You are on page 1of 14

Annual Reviews in Control 47 (2019) 7–20

Contents lists available at ScienceDirect

Annual Reviews in Control


journal homepage: www.elsevier.com/locate/arcontrol

Feedback, dynamics, and optimal control in climate economicsR


Christopher M. Kellett a,∗, Steven R. Weller a, Timm Faulwasser b, Lars Grüne c,
Willi Semmler d
a
School of Electrical Engineering and Computing, University of Newcastle, Callaghan, New South Wales 2308, Australia
b
Institute for Automation and Applied Informatics, Karlsruhe Institute of Technology, 76344 Eggenstein-Leopoldshafen, Germany
c
Mathematisches Institut, Universität Bayreuth, 95440 Bayreuth, Germany
d
New School for Social Research, New York, United States of America, and University of Bielefeld, Germany, and International Institute for Applied Systems
Analysis, Laxenburg, Austria

a r t i c l e i n f o a b s t r a c t

Article history: For his work in the economics of climate change, Professor William Nordhaus was a co-recipient of the
Received 19 December 2018 2018 Nobel Memorial Prize for Economic Sciences. A core component of the work undertaken by Nord-
Revised 19 March 2019
haus is the Dynamic Integrated model of Climate and Economy, known as the DICE model. The DICE
Accepted 3 April 2019
model is a discrete-time model with two control inputs and is primarily used in conjunction with a
Available online 16 May 2019
particular optimal control problem in order to estimate optimal pathways for reducing greenhouse gas
Keywords: emissions. In this paper, we provide a tutorial introduction to the DICE model and we indicate challenges
Optimal control and open problems of potential interest for the systems and control community.
Nonlinear systems
Economics
© 2019 Elsevier Ltd. All rights reserved.
Geophysical dynamics
Climate change

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2. The DICE model and methodology . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1. DICE dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2. DICE “Exogenous” states . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3. Economic model. . . . . . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4. The damages function . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.5. Climate model . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.6. Carbon model. . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7. Welfare maximization . . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.8. The social cost of carbon dioxide (SC-CO2 ) . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.9. DICE, PAGE, and FUND . . . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3. Receding horizon solution to DICE . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4. State and input rate constraints . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5. Systems and control in climate-economy assessment: Opportunities and progress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.1. Uncertainty quantification . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.2. Identification of predictive climate models . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.3. Modular tools for simulation and optimization. ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.4. System theoretic analysis . . . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6. Summary and concluding remarks . . . . . . . . . . . . . . . . ........................ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

R
This paper was not presented at any IFAC meeting.

Corresponding author.
E-mail addresses: chris.kellett@newcastle.edu.au (C.M. Kellett), steven.weller@newcastle.edu.au (S.R. Weller), timm.faulwasser@ieee.org (T. Faulwasser), lars-gruene@uni-
bayreuth.de (L. Grüne), semmlerw@newschool.edu (W. Semmler).

https://doi.org/10.1016/j.arcontrol.2019.04.003
1367-5788/© 2019 Elsevier Ltd. All rights reserved.
8 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Appendix A. Default initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Appendix B. Default parameter values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

1. Introduction The IAM community presently pre-dominantly employs simplis-


tic Monte Carlo-based methods to emulate the impact of paramet-
In the absence of deep and sustained reductions in greenhouse ric uncertainty on the SC-CO2 (IWG, 2013), whilst recognizing that
gas emissions, the overwhelming scientific consensus points to such an approach can lead to contradictory policy advice (Crost &
global warming of several degrees Celsius by 2100. Warming of Traeger, 2013). On the other hand, known resolutions to this major
this magnitude poses profound risks to both human society and deficiency are computationally intractable (e.g., stochastic dynamic
natural ecosystems (IPCC, 2013b). In response to these risks, in programming Crost & Traeger, 2013) and do not easily accommo-
late 2015 at the United Nations Climate Change Conference gov- date enhanced geophysical models. At a time when governments,
ernments around the world committed to urgent reductions in financial bodies (Farid, 2016), business (World Bank, 2014), and
human-caused emissions of greenhouse gases, most notably car- even emissions-intensive industries (BP2, 2017; Guardian, 2016)
bon dioxide (CO2 ), in order to limit the increase in global av- are demanding a price on carbon, it is imperative that the short-
erage temperature to well below 2 ◦ C relative to pre-industrial comings in quantifying uncertainty in SC-CO2 estimates be rec-
levels. tified. Indeed, this is considered to be a problem of the utmost
With global average warming of 1 ◦ C having already been real- importance (National Academies, 2016, 2017; Hope, 2015; Stern,
ized, constraining temperature increases below agreed target lev- 2016).
els will require careful control of future emissions, with the In- Our contribution in this paper is three-fold. First, we provide a
tergovernmental Panel on Climate Change (IPCC) special report on complete and replicable specification of the DICE model, with ac-
Global Warming of 1.5 ◦ C (IPCC, 2018) indicating that remaining companying code available for download at Faulwasser, Kellett, and
below 1.5◦ C will require net-zero CO2 emissions by about 2050. Weller (2018b). Second, we summarize some of our recent work
Complicating this task are large uncertainties regarding the speed and update the numerical results to account for updated param-
and extent of warming in response to elevated atmospheric CO2 eters released by Nordhaus (2017b). Third, we indicate challenges
concentrations, coupled with the need for a policy response that and open problems of potential interest for the systems and con-
balances reduced economic consumption today with avoided (and trol community.
discounted) economic damages of an uncertain magnitude in the The paper is organized as follows. Section 2 provides a tutorial
future. description of the DICE model and of its usage in the context of
To quantify the damages from anthropogenic emissions of heat- computing the Social Cost of Carbon Dioxide. Section 3 describes
trapping greenhouse gases, specifically CO2 , economists model the the benefits of receding horizon control for the DICE model both
dynamics of climate–economy interactions using Integrated Assess- as a numerical solution technique and as a way to investigate the
ment Models (IAMs), which incorporate mathematical models of impact of parametric uncertainty. Section 4 considers the impact
phenomena from both economics and geophysical science. Possi- of placing constraints on the atmospheric temperature and mitiga-
bly the first IAM in the area of climate economics was proposed tion rate constraints. Section 5 indicates potential challenges and
by William Nordhaus in Nordhaus (1975). Subsequently, Nordhaus opportunities of particular relevance for the systems and control
proposed the Dynamic Integrated model of Climate and Economy community. Section 6 provides some brief concluding remarks.
(DICE) in Nordhaus (1992), with regular refinements and param-
eter updates, such as Nordhaus (2008, 2010, 2014, 2017b) and 2. The DICE model and methodology
Nordhaus and Sztorc (2013). Largely for this body of work, Nord-
haus was awarded the 2018 Nobel Memorial Prize in Economic Sci- There are three dominant IAMs used for the calculation of
ences. the Social Cost of Carbon Dioxide (Bonen, Semmler, & Klasen,
A central role for IAMs is to estimate the Social Cost of Carbon 2014; IWG, 2013): the previously mentioned DICE (Nordhaus, 2014;
Dioxide (SC-CO2 ), defined as the dollar value of the economic dam- Nordhaus & Sztorc, 2013), Policy Analysis of the Greenhouse Ef-
age caused by a one metric tonne increase in CO2 emissions to the fect (PAGE) (Hope, 2013), and Climate Framework for Uncertainty,
atmosphere. The SC-CO2 is used by governments, companies, and Negotiation, and Distribution (FUND) (Anthoff & Tol, 2013). As we
international finance organizations as a key quantity in all aspects will describe below, the DICE model and methodology consists of
of climate change mitigation and adaptation, including cost-benefit an optimal control problem for a discrete time nonlinear system. A
analyses, emissions trading schemes, carbon taxes, quantification brief description of PAGE and FUND is provided in Section 2.9.
of energy subsidies, and modelling the impact of climate change on
financial assets, known as the value at risk (EIU, 2015). The SC-CO2 2.1. DICE dynamics
therefore underpins trillions of dollars worth of investment deci-
sions (Hope, 2015). It should be noted that there exist different open-source im-
A commonly used derivation for the SC-CO2 solves an open- plementations of DICE. While Nordhaus maintains an open-source
loop optimal control problem to determine economically optimal GAMS implementation (Nordhaus, 2017c),1 a subset of the authors
CO2 emissions pathways. The open-loop use of IAMs for decision- of this paper have recently published open-source DICE code that
making, however, disregards crucial uncertainties in both geophys- runs in Matlab (Faulwasser et al., 2018b; Kellett, Faulwasser, &
ical and economic models. As a consequence, current SC-CO2 esti- Weller, 2016) (see also Faulwasser, Kellett, & Weller, 2018a).
mates range from US$11 per tonne of CO2 to US$63 per tonne of It is important to note at the outset that there is not a definitive
CO2 or higher, and hence the SC-CO2 fails to reflect the true eco- statement of the DICE model. Rather, there are two primary sources
nomic risk posed by CO2 emissions, seriously compromising the in the form of a user’s manual (Nordhaus & Sztorc, 2013) (updated
accuracy of the SC-CO2 as a price on carbon for the purposes of in Nordhaus, 2017a) and the available code itself (both the manual
climate change mitigation and adaptation (Howarth, Gerst, & Bor-
suk, 2014; Otto, Frame, Otto, & Allen, 2015; Pizer, 2014). 1
We refer to https://www.gams.com for details on GAMS.
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 9

Fig. 1. Block diagram of DICE.

   
MAT (i )
T (i + 1 ) = T T (i ) + BT F2× log2 MAT,1750
+ FEX (i ) , (CLI)

 
M (i + 1 ) = M M (i ) + BM σ (i )(1 − μ(i ))A(i )K (i )γ L(i )1−γ + ELand (i ) , (CAR)

 1
 
K (i + 1 ) = K K (i ) +  1 − θ 1 ( i ) μ ( i ) θ2 A ( i ) K ( i ) γ L ( i ) 1 − γ s ( i ) , (CAP)
1 + a2 TAT (i )a3

 1+L g
a
L (i + 1 ) = L (i ) , (POP)
1+L(i )

A (i )
A (i + 1 ) = , (TFP)
1 − gA exp(−δA (i − 1 ))

 
σ (i + 1 ) = σ (i ) exp −gσ (1 − δσ )(i−1)  , (EI)

and the code are available at Nordhaus, 2017c). Additional expla-


nations and, occasionally, equations can be found in various other We describe each of the modeling blocks in Fig. 1 in
sources including Nordhaus (2014, 2017b). However, these sources Sections 2.2–2.6 below. Here, however, we note that the model is
are not consistent with each other and, in fact, the specification nonlinear and time-varying. The model assumes two control in-
in Nordhaus and Sztorc (2013) is incomplete. Furthermore, there puts: the savings rate s and the mitigation or abatement rate μ.
are some minor inconsistencies between text and equations in The first of these we describe more fully in Section 2.3. The latter
Nordhaus and Sztorc (2013). For completeness, and with the aim is the rate at which mitigation of industrial carbon dioxide emis-
of presenting the DICE model and methodology in a way that sions occurs.
can be independently implemented, we necessarily deviate from The model uses a time-step of 5 years, starting in the year 2015
Nordhaus and Sztorc (2013) and Nordhaus (2017c). However, the for DICE2016 (or 2010 for DICE2013). Take the discrete time index
subsequent impact on the numerical results when using the default i ∈ N, the sampling rate  = 5, and the initial time t0 = 2015 (or
parameters (included in the Appendices) is not significant. 2010) so that
One further note before proceeding to the model description:
t = t0 +  · ( i − 1 ) (1)
while the most recent version of the model is DICE2016 (as used
in, for example, Nordhaus, 2017b), the previous version of the and hence t ∈ {2015, 2020, 2025, . . .}.
model, DICE2013, has been widely used in the literature. Impor-
tantly, the move from DICE2013 to DICE2016 involves essentially 2.2. DICE “Exogenous” states
no structural changes. Rather, the model update involves updates
on initial conditions and most of the model parameters. For ease of The states for population L, total factor productivity A (which
reference, we provide initial conditions and model parameters for is a measure of technological progress), and carbon intensity of
both DICE2013 and DICE2016 in Appendix A and Appendix B (Table economic activity σ , are frequently referred to as “exogenous vari-
2). The Matlab DICE code (Faulwasser et al., 2018b; Kellett et al., ables”. This is due to the fact that they are not influenced by the
2016) implements both parameter sets. states for climate, carbon, or capital, which are frequently referred
The dynamics of the DICE IAM (Nordhaus & Sztorc, 2013) are to as “endogenous variables” (see Fig. 1).
given by equations (CLI)–(EI) and the inter-relationships shown As mentioned, there are some inconsistencies in the published
in Figure 1. Note that the equation labels are descriptive, with literature with regards to the form of these states. For the sake of
(CLI) describing the climate (or temperature) dynamics; (CAR) de- completeness and to remain close to the numerical results gen-
scribing the carbon cycle dynamics; (CAP) describing capital erated by Nordhaus (2017c), we present and use the exogenous
(or economic) dynamics; (POP) providing population dynamics; states as defined in Nordhaus (2017c). Background information on
(TFP) giving the dynamics of total factor productivity; and (EI) de- the parameters and functional form of these expressions can be
scribing the emissions intensity of economic activity. found in Nordhaus (2007, 2008), and Nordhaus and Sztorc (2013).
10 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

 1  
The population model (POP) is referred to as the Hassell Model
Q (i ) = 1 − θ 1 ( i ) μ ( i ) θ2 Y ( i )
(Hassell, 1975). Total factor productivity (TFP) yields a logistic-type 1 + a2 TAT (i )a3
function; i.e., the total factor productivity is monotonically increas-  1
 
ing with a decreasing growth rate. Carbon intensity of economic = 1 − θ1 (i )μ(i )θ2 · A(i )K (i )γ L(i )1−γ , (6)
1 + a2 TAT (i )a3
activity (EI) is similar to total factor productivity in that it is a
monotonically decreasing function with a decreasing decrease rate. where θ 1 is as defined in (2). Observe the component of net eco-
The quantities L(1 ) = L0 and A(1 ) = A0 are prescribed initial condi- nomic output in (CAP).
tions for the global population and total factor productivity in the Net economic output can then be split between consumption
base year. and investment
An estimate for the initial emissions intensity of economic ac- Q (i ) = C (i ) + I (i ) (7)
tivity σ (1 ) = σ0 can be calculated as the ratio of global industrial
emissions to global economic output. The estimate of σ 0 can be and the savings rate is defined as
further refined by estimating the mitigation rate in the base year. I (i )
s (i ) = . (8)
In other words, with base year emissions e0 , base year economic Q (i )
output q0 , and an estimated base year mitigation rate μ0 , we can
The economic dynamics (CAP) are a capital accumulation model
estimate σ0 = q (1−0 μ ) .
e
0 0 where capital depreciates according to
An estimate of the cost of mitigation efforts is given by
K =. (1 − δK ) (9)
pb
θ1 ( i ) = (1 − δ pb )i−1 · σ (i ). (2)
10 0 0 · θ2 and is replenished by investment I in the form of the product of
the savings rate and net economic output; i.e.,
Here, pb represents the price of a backstop technology that can re-
move carbon dioxide from the atmosphere. Note that this equation K (i + 1 ) = K K (i ) +  · I (i )
embeds the assumption that the cost of such technology will de- = K K (i ) +  · Q (i )s(i ) (10)
crease over time (since δ pb ∈ (0, 1)) and will be proportional to the
and substitution of (6) into (10) yields (CAP). The savings rate s is
emissions intensity of economic activity.
the second of the two control inputs.
The remaining two exogenous signals are given by
 
f1 − f0 2.4. The damages function
FEX (i ) = f0 + min f1 − f0 , (i − 1 ) , (3)
tf
One of the two most contentious elements3 in climate
economics is the specification of the damages function (see
ELand (i ) = EL0 · (1 − δEL )i−1 . (4) Bonen et al., 2014). This stems from the inherent difficulty of mod-
eling in an application where experimentation is simply not possi-
The signals FEX and ELand are estimates of the effect of greenhouse ble and the fact that rising temperatures will have different local
gases other than carbon dioxide and the emissions due to land use effects. Hence, different researchers have proposed many different
changes, respectively. damage functions and levels (Bonen et al., 2014). The specific form
Numerical values for all parameters can be found in the ap- of the damages function in DICE, as shown in (CAP), is
pendix as well as in the accompanying code (Faulwasser et al.,
1
2018a). (11)
1 + a2 TAT (i )a3
where, with a3 = 2, the parameter a2 is calibrated to yield a loss
2.3. Economic model
of 2% at 3 ◦ C (see Nordhaus, 2017a for the calibration of this and
other parameters).
We now turn to the economic component of the DICE inte-
While a full discussion of the appropriateness of (11) is beyond
grated assessment model. In summary, the DICE model assumes a
the scope of this article, it is worthwhile noting that it has been
single global economic “capital”. Capital depreciates and is replen-
vigorously argued that the above calibration of 2% loss at 3 ◦ C is
ished by investment. The amount available to invest is some frac-
unreasonably low if it is to be consistent with currently available
tion of the net economic output which can be derived from the
climate science (Stern, 2013). Recent efforts to empirically estimate
gross economic output. This is a standard economic growth model
climate damages can be found in Hsiang et al. (2017).
(see Acemoglu, 2009 for a comprehensive treatment of such mod-
els).
2.5. Climate model
Gross economic output is the product of three terms; the total
factor productivity A; capital K; and labor L approximated by the
The climate or temperature dynamics used in DICE are de-
global population. Additionally, capital and labor contribute at dif-
rived from a two-layer energy balance model (Geoffroy, Saint-
ferent levels given by a constant called the capital elasticity γ ∈ [0,
Martin, Bellon, Voldoire, & Olivié, 2013; Gregory, 20 0 0; Schnei-
1]; that is, gross economic output is given by2
der & Thompson, 1981). In particular, a simple explicit Euler dis-
Y ( i ) = A ( i ) K ( i ) γ L ( i ) 1 −γ . (5) cretization is applied to an established continuous-time energy
balance model to obtain (CLI). However, the implementation in
Note the expression corresponding to gross economic output in Nordhaus (2017c) is not strictly causal in that the atmospheric
(CAR) (see also (20) below). temperature at the next time step depends on the radiative forcing
Net economic output, Q, is gross economic output, Y, reduced at the next time step. This was previously observed in Cai, Judd,
by two factors: (1) climate damages from rising atmospheric tem- and Lontzek (2012) and Calel, Stainforth, and Dietz (2015). We ex-
perature, and (2) the cost of efforts towards mitigation: plicitly provide the derivation of the model here for future ref-
erence. Furthermore, we note that using the causal model below,
2
The quantity in (5) is referred to as a Cobb-Douglas Production Function with
3
Hicks-Neutral technological progress (see Acemoglu, 2009, p. 36, p. 58). The other being the discount rate discussed below.
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 11

as opposed to the version in Nordhaus (2017c), has a negligible carbon doubling. Ignoring the contribution of the exogenous forc-
quantitative impact on the numerical results obtained using the ing FEX , the radiative forcing is given by
model.

MAT (t )
The two layers in the energy balance model are the combined F (t ) = F2× log2 (16)
atmosphere, land surface, and upper ocean (simply referred to as MAT, 1750
the atmospheric layer in what follows) and the lower ocean. We where MAT, 1750 is the atmospheric mass of carbon in the year 1750.
denote these two states by TAT and TLO , respectively, and the zero Doubling the value of atmospheric carbon from pre-industrial
reference is taken as the temperature in the year 1750. With F(t) equilibrium yields radiative forcing of
denoting the radiative forcing at the top of atmosphere due to the

enhanced greenhouse effect, the (continuous-time) dynamics for 2MAT, 1750
F2× log2 = F2× . (17)
these states are given by MAT, 1750
d TAT (t ) The Equilibrium Climate Sensitivity (ECS) is defined as the
CAT = F (t ) − λTAT (t ) − γ (TAT (t ) − TLO (t ) ), (12a)
dt steady-state atmospheric temperature arising from a doubling of
d TLO (t ) atmospheric carbon. Hence, for thermal equilibrium corresponding
CLO = γ (TAT (t ) − TLO (t ) ). (12b) to a doubling of atmospheric carbon, we can combine (12a) and
dt
(16) to see that
Here, CAT and CLO are the heat capacities of the atmospheric

and lower ocean layers and γ is a heat exchange coefficient. The 2MAT, 1750
F2× log2 − λ ECS = 0 (18)
quantity λ is inversely proportional to what is called the Equilib- MAT, 1750
rium Climate Sensitivity (ECS), as discussed in Remark 1 below.
Taking an explicit Euler discretization with time-step  yields or λ = F2× /ECS. In words, λ is the ratio between the radiative
forcing associated with a doubling of atmospheric carbon and the
 equilibrium atmospheric temperature arising from such a doubling.
TAT (i + 1 ) = TAT (i ) + (F (i ) − λTAT (i ) − γ (TAT (i ) − TLO (i ) ))
CAT
(13a) 2.6. Carbon model

TLO (i + 1 ) = TLO (i ) + (γ (TAT (i ) − TLO (i ) )). (13b) Similar to the temperature dynamics, (CAR) is a three-reservoir
CLO
model of the global carbon cycle, with states describing the aver-
.
With T = [TAT TLO ] ∈ R2 , the above4 can be rewritten as age mass of carbon in the atmosphere, MAT , the upper ocean, MUP ,
and the deep or lower ocean, MLO . We denote the carbon states by
T (i + 1 ) = T T (i ) + BT F (i ) (14) .
M = [MAT MUP MLO ] ∈ R3 and the coefficients ζ ii ∈ [0, 1] give the
where diffusion between reservoirs. We define
 
φ11 φ12 ξ1 ζ11 ζ12 0 ξ2
T =. ,
.
BT = (15a) M =. ζ21 ζ22 ζ23 , .
BM = 0 . (19)
φ21 φ22 0
0 ζ32 ζ33 0
and
The mass of atmospheric carbon is driven by CO2 emissions
 due to economic activity5 . This occurs via a nonlinear, time-varying
φ11 = 1 − (λ + γ ) (15b)
CAT function as shown in (CAR), that corresponds to modeled predic-
γ tions of emissions and the emissions intensity of economic activ-
φ12 = (15c) ity. The additional term, ELand , captures emissions due to land use
CAT
changes as given by (4) above. Hence, the total emissions are de-
γ scribed by
φ21 = (15d)
CLO
E (i ) = σ (i )(1 − μ(i ))A(i )K (i )γ L(i )1−γ + ELand (i ). (20)

φ22 = 1 − (15e) Note that this model is conceptually similar to the three reser-
CLO
voir model of the Global Carbon Budget project (Quéré, 2018).

ξ1 = . (15f) However, the three reservoirs used by the Global Carbon Budget
CAT correspond to atmospheric, ocean, and land reservoirs.
Atmospheric temperature rise is driven by radiative forcing, or
the greenhouse effect, at the top of atmosphere and, as shown in 2.7. Welfare maximization
(CLI) has a nonlinear (logarithmic) dependence on the mass of CO2
in the atmosphere, MAT . Greenhouse gases other than CO2 (e.g., While the nine state, two decision variable DICE model (CLI)–
methane, nitrous oxide, and chloroflourocarbons) contribute to the (EI) can be used to predict outcomes based on externally (e.g., ex-
radiative forcing effect, and these are accounted for in the DICE pert) predicted mitigation and savings rates, the inputs, and pre-
model by the exogenously defined signal, FEX in (3). dicted outcomes, are more usually the result of solving an Optimal
Control Problem (OCP). In particular, the DICE dynamics act as con-
Remark 1 (Equilibrium Climate Sensitivity). The parameter λ in straints in a social welfare maximization problem.
(12a) has a specific physical interpretation in terms of the radiative The social welfare W is defined as the discounted sum of (time-
forcing and an experiment involving the doubling of atmospheric varying) utility U which depends on consumption. Consumption is
carbon. Let F2 × > 0 denote the forcing associated with equilibrium derived from (7) and (8) as

C (i ) = Q (i )(1 − s(i )) (21)


4
The implementation in Nordhaus (2017c) replaces F(i) with F (i + 1 ) in (13a).
This could be interpreted as an implicit Euler discretization. However, an implicit
Euler discretization of (12) would lead to very different expressions for the con- 5
Note that the parameter ξ 2 is simply for converting CO2 to carbon (see Table
stants in (15). 2).
12 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

which can be written explicitly in terms of states and inputs


using (6).
The utility is taken as:
  
C ( i ) 1 −α
L (i )
−1
U (C (i ), L(i )) = L(i ) , (22)
1−α

where α ≥ 0 is called the elasticity of marginal utility of consump-


tion. Note that, in the limit as α → 1, the utility is a logarithmic
function of per capita consumption and for α ∈ (0, 1), (22) behaves
qualitatively like a logarithm. For α > 1, since the population L(i)
is bounded (as the solution of (POP)), the utility is also bounded.
Indeed, denoting the upper bound on population by  L (i.e., L(i ) ≤ 
L
for all i), we see that
   1 −α 
C
L
−1 
L
lim L ≤− .
(C/L )→∞ 1−α 1−α
Fig. 2. The Social Cost of Carbon Dioxide (SC-CO2 ) computed from an emissions
The optimal control problem of interest, maximizing the social pulse experiment. (a) Baseline emissions and a pulse in the year 2020 (vertical
welfare, is then given by axis in gigatonnes CO2 ). (b) Consumption pathways resulting from the baseline and
∞ pulse emissions pathways are virtually indistinguishable (vertical axis in millions
.  U (C (i ), L(i ))
W  = max of 2010USD). (c) The difference between the baseline consumption pathway and
s,μ ( 1 + ρ ) i
i=0
the emissions pulse pathway (vertical axis in millions of 2010USD). (d) The 5% dis-
counted difference between the two consumption pathways. Summing the values
subject to (OCP ) in this plot, and normalizing by the size of the emissions pulse, yields the SC-CO2
for 2020.
(CLI ) − (EI ), (3 ) − (4 )
μ(i ), s(i ) ∈ [0, 1], ∀i ∈ N, Table 1
SC-CO2 computed for DICE2016 via Nordhaus (2017c) for selected years
where ρ > 0 is a prescribed discount rate. and different values of the discount rate ρ .

Remark 2 (Discount rates and social time preference). It should Year ρ = 0.005 ρ = 0.015 ρ = 0.03
be mentioned that the numerical value chosen for ρ can have a 2015 US$73.95 US$27.14 US$10.84
significant impact on the results and is a subject of significant dis- 2020 US$89.31 US$32.28 US$12.54
2030 US$124.20 US$44.54 US$16.98
cussion. On the one hand, when analyzing capital investment de-
cisions, discount rates of 7–9% are common (OMB, 2003). On the

other hand, given the extremely long time scales involved in the the incremental change in emissions, ∂∂EW(i ) , and the incremental
climate system and hence the long-term impacts of current emis- change in welfare with respect to the incremental change in con-

sions, the discount rate can also be viewed in the context of inter- sumption, ∂∂CW(i ) . The SC-CO2 is then given by
generational fairness (or social time preference). Here, arguments
have been made for an effective 0% discount rate (Stern, 2013), ∂ W  /∂ E (i )
SC-CO2 (i ) = −10 0 0 ·
though rates of 1–3% are more common. In the results that follow, ∂ W  /∂ C (i )
when not otherwise specified, we use the default value of 1.5% as ∂ C (i )
in Nordhaus (2017c). = −10 0 0 · . (23)
∂ E (i )
2.8. The social cost of carbon dioxide (SC-CO2 ) Note that the factor of 10 0 0 scales the SC-CO2 to 2010 US dollars
per tonne of CO2 , whereas consumption is in trillions of 2010 US
The Social Cost of Carbon Dioxide (SC-CO2 ) in a particular year dollars and emissions are in gigatonnes of CO2 .
is defined as As mentioned above, the discount rate ρ can have a significant
impact on the optimal solution and, hence, the monetary value
“the decrease in aggregate consumption in that year that would of emissions given by the SC-CO2 . Table 1 lists the SC-CO2 for
change the current...value of social welfare by the same amount different years and the different discount rates ρ ∈ {0.005, 0.015,
as a one unit increase in carbon emissions in that year.” 0.03} and considering a finite horizon N = 100. For example, the
(Newbold, Griffiths, Moore, Wolverton, & Kopits, 2013) estimates of the SC-CO2 for the year 2020 range from US$12.55
This can be computed as shown in Fig. 2 where baseline emis- (ρ = 0.03) to US$89.31 (ρ = 0.005).
sions and consumption (Fig. 2(a) and (b)) are defined, e.g., by solv-
ing the optimal control problem (OCP). A pulse of CO2 emissions 2.9. DICE, PAGE, and FUND
is then injected at a particular year (10 GtCO2 in 2020 in Fig. 2(a))
and the aggregate reduction in consumption (Fig. 2(c)) over suc- While many integrated assessment models have been proposed,
ceeding years, appropriately discounted (Fig. 2(d)), is the SC-CO2 the three most commonly used and cited models are DICE, PAGE,
for that year. Note the different time scales between Fig. 2(a) and and FUND. In particular, the U.S. Interagency Working Group made
(b)-(d), which emphasizes that although industrial emissions in use of these three models in deriving its estimates of the SC-CO2
this scenario go to zero shortly after the year 2100, the effects of (IWG, 2013). The PAGE model was used extensively in the Stern
these emissions, even discounted, persist far into the future. Review (Stern, 2007).
These pulse experiments are suggestive of a sensitivity analysis PAGE and FUND are fundamentally different models than DICE.
and, in fact, the computation of the SC-CO2 is given by the ratio In the economics lexicon, PAGE and FUND are “partial equilibrium
of two Lagrange multipliers. Specifically, the Lagrange multipliers models” while DICE is a “general equilibrium model”. Specifically,
of interest are the incremental change in welfare with respect to economic growth is an input in the former type of model, but a
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 13

state (given by the evolution of K) in the latter. As a consequence, of Grüne (2016) to include time-varying systems and discounted
in solving the welfare maximization problem (OCP), DICE generates optimal control problems is the subject of ongoing work and some
optimal emissions and consumption pathways. By contrast, such specific indications are provided at the end of this section and in
pathways must be provided as inputs to PAGE and FUND. Section 6.
The PAGE model divides the world into eight regions and con- It is worth noting that, despite the familiarity of economists
siders four different damages components given by sea level rise, with optimal control methods (e.g., Seierstad & Sydsæter, 1987)
economic damages, non-economic damages, and discontinuities. and despite its significant impact on systems and control, the
This is in contrast to DICE which considers a single global region receding-horizon approach is still largely unknown in the eco-
and a single damages component. Additionally, PAGE looks to in- nomics community (Grüne, Semmler, & Stieler, 2015). However,
corporate uncertainty by repeatedly drawing several parameters from a control point of view the welfare maximization described
from probability distributions. The model is instantiated as an Ex- in (OCP) immediately suggests a receding-horizon approach for
cel spreadsheet and makes use of the proprietary @RISK software at least two reasons: One, it is conjectured that, as in the case
add-in from Palisade to perform the required Monte Carlo calcula- of undiscounted optimal control problems, receding horizon con-
tions. trol likely provides an approximate solution to the infinite hori-
It is important to note that PAGE takes not just economic zon optimal control problem. Two, a receding horizon implemen-
growth (or projected Gross Domestic Product) as an input, but tation provides a natural framework for considering robustness is-
also climate policies (such as the mitigation rate μ) as inputs. In sues by, in particular, separating the ‘plant’ and the model of the
other words, there is no optimization problem associated with the plant used for control purposes (Hafeez, Weller, & Kellett, 2017).8
model. This idea is intuitive from a control point view, yet is not the
FUND is similar in concept to PAGE as a partial equilibrium standard for climate-economy assessment. To the best of the au-
model, but differs in its specifics. FUND considers sixteen ge- thors’ knowledge the earliest application of a receding-horizon
ographic regions and eight damages components. Furthermore, framework to the DICE OCP was presented in Chu, Duncan, Pa-
some of the damages components are dependent on both the tem- pachristodoulou, and Hepburn (2012), which looked at the RICE
perature increase and the rate of temperature rise or CO2 concen- model, while Weller, Hafeez, and Kellett (2015c) considered the
trations, while damages in DICE and PAGE are dependent solely on DICE model and explicitly accounted for uncertainty in emissions
the temperature increase. FUND is coded in C# and is available at and temperature measurements in relation to the SC-CO2 .
Anthoff and Tol. Subsequently, we aim at solving (OCP) in a receding-horizon
Given the fact that neither PAGE nor FUND involve an optimal fashion to the end of computing the SC-CO2 . As mentioned before,
control problem, they do not compute the SC-CO2 as per (23), but the SC-CO2 definition (23) can be read as a quotient of Lagrange
rather do so via the pulse experiment as indicated in Fig. 2. multipliers (or adjoint states). Hence, following our development in
Finally, we note that a regional variant of DICE—called RICE Faulwasser et al. (2018a), we reformulate the DICE dynamics such
(Regional Integrated model of Climate and Economy)—was consid- that the consumption C and the E formally can be regarded as state
ered by Nordhaus in conjunction with the 2010 variant of DICE variables. In turn this implies that the required Lagrange multipli-
(Nordhaus, 2010). The RICE model used the same geophysical ers/adjoints states are readily available upon solving (OCP) using
structure as previously described for DICE, but considered twelve state-of-the-art optimization codes.
global regions by calibrating twelve essentially independent copies We begin by defining the augmented state vector
of the economic model (CAP).  
x˜ = i T M K σ L ATFP ELand FEX ,
3. Receding horizon solution to DICE  
xaux = E (i ) C (i ) μ (i ) s (i ) W (i ) .

As defined, (OCP) is a non-convex infinite-horizon optimal con- Note that x˜(i ) ∈ R12 collects the time index i, the state variables of
trol problem and is thus difficult to solve analytically and nu- (CLI)–(EI), and the sequences (3) and (4). The vector xaux (i ) ∈ R5
merically.6 However, from a systems and control perspective it collects the emissions (20), consumption (21), inputs μ(i) and s(i)
is intuitive to approximate the solution to the infinite-horizon at time i, and the extra state
problem (OCP) by means of a receding-horizon—or model predic-
tive control—approach. Recently the analysis of asymptotic prop- 
i
U (C ( j ), L( j ))
W (i ) =
erties of model predictive control with generic objective func- (1 + ρ )( j−1)
j=1
tionals (that do not explicitly encode a control task) has re-
ceived significant attention under the label economic MPC, cf. which is used to define the objective (social welfare). Moreover,
Rawlings, Mayne, and Diehl (2017) and Faulwasser, Grüne, and using
.  
Müller (2018). Indeed, for a very general class of problems in
the undiscounted time-invariant setting, it can be shown that the x(i ) = x˜(i ) xaux (i )
receding-horizon approach yields a quantifiably accurate approxi-
and the shifted input variables
mation of the infinite-horizon solution that improves as the hori-
zon increases7 (Grüne, 2016). Extending the approximation results .  
w (i ) = μ (i + 1 ) s (i + 1 ) ,

6
Note that Nordhaus (2017c) solves a slightly different problem than (OCP).
Specifically, Nordhaus (2017c) solves over a fixed horizon of 60 or 100 (correspond- that the solution spends close to a specific attractor—i.e., close to the turnpike—
ing to 300 or 500 years), and fixes the savings rate over the last ten time steps to grows with horizon length. Early observations of this phenomenon can be traced
a value close to the turnpike value. This latter element precludes the capital stock back to von Neumann (1938), while the term “turnpike” was coined in 1958 in
from being depleted at the end of the fixed horizon. Conceptually, fixing the hori- Dorfman, Samuelson, and Solow (1958). The concept has received widespread at-
zon a priori rules out discount rates below a certain threshold since numerically tention in economics (Carlson, Haurie, & Leizarowitz, 1991; McKenzie, 1976) and,
significant behavior occurs on long time scales but is not rendered insignificant by more recently, in systems and control (Faulwasser, Korda, Jones, & Bonvin, 2017;
the discounting. Grüne & Müller, 2016; Trélat & Zuazua, 2015).
7 8
Specifically, the analysis of economic MPC schemes leverages so-called turn- Interestingly, in 2015 the EU called for revisiting emission reduction targets
pike properties of OCPs. Turnpike properties are similarity properties of parametric every five years (Steinhauser, 2015), which can also be understood as a feedback
OCPs, whereby for varying horizon lengths and varying initial conditions, the time mechanism.
14 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

we can rewrite the dynamics underlying (OCP) as follows: vk ∈ [0, 1], k = 15, 16 (25e)
x(i + 1 ) = f (x(i ), w(i )), x ( 1 ) = v. (24) w( j ) ∈ [0, 1] × [0, 1], j = 1, . . . , N. (25f)
The first component of the righthand-side function f : R17 × R2 → In order to obtain a receding horizon variant of the original
R17 , f = [ f1 , . . . , f17 ] is given by OCP, we define a second optimization problem as follows
. max x17 (N + 1 )
f1 (x, w ) = x1 + 1, w
(26a)
and the components k = 2, . . . , 12 are given by (CLI)–(EI), (3), and subject to
(4). For k = 13 we obtain from (20) x( j + 1 ) = f (x( j ), w( j )), j = 1, . . . , N (26b)
E (i + 1 ) = f13 (x(i ), w(i )) x ( 1 ) = x ( 2 | i − 1 ) (26c)
= (σ (i + 1 )(1 − μ(i + 1 ))Y (i + 1 ) + ELand (i + 1 ) ) w( j ) ∈ [0, 1] × [0, 1], j = 1, . . . , N, (26d)
which differs from OCP (25) in that the initial condition x(1) is
= ( f8 (x(i ), w(i )) · (1 − w1 (i )) · f10 (x(i ), w(i ))
 f9 (x(i),w(i)) 1−γ available from the previous optimization via the variable x (2 | i −
· f7 (x(i ), w(i ))γ · 10 0 0
1 ).9 This OCP is to be solved for i = 2, . . . , Nsim , where Nsim is
the desired simulation horizon. Consequently, the extra decision
+ EL0 · (1 − δEL )i ). variable v is not required, since x13 (1 ) = E  (2 | i − 1 ) and x14 (1 ) =
In other words, we can rewrite the emissions explicitly as a state C  ( 2 | i − 1 ).
using (CAP)–(TFP) to expand f7 , f8 , f9 , and f10 . Immediately from Solving either OCP (25) or OCP (26), we obtain the following
the above, we obtain the initial emissions data:

• The optimal state trajectory x ( j ), j = 1, . . . , N + 1, which con-
 1 − γ
E (1 ) =  x8 (1 )(1 − x15 (1 ))x10 (1 )x8 (1 )γ x109 (010) + EL0 . tains the savings rate and the mitigation rate as
μ ( j ) = x15 ( j ) and s ( j ) = x16 ( j ).
Similarly, we may rewrite the consumption state equation as
• The optimal adjoint variables λC ( j ) and λE ( j ) which are given
C (i + 1 ) = f14 (x(i ), w(i )) by the Lagrange multipliers associated to the equality con-

straints implied by the dynamics of E ( j ) = x13 ( j ) and C ( j ) =
1 − θ 1 ( i + 1 ) μ ( i + 1 ) θ2
= Y (i + 1 )(1 − s(i + 1 )) x14 ( j ).10
1 + a2 TAT (i + 1 )a3

Hence, the SC-CO2 at time j is obtained by
1 − θ 1 ( i + 1 ) w 1 ( i ) θ2
= · f10 (x(i ), w(i )) ∂ W/∂ E ( j ) λ ( j )
1 + a2 f2 (x(i ), w(i ))a3 SC-CO2 ( j ) = −10 0 0 · = −10 0 0 · E .

1 −γ ∂ W/∂ C ( j ) λC ( j )
f9 (x(i ), w(i )) Finally, the receding-horizon approximation of (OCP) is summa-
· f7 (x(i ), w(i ))γ ·
10 0 0 rized in Algorithm 1.
· (1 − w2 (i ))
Algorithm 1 MPC-DICE
with initial condition given by 1: Input: simulation horizon Nsim , prediction horizon N,

if i == 1 then
1 − θ1 (1 )x15 ( 1 ) θ2 2:
C (1 ) =  · x10 (1 ) · x8 (1 )γ 3: Solve OCP (25)
1 + a2 x2 ( 1 )a3
4: Set x (1 ) = x (1|1 ), λE (1 ) = λE (1|1 ), λC (1 ) = λC (1|1 ).
 x 8 ( 1 ) 1 − γ
· 10 0 0
· (1 − x16 (1 )). 5: for i = 2, . . . , Nsim do
6: Solve OCP (26) for x(1 ) = x (1 | i − 1 ).
The final three states are given by
7: Set x (i ) = x (2 | i − 1 ), λE (i ) = λE (2 | i − 1 ), λC (i ) = λC (1 | i − 1 ).
x15 (i + 1 ) = w1 (i ), x15 (1 ) = v15 8: Return x( j ), λE ( j ) and λC ( j ), j = 0, . . . , Nsim .
x16 (i + 1 ) = w2 (i ), x16 (1 ) = v16
U (x12 (i ), x9 (i ))
x17 (i + 1 ) = x17 (i ) + , x17 (1 ) = 0. Remark 3 (Open source code). MPC-DICE (Faulwasser et al.,
(1 + ρ )(i−1)
2018a) is an open-source Matlab implementation of DICE which
Observe that the initial condition x14 (1 ) = C (1 ) depends on the provides parameter sets for both DICE2013 and DICE2016. Specif-
(unshifted) inputs at time i = 1; i.e. it depends on μ(1 ) = x15 (1 ) ically, MPC-DICE provides an implementation of the receding hori-
and s(1 ) = x16 (1 ). Likewise the initial condition x13 (1 ) = E (1 ) de- zon reformulation described above. It uses CasADi (Andersson,
pends on μ(1 ) = x15 (1 ). Åkesson, & Diehl, 2012), which comes with IPOPT (Wächter &
To handle this dependence in the optimization, we introduce Biegler, 2006) as an NLP solver, to solve (OCP). The relatively sim-
the auxiliary decision variable v ∈ R17 and the additional constraint ple CasADi syntax enables extensions of the DICE OCP, some of
x(1 ) = v. Now, we can summarize the equivalent (finite-horizon) which we will describe in Section 4. The code is available at
reformulation of (OCP) based on the augmented dynamics (24) as Faulwasser et al. (2018b).
follows
Fig. 3 shows simulation results obtained with MPC-DICE for the
max x17 (N + 1 ) (25a) 2016 parameter set for different prediction horizons N ∈ {10, 20,
w, v

subject to
9
Here, whenever helpful, we employ the common MPC notation convention that
x( j + 1 ) = f (x( j ), w( j )), j = 1, . . . , N (25b) x (2 | i − 1 ) refers to the second element of the state prediction computed at time
i − 1.
x (1 ) = v (25c) 10
The Lagrange multipliers are typically provided by modern NLP solvers such as
IPOPT (Wächter & Biegler, 2006), which is used in the open-source DICE implemen-
vk = xk (1 ), k ∈ {1, . . . , 17} \ {15, 16} (25d) tation (Faulwasser et al., 2018a).
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 15

Fig. 3. Results for MPC-DICE 2016 for varying prediction horizons using Algorithm 1 with N ∈ {10, 20, 40, 60} (black, black-dashdot, blue, blue-dashdot) and the open-loop
solution to (OCP) for N = 120 (red-dashdot).

40, 60} and an MPC-DICE simulation horizon Nsim = 40 in com- of warming represents a dangerous threshold. One approach to
parison to the solution of (OCP) with N = 120 of which we plot addressing this is to modify the model directly; for example by
the first 40 steps. Fig. 3a shows temperature increase, emissions as changing the climate damages function (11) to reflect the consen-
well as mitigation rate and savings rate. Fig. 3b shows the corre- sus that damages at 2 ◦ C are expected to be significantly higher
sponding SC-CO2 trajectories. As one can see, for increasing pre- than a loss of 0.9% of global economic output. It is also possible to
diction horizons the receding-horizon input and state trajectories consider a different welfare function that not only places a value
both converge towards the infinite-horizon solution; approximated on consumption but also values environmental “services” (such as
here by computing a long horizon solution (N = 120). This approx- clean air and water) (Sterner & Persson, 2008) or accounts for the
imation property can also be observed in Fig. 3b. Hence we conjec- cost of adaptation to climate change (Atolia, Loungani, Maurer, &
ture that under suitable assumptions the approximation properties Semmler, 2018).
of MPC, which are established for time-invariant and undiscounted A third approach, as done in Nordhaus (2008, pp. 69–73), is to
OCPs in Grüne (2016), also hold for time-varying and discounted add a constraint on the temperature increase to (OCP). As men-
problems. tioned in (Nordhaus, 2008, pp. 69–73), the purely economic case
In fact, the theoretical results supporting this conjecture are for imposing a hard limit is somewhat unjustified as it effectively
reasonably mature where Grüne et al. (2015) shows that reced- implies an infinite cost of exceeding the constraint. However, also
ing horizon control yields approximate optimal solutions for dis- as discussed in Nordhaus (2008, pp. 69–73), a hard constraint can
counted problems if the turnpike property holds. Furthermore, it represent a tipping point where the climate damages dramati-
follows from a combination of Gaitsgory, Grüne, Höger, Kellett, and cally increase, for example due to adverse climatic effects that are
Weller (2018) and Grüne, Müller, Kellett, and Weller (2018) that not captured in the simple DICE climate model. Here, following
strict dissipativity implies the turnpike property for discounted Weller, Hafeez, and Kellett (2015b), we place an upper limit on
problems, provided the discount factor is sufficiently close to one. the atmospheric temperature rise and investigate what this then
Note that the default DICE discount rate of 1.5% corresponds to requires of the control inputs.
a discount factor of approximately 0.985. Hence, while these re- Consider the state constraint
sults for discounted optimal control do not yet accommodate time-
TAT (i ) ≤ TAT,max , ∀i ∈ N. (27a)
varying systems or cost functions, the primary difficulty lies in
checking the appropriate assumptions for complicated models such Moreover, in the economics literature the value for the mitigation
as the DICE model. rate at time i = 1 is usually fixed, with

4. State and input rate constraints μ(1 ) = μ0 = 0.03, (27b)


an estimate of the global greenhouse gas emissions abatement or
The welfare maximization problem as posed in Section 2.7 con- mitigation rate in the base year of 2015 (or μ(1 ) = 0.039 for the
siders only input magnitude constraints and the dynamics. How- DICE2013 base year of 2010) (Nordhaus, 2017c).
ever, in view of the reports of the IPCC, the overwhelming scien- Fig. 4a shows the corresponding results for MPC-DICE 2016 con-
tific consensus is that temperature increase should be limited to sidering (27) and
2 ◦ C (IPCC, 2013b) and preferably to 1.5◦ C (IPCC, 2018). Inspection
of Fig. 3a reveals that straightforward maximization of social wel- TAT,max ∈ {2.36 ◦C, 3 ◦C }.
fare might lead to much higher values of temperature increase in Not surprisingly the tighter temperature constraint of 2.36 ◦ C re-
the order of 3 − 4 ◦ C.11 quires a drastic and fast reduction of emissions, which would im-
This indicates that there is an inconsistency between the model ply a steep increase of the mitigation rate in the near future. In-
(or the chosen parameters) and the scientific consensus that 2 ◦ C terestingly the savings rate is not affected by the temperature con-
straint. Moreover, we remark that TAT,max = 2.36 ◦ C is the lowest
11
In climate physics the temperature increase is also referred to as the tempera- heuristically determined value of the temperature constraint for
ture anomaly. which (OCP) is feasible.
16 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

Fig. 4. Results for MPC-DICE 2016 with additional constraints.

However, the steep increase of the mitigation rate shown in ters would be a valuable contribution to improved estimates of the
Fig. 4a might be difficult to realize on a policy level. This moti- SC-CO2 .
vates analyzing whether constraints on the increase of the mitiga-
tion rate are compatible with a temperature constraint of 2 − 3 ◦ C.
To this end, in Weller et al. (2015b) (for DICE2013) we considered
5. Systems and control in climate-economy
both rate and growth constraints on μ. The rate constraint takes
assessment: Opportunities and progress
the form
| μ ( i + 1 ) − μ ( i ) | ≤ μ , ∀ i ∈ N (28a) The DICE model interconnects geophysical and socioeconomic
while the growth constraint is given by dynamics, the structure and parameters of which are clearly sub-
ject to a vast array of uncertainties. Quantifying the implications
μ (i + 1 ) − μ (i )
≤ μ , ∀i ∈ N. (28b) of these uncertainties on SC-CO2 estimates and on related pol-
μ (i ) icy advice is consequently an issue of major importance to poli-
Note that the growth constraint (28b) is motivated by the argu- cymakers. Given the wide range of tools developed by the systems
ment that abatement technologies and markets will increase year- and control community for handling and quantifying uncertainty,
on-year rather than in equal increments. many challenges and research opportunities exist for this commu-
Results for this setting using MPC-DICE with Nsim = 40 and N = nity within climate–economy assessment. In this section, we iden-
60 are depicted in Fig. 4b. We consider the temperature bound of tify several key avenues of research opportunity, and describe pre-
3 ◦ C according to (27), the rate bound μ = 0.1 and the growth vious work undertaken in the context of those directions.
bound μ = 0.53. Note that (27b) is needed to ensure that the con-
straints (28) are defined at i = 1.
As one can see the temperature trajectories do not differ much
5.1. Uncertainty quantification
for the different constraints (28). However, not surprisingly, the
mitigation rates behave very differently as do the emissions.
In early 2017, the U.S. National Academies released an ex-
It is worth noting that μ = 0.53 is an experimentally deter-
tensive and influential report on improving estimates of the SC-
mined threshold whereby lower values render the optimal con-
CO2 (National Academies, 2017). Many of the recurring themes in
trol problem infeasible. In this extreme case, the optimal solution
National Academies (2017) would be familiar to the systems and
shows the savings rate effectively taking a value of zero for sev-
control community, particularly around the quantification of un-
eral years. Conceptually, this corresponds to an extended period of
certainty and its impacts.
non-investment in the capital stock K so as to reduce emissions by
The climate economics literature distinguishes between para-
explicitly reducing economic activity.
metric uncertainty (Anderson, Borgonovo, Galeotti, & Roson, 2014;
Remark 4 (Feasibility of temperature constraints). Given that Botzen & van den Bergh, 2012; Grüne, Greiner, & Semmler, 2010;
it is considered feasible that global atmospheric temperature rise IWG, 2013) and structural uncertainty (Gillingham, 2015; Guivarch
could yet be limited to 1.5 ◦ C, it is interesting to note that the de- & Pottier, 2018; Moore & Diaz, 2015; Nordhaus, 2017b) in IAMs
fault parameters of DICE2016 do not even yield a feasible solution such as DICE. Parametric uncertainty is uncertainty about the value
to (OCP) for the higher 2 ◦ C limit. Indeed, Nordhaus (2017b) notes of various parameters within an IAM module, e.g. F2 × or T in
that while the 2 ◦ C limit was achievable in DICE2013, this is not (CLI). Structural uncertainty, on the other hand, refers to uncer-
the case for DICE2016. However, only limited information on the tainty regarding the functional form of the equations comprising
calibration of the model (i.e., the parameter choices shown in the IAM. As one example of structural uncertainty, consider carbon
Appendix B) has been provided. In Section 5 below, we indicate cycle feedbacks—currently neglected in DICE—wherein rising sur-
some recent work on improving the transparency of the tempera- face temperatures lead to thawing of carbon-rich permafrost and
ture model parameters. Similar work on the carbon cycle parame- the consequent release of methane, itself a potent greenhouse gas.
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 17

5.2. Identification of predictive climate models Assessment Report (AR5) of the IPCC (see IPCC, 2013a, p. 818,
Table 9.5).
The geophysical models presented in Section 2.1 are clearly sig-
nificant simplifications of reality, with the climate (temperature) 5.3. Modular tools for simulation and optimization
model having been originally proposed in Schneider and Thomp-
son (1981). While low-order models are necessary to efficiently The report National Academies (2017) also recommends in-
solve the optimal control problem (OCP), it is possible to con- creasing transparency around the models used, and maintain-
struct improved higher order models. In particular, a large num- ing modular models to allow for advances in any particular
ber of supercomputer-based, atmosphere–ocean general circulation model to be easily incorporated. As an example of this latter
models (AOGCMs) have been developed by a number of climate topic, in Faulwasser, Nydestedt, Kellett, and Weller (2018d) we re-
modeling centres, providing very high spatio-temporal resolution. placed the standard DICE geophysical model (i.e., (CLI)–(CAR)) with
Furthermore, many of these AOGCMs participate in the Coupled a state-of-the-art reduced-order geophysical model termed FAIR
Model Intercomparison Project (CMIP) (Taylor, Stouffer, & Meehl, (Millar, Nicholls, Friedlingstein, & Allen, 2017). In this context it
2012), which effectively provides input–output data for a number is worth noting that the original reference for the FAIR model
of AOGCMs. (Millar et al., 2017) does not highlight the fact that FAIR is a sys-
With such input–output data available, standard system identi- tem of differential algebraic equations (DAEs), which should be ac-
fication tools can be applied. In Weller, Schulz, and Ninness (2014), counted for in developing simulation code.
for example, we derived fourth-order linear time-invariant models However, as of now, when it comes to uncertainty quantifi-
from the CMIP3 (CMIP, Phase 3) data set. In particular, 12 AOGCMs cation combined with dynamic optimization, no widely accepted
from the CMIP3 ensemble were identified for which linear, time- open-source tools that go beyond various sampling techniques ex-
invariant (LTI) models of order 4 were able to very closely approx- ist. Hence, there is a need for tailoring and implementing the pow-
imate surface temperature projections under each of the four Rep- erful methods developed by the systems and control community to
resentative Concentration Pathway (RCP) emission scenarios in the climate-economy assessment.
AR5 assessment report (see IPCC, 2013b, p. 45, Box SPM.1). Addi-
tionally, these fourth order models were shown to closely match
the ECS and TCR values reported in the AR5 assessment report 5.4. System theoretic analysis
(Hafeez, Weller, & Kellett, 2015).
The LTI models identified in Weller et al. (2014) are suitable In addition to the research questions mentioned above and
for application in feedback-based approaches to mitigation; see posed in National Academies (2017), the framework of discounted
for example Weller and Schulz (2014) in which these models are optimal control (i.e., where the cost function involves a discount
applied in an optimal control-based approach to geoengineering factor) is one which has received less attention in the systems
of the climate based on solar radiation management (SRM). In and control community than the usual undiscounted framework.
Weller, Hafeez, and Kellett (2015a), we considered (CLI)–(CAP) in Several recent results (Gaitsgory et al., 2018; Gaitsgory, Grüne, &
which the climate model (CLI) is replaced by each of 12 fourth- Thatcher, 2015; Grüne, Kellett, & Weller, 2017; Grüne et al., 2018)
order models derived in Weller et al. (2014). The range of esti- indicate that the connections between strict dissipativity, turn-
mated SC-CO2 values for 2015 obtained using this method span pike properties, and numerically accurate approximations via MPC,
US$10.20–$58.20/tCO2 depending on the specific CMIP3 model, which are known for undiscounted optimal control (as reported in
with an ensemble mean SC-CO2 of US$22.90/tCO2. This wide range Grüne, 2016) also hold in the discounted setting. However, check-
of values highlights the substantial variability in estimates of the ing the necessary assumptions to use results in particular applica-
SC-CO2 arising from scientific uncertainty in the climatic response tions, such as for the DICE model, remains a difficult problem.
to net radiative forcing, all other components of the IAM being Moreover, the fact that the receding-horizon solution to
held constant. DICE approximates long/infinite horizon solutions quite well (see
In recognizing the numerous uncertainties inherent in estima- Hafeez et al., 2017, Fig. 2) gives rise to the conjecture that the DICE
tion of the SC-CO2 , the National Academies report recommends OCP exhibits a time-varying turnpike phenomenon. However, a for-
that research effort on the SC-CO2 be focused on “incorporat- mal analysis remains to be done.
ing the most important sources of uncertainty, rather than seek- Finally, recall that Nordhaus also proposed a regionally dis-
ing to incorporate all possible sources of uncertainty” (National tributed variant of DICE named RICE, wherein several economic
Academies, 2017, p. 54). One approach along these lines avoids di- regions (US, EU, China,...) are considered (Nordhaus, 2010). From a
rect appeal to strongly geophysically-inspired climate models, in- systems and control perspective RICE raises many interesting prob-
stead capturing climate behaviour simply via three scalar parame- lems ranging from distributed implementation to game-theoretic
ters: the transient climate response (TCR), the equilibrium climate frameworks.
sensitivity (ECS), and F2 × .
Here we recall ECS as the steady-state atmospheric temperature 6. Summary and concluding remarks
arising from a doubling of atmospheric carbon, F2 × as the associ-
ated downward radiative forcing at top-of-atmosphere for doubled The overwhelming scientific consensus is that avoiding the
atmospheric carbon (see Remark 1), and define TCR as the tem- worst potential effects of anthropogenic climate change require
perature change at the time of CO2 doubling under a scenario in achieving economy-wide net-zero greenhouse gas emissions by the
which CO2 concentrations increase by 1% yr−1 . middle of this century. Such a significant economic transition will
In Hafeez, Weller, and Kellett (2016) we proposed an require a suite of policy responses, many of which will rely on
optimization-based methodology for computing the parameters of a price on greenhouse gas emissions (Kellett, Aydos, Rudolph, &
a climate model in such a way that the resulting model exhibits Weller, 2018). Estimates of the SC-CO2 provide guidance on the
a specified TCR. The results reported in Hafeez et al. (2016) en- range of prices.
able policymakers using DICE—which specifies the TCR param- In this paper, we have provided a complete tutorial descrip-
eter only indirectly—to compute optimal CO2 emissions path- tion of the DICE model, one of the most widely used IAMs for
ways which directly reflect the reported TCR of state-of-the-art estimation of the SC-CO2 and have indicated some work already
AOGCM climate models documented in the most recent (Fifth) undertaken to improve SC-CO2 estimates and indicated where we
18 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

The parameters for calculating σ0 =


e0
believe the systems and control community can make important :
q0 (1−μ0 )
contributions.
e0 q0 μ0
2013R 33.61 63.69 0.039
Acknowledgements 2016R 35.85 105.5 0.03

The authors are supported by the Australian Research Council


under Discovery Project DP180103026. TF acknowledges financial Appendix B. Default parameter values
support from the Daimler und Benz Stiftung.
Rather than the cost function in (OCP), Nordhaus has used the
Appendix A. Default initial conditions scaled cost function

TAT (0 ) TLO (0 ) K (0 ) 
N+1
U (C (i ), L(i ))
(scale2 ) + (scale1 ) · max . (29)
2013R 0.8 0.0068 135 s,μ
i=1
( 1 + ρ ) i
2016R 0.85 0.0068 223
This obviously has no impact on the solution of (OCP). These val-
MAT (0 ) MUP (0 ) MLO (0 ) ues are chosen so that the optimal value function has a numerical
2013R 830.4 1527 10010 value consistent with economic intuition.
2016R 851 460 1740

Table 2
Default parameter values for DICE 2013 and DICE 2016.

Parameter Value Value Equations Unit


DICE2013R DICE2016R

 5 5 (1) years
t0 2010 2015 (1) year
N 60 100 (25), (26) time steps
μ0 0.039 0.03 (27b)
Climate diffusion parameters
φ 11 0.8630 0.8718 (CLI), (15)
φ 12 0.0086 0.0088 (CLI), (15)
φ 21 0.025 0.025 (CLI), (15)
φ 22 0.975 0.975 (CLI), (15)
Carbon cycle diffusion parameters
ζ 11 0.912 0.88 (CAR), (19)
ζ 12 0.03833 0.196 (CAR), (19)
ζ 21 0.088 0.12 (CAR), (19)
ζ 22 0.9592 0.797 (CAR), (19)
ζ 23 0.0 0 03375 0.001465 (CAR), (19)
ζ 32 0.00250 0.007 (CAR), (19)
ζ 33 0.9996625 0.99853488 (CAR), (19)
Other geophysical parameters
η 3.8 3.6813 (CLI), Remark 1 W/m2
ξ1 0.098 0.1005 (CLI), (15f)
ξ2 12/44 12/44 (CAR), (19) GtC/GtCO2
MAT,1750 588 588 (CLI), Remark 1 GtC
f0 0.25 0.5 (3) W/m2
f1 0.70 1.0 (3) W/m2
tf 18 17 (3) time steps
EL0 3.3 2.6 (4) GtCO2 /yr
δ EL 0.2 0.115 (4)
Socioeconomic parameters
γ 0.3 0.3 (CAP), (5)
θ2 2.8 2.6 (CAP), (6) n/a
a2 0.00267 0.00236 (CAP), (6)
a3 2 2 (CAP), (6) n/a
δK 0.1 0.1 (CAP), (9)
α 1.45 1.45 (22)
ρ 0.015 0.015 (OCP)
L0 6838 7403 (POP) millions people
La 10500 11500 (POP) millions people
g 0.134 0.134 (POP)
A0 3.80 5.115 (TFP)
gA 0.079 0.076 (TFP)
δA 0.006 0.005 (TFP)
σ0 0.5491 0.3503 (EI) GtC / trillions 2010USD
gσ 0.01 0.0152 (EI)
δσ 0.001 0.001 (EI)
pb 344 550 (2) 2010USD/tCO2
δ pb 0.025 0.025 (2)
scale1 0.016408662 0.030245527 (29) n/a
scale2 3855.106895 10993.704 (29) 2010USD
C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20 19

References Grüne, L., Müller, M. A., Kellett, C. M., & Weller, S. R. (2018). Strict dissipativity for
discrete time discounted optimal control problems. Preprint available at https://
ABC Rural (2017). BP Australia says pricing carbon one of the optimum methods of epub.uni-bayreuth.de/3738/1/discounted_dissipativity.pdf.
reducing emissions. 17 February, http://ab.co/2lg4liZ. Grüne, L., Semmler, W., & Stieler, M. (2015). Using nonlinear model predictive con-
Acemoglu, D. (2009). Introduction to Modern Economic Growth. Princeton University trol for dynamic decision problems in economics. Journal of Economic Dynamics
Press. and Control, 60, 112–133.
Anderson, B., Borgonovo, E., Galeotti, M., & Roson, R. (2014). Uncertainty in climate Guivarch, C., & Pottier, A. (2018). Climate damage on production or on growth: What
change modeling: Can global sensitivity analysis be of help? Risk Analysis, 34(2), impact on the social cost of carbon? Environmental Modeling & Assessment, 23,
271–293. 117–130.
Andersson, J., Åkesson, J., & Diehl, M. (2012). CasADi: A symbolic package for au- Hafeez, S., Weller, S. R., & Kellett, C. M. (2015). Steady-state and transient dy-
tomatic differentiation and optimal control. In Recent Advances in Algorithmic namic behavior of simple climate models for application in integrated assess-
Differentiation (pp. 297–307). Springer. ment models. In Proceedings Australian Control Conference (pp. 269–273). Aus-
Anthoff, D., & Tol, R. S. Fund. http://www.fund-model.org. tralia: Gold Coast. 5–6 November.
Anthoff, D., & Tol, R. S. (2013). The uncertainty about the social cost of car- Hafeez, S., Weller, S. R., & Kellett, C. M. (2016). Transient climate response in the
bon: A decomposition analysis using FUND. Climatic Change, 117(3), 515– DICE integrated assessment model of climate-economy. In Proceedings of the
530. Australian Control Conference (pp. 282–287). Australia: Newcastle. 3–4 Novem-
Atolia, M., Loungani, P., Maurer, H., & Semmler, W. (2018). Optimal control of a global ber.
model of climate change with adaptation and mitigation. International Monetary Hafeez, S., Weller, S. R., & Kellett, C. M. (2017). Impact of climate model paramet-
Fund. IMF Working Paper WP/18/270. ric uncertainty in an MPC implementation of the DICE integrated assessment
Bonen, A., Semmler, W., & Klasen, S. (2014). Economic damages from climate model. In Proceedings of the IFAC World Congress. France: Toulouse.
change: A review of modeling approaches. Schwartz Center for Economic Anal- Hassell, M. P. (1975). Density-dependence in single-species populations. Journal of
ysis and Department of Economics, The New School for Social Research. Working Animal Ecology, 44(1), 283–295.
Paper Series. Hope, C. (2013). Critical issues for the calculation of the social cost of CO2 : Why the
Botzen, W. J. W., & van den Bergh, J. C. J. M. (2012). How sensitive is Nordhaus to estimates from PAGE09 are higher than those from PAGE2002. Climatic Change,
Weitzman? Climate policy in DICE with an alternative damage function. Eco- 117(3), 531–543.
nomics Letters, 117, 372–374. Hope, C. (2015). The $10 trillion value of better information about the transient
Cai, Y., Judd, K. L., & Lontzek, T. S. (2012). Open science is necessary. Nature Climate climate response. Philosophical Transactions of the Royal Society A, 373(2054),
Change, 2, 299. 20140429.
Calel, R., Stainforth, D. A., & Dietz, S. (2015). Tall tales and fat tails: The science and Howarth, R. B., Gerst, M. D., & Borsuk, M. E. (2014). Risk mitigation and the social
economics of extreme warming. Climatic Change, 132, 127–141. cost of carbon. Global Environmental Change, 24, 123–131.
Carlson, D. A., Haurie, A., & Leizarowitz, A. (1991). Infinite Horizon Optimal Control: Hsiang, S., Kopp, R., Jina, A., Rising, J., Delgado, M., Mohan, S., et al. (2017). Estimat-
Deterministic and Stochastic Systems. Springer Verlag. ing economic damage from climate change in the United States. Science, 356,
Chu, B., Duncan, S., Papachristodoulou, A., & Hepburn, C. (2012). Using economic 1362–1369.
model predictive control to design sustainable policies for mitigating climate IPCC (2013a). Evaluation of climate models. In T. F. Stocker, et al. (Eds.), Climate
change. In Proceedings of 51st IEEE Conference on Decision and Control. Hawaii, Change 2013: The Physical Science Basis. Contribution of Working Group I to the
USA: Maui. Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cam-
Crost, B., & Traeger, C. P. (2013). Optimal climate policy: Uncertainty versus Monte bridge, UK: Cambridge University Press.
Carlo. Economics Letters, 120(3), 552–558. IPCC (2013b). Summary for policymakers. In T. F. Stocker, et al. (Eds.), Climate
Dorfman, R., Samuelson, P. A., & Solow, R. M. (1958). Linear Programming and Eco- Change 2013: The Physical Science Basis. Contribution of Working Group I to the
nomic Analysis. New York: McGraw-Hill. Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cam-
Farid, M., et al. (2016). After Paris: Fiscal, macroeconomic, and financial im- bridge, UK: Cambridge University Press.
plications of climate change. International Monetary Fund staff discussion IPCC (2018). Global warming of 1.5◦ C. Intergovernmental Panel on Climate Change.
note SDN/16/01. Available http://www.economistinsights.com/financial-services/ Interagency Working Group (IWG) on Social Cost of Carbon U.S. Government (2013).
analysis/cost-inaction. Technical update of the social cost of carbon for regulatory impact analysis - under
Faulwasser, T., Grüne, L., & Müller, M. (2018). Economic nonlinear model predictive Executive Order 12866.
control: Stability, optimality and performance. Foundations and Trends in Systems Kellett, C. M., Aydos, E., Rudolph, S., & Weller, S. R. (2018). The social cost
and Control, 5(1), 1–98. of carbon dioxide: Policy and methods for pricing greenhouse gas emis-
Faulwasser, T., Kellett, C. M., & Weller, S. R. (2018a). MPC-DICE: An open-source Mat- sions. In T. R. Finlayson (Ed.), Our Changing World in the South Pacific: Aus-
lab implementation of receding horizon solutions to DICE. In Proceedings of the tralasian and German Perspectives. Australian Association of von Humboldt
1st IFAC Workshop on Integrated Assessment Modeling for Environmental Systems. Fellows.
Italy: Brescia. Kellett, C. M., Faulwasser, T., & Weller, S. R. (2016). DICE2013r-mc: A Matlab/CasADi
Faulwasser, T., Kellett, C. M., & Weller, S. R. (2018b). MPC-DICE: Model predictive implementation of vanilla DICE2013r. Code available at http://bit.ly/2lpdeqp.
control – dynamic integrated model of climate and economy. Available at https: McKenzie, L. W. (1976). Turnpike theory. Econometrica: Journal of the Econometric
//github.com/cmkellett/MPC-DICE. Society, 44(5), 841–865.
Faulwasser, T., Korda, M., Jones, C. N., & Bonvin, D. (2017). On turnpike and dissipa- Millar, R. J., Nicholls, Z. R., Friedlingstein, P., & Allen, M. R. (2017). A
tivity properties of continuous-time optimal control problems. Automatica, 81, modified impulse-response representation of the global near-surface
297–304. air temperature and atmospheric concentration response to carbon
Faulwasser, T., Nydestedt, R., Kellett, C. M., & Weller, S. R. (2018d). Towards a dioxide emissions. Atmospheric Chemistry and Physics, 17(11), 7213–
FAIR-DICE IAM: Combining DICE and FAIR models. In Proceedings of the 1st IFAC 7228.
Workshop on Integrated Assessment Modeling for Environmental Systems. Italy: Moore, F. C., & Diaz, D. B. (2015). Temperature impacts on economic growth warrant
Brescia. stringent mitigation policy. Nature Climate Change, 5, 127–131.
Gaitsgory, V., Grüne, L., Höger, M., Kellett, C. M., & Weller, S. R. (2018). Stabilization National Academies of Sciences, Engineering, and Medicine (2016). Assessment of
of strictly dissipative discrete time systems with discounted optimal control. Au- Approaches to Updating the Social Cost of Carbon: Phase 1 Report on a Near-Term
tomatica, 93, 311–320. Update. Washington, D.C.: The National Academies Press.
Gaitsgory, V., Grüne, L., & Thatcher, N. (2015). Stabilization with discounted optimal National Academies of Sciences, Engineering, and Medicine (2017). Valuing Climate
control. Systems & Control Letters, 82, 91–98. Damages: Updating Estimation of the Social Cost of Carbon Dioxide. The National
Geoffroy, O., Saint-Martin, D., Bellon, G., Voldoire, A., Olivié, D. J. L., et al. (2013). Academies Press.
Transient climate response in a two-layer energy-balance model. Part II: Rep- Newbold, S. C., Griffiths, C., Moore, C., Wolverton, A., & Kopits, E. (2013). A rapid
resentation of the efficacy of deep-ocean heat uptake and validation for CMIP5 assessment model for understanding the social cost of carbon. Climate Change
AOGCMs. Journal of Climate, 26(6), 1859–1876. Economics, 4(1).
Gillingham, K., et al. (2015). Modeling uncertainty in climate change: A multi-model Nordhaus, W. D. (1975). Can we control carbon dioxide? Technical report WP-75-63
comparison. National Bureau of Economic Research. Working paper 21637 Austria, Laxenburg: International Institute for Applied Systems Analysis.
Gregory, J. M. (20 0 0). Vertical heat transports in the ocean and their ef- Nordhaus, W. D. (1992). An optimal transition path for controlling greenhouse gases.
fect on time-dependent climate change. Climate Dynamics, 16(7), 501– Science, 258(20), 1315–1319.
515. Nordhaus, W. D. (2007). Accompanying notes and documentation on development of
Grüne, L. (2016). Approximation properties of receding horizon optimal control. DICE-2007 mode: Notes on DICE-2007.v8 of September 21, 2007. New Haven, CT:
Jahresbericht der Deutschen Mathematiker-Vereinigung, 118, 3–37. Yale University. Available http://nordhaus.econ.yale.edu/Accom_Notes_100507.
Grüne, L., Greiner, A., & Semmler, W. (2010). Growth and climate change: Thresh- pdf.
olds and multiple equilibria. In J. C. Cuaresma, T. Palokangas, & A. Tarasyev Nordhaus, W. D. (2008). A Question of Balance: Weighing the Options on Global Warm-
(Eds.), Dynamic Systems, Economic Growth, and the Environment. Springer Pub- ing Policies. Yale University Press.
lishing House. Nordhaus, W. D. (2010). Economic aspects of global warming in a post-Copenhagen
Grüne, L., Kellett, C. M., & Weller, S. R. (2017). On the relation between turnpike environment. Proceedings of National Academy of Sciences of the USA (PNAS),
properties for finite and infinite horizon optimal control problems. Journal of 107(26), 11721–11726.
Optimization Theory and Applications, 173(3), 727–745. Nordhaus, W. D. (2014). Estimates of the social cost of carbon: Concepts and results
Grüne, L., & Müller, M. A. (2016). On the relation between strict dissipativity and from the DICE-2013R model and alternative approaches. Journal of the Associa-
turnpike properties. Systems & Control Letters, 90, 45–53. tion of Environmental and Resource Economists, 1(1/2), 273–312.
20 C.M. Kellett, S.R. Weller and T. Faulwasser et al. / Annual Reviews in Control 47 (2019) 7–20

Nordhaus, W. D. (2017a). Projections and uncertainties about climate change in an Taylor, K. E., Stouffer, R. J., & Meehl, G. A. (2012). An overview of CMIP5 and the
era of minimal climate policies. NBER Working Paper Series, no. 22933. http:// experiment design. Bulletin of the American Meteorological Society, 93(4), 485–
www.nber.org/papers/w22933. 498.
Nordhaus, W. D. (2017b). Revisiting the social cost of carbon. Proceedings of National The Guardian (2016). AGL boss: Regardless of climate science, it’s time to drop the
Academy of Sciences of the USA (PNAS), 114(7), 1518–1523. 14. ‘emissions business’. https://bit.ly/2GgFx8t, 24 February.
Nordhaus, W. D. (2017c). Scientific and economic background on DICE models. https: The Economist Intelligence Unit (EIU) (2015). The cost of inaction: Recognising the
//sites.google.com/site/williamdnordhaus/dice-rice. value at risk from climate change. Available http://bit.ly/2kKbOJU.
Nordhaus, W. D., & Sztorc, P. (2013). DICE 2013r: Introduction and user’s manual (2nd Trélat, E., & Zuazua, E. (2015). The turnpike property in finite-dimensional nonlinear
ed.). optimal control. Journal of Differential Equations, 258, 81–114. In press.
Office of Management and Budget, & U.S. Federal Government (2003). OMB Cir- von Neumann, J. (1938). Über ein ökonomisches Gleichungssystem und eine Verall-
cular A-4 of September 17. (Regulatory Analysis). Available at https://www. gemeinerung des Brouwerschen Fixpunktsatzes. In K. Menger (Ed.), Ergebnisse
transportation.gov/regulations/omb-circular-no-4-0. eines Mathematischen Seminars.
Otto, F. E., Frame, D. J., Otto, A., & Allen, M. R. (2015). Embracing uncertainty in World Bank (2014). 73 countries and over 1000 businesses speak out in support of a
climate change policy. Nature Climate Change, 5, 917–920. price on carbon. Available http://www.worldbank.org/en/news/feature/2014/09/
Palisade @RISK https://www.palisade.com/risk/. 22/governments- businesses- support- carbon- pricing.
Pizer, W., et al. (2014). Using and improving the social cost of carbon. Science, 346, Wächter, A., & Biegler, L. T. (2006). On the implementation of primal-dual interior
1189–1190. point filter line search algorithm for large-scale nonlinear programming. Math-
Quéré, C. L., et al. (2018). Global carbon budget 2018. Earth System Science Data, 10, ematical Programming, 106(1), 25–57.
2141–2194. Weller, S. R., Hafeez, S., & Kellett, C. M. (2015a). Estimates of the social cost of car-
Rawlings, J. B., Mayne, D. Q., & Diehl, M. (2017). Model Predictive Control: Theory, bon using climate models derived from the CMIP3 ensemble. In Proceedings of
Computation, and Design. Madison, WI: Nob Hill Publishing. ASEAN–Australian Engineering Congress on Innovative Technologies for Sustainable
Schneider, S. H., & Thompson, S. L. (1981). Atmospheric CO2 and climate: Impor- Development and Renewable Energy (AAEC 2015), Singapore (pp. 11–13).
tance of the transient response. Journal of Geophysical Research: Oceans, 86(C4), Weller, S. R., Hafeez, S., & Kellett, C. M. (2015b). Feasibility of 2 ◦ C as a post-2020
3135–3147. warming threshold via rate-constrained optimal control. In Proceedings IEEE
Seierstad, A., & Sydsæter, K. (1987). Optimal Control Theory with Economic Applica- Conference Control Applications (CCA 2015) (pp. 1117–1123). Sydney, Australia,
tions. North-Holland. 21–23 September.
Steinhauser, G. (2015). EU calls for five-year emission target renewal. https://on.wsj. Weller, S. R., Hafeez, S., & Kellett, C. M. (2015c). A receding horizon control approach
com/2LjE3Jo. to estimating the social cost of carbon in the presence of emissions and tem-
Stern, N. (2007). The Economics of Climate Change: The Stern Review. Cambridge Uni- perature uncertainty. Proceedings of 54th IEEE Conference on Decision and Control
versity Press. (CDC), Osaka, Japan.
Stern, N. (2013). The structure of economic modeling of potential impacts of cli- Weller, S. R., & Schulz, B. P. (2014). Geoengineering via solar radiation manage-
mate change: Grafting gross underestimation of risk onto already narrow sci- ment as a feedback control problem: Controller design for disturbance rejection.
ence models. Journal of Economic Literature, 51(3), 838–859. In Proceedings of 2014 Australian Control Conference (AUCC 2014) (pp. 101–106).
Stern, N. (2016). Current climate models are grossly misleading. Nature, 530, Canberra, Australia, 17–18 November.
407–409. Weller, S. R., Schulz, B. P., & Ninness, B. M. (2014). Identification of linear cli-
Sterner, T., & Persson, U. M. (2008). An even Sterner Review: Introducing relative mate models from the CMIP3 multimodel ensemble. In Proceedings of 2014
prices into the discounting debate. Review of Environmental Economics and Pol- IFAC World Congress (pp. 10875–10881). Cape Town, South Africa, 24–29
icy, 2, 61–76. August.

You might also like