Climatic Change (2012) 110:123–141 DOI 10.1007/s10584-011-0072-2

Economics- and physical-based metrics for comparing greenhouse gases Daniel J. A. Johansson

Received: 28 April 2009 / Accepted: 5 April 2011 / Published online: 3 May 2011 © Springer Science+Business Media B.V. 2011

Abstract A range of alternatives to the Global Warming Potential (GWP) have been suggested in the scientific literature. One of the alternative metrics that has received attention is the cost-effective relative valuation of greenhouse gases, recently denoted Global Cost Potential (GCP). However, this metric is based on complex optimising integrated assessment models that are far from transparent to the general scientist or policymaker. Here we present a new analytic metric, the CostEffective Temperature Potential (CETP) which is based on an approximation of the GCP. This new metric is constructed in order to enhance general understanding of the GCP and elucidate the links between physical metrics and metrics that take economics into account. We show that this metric has got similarities with the purely physical metric, Global Temperature change Potential (GTP). However, in contrast with the GTP, the CETP takes the long-term temperature response into account.

1 Introduction Policy discussions concerning mitigation of climate change often focus on abatement of CO2 emissions. However, as is well known substantial economic gains is achieved if other greenhouse gases are included in the policy portfolio (Reilly et al. 1999; Hayhoe et al. 2000; Van Vuuren et al. 2006; Weyant et al. 2006). Different greenhouse gases have different physical properties, both in terms of their effect on the radiative balance of the earth system and in terms of atmospheric

D. J. A. Johansson (B) Division of Physical Resource Theory, Department of Energy and Environment, Chalmers University of Technology, Gothenburg, Sweden e-mail: [email protected] D. J. A. Johansson Environmental Economics Unit, Department of Economics, School of Business, Economics and Law, University of Gothenburg, Gothenburg, Sweden

124

Climatic Change (2012) 110:123–141

lifetimes. Therefore, to facilitate the implementation of a multigas approach to climate change mitigation, a way of comparing reductions in emissions of different greenhouse gases is needed. For this purpose, the Intergovernmental Panel on Climate Change (IPCC) and the United Nations Framework Convention on Climate Change (UNFCCC) have adopted the use of Global Warming Potentials (GWP), originally developed in Shine et al. (1990). Basically, the GWP is constructed to ensure equivalence in time-integrated radiative forcing of emission pulses over a pre-specified time horizon, usually 20, 100, or 500 years. As the Kyoto protocol stands, GWPs calculated over a time period of 100 years should be used to facilitate the trade-offs between different greenhouse gases (UNFCCC 1997). However, the choice of using GWP in the first place is rather arbitrary, as is the time horizon over which the GWPs are calculated. There is no clear scientific reason why to choose GWP instead of any other metric or one time horizon over another when calculating the GWP value.1 The GWP metric has been criticized by both climate scientists and economists. Choosing to construct the GWP so as to ensure equivalence in time-integrated radiative forcing for equal CO2 equivalent emissions of different species implies, in general, that equivalence will not be obtained for other end-points, such as temperature change or sea level rise, which may be considered more relevant endpoints (Reilly et al. 1999; O’Neill 2000; Smith and Wigley 2000).2 Based on the critique of the GWP, a range of different metrics, purely physicalbased metrics as well as metrics that also take economics into account, have been proposed, see for example Michaelis (1992), Eckaus (1992), Schmalensee (1993), Reilly and Richards (1993), Hammitt et al. (1996), Kandlikar (1996), Wigley (1998), Manne and Richels (2001), Shine et al. (2005) and Tanaka et al. (2009) and the review articles by Fuglestvedt et al. (2003, 2010). In both the research community and the climate policy community there has been a renewed interest in reconsidering the GWP (UNFCCC 2008; Plattner et al. 2009). Two of the alternatives to the GWP that have gained attention recently are the Global Temperature change Potential (GTP) and the Global Cost Potential (GCP). The basic principle of the GTP is to take the ratio of the temperature response after a specific period of time, , following an emission pulse of a specific greenhouse gas, X, to the temperature response after years of an emissions pulse of CO2 , see Shine et al. (2005, 2007). The GCP is based on a cost-effectiveness approach, and the metric is obtained by using optimising climate-economy models with exogenously set climate targets, such as the 2◦ C limit endorsed in the Copenhagen Accord (UNFCCC 2009). The GCP is defined as the ratio of the cost-effective price on emissions (i.e., the shadow price of the emissions) of greenhouse gas X to the cost-effective price on emissions of carbon dioxide, see for example Eckaus (1992), Michaelis (1992), Kandlikar (1996), Manne and Richels (2001), O’Neill (2003) and Johansson et al.

1 Eventhough

the time horizon is arbitrary, one may argue that there is a relation between the time horizon, T, and discount rate, r. For example one may argue that there is relation such as r = 1/T. Other approaches to this relationship are also discussed in Fuglestvedt et al. (2003).

2 However,

there is an exception. Given an integration time horizon that approaches infinity, GWP will overlap with an end-point metric that is the ratio of equilibrium temperature change for sustained emissions, see Fisher et al. (1990), O’Neill (2000) and Shine et al. (2005).

Climatic Change (2012) 110:123–141

125

(2006, 2008). Shine et al. (2007) observed and Tol et al. (2008) argued that the GTP and the GCP share many traits. The GCP approach should not be confused with another economic approach to metrics based on climate damages. In that approach the metric is based on the economic valuation of marginal climate damages of emissions of different gases, see for example Schmalensee (1993), Hammitt et al. (1996), and Kandlikar (1996). The aim of this paper is two-fold:3 – –

to show in a “simple” analytical framework that the GCP approach can be approximated with purely physical information and the discount rate; to show that this approximation of GCP has, under idealised conditions, similarities with the GTP metric as suggested by Shine et al. (2005, 2007) and the formulation of GWP as suggested by Lashof and Ahuja (1990). The formulation of GWP that Lashof and Ahuja (1990) suggested and that we refer to here is reminiscent of the IPCC’s version of GWP, but the integration time horizon approaches infinity, and the future radiative forcing of an emission pulse is discounted using a constant discount rate.

In Section 2 we develop a simple analytical framework in which the GCP is analysed and an approximation of it, the Cost-Effective Temperature Potential (CETP), is suggested. In Section 3 we further analyse the CETP given idealised assumptions, i.e., assumptions in line with those used for calculating the GWP and GTP. In Section 4 we present a simple climate-economy model that will be used to evaluate the CETP and the GTP with numerically obtained GCPs. In Section 5 we present the results and further analyse the differences between the GCP, CETP, and GTP. In Section 6 we discuss the policy implications of the results and conclude.

2 Analytical approach to an approximation of the GCP To analyse the GCP metric and construct an approximation, we use an integrated climate-economy model in which a limit on the global mean surface temperature is to be met at the lowest possible abatement cost in net present value terms. The basic model set up has similarities with the simple analytical approach in Eckaus (1992) but is comprehensive enough to capture the essence of the results in Manne and Richels (2001). Climate damages are not considered since the GCP metric is based on a cost-effectiveness approach and not a cost-benefit approach. We consider only two greenhouse gases, carbon dioxide and a generic gas, X. Generalizing so as to incorporate a range of different greenhouse gases would be straightforward. However, it would not change the nature of the results we are focused on and would make the model less transparent. We leave this for future work.

3 The

aim of this paper relates to several of the recommendations from the IPCC Expert Meeting on the Science of Alternative Metrics (Plattner et al. 2009). For example, our analysis presents an approach on how the long-term outcome for the post target period for GTPs can be dealt with, and we analyse how well physical based metrics approximate metrics based on both economics and physics.

126

Climatic Change (2012) 110:123–141

Given a limit on the global average surface temperature and the aim of minimization of the net present value abatement cost a general, albeit simple, model can be expressed as the following constrained optimisation problem:     t˜  CCO2 ,t aCO2 ,t + C X,t a X,t Min (1) (1 + r)t t=0 s.t.

 Tt = f ECO2 ,t − aCO2 ,t , ..., ECO2 ,0 − aCO2 ,0 , E X,t − a X,t , ...,  0 , S0X , T0 E X,0 − a X,0 , SCO 2

(2)

Tt ≤ Tˆ

(3)

CCO2,t and C X,t refer to the abatement cost functions for reducing CO2 and gas X in time step t, respectively, whereas a CO2, t and a X,t refer to the level of abatement of CO2 and gas X, respectively, in time step t. t˜ is the planning time horizon used in the model. The abatement cost is discounted at discount rate r. When analysing longterm issues such as climate change, the approach to discounting is critical, see for example Lind (1982) and Dasgupta (2008). We have chosen the most straightforward and simple approach: we use a discount rate that is constant over time. For now we leave the choice of discount rate open and discuss possible numerical values in Section 4. The function f , shown in Eq. 2, is a temperature model in which the global mean temperature at time t depends on past emissions (where the emissions for each greenhouse gas are equal to reference emissions minus abatement), the initial 0 atmospheric stocks (SCO and S0X ) of the greenhouse gases, and initial global mean 2 surface temperature above the pre-industrial level (T0 ). Implicitly included in this temperature model are models that keep track of the gas cycles, the radiative forcing of different gases and a model used for calculating the responding change in the global average surface temperature, e.g. by the use of an energy balance model. In Section 3 we make explicit assumptions for this temperature model, but for now we keep it more general. Equation 3 is a policy constraint implying that the increase in global mean surface temperature has to be kept below Tˆ ◦ C. The abatement cost functions for both gases are positive, increasing in abatement, 2 C(a) and convex, i.e., C (a) > 0, ∂C(a) > 0, ∂ ∂a ≥ 0. 2 ∂a If we solve for the first order conditions of the constrained optimisation problem presented in 1–3 by stating the Lagrangian we get: t˜

 ∂CCO2 ,t ∂ Tτ ητ = τ −t · ∂aCO2 ,t ∂a + r) (1 CO2 ,t τ =t

(4)

and t˜

 ∂ Tτ ητ ∂C X,t = τ −t · ∂a X,t ∂a + r) (1 X,t τ =t

(5)

where ητ is the Lagrange multiplier (expressed in current value terms) associated with the constraint on the global annual mean surface temperature in Eq. 3.

Climatic Change (2012) 110:123–141

127

Optimality conditions in Eqs. 4 and 5 imply that the optimal marginal abatement cost in each time step equals the sum up to t˜ of the discounted Lagrange multiplier (shadow price) associated with the temperature constraint multiplied by the marginal temperature response of abatement. Note that the Lagrange multiplier associated with the temperature constraint is negative or zero; the marginal temperature response of abatement is also negative, and as a consequence the right hand sides of Eqs. 4 and 5 are positive. The Lagrange multiplier ητ of the temperature constraint is zero as long as the constraint in Eq. 3 is not binding. The constraint will be non-binding as long as the global average surface temperature is less than the limit in Eq. 3. When the temperature meets the limit, at time t¯, the constraint starts to bind and the Lagrange multiplier ητ takes on a non-zero value. This implies that the temperature response prior to year t¯ does not have any influence on the optimal marginal abatement cost, presented in Eqs. 4 and 5. Hence, it is only the temperature response beyond the date of stabilisation, i.e., at time t¯ and beyond, that has on effect on the optimal marginal abatement cost. In order to continue the analysis we have to define what we mean by a costeffective price on emissions. The cost-effective price on emissions is the shadow price on emissions generated in the model. The shadow price on emissions of a specific gas and in a specific year is a measure on how much one should be willing to pay to emit an additional unit of that gas in that year. The cost-effective price on emissions would in our simple model be equal to the optimal marginal abatement cost presented in Eqs. 4 and 5.4 Hence, the cost-effective price on CO2 emissions (equivalent for emissions of X) is: PriceCO2 (t) =

t˜  τ =t

∂ Tτ ητ τ −t · ∂aCO2 ,t (1 + r)

(6)

Therefore, the GCP approach as suggested by Manne and Richels (2001) and many others can be summarised by the following ratio: t˜ 

PriceX (t) τ =t = GCPX (t) = t˜ PriceCO2 (t)  τ =t

∂ Tτ ∂a X,t

∂ Tτ ∂aCO2 ,t

ητ (1+r)τ −t

·

(7) ·

ητ (1+r)τ −t

Note, though, that GCPs do not directly depend on the abatement cost assumptions but only indirectly via the abatement cost affecting the Lagrange multiplier ητ . Unfortunately the ratio in Eq. 7 does not make us much more content since we do not know much about ητ . We know that ητ is zero as long as the temperature constraint 3 is not binding and that ητ takes on a negative value in time periods when the constraint is binding. ητ depends then in a complex manner on a range of parameter assumptions such abatement cost assumptions, initial conditions, baseline

4 The

cost-effective price on the emissions is also the level an optimal tax on emissions (or permit price generated in a cap-and-trade system) should have in order to induce the necessary abatement at the lowest social cost possible in a society where the production and consumption decisions are taken in order to maximize profit and utility, respectively, see for example Baumol and Oates (1988) and Sterner (2002).

128

Climatic Change (2012) 110:123–141

emissions etc. Digging into this is outside the scope of this study and it will not be discussed further, however, related discussions on this in the context of climate metrics can be found in Eckaus (1992) and Reilly and Richards (1993), see also Appendix. Since our aim is to find a simple metric that approximates the GCP we suggest the following metric, which we call the Cost-Effective Temperature Potential (CETP): t˜ CETPX (t) =



t˜

ECO ,t Tτ 2

ECO2 ,t



t˜ 

E

Tτ x,t −r(τ −t) e dτ E X,t

≈ e−r(τ −t) dτ

τ =t¯ t˜  τ =t¯

∂ Tτ ∂a X,t

·

∂ Tτ ∂aCO2 ,t

1 (1+r)τ −t

·

for t ≤ t¯

(8)

for t > t¯

(9)

1 (1+r)τ −t

and t˜ CETPX (t) =

t

t˜ t

E

t˜ 

E

Tτ x,t −r(τ −t) e dτ E X,t

ECO ,t Tτ 2

ECO2 ,t

e−r(τ −t) dτ



τ =t t˜  τ =t

∂ Tτ ∂a X,t

∂ Tτ ∂aCO2 ,t

·

1 (1+r)τ −t

·

1 (1+r)τ −t

where Tτ x,t is the temperature response in time τ following a small emissions ECO ,t pulse, E X,t , of gas X in time t, and where Tτ 2 is the temperature response in time τ following a small emissions pulse, ECO2 ,t , of CO2 in time t. t¯ is the year the global average surface temperature stabilizes at the climate policy limit and t˜ is the planning time horizon. Hence, the integration starts from t¯ in Eq. 8 since the temperature response prior to stabilisation does not have any affect on the CETP metric value or the GCP metric value; see above for discussion. Equation 8 is valid for an emission pulse that takes place prior to the stabilisation of the global average surface temperature, while Eq. 9 is valid for an emissions pulse that takes place after the global average surface temperature has stabilised at the policy limit. The most important step between Eq. 7, which shows the GCP, and Eqs. 8 and 9, which show the CETP, is that we dropped the Lagrange muliplier ητ . This is an important simplification; in most of the remaining parts of the paper we evaluate this simplification. It is virtually impossible to solve for ητ analytically and as is seen in the numerical tests in Section 5 omitting ητ shows that the suggested approximation works well for the assumptions we have in the numerical model. Eckaus (1992) discussed a related case but related to radiative forcing targets, instead of temperature targets in an analysis of alternatives to the GWP. Although he criticized an approximation similar to the one here we want to take the analysis one step forward, make approximations and add to the understanding of the relationship between physical metrics and economic metrics. Further, as one of the main aims of the paper is to find an approximation of the GCP that captures its essence but is simple enough to be captured in an analytical formula, simplifications have to be made. By the suggested approximation of the GCP, the metric could be approximated with a formula that only requires physical information, an assumption on the date of stabilisation, and the discount rate. In addition, given the rather

Climatic Change (2012) 110:123–141

129

restrictive assumption that ητ is constant beyond the date of stabilisation, t¯, the CETP would be identical to the GCP since ητ would in this case cancel out.5 Other important simplifications and changes we made in going from from the GCP to the CETP are: •



In Eqs. 8 and 9 we went from discrete time to continuous time and analyses the temperature response of a small positive emissions pulse instead of the marginal temperature response of abatement. This is done to enhance the tractability of the metric without losing anything essential. The integration starts from t¯ and not t in Eq. 8. The reasons for this is that prior to t¯ the Lagrange multiplier ητ is zero and as discussed above the temperature response prior to t¯ does not have any affect on the cost-effective price on the emissions nor on the GCP. For these reasons, the temperature response prior to t¯ should not have any impact on the CETP either. Equation 9 shows the CETP for the time period after stabilisation, so in this case the integration is from t.

In order to analyze the approximation errors, gain insights in the important factors that determine the numeric value for the GCP and CETP metrics, and elucidate these two metrics’ links with the GTP and GWP metrics, we will derive an explicit solution of the CETP in Section 3.

3 Analytical solution of the CETP under linearized conditions We will first derive an expression for the Absolute CETP (ACETP). Since the aim is to find a metric at a similar level of complexity as the GWP and GTP, we make similar assumptions concerning linearity in radiative efficiency, atmospheric decay rates and the temperature response function. We use a simple, one-box energy balance model, see for example Shine et al. (2005) and Andrews and Allen (2008), to assess the change in the global annualmean surface temperature: •

Tτ =

Tτ Fτ − H λH

(10)

where H is the equivalent heat capacity of the climate system, λ is the climate sensitivity, τ is time, Fτ the time dependent radiative forcing and Tτ the temperature response. The climate sensitivity, λ, is assumed being a constant that is independent of the particular greenhouse gas or aerosol causing an effect on radiative forcing. By using the energy balance model presented in Eq. 10 the temperature response in year τ of a unit emission pulse of agent X in year t with a constant atmospheric

Lagrange multiplier ητ would be constant if the system would be in a steady state after the stabilisation had taken place.

5 The

Climatic Change (2012) 110:123–141

Temperature (K)

130

t

t

~ t

Temperature response prior to stabilisation Component 1: Relaxation of the temperatue response built up prior to stabilization Component 2: Temperature response from the remaining forcing at the stabilization Full temperature response beyond stabilization

Fig. 1 The figure illustrates the integration shown in Eq. 12. It shows the temperature response of an emission pulse in year t, and how the integration of the temperature response beyond the year, t¯, i.e., the year in which the temperature is stabilised at the temperature limit, is divided into two components; the relaxation of the temperature that is built up prior to the stabilization and the built up of additional temperature change from the forcing of the emissions pulse that is remaining in the stabilisation year and beyond. t˜ is the planning time horizon used in the model

decay rate of δ X and radiative efficiency I X can be described by, see Shine et al. (2005):   −(τ −t) IX  e−(τ −t)δ X − e λH  1 TτEx,t = for τ ≥ t (11) H λH − δ X Assuming that E X,t (t) = 1 kg, and that the planning time horizon, t˜, approaches infinity the integrals in the numerator and denominator in Eqs. 8 and 9 can be solved. The assumption of an infinite planning time horizon implies that all future temperature impacts beyond the time of stabilisation (t¯ and beyond) of an emission pulse at t, are taken into account in the metric. Further, when the integrals in Eqs. 8 and 9 are solved we utilise the assumption that the temperature response of an emission pulse is modelled as a linear system. Due to this linearity assumption, the temperature response of an emission pulse can be divided into components that later can be added to give the total temperature response, see Fig. 1.6 We divide the temperature response into two components: 1. The temperature response prevailing in the stabilization year t¯ and the decline in temperature beyond year t¯ given the assumption that the remaining forcing prevailing when the global average surface temperature stabilizes at the policy limit in year t¯ would suddenly be removed. This would cause a decline in the temperature beyond year t¯ that can be described by the temperature response time of λH. 2. The additional temperature response beyond the stabilisation year t¯ that is generated by the forcing prevailing in year t¯ and beyond. The time dynamics of the temperature response for this component will depend on both the atmospheric life time of the forcing agent δ X and the temperature response time λH.

6 This

principle is usually referred to as the Superposition principle.

Climatic Change (2012) 110:123–141

131

The division of the temperature response into two different components is done in order to get the CETP metric formula in a form that shows its relationship with the GTP and GWP. By summing these two components, under the assumption of linearity, the full temperature response beyond t¯ is obtained, see Fig. 1. Given these assumptions and assuming an emission pulse occurring prior to the year in which the temperature stabilises at the policy limit, the ACETPX can be written as: t˜

E

Tτ x,t −r(τ +t) e dτ E X,t

ACETPX (t) = lim

t˜→∞



∞ =

H



∞ + t¯



 −(τ −t¯)  −(t¯−t) IX  e−(t¯−t)δ X − e λH · e λH · e−(τ −t)·r dτ − δX

1 λH

 −(τ −t¯) I X · e−δ X (t¯−t)  −(τ −t¯)δ X  e  1 − e λH · e−(τ −t)·r dτ H λH − δ X

IX  = H r+

1 λH



1 e−(t¯−t)(r+δ X ) − e−(t¯−t)(r+ λH ) e−(t¯−t)(r+δ X ) + 1 (r + δ X ) − δX λH

for t ≤ t¯ (12)

For an emission pulse occurring after the year in which the global average surface temperature stabilises at the policy limit, the ACETPX can be written as: ACETPX (t) =

 H r+

IX  1 λH (r

+ δX )

for t > t¯

(13)

The ACETP for carbon dioxide is slightly more complex since the atmospheric perturbation lifetime of carbon dioxide cannot be described accurately by simple exponential decay, instead several decay rates have to be considered for the impulse response, see for example Maier-Reimer and Hasselmann (1987). In here we use the four term impulse response function given in Forster et al. (2007). Accordingly, the ACETP for carbon dioxide prior to the date of stabilisation can be written as:

1 ICO2 e−(t¯−t)r e−(t¯−t)r − e−(t¯−t)(r+ λH )   α0 ACETPCO2 (t) = + 1 1 r H r + λH λH

1 3  e−(t¯−t)(r+δi ) e−(t¯−t)(r+δi ) − e−(t¯−t)(r+ λH ) for t ≤ t¯ αi + + 1 (r + δi ) − δi λH i=1 (14) and for the time period after the date of stabilisation the ACETP for carbon dioxide is:

3 ICO2 α0  αi   for t > t¯ (15) ACETPCO2 (t) = + 1 r (r + δi ) H r + λH i=1

132

Climatic Change (2012) 110:123–141

By combining Eq. 12 with Eq. 14 we get the CETP for agent X prior to the date of temperature stabilisation: CETPX (t) =

ACETPX (t) ACETPCO2 (t) e−(t¯−t)δ X −e−(t−t) λH 1 λH −δ X 1

¯

=

IX ICO2

 α0



1−e−(t−t) λH 1

¯

+

1 H

+

1 r

3 



αi

i=1

+

e−(t¯−t)δ X (r+δ X )

e−(t¯−t)δi −e−(t−t) λH 1 λH −δi 1

¯

+

e−(t¯−t)δi (r+δi )



for t ≤ t¯

(16) While the CETP of an emission pulse of gas X after the date of stabilisation is: CETPX (t) =

1 (r+δ X ) 3 

IX ICO2

α0 r

+

i=1

=

αi (r+δi )

AGWPX = GWPX AGWPCO2

for t > t¯

(17)

Interestingly, as is seen in Eq. 17, the CETP after stabilisation when an infinite planning time horizon is assumed is identical to the GWP with discounting and infinite time horizon as formulated by Lashof and Ahuja (1990). The case that the GWP with discounting may come out of economic models under certain circumstance have also been found in other studies, see Eckaus (1992) and Reilly and Richards (1993). The response time of the energy balance model (which is equal to λH) cancels out in Eq. 17. The CETP prior to stabilisation as seen in Eq. 16 shares features with both Lashof and Ahuja’s GWP and the GTP as formulated by Shine et al. (2005, 2007). GTP for X evaluated at time t¯, i.e., the year that the temperature hits the policy limit, is: GTPX (t) =

AGTPX (t) AGTPCO2 (t) e−(t¯−t)δ X −e−(t−t) λH 1 λH −δ X 1

¯

=

IX ICO2

 α0

1−e−(t−t) λH ¯

1



+

1 λH

3 



αi

i=1

e−(t¯−t)δi −e−(t−t) λH 1 λH −δi ¯

1



for t ≤ t¯

(18)

Equation 16 for the CETP can be rewritten as: CETPX (t) =

¯

¯

¯

¯

t AGTPt,t→ · H + AGWPt,Xt→∞ X t t,t→∞ AGTPt,t→ CO2 · H + AGWPCO2

for t ≤ t¯

(19)

where t, t → t¯ means an emission pulse in year t and that the AGTP is calculated for a time of horizon t¯ minus t year, and where t, t¯ → ∞ means an emission pulse in year t and where the AGWP with discounting is calculated for the period t¯ → ∞ for the fraction of the pulse still present in the atmosphere after t¯ minus t years. Finally, it is worth noting that the CETP and GTP converge when the discount rate is increased. The long-term temperature response taking place beyond stabilisation has a smaller impact on the CETP metric value when the discount rate is high. A similar result is observed to also hold for the GCP in the numerical results presented in Section 5.

Climatic Change (2012) 110:123–141

133

4 Numerical model assumptions In order to compare the CETP and GTP with the GCP obtained from an optimising climate-economy model, we have used a simple dynamic climate-economy model that includes CH4 and CO2 .7 The numerical model is an updated and simplified version of the MiMiC model presented and used in Johansson et al. (2006, 2008). The main difference between the model used here and the versions in Johansson et al. (2006, 2008) is that the concentration radiative forcing relationships have been linearized around concentration levels prevailing in 2005 for the model runs presented in this paper. This assumption makes the cost-effective price on emissions, and consequently the GCP and CETP, commensurable with the GTP and the GWP, since linearity is assumed in the two latter metrics.8 This is of course a very crude simplification of the geophysical system, but it yields a model complicated enough to show the approximation made by the CETP in comparison with the GCP and to analyse the differences as well as similarities between the GCP, CETP, and GTP. The values for the radiative efficiency for each gas are taken from Forster et al. (2007). For CH4 we add 40% to the direct radiative efficiency due to the indirect effects of CH4 on the level of tropospheric ozone and stratospheric water, in line with Forster et al. (2007). CO2 concentrations are modelled by a linear impulse representation of the Bern carbon cycle model with numerical values from Forster et al. (2007). CH4 concentrations are modelled by using a global mean mass-balance equation with an average atmospheric life time of 9 years. So as to keep the climate part of the model linear we do not take into account that the atmospheric adjustment time for CH4 is different than the average atmospheric life time. For the time-discrete version of the one-box energy balance model shown in Eq. 10, we assume that the climate sensitivity is 3◦ C for a doubling of the CO2 concentration and that the equivalent heat capacity of the ocean-land-atmosphere system is equivalent to a 500 m deep ocean covering 71% of the earth’s surface. These assumptions on the climate sensitivity and the equivalent heat capacity results in an e-folding time of the global annual mean temperature response to about 36 years. This is roughly in line with the e-folding times obtained in the simple climate model MAGICC for a climate sensitivity of 3◦ C, see Richels et al. (2007), and in line with the e-folding times estimated from transient runs of large-scale AOGCMs (Andrews and Allen 2008). Note that the calibration of the equivalent heat capacity and the choice of using or not using an EBM that more accurately represents the heat uptake by the climate system are important for the results since these choices have an influence on the temperature response following an emission pulse and of the emissions stabilisation pathways generated in the model, see Johansson (2010). Shine et al. (2005, 2007) chose to calibrate their one-box EBM with a, compared to the one

7 Other

greenhouse gases and other forcers are also taken into account in the model, but we only present results related to CH4 and CO2 . For information on how other contributors to radiative forcing are handled see Johansson et al. (2006) and Johansson (2010).

8A

constant background atmosphere is assumed when calculating the GWP and the GTP. However, this gives identical results to those obtained assuming linearity and a non-constant background atmosphere.

134

Climatic Change (2012) 110:123–141

used in here, radically different value for the equivalent heat capacity. Consequently our numerical GTP value for CH4 differs from their value. The model runs from 2000 to 2400, with annual time-steps. Abatement of emissions is allowed after year 2010. Baseline emissions of the well-mixed GHGs in the period 2020–2100 are taken from the IPCC IS92a scenario (IPCC 1992). Years up to 2006 are based on historical data, and emissions estimates for 2007 to 2019 are obtained by using linear interpolation. After 2100 the baseline emissions are assumed to first grow slightly and then to stabilise and then, in the case of CO2 , eventually decline due to fossil fuel resource limitations. The model chooses abatement levels for the greenhouse gases in order to minimize the net present value abatement cost to meet a stabilisation limit for the global average surface temperature at 2◦ C above the pre-industrial level. Abatement costs are modelled with the aid of abatement cost functions, see Johansson (2010) for details. The choice of discount rate for long-term analysis is essential. This issue has been at the centre of the climate debate for a long time and was recently amplified in the aftermath of the Stern Review (Stern 2007). We will not here delve into the issue of what constitutes a proper long-term discount rate. Rather, we take note of the difficulties in choosing a long-term discount rate and analyse the model for different discount rates. In the base case a discount rate of 4% per year is used. Two alternative discount rates are also tested for, 1% and 7% per year. To generate the results presented in Section 5 we first run the numerical model in order to get the GCP values and obtain the year at which the temperature stabilizes at the 2◦ C limit. The stabilisation year is fed into the formulas for calculating the GTP, Eq. 18, and the CETP, Eqs. 16 and 17. We limit the results to the relative valuation of methane to carbon dioxide and leave the valuation of other greenhouse gases for future analysis.

5 Results—comparison and illustration The results presented in this section are for the three different assumptions on the discount rate. The choice of discount rate will affect the year at which the global average surface temperature stabilizes at the policy limit, which is in the cases studied here, a limit on the global average surface temperature at 2◦ C above the preindustrial level. The higher the discount rate, the more the abatement is postponed so as to minimize the net present value abatement costs. As a consequence, the year at which the 2◦ C limit is met in the model will take place earlier the higher the discount is (due to the larger emissions early on when the discount rate is high). 5.1 Base discount rate case Given a discount rate of 4%/year the stabilisation limit is met by 2103. The general pattern for all three metrics is that the relative value of CH4 to CO2 is relatively low when the stabilisation of the global average surface temperature is far off in time, see Fig. 2. As time moves closer to the stabilisation year the metric value for CH4 increases in all three cases. The reason for this pattern is that CH4 has a short atmospheric life time in comparison to CO2 . When emitting CH4 far in time from the

Fig. 2 The figure shows the relative valuation of methane to carbon dioxide for three different metrics, i.e., GTP, GCP, and CETP, as a function of time given a discount rate of 4%/year

Relative valuation of CH4 to CO2

Climatic Change (2012) 110:123–141

135 120 GCP 100 80

CETP GTP

60 40 20 0 2010

2030

2050

2070

2090

2110

2130

2150

Time (year)

stabilisation year only a small fraction of the maximum temperature response still prevails when stabilisation occurs. Due to the long atmospheric life time of a CO2 perturbation the temperature response following emissions of CO2 is much more persistent than for CH4 . As seen in Fig. 2, the CETP approximates the GCP rather well. As can also be seen in Fig. 2, the GTP is higher than both the GCP and the CETP. The reason for this is that the GTP does not take into account the long-term temperature response that occurs after stabilisation of the global average surface temperature, while this is reflected in the GCP and CETP. As discussed above, the temperature response for a CO2 emission pulse is more persistent than for a CH4 emission pulse. For this reason, the temperature response that takes place beyond the year at which the temperature stabilises is more important for the ACETP for CO2 than for CH4 . As a consequence the CETP metric value for CH4 is lower than the GTP metric value for CH4 . This can also be observed in Eqs. 16 and 18. The principle is much the same for the GCP as for CETP, although GCP also depends on the Lagrange multiplier of the temperature constraint. The reason why the GCP and CETP diverge in the decades just prior to stabilisation is that the Lagrange multiplier (shadow price) associated with the temperature stabilisation constraint, Eq. 3, takes on a high negative value the first date(s) the constraint bites, and thereafter becomes more stable, see Appendix. The relatively (in absolute terms) large initial Lagrange multiplier associated with the temperature constraint is a result of the thermal inertia of the climate system.9 A consequence of this (in absolute terms) large initial Lagrange multiplier is that it pushes the emission price for all gases upward prior to stabilisation. In addition, this effect on the emission price is more important for short-lived gases than for long-lived gases. The reason for this is that the emission prices of long-lived greenhouse gases are relatively more affected by the Lagrange multiplier associated with the temperature constraint in the long-run. As a consequence the GCP metric value for CH4 is higher than the CETP metric value prior to the year of stabilisation. The reason why the GCP is above the CETP after the year of stabilisation is that the Lagrange multiplier associated with the temperature constraint falls (in

9 See

den Elzen and van Vuuren (2007) and Johansson (2010) for analyses of radiative forcing overshoots and temperature stabilisation; although they do not analyse shadow prices.

136

Climatic Change (2012) 110:123–141

absolute terms) slightly with time after stabilisation has occurred, see Appendix. This feature will increase the CETP of a short-lived gas compared to when the Lagrange multiplier associated with the temperature constraint would have been constant over time. A constant Lagrange multiplier associated with the temperature constraint would have resulted in that the GCP and CETP would in principle be identical, as discussed in Section 2. 5.2 Low discount rate case With a low discount rate of 1% per year the stabilisation limit is not met in the model until 2131. As a consequence of the later stabilisation year, the GCP, CETP, and GTP values for CH4 are all lower early this century than they are in the base case with a 4% discount rate. An important reason is simply the longer time span left to the stabilisation year. In addition, both GCP and CETP depend on the discount rate, and as discussed above the temperature response of a CO2 emission pulse is more persistent than that of a CH4 emission pulse. Taken together, this implies that a low discount rate would lower the metric value for CH4 when GCP and CETP are used. The reason for this is that the lower the discount rate is the more important the long-term temperature response becomes for these two metrics. The importance of this effect can be seen in the difference in the valuation of CH4 using GTP or GCP/CETP and that the difference is more pronounced when using a low discount rate compared to when using a more midrange discount rate, compare Figs. 2 and 3. This result can also be understood by examining Eq. 16 where it is clear that the long-term response becomes more important for the CETP when a low discount rate is used. Note that the CETP approximates the GCP well under the assumption of a low discount rate while the GTP does not. 5.3 High discount rate case With a high discount rate set to 7% per year the stabilisation limit is met in the model in 2091. As is discussed in the last paragraph of Section 3, the CETP (and also the GCP) converges toward the GTP as the discount rate becomes large. This is clearly

120 Relative valuation of CH4 to CO2

Fig. 3 The figure shows the relative valuation of methane to carbon dioxide for three different metrics, i.e., GTP, GCP, and CETP, as a function of time given a discount rate of 1%/year

GCP 100 80

CETP GCP

60 40 20 0 2010

2030

2050

2070

2090

Time (year)

2110

2130

2150

Fig. 4 The figure shows the relative valuation of methane to carbon dioxide for three different metrics, i.e., GTP, GCP, and CETP, as a function of time given a discount rate of 7%/year

Relative valuation of CH4 to CO2

Climatic Change (2012) 110:123–141

137 120 GCP 100 80

CETP GTP

60 40 20 0 2010

2030

2050

2070

2090

2110

2130

2150

Time (year)

visible in Fig. 4, where relative valuation of CH4 to CO2 is more similar for GTP, CETP, and GCP when the discount rate is 7% than when it is 1% or 4%.

6 Discussion and conclusion Shine et al. (2007) and Tol et al. (2008) observed that the GTP and GCP metrics share many features. The analysis in this paper substantiates those observations. We also show that the GCP approach does not in a direct way depend on abatement cost assumptions. In addition, we suggest a new metric, CETP, which is based on a approximation of the GCP. This metric turns out to be related to the GTP and the formulation of GWP with discounting as suggested by Lashof and Ahuja (1990). Not only does this metric provide a better approximation of the GCP than the GTP does, it also gives an estimate of a trade-off ratio in the long-term; GTP fails to do so. Furthermore, we have shown that both the GCP and the CETP depend on the discount rate, while the GTP does not. However, the GCP and the CETP converge to the GTP when high discount rates are assumed. Based on this, one may argue that GTP is a good approximation of the GCP given high discount rates. However, applying high discount rates on impacts occurring over hundreds of years can be questioned on the ground of intergenerational equity, see for example Azar and Sterner (1996) and Stern (2007). An important extension to the analysis performed in this paper would be to test the CETP’s ability to approximate the GCP under more general circumstances than studied here and use several different optimising climate-economy models to estimate the GCP. Finally, it is well known that the choice of metric is not only a scientific issue, but is in the end a question of political choice and policy objective. The relevant endpoints and/or policy targets are matters of value judgments, not science. However, policymakers have historically viewed the global warming potentials as a purely scientific issue and not an issue for policy (Shackley and Wynne 1997; Smith 2003). Nevertheless, a “perfect” metric cannot be constructed as long as the policy objectives are not clearly decided upon. The Climate Convention (UNFCCC 1992) and the Copenhagen Accord (UNFCCC 2009) provide some guidance on what the overall policy objective is. It is clear from the UNFCCC that “stabilization

138

Climatic Change (2012) 110:123–141

of greenhouse gas concentrations in the atmosphere at a level that would prevent dangerous anthropogenic interference with the climate system” is the overall target. The focus on long-term stabilisation has been reinforced by the 2◦ C limit stated in the Copenhagen Accord (UNFCCC 2009). Further, it is stated in UNFCCC (1992) that “policies and measures to deal with climate change should be cost-effective so as to ensure global benefits at the lowest possible cost.” With these policy goals in mind, it is clear that the GCP and consequently the CETP are suitable tools to make comparisons of emissions of different greenhouse gases possible. GTP may also do the job quite well due to the similarities with the GCP and the CETP, but as seen in Section 5 this depends very much on the discount rate used. However, in UNFCCC (1992) it is also stated that the “stabilization of greenhouse gas concentrations” should be “achieved within a time frame sufficient to allow ecosystems to adapt naturally to climate change.” Hence, limitation on the rate of climate change is an aim in itself. All three metrics as they are presented here fail to incorporate any aspects of the rate of climate change. Constraints on the rate of change on, for example, the temperature can be taken into account in the construction of the GCP, as it is done in Manne and Richels (2001) and Johansson et al. (2006). However, such an approach has to depend on complex numerical models and the transparency gained with an analytical approach such as used for the GTP or the CETP would be lost. Finally, even though the current GWP integrated over 100 years does not seem to be constructed with the policy goals of the UNFCCC in mind, it tends to perform quite well and its performance seems to be robust to a range of policy alternatives and socio-economic and geophysical uncertainties, see Johansson et al. (2006). Whether changing metrics is worthwhile is consequently still an open question and our intention here was not to provide an answer concerning that. Rather, the intention was to clarify the relationship between the GCP and GTP and to suggest an analytical approximation of the GCP, the Cost-Effective Temperature Potential, CETP. Acknowledgements I would like to thank Brian O’Neill and Paulina Essunger as well as the reviewers, Keith Shine and two anonymous, for constructive comments. Funding from the Swedish Energy Agency and Göteborg Energi Research Foundation is gratefully acknowledged. This study was performed in part during Daniel Johansson’s stay as Guest Research Scholar at IIASA autumn 2007. He would like to thank IIASA for their hospitality.

Appendix—The Lagrange multiplier associated with the temperature constraint The Lagrange multiplier associated with the temperature constraint in the optimization model is zero as long as the constraint, in Eq. 3, is slack. When the constraint starts to bind the Lagrange multiplier takes initially on a large negative value and then become more stable, see Fig. 5. The value of the Lagrange multiplier depends on parameter assumptions in the model and the setup of control and state variables and equations. The particular result that the Lagrange multiplier take on a large negative value initially and then becomes more stable is dependent on the inertia caused by the thermal inertia in the energy balance model coupled with the time dynamics of the gas cycles. If the thermal inertia of the climate system would be zero in the model used here the Lagrange multiplier associated with the temperature constraint would be relatively stable in periods when the constraint is binding. Although note that

Fig. 5 The figure shows the value of the Lagrange multiplier associated with the temperature constraint in the model given a discount rate of 4%/year and presented in current value terms

Lagrange multiplier associated with the temperature constraint (Trillion US$/K)

Climatic Change (2012) 110:123–141

139 0 2000 -100

2050

2100

2150

2200

2250

2300

-200 -300 -400 -500 -600 -700

Time (year)

the fact that the Lagrange multiplier associated with the temperature constraint is relatively stable when the constraint is binding may not hold under more general circumstances, see Eckaus (1992) for a related discussion. The Lagrange multiplier associated with the temperature constraint take on different values than what is shown in Fig. 5 when a different discount rate than 4% is used in the optimization. However, the qualitative nature of the Lagrange multiplier remains the same, i.e., the Lagrange multiplier takes initially on a large negative value once the constraint become binding and then becomes more stable.

References Andrews DR, Allen MR (2008) Diagnosis of climate models in terms of transientclimate response and feedback response time. Atmos Sci Lett 9:7–12 Azar C, Sterner T (1996) Discounting and distributional considerations in the context of global warming. Ecol Econ 19(2):169–184 Baumol WJ, Oates WE (1988) The theory of environmental policy, 2nd edn. Cambridge University Press, Cambridge Dasgupta P (2008) Discounting climate change. J Risk Uncertainty 37(2–3):141–169 den Elzen MGJ, van Vuuren DP (2007) Peaking profiles for achieving long-term temperature targets with more likelihood at lower costs. PNAS 104(46):17931–17936 Eckaus RS (1992) Comparing the effects of greenhouse gas emissions on global warming. Energy J 13:25–35 Fisher DA, Hales CH, Wang W-C, Ko MKW, Dak SN (1990) Model calculations of the relative effects of CFCs and their replacements on global warming. Nature 344:513–516 Forster P, Ramaswamy V, Artaxo P, Berntsen T, Betts R, Fahey DW, Haywood J, Lean J, Lowe DC, Myhre G, Nganga J, Prinn R, Raga G, Schulz M, Van Dorland R (2007) Changes in atmospheric constituents and in radiative forcing. In: Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, Tignor M, Miller HL (eds) Climate change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change. Cambridge University Press, Cambridge Fuglestvedt JS, Berntsen TK, Godal O, Sausen R, Shine KP, Skodvin T (2003) Metrics of climate change: assessing radiative forcing and emission indices. Clim Change 58:267–331 Fuglestvedt JS, Shine KP, Berntsen T, Cook J, Lee JS, Stenke A, Bieltvedt Skeie R, Velders G, Waitz I (2010) Transport impacts on atmosphere and climate: metrics. Atmos Environ 44:4648–4677 Hammitt JK, Jain AK, Adams JL, Wuebbles DJ (1996) A welfare-based index for assessing environmental effects of greenhouse-gas emissions. Nature 381:301–303 Hayhoe K, Jain A, Pitcher H, MacCracken C, Gibbs M, Wuebbles D, Harvey R, Kruger D (2000) Climate change policy - costs of multigreenhouse gas reduction targets for the USA. Science 286:905–906

140

Climatic Change (2012) 110:123–141

IPCC (1992) In: Houghton JT, Callander BA, Varney SK (eds) 1992 IPCC supplement. Cambridge University Press, Cambridge Johansson DJA, Persson UM, Azar C (2006) The cost of using global warming potentials: analysing the trade off between CO2 , CH4 , and N2 O. Clim Change 77:291–309 Johansson DJA (2010) Temperature stabilization, ocean heat uptake and radiative forcing overshoot profiles. Clim Change. doi:10.1007/s10584-010-9969-4 Johansson DJA, Persson UM, Azar C (2008) Uncertainty and learning: implications for the trade-off between short-lived and long-lived greenhouse gases. Clim Change 88:293–308 Kandlikar M (1996) Indices for comparing greenhouse gas emissions: integrating science and economics. Energy Econ 18:265–281 Lashof DA, Ahuja DR (1990) The relative contributions of greenhouse gas emissions to global warming. Nature 344:529–531 Lind RC (ed) (1982) Discounting for time and risk in energy policy. Johns Hopkins University Press, Baltimore Manne AS, Richels RG (2001) An alternative approach to establishing trade-offs among greenhouse gases. Nature 410(6829):675–677 Maier-Reimer E, Hasselmann K (1987) Transport and storage of CO2 in the ocean – an inorganic ocean-circulation carbon cycle model. Climate Dyn 2(2):63–90 Michaelis P (1992) Global warming: efficient policies in the case of multiple pollutants. Environ Resour Econ 2:61–77 O’Neill BC (2000) The jury is still out on global warming potentials. Clim Change 44:427–443 O’Neill BC (2003) Economics, natural science, and the costs of global warming potentials. Clim Change 58(3):251–260 Plattner G-K, Stocker T, Midgley P, Tigno M (2009) IPCC expert meeting on the science of alternative metrics Reilly JM, Richards KR (1993) Climate change damage and the trace gas index issue. Environ Resour Econ 3:41–61 Reilly J, Prinn R, Harnisch J, Fitzmaurice J, Jacoby H, Kicklighter D, Melillo J, Stone P, Sokolov A, Wang C (1999) Multi gas assessment of the Kyoto protocol. Nature 401:549–555 Richels RG, Manne AS, Wigley TML (2007) Moving beyond concentrations: the challenge of limiting temperature change. In: Schlesinger ME, Kheshgi HS, Smith J, Chesnaye FDL, Reilly JM, Wilson T, Kolstad C (eds) Human-induced climate change: an interdisciplinary assessment. Cambridge University Press, Cambridge Schmalensee R (1993) Comparing greenhouse gases for policy purposes. Energy J 14:245–255 Shackley S, Wynne B (1997) Global warming potentials: ambiguity or precision as an aid to policy? Clim Res 8:89–106 Shine K, Derwent RG, Wuebbles DJ, Morcrette JJ (1990) Radiative forcing of climate. In: Houghton JT et al (eds) Climate change: the IPCC scientific assessment. Cambridge University Press, Cambridge Shine KP, Fuglestvedt JS, Hailemariam K, Stuber N (2005) Alternatives to the global warming potential for comparing climate impacts of emissions of greenhouse gases. Clim Change 68(3):281–302 Shine KP, Berntsen TK, Fuglestvedt JS, Bieltvedt SR, Stuber N (2007) Comparing the climate effect of emissions of short and long lived climate agents. Philos Trans R Soc A 365(1856):1903–1914 Smith SJ (2003) The evaluation of greenhouse gas indices. Clim Change 58(3):261–265 Smith SJ, Wigley TML (2000) Global warming potentials: 1. Climatic implications of emissions reductions. Clim Change 44:445–457 Stern NH (2007) The economics of climate change: the stern review. Cambridge University Press, Cambridge Sterner T (2002) Policy instruments for environmental and natural resource management. RFF, Mishawaka Tanaka K, O’Neill BC, Rokityanskiy D, Obersteiner M, Tol RSJ (2009) Evaluating global warming potentials with historical temperature. Clim Change 96(4):443–466 Tol RSJ, Berntsen TK, O’Neill BC, Fuglestvedt JS, Shine KP, Balkanski Y, Makra L (2008) Metrics for aggregating the climate effect of different emissions: a unifying framework. ESRI Working paper No. 257 UNFCCC (1992) United Nations Framework Convention on Climate Change UNFCCC (1997) Kyoto protocol to the United Nations Framework Convention on Climate Change UNFCCC (2008) Consideration of relevant methodological issues. Document: FCCC/KP/AWG/ 2008/L.14 UNFCCC (2009) Copenhagen Accord

Climatic Change (2012) 110:123–141

141

Van Vuuren DP, Weyant J, de la Chesnaye F (2006) Multi-gas scenarios to stabilize radiative forcing. Energy Econ 28:102–120 Weyant JP, de la Chesnaye F, Blanford GJ (2006) Overview of EMF-21: multigas mitigation and climate policy. Energy Journal, Special Issue #3 – Multi-greenhouse gas mitigation and climate policy Wigley TML (1998) The Kyoto Protocol: CO2 , CH4 and climate implications. Geophys Res Lett 25(13):2285–2288

Economics- and physical-based metrics for comparing ... - Springer Link

May 3, 2011 - Here we present a new analytic metric, the Cost- ... Environmental Economics Unit, Department of Economics, School of Business, ...... mind, it is clear that the GCP and consequently the CETP are suitable tools to make.

371KB Sizes 1 Downloads 239 Views

Recommend Documents

Ecological economics: themes, approaches, and ... - Springer Link
traditional environmental (and resource) economics. (ERE) ... Reg Environ Change (2001) 2:13±23. 13 .... ronmental products and services back to solar energy,.

Learning Contextual Metrics for Automatic Image ... - Springer Link
Mi) dx. (6). In our metric learning method, we regularize each metric by Euclidean distance function. So we add the following single regularization term into our ...

Two models of unawareness: comparing the object ... - Springer Link
Dec 1, 2010 - In this paper we compare two different approaches to modeling unawareness: the object-based approach of Board and Chung (Object-based unawareness: theory and applications. University of Minnesota, Mimeo, 2008) and the subjective-state-s

Two models of unawareness: comparing the object ... - Springer Link
Dec 1, 2010 - containing no free variables.3 We use OBU structures to provide truth conditions only ..... can envisage an extension where unawareness of properties is also modeled. ..... are the identity when domain and codomain coincide.

The Incredible Economics of Geoengineering - Springer Link
Dec 6, 2007 - As I shall explain in this paper, its future application seems more likely than not. ... because the incentives for countries to experiment with ...

MAJORIZATION AND ADDITIVITY FOR MULTIMODE ... - Springer Link
where 〈z|ρ|z〉 is the Husimi function, |z〉 are the Glauber coherent vectors, .... Let Φ be a Gaussian gauge-covariant channel and f be a concave function on [0, 1].

Conflict and Health - Springer Link
Mar 14, 2008 - cle.php?art_id=5804]. May 30, 2006. 21. Tin Tad Clinic: Proposal for a Village-Based Health Care. Project at Ban Mai Ton Hoong, Fang District, ...

Exploiting Graphics Processing Units for ... - Springer Link
Then we call the CUDA function. cudaMemcpy to ..... Processing Studies (AFIPS) Conference 30, 483–485. ... download.nvidia.com/compute/cuda/1 1/Website/.

Evidence for Cyclic Spell-Out - Springer Link
Jul 23, 2001 - embedding C0 as argued in Section 2.1, this allows one to test whether object ... descriptively head-final languages but also dominantly head-initial lan- ..... The Phonology-Syntax Connection, University of Chicago Press,.

Tinospora crispa - Springer Link
naturally free from side effects are still in use by diabetic patients, especially in Third .... For the perifusion studies, data from rat islets are presented as mean absolute .... treated animals showed signs of recovery in body weight gains, reach

Chloraea alpina - Springer Link
Many floral characters influence not only pollen receipt and seed set but also pollen export and the number of seeds sired in the .... inserted by natural agents were not included in the final data set. Data were analysed with a ..... Ashman, T.L. an

GOODMAN'S - Springer Link
relation (evidential support) in “grue” contexts, not a logical relation (the ...... Fitelson, B.: The paradox of confirmation, Philosophy Compass, in B. Weatherson.

Bubo bubo - Springer Link
a local spatial-scale analysis. Joaquın Ortego Æ Pedro J. Cordero. Received: 16 March 2009 / Accepted: 17 August 2009 / Published online: 4 September 2009. Ó Springer Science+Business Media B.V. 2009. Abstract Knowledge of the factors influencing

Quantum Programming - Springer Link
Abstract. In this paper a programming language, qGCL, is presented for the expression of quantum algorithms. It contains the features re- quired to program a 'universal' quantum computer (including initiali- sation and observation), has a formal sema

BMC Bioinformatics - Springer Link
Apr 11, 2008 - Abstract. Background: This paper describes the design of an event ontology being developed for application in the machine understanding of infectious disease-related events reported in natural language text. This event ontology is desi

Isoperimetric inequalities for submanifolds with ... - Springer Link
Jul 23, 2011 - if ωn is the volume of a unit ball in Rn, then. nnωnVol(D)n−1 ≤ Vol(∂D)n and equality holds if and only if D is a ball. As an extension of the above classical isoperimetric inequality, it is conjectured that any n-dimensional c

Probabilities for new theories - Springer Link
where between 0 and r, where r is the prior probability that none of the existing theories is ..... theorist's internal programming language"(Dorling 1991, p. 199).

A Process Semantics for BPMN - Springer Link
Business Process Modelling Notation (BPMN), developed by the Business ..... In this paper we call both sequence flows and exception flows 'transitions'; states are linked ...... International Conference on Integrated Formal Methods, pp. 77–96 ...

Unsupervised Learning for Graph Matching - Springer Link
Apr 14, 2011 - Springer Science+Business Media, LLC 2011. Abstract Graph .... tion as an integer quadratic program (Leordeanu and Hebert. 2006; Cour and Shi ... computer vision applications such as: discovering texture regularity (Hays et al. .... fo

Candidate quality - Springer Link
didate quality when the campaigning costs are sufficiently high. Keywords Politicians' competence . Career concerns . Campaigning costs . Rewards for elected ...

Mathematical Biology - Springer Link
Here φ is the general form of free energy density. ... surfaces. γ is the edge energy density on the boundary. ..... According to the conventional Green theorem.