information
Document, http://www.climatechange.govt.nz
New Zealand Climate Change Programme, 2001b, Climate Change Impacts on New Zealand,
http://www.climatechange.govt.nz
Newell, Peter, ca. 1997, Climate Politics in Western Europe: Regional and Global Dimensions, Climate
Network Europe, http://www.ecouncil.ac.cr/rio/focus/report/english/climate.htm
OECD, 1999, Environmental Data Compendium 1999, OECD, Paris
OECD, 1996, Environmental performance reviews: New Zealand, (excerpts)
OECD, 1999, Nominal tax rates of environmentally related taxes - by country (individual print-outs for
Australia, Denmark, Ireland, Japan, Netherlands, New Zealand, Sweden, UK),
http://www.oecd.org/env/policies/taxes/index.htm
OECD, 2000, Ancillary Benefits and Costs of Greenhouse Gas Mitigation; Proceedings of an IPCC CoSponsored Workshop, 27-29 March 2000, Washington D.C.
Parliamentary Commissioner for the Environment , 1992, Sustainable Energy Management in New
Zealand: improvements required in Government policy
Parliamentary Commissioner for the Environment , 2000, Getting More from Less: A review of progress on
energy efficiency and renewable energy initiatives in New Zealand, PCE February 2000
Schloerer, Jan, 1997, Climate change: some basics, University of Ulm,
ftp://rtfm.mit.edu/pub/usenet/news.answers/sci/climate-change/basics
Soon, Willie, Baliunas, Sallie L., Robinson, Arthur B., and Robinson, Zachary W., 1999, Environmental
effects of increase atmospheric carbon dioxide, Climate Research 13:149-164
Tiwari, Dirgha Nidhi, 2000, Towards a Framework for the Implementation of the Clean Development
Mechanism in the Agricultural Sector of Developing Countries, In FAO 2000, Two essays on climate
change and agriculture; a developing country perspective, FAO Economic & Social development Paper
no. 145,FAO, Rome
Transport and Environment Select Committee, 1998, Inquiry into the Environmental Effects of Road
Transport, New Zealand House of Representatives, Shoulder no. I.12B
UMR Research Ltd., 2001, Climate Change Issues: A Study of Public Awareness and Level of Concern,
New Zealand Climate Change Programme, http://www.climatechange.govt.nz
UNFCCC Secretariat, 1997, Kyoto Protocol to the United Nations Framework Convention on Climate
Change, http://www.unfccc.de/resource/docs/convkp/kpeng.html
UNFCCC Secretariat, 1999, New Zealand: Report on the in-depth review of the second national
communication of New Zealand, http://www.unfccc.int/
129
The greenhouse effect and climate change
Parliamentary Library, August 2001
UNFCCC Secretariat, 2000, National Communications from Parties included in Annex I to the Convention:
Greenhouse Gas Inventory Data from 1990 to 1998, FCCC/SBI/2000/11, 5 September 2000
UNFCCC Secretariat, 2001a, Report of the Conference of the Parties on the First part of its Sixth session,
held at The Hague from 13 to 25 November 2000: Addendum, Part Two: Action Taken by the Conference
of the Parties at the First Part of its Sixth Session, document FCCC/CP/2000/Add.2, 4 April 2001,
http://www.unfccc.de/
UNFCCC Secretariat, 2001b, Data tables - emissions 1990 and 1998, via
http://www.unfccc.de/resource/ghg/tempemis2.html
UNFCCC Secretariat, 2001c, Decision 5/CP.6, Implementation of the Buenos Aires Plan of Action
(unedited version), FCCC/CP/2001/L.7, http://www.unfccc.int/resource/docs/cop6secpart/l07.pdf
UNFCCC Secretariat, 2001d, Press Release: Governments adopt Bonn agreement on Kyoto Protocol
rules, 23 July 2001, http://www.unfccc.de
UNFCCC Secretariat, 2001e, Review of the implementation of commitments and of other provisions of the
Convention: Preparations for the First Session of the Conference of the parties serving as the meeting of
the Parties to the Kyoto protocol (Decision 8/CP.4), Decision 5/CP.6: Implementation of the Buenos Aires
Plan of Action , FCCC/CP/2001/L.7, 24 July 2001, http://www.unfccc.de
UNFCCC Secretariat, 2001f, Preparations For The First Session Of The Conference Of The Parties
Serving As The Meeting Of The Parties To The Kyoto Protocol (Decision 8/CP.4), Matters Relating To
Land-use, Land-use Change and Forestry: Draft Decision proposed by the Co-Chairmen of the negotiating
group, Draft Decision-/CP.6, Land-use, land-use change and forestry,
http://www.unfccc.int/resource/docs/cop6secpart/l11.pdf
United Nations Conference on Trade and Development (UNCTD), 1998, Greenhouse Gas Emissions
Trading: defining the principles, modalities, rules and guidelines for verification, reporting and
accountability, http://www.unctad.org/en/pub/
United Nations Environment Programme (UNEP), 1998, Press Kit: Buenos Aires 1998, Fourth session of
the Conference of the Parties UNFCCC 2-13 November 1998,
http://www.unep.ch/iuc/submenu/press/climate/cop4kit.htm
US Environmental Protection Agency, 1999, Progress Report on the EPA Acid Rain program, EPA430-R99-011, November 1999, http://www.epa.gov/acidrain
World Resources Institute, UN Development Programme, UN Environment Programme, and the World
Bank , 2000, World Resources 2000-2001 People & Ecosystems: The Fraying Web of Life, World
Resources Institute, Washington DC, http://www.wri.org/wr2000
Selected websites
IPPC home page (electronic copies of reports): http://www.ipccc.ch
New Zealand Government Climate Change programme: http://www.climatechange.govt.nz
Ministry for the Environment: http://www.mfe.govt.nz/issues/climate.htm
New Zealand climate data and model projections: http://www.katipo.niwa.cri.nz/ClimateFuture
Ministry of Economic Development (energy data, emissions trading):
http://www.med.govt.nz/ers/environment.html
130