Skip to main content

Advertisement

Log in

Dynamics of the coupled human–climate system resulting from closed-loop control of solar geoengineering

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

If solar radiation management (SRM) were ever implemented, feedback of the observed climate state might be used to adjust the radiative forcing of SRM in order to compensate for uncertainty in either the forcing or the climate response. Feedback might also compensate for unexpected changes in the system, e.g. a nonlinear change in climate sensitivity. However, in addition to the intended response to greenhouse-gas induced changes, the use of feedback would also result in a geoengineering response to natural climate variability. We use a box-diffusion dynamic model of the climate system to understand how changing the properties of the feedback control affect the emergent dynamics of this coupled human–climate system, and evaluate these predictions using the HadCM3L general circulation model. In particular, some amplification of natural variability is unavoidable; any time delay (e.g., to average out natural variability, or due to decision-making) exacerbates this amplification, with oscillatory behavior possible if there is a desire for rapid correction (high feedback gain). This is a challenge for policy as a delayed response is needed for decision making. Conversely, the need for feedback to compensate for uncertainty, combined with a desire to avoid excessive amplification of natural variability, results in a limit on how rapidly SRM could respond to changes in the observed state of the climate system.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. A quote attributed to Ron Prinn in Morton (2007).

  2. The same oscillatory behavior can be observed if one impatiently adjusts the knobs in an unfamiliar shower: if there is time delay, then a large response to water that is either too cold or too hot will result in overcompensation before the system has responded to the current settings. In aircraft, this phenomenon is referred to as “Pilot-induced oscillation” (PIO).

References

  • Åström KJ, Murray RM (2008) Analysis and design of feedback systems. Princeton, New York

  • Bode HW (1945) Network analysis and feedback amplifier design. Van Nostrand, New York

    Google Scholar 

  • Caldeira K, Myhrvold N (2013) Projections of the pace of warming following an abrupt increase in atmospheric carbon dioxide concentration (submitted)

  • Crutzen PJ (2006) Albedo enhancement by stratospheric sulfur injections: a contribution to resolve a policy dilemma? Clim Chang 77:211–219

    Article  Google Scholar 

  • Fraedrich K, Luksch U, Blender R (2004) 1/f model for long-time memory of the ocean surface temperature. Phys Rev E 70(037301)

  • Franklin GF, Powell JD, Workman ML (1997) Digital control of dynamic systems. Addison-Wesley, New York

    Google Scholar 

  • Govindasamy B, Caldeira K (2000) Geoengineering Earth’s radiation balance to mitigate CO2-induced climate change. Geophys Res Lett 27:2141–2144

    Article  Google Scholar 

  • Hansen J, Lacis A, Rind D, Russell G, Stone P, Fung I, Ruedy R, Lerner J (1984) Climate sensitivity: analysis of feedback mechanisms. In: Climate processes and climate sensitivity, vol 29 of geophysical monograph, pp 130–163. Am Geophys Union

  • Hansen J et al (2005) Efficacy of climate forcings. J Geophys Res 110(D18104)

  • Held IM, Winton M, Takahashi K, Delworth T, Zeng F, Vallis GVK (2010) Probing the fast and slow components of global warming by returning abruptly to preindustrial forcing. J Clim 23:2418–2427

    Article  Google Scholar 

  • IPCC (2007) Climate change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change

  • Jarvis A, Leedal D (2012) The geoengineering model intercomparison project (GeoMIP): a control perspective. Atmos Sci Lett 13:157–163

    Article  Google Scholar 

  • Jarvis A, Leedal D, Taylor CJ, Young P (2009) Stabilizing global mean surface temperature: a feedback control perspective. Environ Model Softw 24:665–674

    Article  Google Scholar 

  • Jones C (2003) A fast ocean GCM without flux adjustments. J Atmos Ocean Technol 20:1857–1868

    Article  Google Scholar 

  • Keith D (2000) Geoengineering the climate: history and prospect. Annu Rev Energy Environ 25:245–284

    Article  Google Scholar 

  • Kravitz B, Robock A, Boucher O, Schmidt H, Taylor KE, Stenchikov G, Schulz M (2011) The geoengineering model intercomparison project (GeoMIP). Atmos Sci Lett 12:162–167

    Article  Google Scholar 

  • Kravitz B, Caldeira K, Boucher O, Robock A, Rasch PJ, Alterskjær K, Bou Karam D, Cole JNS, Curry CL, Haywood JM, Irvine PJ, Ji D, Jones A, Lunt DJ, Kristjánsson JE, Moore J, Niemeier U, Schmidt H, Schulz M, Singh B, Tilmes S, Watanabe S, Yang S, Yoon J-H (2013) Climate model response from the Geoengineering Model Intercomparison Project (GeoMIP). J Geophys Res (submitted)

  • Latham J (1990) Control of global warming? Nature 347:339–340

    Article  Google Scholar 

  • Lebedoff SA (1988) Analytic solution of the box diffusion model for a global ocean. J Geophys Res 93(D11):14243–14255

    Google Scholar 

  • Li S, Jarvis A (2009) Long run surface temperature dynamics of an A-OGCM: the HadCM3 4 × CO2 forcing experiment revisited. Clim Dyn 33:817–825

    Article  Google Scholar 

  • Lunt DJ, Ridgwell A, Valdes PJ, Seale A (2008) Sunshade World: a fully coupled GCM evaluation of the climatic impacts of geoengineering. Geophys Res Lett 35:L12710

    Article  Google Scholar 

  • MacMartin DG, Keith DW, Kravitz B, Caldeira K (2013) Management of trade-offs in geoengineering through optimal choice of non-uniform radiative forcing. Nat Clim Chang 3:365–368

    Google Scholar 

  • MacMynowski DG, Tziperman E (2010) Testing and improving ENSO models by process rather than by output, using transfer functions. Geophy Res Lett 37(L19701)

  • MacMynowski DG, Keith DW, Caldeira K, Shin H-J (2011a) Can we test geoengineering? Energy Environ Sci 4:5044–5052

  • MacMynowski DG, Shin H-J, Caldeira K (2011b) The frequency response of temperature and precipitation in a climate model. Geophys Res Lett 38:L16711

    Article  Google Scholar 

  • Meinshausen M, Smith SJ, Calvin KV, Daniel JS, Kainuma MLT, Lamarque J-F, Matsumoto K, Montzka SA, Raper SCB, Riahi K, Thomson AM, Velders GJM, van Vuuren D (2011) The RCP greenhouse gas concentrations and their extension from 1765 to 2300. Clim Chang 109:213–241

    Google Scholar 

  • Morantine M, Watts RG (1990) Upwelling diffusion climate models: analytical solutions for radiative and upwelling forcing. J Geophys Res 95(D6):7563–7571

    Article  Google Scholar 

  • Moreno-Cruz J, Ricke K, Keith DW (2011) A simple model to account for regional inequalities in the effectiveness of solar radiation management. Clim Chang 110(3–4):649–668. doi:10.1007/s10584-011-0103-z

    Article  Google Scholar 

  • Morton O (2007) Climate change: Is this what it takes to save the world? Nature 447:132–136

    Article  Google Scholar 

  • Ricke KL, Granger Morgan M, Allen MR (2010) Regional climate response to solar-radiation management. Nat Geosci 3:537–541

    Article  Google Scholar 

  • Roe GH, Baker MB (2007) Why is climate sensitivity so unpredictable? Science 318:629–632

    Article  Google Scholar 

  • Sutton RT, Dong B, Gregory JM (2007) Land/sea warming ratio in response to climate change: IPCC AR4 model results and comparison with observations. Geophys Res Lett 34(L02701)

  • Watterson IG (2000) Interpretation of simulated global warming using a simple model. J Clim 13:202–215

    Article  Google Scholar 

Download references

Acknowledgments

Ben Kravitz is supported by the Fund for Innovative Climate and Energy Research. The Pacific Northwest National Laboratory is operated for the U.S. Department of Energy by Battelle Memorial Institute under contract DE-AC05-76RLO1830. Peter Thompson of Systems Technology Inc. provided assistance with the content of Appendix 2.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Douglas G. MacMartin.

Appendices

Appendix 1: Time-domain calculations with box-diffusion model

Equations (12) can be solved using the Laplace transform. From (2), the temperature in the deep ocean satisfies

$$ T_d(s,z)=M(s)e^{-\sqrt{s/\kappa}z} $$
(14)

for some function M(s), where s is the Laplace variable. Subtituting this into the Laplace transform of (1) and solving for M yields (3).

From Laplace transform tables, the response of a semi-infinite diffusion model (Eq. (3) with C = 0) to a unit step change in radiative forcing at t = 0 is

$$ g_{sd}(t)=\frac{1}{\lambda}\left(1-e^{t/\tau}\hbox{erfc}(\sqrt{t/\tau})\right) $$
(15)

where erfc denotes the complementary error function. With the surface layer included, the step response can be obtained by first factoring G(s) in Eq. (3) as

$$ G(s)=\frac{1/\xi}{\sqrt{s}+b}-\frac{1/\xi}{\sqrt{s}+a} $$
(16)

similar to the derivation in Lebedoff (1988) or Morantine and Watts (1990), where we introduce

$$ \xi=\sqrt{\beta^2/\kappa-4C\lambda}=\lambda\sqrt{\tau-4C/\lambda} $$
(17)

and \(a,b=(\beta/\sqrt{\kappa}\pm\xi)/(2C)\) satisfying (aξ)−1 − (bξ)−1 = λ −1. As in (15), the step response is then

$$ g_{bd}(t)=1/\lambda-1/(b\xi)e^{b^2t}\hbox{erfc} (b\sqrt{t})+1/(a\xi)e^{a^2t}\hbox{erfc}(a\sqrt{t}) $$
(18)

For 4Cλ τ as here, then a 2λ 2 τ/C 2 and b 2≃ 1/τ, and the first two terms in Eq. (18) are approximately the same as the step response of the semi-infinite diffusion model in Eq. (15), while the final term provides a correction for small t/τ. In calculating this final term, note that for \(a\sqrt{t}\gg 1\) then

$$ e^{a^2t}\hbox{erfc}(a\sqrt{t})\simeq \frac{(2/\sqrt{\pi})}{(a\sqrt{t}+\sqrt{a^2t+2})} $$
(19)

The simulations here consider only the average temperature over each year. For semi-infinite diffusion, the average temperature in the nth year after a step change in radiative forcing is

$$ \int\limits_{t=n-1}^{t=n}g_{sd}(t)dt= \frac{1}{\lambda}-\frac{1}{\lambda}\left(2\sqrt{\tau t/\pi}+\tau e^{t/\tau}\hbox{erfc}(\sqrt{t/\tau})\right)_{n-1}^n $$
(20)
$$ =\frac{1}{\lambda}[1-q(\tau^2;n)] $$
(21)

where this defines the function q(a;n). Then for the box-diffusion model,

$$ h(n)=\int\limits_{t=n-1}^{t=n}g_{bd}(t)dt= \frac{1}{\lambda}+\frac{1}{\xi}\left[\frac{q(a;n)}{a}-\frac{q(b;n)}{b}\right] $$
(22)

Simulations here also assume that the radiative forcing is held constant over each year. Given a sequence of forcings f(k) applied during year k, then since the sequence h(n) gives the annual-average temperatures due to a unit step forcing starting at k = 0, the temperature response to the sequence f(k) can be expressed as

$$ T(n)=\sum_{k=1}^{n}h(k)f(n-k)=\sum_{k=0}^{n-1}h(n-k)f(k) $$
(23)

Appendix 2: Frequency-domain calculations

The dynamic response of the climate system with feedback, shown in Figs. 5, 6, 7 and 8, can be understood and an approximate prediction made using G(s) in Eq. (3), K(s) in Eq. (10), and approximating the effects of the N-year averaging with the Laplace transform of a pure time delay, e Ns (obtained from the Laplace transform of \(\tilde{y}(t)=y(t-N)\)).

However, more accurate calculations of the frequency response requires taking into account that updates of solar forcing are only made at discrete time-intervals, not as a continuous function of time. The feedback system including this detail is shown in Fig. 14, where the block G(s) describes the continuous-time evolution of the climate response to forcing as before, A(s) represents the averaging of the output over the past N years, which is then sampled every N years, and Z(s) is a “zero-order hold” that describes the assumption that the solar reduction computed at every N-year decision point is held constant over the subsequent N years. The discrete-time PI control law

$$ u(k)=k_py(k)+k_i\sum_{n=0}^{n=k}y(n) $$
(24)

is represented by its z-transform, K(z). Analysis details can be found in any discrete-time controls textbook (e.g. Franklin et al. 1997); here we simply provide the required formulae used in computing the results herein.

Fig. 14
figure 14

Block diagram of geoengineering feedback, as in Fig. 2, but with additional detail required for accurately predicting dynamics. The response of the climate system G(s) is averaged over the previous N years [A(s)], sampled, and the actual feedback law implemented in discrete-time [K(z)] rather than continuous-time. The desired SRM forcing at each discrete decision point in time is assumed to be held constant for the next N years, until the next sample is made. The system within the dashed box is sampled, resulting in aliasing

For frequencies less than the Nyquist frequency (half the sampling rate), then the Laplace transform of the discrete-time control law K(z) can be obtained by setting z = e Ns, yielding

$$ K(s)=k_p+k_i\frac{N}{1-e^{-Ns}} $$
(25)

The Laplace transform of the N-year averaging process is

$$ A(s)=\frac{1-e^{-Ns}}{Ns} $$
(26)

which is used in all calculations here, and at frequencies small compared to 1/N, behaves similarly to a pure time delay of N/2 years. Maintaining a constant value of the applied radiative forcing for N years (a zero-order-hold) yields Z(s) = A(s), so that A(s)Z(s) has an effect similar to that of an N-year time delay.

Finally, note that sampling the continuous-time system G az (s) = A(s)G(s)Z(s) at N-year intervals results in aliasing. That is, temperature variations with frequency f and variations at frequency 1/N − f are indistinguishable in the sampled signal. (There are an infinite sequence of indistinguishable frequencies, but only the first is significant in predicting the response.) Thus at frequencies below the Nyquist frequency, the system G s (s) within the dashed lines of Fig. 14 is approximately

$$ G_s(i\omega)=G_{az}(i\omega)+G_{az}(2\pi/N-i\omega) $$
(27)
$$ =G_{az}(i\omega)+G_{az}^*(i\omega-2\pi/N) $$
(28)

with \((\cdot)^*\) denoting complex-conjugate. The loop transfer function in Figs. 3 and 4 and the subsequent calculations of the sensitivity function are obtained using K(s) in (25) and G s (s) in (28), where the latter depends not only on G(s) in (3) but also on A(s) and Z(s) in (26).

Rights and permissions

Reprints and permissions

About this article

Cite this article

MacMartin, D.G., Kravitz, B., Keith, D.W. et al. Dynamics of the coupled human–climate system resulting from closed-loop control of solar geoengineering. Clim Dyn 43, 243–258 (2014). https://doi.org/10.1007/s00382-013-1822-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-013-1822-9

Keywords

Navigation