Investigating climate tipping points under various emission reduction and carbon capture scenarios with a stochastic climate model††thanks: Accepted for publication in Proc. Roy. Soc. A
Abstract
We study the mitigation of climate tipping point transitions using an energy balance model. The evolution of the global mean surface temperature is coupled with the CO2 concentration through the green house effect. We model the CO2 concentration with a stochastic delay differential equation (SDDE), accounting for various carbon emission and capture scenarios. The resulting coupled system of SDDEs exhibits a tipping point phenomena: if CO2 concentration exceeds a critical threshold (around ppm), the temperature experiences an abrupt increase of about six degrees Celsius. We show that the CO2 concentration exhibits a transient growth which may cause a climate tipping point, even if the concentration decays asymptotically. We derive a rigorous upper bound for the CO2 evolution which quantifies its transient and asymptotic growths, and provides sufficient conditions for evading the climate tipping point. Combining this upper bound with Monte Carlo simulations of the stochastic climate model, we investigate the emission reduction and carbon capture scenarios that would avert the tipping point.
1 Introduction
It is highly likely that anthropogenic greenhouse gases are responsible for more than half of the increase in the global mean surface temperature between 1951 to 2010 [1]. Therefore, reducing the atmospheric concentration of greenhouse gases must be a central component of any climate change mitigation strategy. This reduction can be achieved in two ways. One involves reducing CO2 emissions through alternative energy sources, increased fuel efficiency, and reduced consumption [2]. A second approach is expanding carbon sinks, e.g., by employing carbon capture and storage (CCS) technologies, which capture CO2 from large emitting sources (such as factories) or directly from the atmosphere and prepares it for long-term storage [3, 4].
What complicates these matters is the possible existence of climate tipping points, i.e., climate regimes where small changes significantly alter the future state of the system [5, 6, 7]. This tipping behavior can involve the sudden disruption of climatological and ecological processes such as melting of ice sheets or large-scale death of rainforests [8]. Even if emission reduction and carbon capture strategies lead to long-term CO2 reduction, its transient growth can trigger a climate tipping point with adversely irreversible impact. Therefore, mitigation strategies not only have to ensure long-term reduction of greenhouse gasses, but also ensure that their transient response remains below the critical tipping point levels.
The purpose of the present study is to quantify emission reduction and carbon capture scenarios that mitigate climate tipping points. To this end, we use a Budyko–Sellers-type model of the climate [9, 10]. This energy balance model assumes that the Earth’s radiation output is balanced by radiation input, and represents key aspects of glacial-interglacial climate transitions [11]. It relies on the hypothesis that glacial cycles are triggered by variations in the Earth’s orbit which alter the incoming solar radiation to the earth. These orbital cycles are amplified by limited feedback between greenhouse gases and temperature [12]. The resulting governing equations are a one-way coupling between the global mean surface temperature and CO2 concentration, where the temperature is affected by CO2 concentration through the greenhouse effect.
We model the evolution of the CO2 concentration by a stochastic delay differential equation (SDDE), which takes into account the carbon emission and capture rates. This model allows us to parameterize the possible reduction in CO2 emission rates as well as the capability of carbon sinks to absorb atmospheric CO2. In particular, we show that, even when the emission rates are reduced, the CO2 concentration exhibits a transient growth which may instigate a climate tipping point. We investigate the response of global mean surface temperature to various emission reduction and carbon capture scenarios, identifying the scenarios which will mitigate the climate tipping point transition.
1.1 Related work
The study of climate tipping points can be divided into three broad categories: modeling, prediction, and mitigation.From the modeling perspective, it is important to accurately parameterize various contributing factors such as green house gasses, water vapor, clouds, and ice sheets [13, 14, 15, 16]. Although here we use a very simple energy balance model, it takes into account the main culprit, i.e., CO2 emissions. Furthermore, we choose the model parameters such that the resulting climate sensitivity agrees with the available estimates from more elaborate models.
Prediction is concerned with early warning signs embedded in observational data that may indicate an upcoming climate tipping point [17]. Several such precursors have been proposed, including critical slowing down [18] and increased variability [19]. Moreover, it has been observed that the stochastic component of the system changes its characteristics close to a tipping point. For instance, Held and Kleinen [20] and Livina and Lenton [21] model the North Atlantic thermohaline circulation and note that the stochastic component of the data changes from white noise to red noise. The same transition is observed by Prettyman et al. [22] who studied tropical cyclones.
From the mitigation standpoint, it is widely accepted that a significant reduction in CO2 emissions is a necessity [23, 24, 7]. A complementary solution is to remove CO2 from the atmosphere using carbon capture technologies [25]. For instance, studying the Atlantic thermohaline circulation (THC), Bahn et al. [26] find that a drastic and fast CO2 reduction is required to avoid disrupting the THC. However, it is argued by Ritchie et al. [27] that crossing a threshold will not necessarily result in a simultaneous extreme change in the climate system. They find that the determining factor is how far the tipping point threshold is exceeded and for how long.
Previous studies mainly focus on the asymptotic climate state as a result of increasing CO2 concentration. In contrast, a main focus of our work is the transient climate dynamics in response to various emission reduction and carbon capture scenarios. In particular, we show that, under certain emission reduction and carbon capture scenarios, the CO2 concentration exhibits a transient growth large enough to trigger a climate tipping point, even though the concentration decays asymptotically. We note that, beyond emission reduction and carbon capture, geoengineering ideas have been proposed [28]. These methods seek to increase the reflected energy of the Sun by releasing aerosol particles into the stratosphere. These geoengineering ideas are beyond the scope of the present work and are not considered here.
1.2 Outline of this paper
This paper is organized as follows. We discuss the climate model and its parameters in sections 2 and 3. In section 4, we derive a rigorous upper bound for the CO2 levels under various emission reduction and carbon capture scenarios. We discuss the temperature variations under each scenario in section 5 using direct numerical simulations. Section 6 contains our concluding remarks.
2 Stochastic climate model
We use two stochastic differential equations for the climate system, modeling the global mean surface temperature, , and average concentration of CO2, , in the atmosphere. The equation for temperature is a Budyko-Sellers-type model, derived from the balance between incoming and outgoing radiations [9, 10]. The resulting temperature model reads,
(1) |
where the temperature is in units of Kelvin, is the concentration of CO2 in parts per million (ppm), and is the thermal inertia in units of . The term represents short-wave radiation from the Sun, where is the solar input in units of . The multiplier denotes temperature dependent albedo that accounts for the light reflecting off the Earth surface. The term models the effect of greenhouse gases, where (in units of ) is the direct forcing of CO2 and determines the sensitivity of the climate equilibrium [29]. The parameter represents the trapping of outgoing radiations by greenhouse gases [11]. The constant denotes the preindustrial concentration of CO2 in units of ppm. Finally, the last term represents long-wave radiation from the Sun, where represents the outgoing long wave radiation that is modified by the emissivity .
Our choice of the temperature dependent albedo function shown in Fig. 1 is similar to [30], although we use a different set of parameters as reported in Table 1. In particular, we define
(2) |
which transitions smoothly between albedo parameters and , where represents the current global albedo, while represents the global albedo of the Earth. The case corresponds to the melting of ice on the Earth’s surface, whereas the case represents the formation of ice. Here, we only consider the case which conforms to the present trends.
The melting of ice depends on the temperature levels. The function is chosen to reflect this temperature dependence, so that transitions smoothly between the albedo at temperature , to the threshold of at temperature . More precisely, we define
(3) |
where
(4) |
is a smooth approximation of the unit Heaviside function. The parameter controls the transition rate between temperatures and .

Following [30], we model the unresolved subgrid processes by a stochastic term , where is a constant amplitude and denotes the standard Wiener process. The subscript is used to distinguish the stochastic processes affecting the temperature from those affecting CO2, to be described shortly. Adding this stochastic term, the temperature model reads
(5) |
where is defined in equation (1). Note that, although the driving stochastic force is a Gaussian white noise, the response itself is non-Gaussian and non-white owing to the nonlinear nature of .
To close the equations, we also need to model the evolution of CO2. Dijkstra and Viebahn [29] prescribe as an explicit function of time. Ashwin and von der Heydt [30] model the CO2 evolution as a random walk confined to an interval . To mimic the current trends and to allow for possible emission reduction and carbon capture, we model the CO2 evolution as a stochastic delay differential equation,
(6) |
where is a time-dependent emission rate, is a time-independent carbon capture rate, and is a time delay. The uncertainties in CO2 concentration are modeled with the stochastic term , where is a constant noise intensity and is a standard Wiener process.
The source term in equation (6) models the CO2 emission into the atmosphere. We allow for a time-dependent rate to model reduction or enhancement of the CO2 emissions. The sink term, on the other hand, models the CO2 captured from the atmosphere either through natural phenomena (e.g., photosynthesis [31]) or through artificial technologies (e.g., chemical looping [32]). For the CO2 sinks, we allow for a time delay to model the time between carbon capture and its effect being felt in the atmospheric concentration. This type of delay is typical in control theory, where there is often a lag between a modifying action and its actualization [33, 34]. If this delay is non-existent or negligible, one can set . Here, we set year, representing the time it takes for CO2 data to be updated and the carbon capture strategies adjusted correspondingly. Nonetheless, we have varied the time delay in the interval years (not shown here) and only observed variations of approximately ppm in the CO2 evolution, which do not significantly alter the results.
Equations (5) and (6) form a closed set of equations modeling the global mean temperature and CO2 concentration . Our main goal here is to investigate whether a combination of emission reduction and carbon capture may avert a possible upcoming climate tipping point. To this end, we consider various scenarios in terms of emission reduction and carbon capture as discussed in section 3.
3 Tipping points and model parameters
The parameter values used in the model are listed in Table 1 and their choice is discussed in Appendix B. For a prescribed CO2 concentration, the temperature model (5) exhibits three distinct regimes as shown in Fig. 2. Regime 1: If ppm, the temperature has a single stable equilibrium satisfying .
Parameter | Value | Units | Parameter | Value | Units |
---|---|---|---|---|---|
Regime 2: For CO2 concentrations a bifurcation takes place, whereby two additional equilibria are born. One of the new equilibria is unstable (dashed line in Fig. 2) whereas the other one is stable. As a result, in this regime the temperature model is bistable. Regime 3: If , the model switches back to a single stable equilibrium. However, in this regime, the global mean temperature is significantly higher than the equilibrium temperature in regime 1.

In our model, the CO2 concentration is not constant, instead it evolve according to the SDDE (6). We choose the initial conditions C and ppm which reflect the current global mean temperature and CO2 concentration, respectively [35, 36, 37]. The model parameters, listed in Table 1, are chosen such that the current climate lies in bistable regime 2 near the lower branch of equilibria. If the current trends continue, i.e., the CO2 concentration keeps increasing, the temperature also continues to increase gradually. Eventually, one reaches the tipping point where a slight increase in CO2 concentration leads to a dramatic increase in the temperature by about six degrees in Celsius. Our goal is to determine the emission reduction and carbon capture scenarios that would avert this catastrophic climate tipping point.
To this end, we allow for a time-dependent emission rate as shown in figure 3. It contains three adjustable periods. First is a period of inaction, up to time , where the emission rate remains constant at its current level . It is followed by a reduction period, where the emission rate decreases linearly toward its terminal value . This reduction continues for a period of years until it plateaus at time . The period constitutes the terminal stage where the emission rate remains at the constant level .

Therefore, the emission rate has the general form
(7) |
where for an integer . The current emission rate is estimated from the data in Ref. [38]. They estimate that there are gigatons of carbon emissions each year from industrial processes and land usage. Converting this to concentration yields ppm added to the atmosphere each year. The current emission rate is then estimated from ppm/s, where ppm is the current CO2 concentration. This yields . In the following sections, we study the evolution of the global mean temperature under various choices of the remaining free parameters, , and .
To conclude the choice of parameters, we need to specify the parameters and in the CO2 model (6). Recall that the parameter denotes the capacity of carbon sinks to remove CO2 from the atmosphere. Assuming that carbon capture takes place only due to natural phenomena, the value of can be estimated from the data in Ref. [38]. They estimate that approximately gigatonnes of carbon are taken from the atmosphere each year from natural processes such as photosynthesis or absorption by the oceans. Converting this to concentration yields ppm removed from the atmosphere each year. The parameter , then can be estimated by ppm/s, which yields .
Finally, we allow for a delay in the CO2 sinks, which can be set to zero if no such delay exists. We have not been able to determine this parameter from the available literature. The results in Section 5 are reported for the delay time equal to one year. However, we examined delay times up to three years and did not observe a significant effect on the results.
4 Transient and asymptotic growth of CO2
The system of equations (5) and (6) is a one-way coupling between the temperature and the CO2 concentration , whereby the CO2 concentration evolves independently of the temperature. As discussed in section 3, if the CO2 concentration increases beyond ppm, the climate system undergoes a tipping point transition leading to an abrupt increase of the global mean surface temperature (see Fig. 2).
The main objective of this paper is to determine emission reduction and carbon capture scenarios that ensure the mitigation of this tipping point. Extreme events, such as tipping point transitions, have been the subject of much research, with an emphasis on their causal mechanisms [39, 40, 41], probabilistic quantification [42, 43, 44, 45], and data-driven prediction [46, 47, 48, 18]. Only recently, control strategies for mitigating extreme events have been proposed [46, 34, 49]. In particular, Farazmand [34] proposes a time-delay feedback control for mitigating noise-induced transitions in multistable systems. This control strategy relies on the stationary equilibrium density of the system. Given the time-dependent emission rate and the linearity of the CO2 model (6), it does not possess such an equilibrium density. Consequently, the framework of Ref. [34] is not applicable here.
Therefore, we take a different approach here and derive a quantitative upper bound for the CO2 concentration which determines whether the climate tipping point transitions can be averted. To this end, we consider the CO2 model (6) without the stochastic term,
(8) |
Note that (8) is a delay differential equation which requires the initial condition to be specified as a function over the interval . Since we consider a relatively short delay of one year, the CO2 concentration does not change significantly over this time interval. As a result, we assume the constant initial condition for all , where ppm is the current level of CO2 concentration. The following theorem provides an upper bound for the solutions . Adding the stochastic term only leads to small fluctuations around this upper bound, without fundamentally altering the results.
Theorem 4.1.
Let , and . Assume that is locally integrable and that is a measurable, locally integrable function solving the delay differential equation (8). Then
(9) |
where
(10) |
Proof.
See Appendix A. ∎
Recall that the climate tipping point occurs if the CO2 levels exceed ppm. Therefore, ensuring that the upper bound (9) is uniformly below is a sufficient condition for avoiding the tipping point. This upper bound encapsulates the competition between CO2 emission rate and the carbon capture rate . The negative-valued function , appearing in the exponent of the upper bound, is proportional to the carbon capture rate . But it also depends on the emission rate due to the delay . This leads to a non-trivial dependence of the upper bound on the carbon emission and capture rates. In section 5, we investigate the shape of this upper bound for various parameter values.
The upper bound (9) also provides a sufficient condition for the asymptotic decay of CO2 concentration.
Corollary 4.1.
Assume the conditions of Theorem 4.1. If the carbon capture rate satisfies , then we have
Proof.
We emphasize that, to avert the climate tipping point, it is not sufficient for the CO2 concentration to decay asymptotically. As we show in Section 5, even a transient growth of CO2 that exceeds the critical threshold will lead to a tipping point transition.
5 Results and discussion
In this section, we present the numerical results obtained from the model (5)-(6). First, we focus on the CO2 model (6) and investigate the behavior of its upper bound (9). Recall that to avoid the climate tipping point it is sufficient for this upper bound to remain below the critical level .
Figure 4 shows the upper bound (9) with the delay time year, emission reduction time span years, and the carbon capture rate . The initial CO2 level is set at ppm, which is the estimated CO2 concentration in the year 2019 [37]. Figure 4 represents the optimistic scenario where emission reductions begin immediately () and continue to decrease linearly for years. The figure shows three terminal emission rates with . In every case, we observe a transient growth of CO2 levels. But none of them reach the tipping point and they decay asymptotically.


On the other hand, figure 4 shows the case where the transient growth surpasses the tipping point . In this figure, we fix the terminal CO2 emission rate at with the time window for reaching this rate being years. We vary the number of years before the linear reduction in emission rate begins. If reductions begin in years, the tipping point can still be avoided. However, if years, the tipping point may be reached and consequently the global mean surface temperature may sharply increase by about six degrees. These observations accentuate the need for immediate action on carbon emission reduction.
We emphasize that in every case shown in figure 4 the CO2 levels decay asymptotically. However, the more germane factor is whether the transient growth of CO2 would exceed the climate tipping point .

Next we consider the full model, i.e., the couple system of equations (5)-(6). Recall that this is a one-way coupling, where the temperature is affected by the CO2 concentration through the green house effect. In contrast, the CO2 concentration evolves independently of the temperature.
This system of stochastic delay differential equations is integrated numerically using the predictor-corrector scheme developed by Cao et al. [50]. The numerical integration is carried out after nondimensionalizing the equations by defining , , and , where ppm denotes the preindustrial CO2 level, K denotes the emissivity threshold given by Ashwin and von der Heydt [30], and s is an arbitrary time scale (approximately one-third of a year). The time step of the numerical integrator in the non-dimensional time is , which is equivalent to one day in dimensional time.
To investigate the behavior of the climate system with regards to various emission reduction scenarios, we simulate the model over a 200 year time span for a range of parameters and . Recall that determines the terminal emission rate and determines the time it takes to reach this plateau (see figure 3). For each set of parameters , we simulate realizations of the stochastic climate model and compute the ensemble average of the global mean surface temperature and the CO2 concentration. In all simulations, the initial conditions are set to Celsius and ppm, which reflect the estimated values in the year 2019 [35, 36].
Figure 5 shows the ensemble means of CO2 concentration and temperature as a function of time and the parameters . This figure has three panels as described below:
Figures 5(a)-(b) correspond to the scenario where emission reductions begin immediately (). In panel (a), the carbon capture rate is equal to the empirical value . In this case, the CO2 concentration increases transiently, but does not reach its critical value . As a result, no tipping point phenomena is observed. In fact, the temperature does not exceed degrees Celsius and decays towards degrees asymptotically.
Panel (b) is identical to (a) except that we use a lower carbon capture rate . The carbon capture rate may decrease over time due to various factors, but most prominently due to deforestation [51, 52]. In this case, the CO2 concentration still undergoes a transient growth before eventually decreasing. Because of the lower carbon capture rate, however, the transient growth surpasses the critical threshold , and leads to a drastic increase in the global mean surface temperature. This tipping point does not occur for all carbon emission scenarios. If the transition period is short enough and the terminal emission rate is small enough, the tipping point can still be mitigated.
In fact, as shown in figure 6(a), there is a nonlinear boundary in the parameter space which separates the tipping point transitions from its successful mitigation. For , the tipping point occurs regardless of the transition time . Therefore, the terminal carbon emission rate has to reduce to at least one-third of its current value to mitigate the climate tipping point. Larger allows for longer transition period . In figure 6(b), we show the ensemble average of the temperature for three parameter values . The shaded areas mark one standard deviation from the mean. The point on the boundary has a large standard deviation because the stochastic process can kick the trajectory towards the tipping regime or away from it. The parameters and lead to the tipping behavior with high probability, where the temperature rises to about Celsius. In contrast, successfully averts the climate tipping point, after a slight transient increase in the global mean surface temperature.

Finally, figure 5(c) shows the case where emission reduction is delayed by 25 years (), and the carbon capture rate stays at its current level (). In this case, CO2 concentration again exhibits a transient growth past the critical value , and as a result, a drastic increase in global temperature ensues. Although the CO2 concentration reduces asymptotically to as low as ppm, the transient tipping point will have dire consequences, such as sea level rise and droughts. This observation reaffirms the need for immediate action towards significant emission reduction. Otherwise, the only alternative would be to significantly increase the CO2 sinks , through artificial carbon capture technologies.
6 Conclusions
We considered a stochastic climate model as a one-way coupling between CO2 concentration and the global mean surface temperature . The CO2 concentration is governed by a stochastic delay differential equation, allowing for the modeling of various emission reduction and carbon capture scenarios. The temperature satisfies a stochastic differential equation derived from energy balance (Budyko–Sellers model). The temperature is coupled with the CO2 equation to model the effect of greenhouse gasses.
The model has a tipping point behavior: if the CO2 concentration exceeds the critical value of ppm, the global mean surface temperature will increase abruptly by about six degrees Celsius. The CO2 model exhibits transient growth which allows for reaching this tipping point in finite time, even when CO2 decreases asymptotically. We derived a tight upper bound for the CO2 concentration, which provides sufficient conditions for mitigating the climate tipping point (see Theorem 4.1). This upper bound depends on several parameters such as the CO2 emission rate , the carbon capture rate , and the delay time . However, since the upper bound is explicitly known, it can be analyzed numerically with low computational cost in order to determine the emission reduction and carbon capture scenarios which would mitigate the climate tipping point.
We examined various emission reduction scenarios by combining our analytic results with Monte Carlo simulations of the climate model. In particular, we find that the climate tipping point in our model can be averted if CO2 emission reductions begin immediately and the emission rate decreases to one-third of its current level within 50 years. However, if these reductions are delayed by as much as ten years, the transient growth of the CO2 concentration will exceed the tipping point value, leading to a drastic and abrupt increase in the global mean surface temperature. In the latter case, increasing the carbon capture rate could still avert the tipping point.
It remains to be seen whether our conclusions carry over to more sophisticated climate models, such as box models or general circulation models. We will investigate such models in future work. Although the temperature model would be more complex, our analysis of the CO2 concentration can be used to derive sufficient conditions that guarantee tipping point evasion. Furthermore, deriving an invariant probability distribution for the CO2 model is of great interest as it would quantify the probability of transitions which in turn leads to necessary and sufficient conditions for mitigating the climate tipping point. Finally, the CO2 model itself can be improved, e.g., by using data-driven discovery methods to obtain simple models that agree with observational data [53].
Data availability
The data and code for reproducing the results of this manuscript are publicly available at https://github.com/mfarazmand/ClimateMitigation
Acknowledgments
We are grateful to Prof. Tamás Tél for recommending reference [30], and Prof. Pedram Hassanzadeh for fruitful conversations. This work was partially supported by the National Science Foundation grant DMS-1745654.
Appendix A Proof of Theorem 4.1
We begin by writing equation (8) as the integral equation,
(11) |
We derive a tight upper bound for using a modified version of a Grönwall-type inequality for delay differential equations given in [54]. We first prove a result with , which we later use to address the general case where the emission rate is non-zero.
Proposition A.1.
Let , , , and be locally integrable. Let be Borel measurable and locally bounded such that
(12) |
and be any locally integrable function that satisfies the inequality,
(13) |
Then
where
Proof.
Note that in the context of our climate model, we have . Györi and Horváth [54] proved a similar result with a plus sign in front of the integral in Eq. (12). We generalize their result to the case with a minus sign in front of the integral as required by our climate model. We first define
Inequality (12) implies
The inequality (13) for then yields
Defining,
we have for all . We now prove that, in fact, for all .
Let . By the definition of , we have
Since is continuous on , there exists such that and
Assume there exists such that . Since is continuous on , we can assume that for all .
It then follows from that
This contradicts . Therefore, we must have for all Since is arbitrarily close to and is continuous, we have for all .
Finally, implies
∎
Now we return to equation (11) and derive an upper bound for for the general case with .
Theorem A.1.
Let , , and , be locally integrable. Let be Borel measurable and locally bounded such that
(14) |
and be any locally integrable function satisfying
(15) |
Then
where
Proof.
Although this theorem provides an upper bound for equation (8), the issue of finding a suitable choice of remains. To obtain a tight upper bound, must be strictly negative. To this end, we choose
which is negative definite and satisfies the inequality (15). Recall that we assumed the constant initial condition for all (cf. equation (8)). This observation, together with the fact that is negative, yields . Therefore the upper bound in Theorem A.1 simplifies to
which is the desired result of Theorem 4.1.
Appendix B Model parameters
Our model parameters mostly agree with those chosen by Dijkstra and Viebahn [29]. However, there are a few differences that we justify in this section. These changes are mostly made so that our model parameters more closely approximate the current climate state.
The main differences appear in the albedo function (2). Dijkstra and Viebahn [29] use the initial albedo value which does not agree with the estimates of the current albedo of the Earth. We instead use which agrees with the value inferred from empirical satellite observations [55, 56, 57]. The temperature dependent albedo (2) requires albedo values to be paired with a corresponding temperature threshold. A reasonable choice is to pair the current albedo of Earth with the temperature K, which roughly corresponds to the current global mean surface temperature.
The terminal albedo value , with the corresponding threshold temperature K, is chosen to reflect the state of the Earth after significant amounts of warming. We point out that the terminal albedo is most likely unrealistic since it will require a significant reduction in cloud coverage in addition to sea level rise. However, the model does not allow for much larger values of without losing the climate tipping point. As seen in figure 7, for any , the fold in the bifurcation curve ceases to exist and, as a result, the irreversible climate tipping point vanishes.

Dijkstra and Viebahn [29] use the smoothness parameter which leads to a rapid change between the initial albedo and the terminal albedo . Following Ashwin and von der Heydt [30], we use which corresponds to a more gradually decreasing albedo as shown in figure 1.
We use the CO2 forcing parameter . The direct effect of CO2 forcing yields the lower value . However, this value neglects other effects such as ice albedo and water vapor [11, 30], and results in a climate sensitivity that is much lower than the range given in AR5 [1]. We note that Ashwin and von der Heydt [30] use , but they allow for a temperature-dependent emissivity function to account for the water vapor feedback. With CO2 forcing parameter , the transient climate response (TCR) of our model is about C and its equilibrium climate sensitivity (ECS) is about C. Note that both the TCR and ECS of our model are only slightly larger than the current likely range given in AR5 [1], which are C and C, respectively.
References
- Pachauri and Meyer [2014] R. K. Pachauri and L. A. Meyer. Climate change 2014: Synthesis report. contribution of working groups i, ii and iii to the fifth assessment report of the intergovernmental panel on climate change. 2014.
- Stocker et al. [2013] T. F. Stocker, G.-K. D. Qin, M. Plattner, S. K. Tignor, J. Allen, A. Boschung, Y. Nauels, V. Xia, Bex, and P.M. Midgley (eds.). Summary for policymakers. in: Climate change 2013: The physical science basis. contribution of working group i to the fifth assessment report of the intergovernmental panel on climate change. 2013.
- Gibbins and Chalmers [2008] J. Gibbins and H. Chalmers. Carbon capture and storage. Energy Policy, 36(12):4317–4322, 2008.
- Haszeldine [2009a] R. S. Haszeldine. Carbon capture and storage: How green can black be? Science, 325(5948):1647–1652, 2009a.
- Drijfhout et al. [2015] S. Drijfhout, S. Bathiany, C. Beaulieu, V. Brovkin, M. Claussen, C. Huntingford, M. Scheffer, G. Sgubin, and D. Swingedouw. Catalogue of abrupt shifts in intergovernmental panel on climate change climate models. Proc. Natl Acad, Sci, 112(43), 2015.
- Lenton et al. [2008] T. M. Lenton, H. Held, E. Kriegler, J. W. Hall, W. Lucht, S. Rahmstorf, and H. J. Schellnhuber. Tipping elements in the earth’s climate system. Proceedings of the National Academy of Sciences, 105(6):1786–1793, 2008. doi: 10.1073/pnas.0705414105.
- Schneider et al. [2019] T. Schneider, C. M. Kaul, and K. G. Pressel. Possible climate transitions from breakup of stratocumulus decks under greenhouse warming. Nature Geoscience, 12(3):163–167, 2019.
- Lenton [2011] T. Lenton. Early warning of climate tipping points. Nature Clim Change, 1:201–209, June 2011.
- Budyko [1969] M. I. Budyko. The effect of solar radiation variations on the climate of the earth. Tellus, 21(5):611–619, 1969.
- Sellers [1969] W. D. Sellers. A global climatic model based on the energy balance of the earth-atmosphere system. J. Appl. Meteorol., 8(3):392–400, 1969.
- Hogg [2008] A. M. Hogg. Glacial cycles and carbon dioxide: A conceptual model. Geophys. Res. Lett., 35(L01701), 2008.
- McGuffie and Henderson-Sellers [2005] K. McGuffie and A. Henderson-Sellers. A Climate Modelling Primer. Chichester, West Sussex. J. Wiley, 2005.
- Schneider and Dickinson [1974] S. H. Schneider and R. E. Dickinson. Climate modeling. Reviews of Geophysics, 12(3):447–493, 1974. doi: https://doi.org/10.1029/RG012i003p00447.
- Oerlemans and van der Veen [1984] J. Oerlemans and C. J. van der Veen. Ice sheets and climate, volume 21. Springer, 1984.
- Dessler et al. [2013] A. E. Dessler, M. R. Schoeberl, T. Wang, S. M. Davis, and K. H. Rosenlof. Stratospheric water vapor feedback. Proceedings of the National Academy of Sciences, 110(45):18087–18091, 2013. doi: 10.1073/pnas.1310344110.
- Zelinka et al. [2020] M. D. Zelinka, T. A. Myers, D. T. McCoy, S. Po-Chedley, P. M. Caldwell, P. Ceppi, S. A. Klein, and K. E. Taylor. Causes of higher climate sensitivity in cmip6 models. Geophysical Research Letters, 47(1):e2019GL085782, 2020. doi: https://doi.org/10.1029/2019GL085782.
- Scheffer et al. [2012] M. Scheffer, S. R. Carpenter, T. M. Lenton, J. Bascompte, W. Brock, V. Dakos, J. van de Koppel, I. A. van de Leemput, Simon A. Levin, E. H. van Nes, M. Pascual, and J. Vandermeer. Anticipating critical transitions. Science, 338(6105):344–348, 2012. doi: 10.1126/science.1225244.
- Dakos et al. [2008] V. Dakos, M. Scheffer, E. H. van Nes, V. Brovkin, V. Petoukhov, and H. Held. Slowing down as an early warning signal for abrupt climate change. Proceedings of the National Academy of Sciences, 105(38):14308–14312, 2008. doi: 10.1073/pnas.0802430105.
- Lenton et al. [2012] T. M. Lenton, V. N. Livina, V. Dakos, E. H. van Nes, and M. Scheffer. Early warning of climate tipping points from critical slowing down: comparing methods to improve robustness. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 370(1962):1185–1204, 2012. doi: 10.1098/rsta.2011.0304.
- Held and Kleinen [2004] H. Held and T. Kleinen. Detection of climate system bifurcations by degenerate fingerprinting. Geophys. Res. Lett., 31(L23207), 2004.
- Livina and Lenton [2007] V. Livina and T. Lenton. A modified method for detecting incipient bifurcations in a dynamical system. Geophys. Res. Lett., 34, 02 2007.
- Prettyman et al. [2018] J. Prettyman, T. Kuna, and V. Livina. A novel scaling indicator of early warning signals helps anticipate tropical cyclones. Europhysics Letters, 121(1):10002, 2018.
- Cai et al. [2016] Y. Cai, T. M. Lenton, and T. S. Lontzek. Risk of multiple interacting tipping points should encourage rapid co2 emission reduction. Nature Climate Change, 6(5):520–525, 2016. doi: 10.1038/nclimate2964.
- Anderson et al. [2019] C. M. Anderson, R. S. DeFries, R. Litterman, P. A. Matson, D. C. Nepstad, S. Pacala, W. H. Schlesinger, M. R. Shaw, P. Smith, C. Weber, and C. B. Field. Natural climate solutions are not enough. Science, 363(6430):933–934, 2019. doi: 10.1126/science.aaw2741.
- Haszeldine [2009b] R. S. Haszeldine. Carbon capture and storage: how green can black be? Science, 325(5948):1647–1652, 2009b.
- Bahn et al. [2011] O. Bahn, N. R. Edwards, R. Knutti, and T. F. Stocker. Energy policies avoiding a tipping point in the climate system. Energy Policy, 39(1):334–348, 2011.
- Ritchie et al. [2019] P. Ritchie, Ö. Karabacak, and J. Sieber. Inverse-square law between time and amplitude for crossing tipping thresholds. Proc. R. Soc. A., 475, 2019.
- Keith [2000] D. W. Keith. Geoengineering the climate: History and prospect. Annual Review of Energy and the Environment, 25(1):245–284, 2000. doi: 10.1146/annurev.energy.25.1.245.
- Dijkstra and Viebahn [2015] H. A. Dijkstra and J. P. Viebahn. Sensitivity and resilience of the climate system: A conditional nonlinear optimization approach. Communications in Nonlinear Science and Numerical Simulation, 22(1):13 – 22, 2015. doi: https://doi.org/10.1016/j.cnsns.2014.09.015.
- Ashwin and von der Heydt [2019] P. Ashwin and A. S. von der Heydt. Extreme sensitivity and climate tipping points. Journal of Statistical Physics, 179(5-6):1531–1552, Nov 2019.
- Sedjo and Sohngen [2012] R. Sedjo and B. Sohngen. Carbon sequestration in forests and soils. Annual Review of Resource Economics, 4(1):127–144, 2012. doi: 10.1146/annurev-resource-083110-115941.
- Song et al. [2018] C. Song, Q. Liu, N. Ji, S. Deng, J. Zhao, Y. Li, Y. Song, and H. Li. Alternative pathways for efficient co2 capture by hybrid processes—a review. Renewable and Sustainable Energy Reviews, 82:215 – 231, 2018. doi: https://doi.org/10.1016/j.rser.2017.09.040.
- Dorf and Bishop [2011] R. C. Dorf and R. H. Bishop. Modern control systems. Pearson, 12 edition, 2011.
- Farazmand [2020] M. Farazmand. Mitigation of tipping point transitions by time-delay feedback control. Chaos: An Interdisciplinary Journal of Nonlinear Science, 30(1):013149, Jan 2020. doi: 10.1063/1.5137825.
- NASA [2021] NASA. Global temperature. https://data.giss.nasa.gov/gistemp/graphs/graph_data/Global_Mean_Estimates_based_on_Land_and_Ocean_Data/graph.txt, August 2021. URL https://data.giss.nasa.gov/gistemp/graphs/graph_data/Global_Mean_Estimates_based_on_Land_and_Ocean_Data/graph.txt.
- Hansen et al. [2010] J. Hansen, R. Ruedy, M. Sato, and K Lo. Global surface temperature change. Rev. Geophys, 48(RG4004), 2010.
- noa [2021] NOAA Global Monitoring Laboratory and Scripps Institute of Oceanography. https://www.esrl.noaa.gov/gmd/ccgg/trends/, 2021. Accessed: 2020-11-18.
- Friedlingstein et al. [2019] P. Friedlingstein, M. W. Jones, M. O’Sullivan, R. M. Andrew, J. Hauck, G. P. Peters, …., and S. Zaehle. Global carbon budget 2019. Earth System Science Data, 11(4):1783–1838, 2019. doi: 10.5194/essd-11-1783-2019.
- Ansmann et al. [2013] G. Ansmann, R. Karnatak, K. Lehnertz, and U. Feudel. Extreme events in excitable systems and mechanisms of their generation. Phys. Rev. E, 88(5):052911, 2013.
- Farazmand and Sapsis [2017a] M. Farazmand and T. P. Sapsis. A variational approach to probing extreme events in turbulent dynamical systems. Science Advances, 3(9):e1701533, 2017a. doi: 10.1126/sciadv.1701533.
- Farazmand and Sapsis [2019a] M. Farazmand and T. P. Sapsis. Extreme Events: Mechanisms and Prediction. Applied Mechanics Reviews, 71(5), 2019a. doi: 10.1115/1.4042065.
- E and Vanden-Eijnden [2006] W. E and E. Vanden-Eijnden. Towards a theory of transition paths. Journal of statistical physics, 123(3):503, 2006.
- E and Vanden-Eijnden [2010] W. E and E. Vanden-Eijnden. Transition-path theory and path-finding algorithms for the study of rare events. Annual review of physical chemistry, 61, 2010.
- Dematteis et al. [2018] G. Dematteis, T. Grafke, and E. Vanden-Eijnden. Rogue waves and large deviations in deep sea. Proc. Natl. Acad. Sci., 115(5):855–860, 2018. doi: 10.1073/pnas.1710670115.
- Mohamad et al. [2016] M. A. Mohamad, W. Cousins, and T. P. Sapsis. A probabilistic decomposition-synthesis method for the quantification of rare events due to internal instabilities. J. Comput. Phys., 322:288 – 308, 2016. doi: 10.1016/j.jcp.2016.06.047.
- Bialonski et al. [2015] S. Bialonski, G. Ansmann, and H. Kantz. Data-driven prediction and prevention of extreme events in a spatially extended excitable system. Physical Review E, 92(4):042910, 2015.
- Farazmand and Sapsis [2017b] M. Farazmand and T. P. Sapsis. Reduced-order prediction of rogue waves in two-dimensional deep-water waves. J. Comput. Phys., 340:418 – 434, 2017b.
- Scheffer [2010] M. Scheffer. Complex systems: foreseeing tipping points. Nature, 467(7314):411, 2010. doi: 10.1038/467411a.
- Farazmand and Sapsis [2019b] M. Farazmand and T. P. Sapsis. Closed-loop adaptive control of extreme events in a turbulent flow. Phys. Rev. E, 100:033110, 2019b. doi: 10.1103/PhysRevE.100.033110.
- Cao et al. [2015] W. Cao, Z. Zhang, and G. Karniadakis. Numerical methods for stochastic delay differential equations via the Wong–Zakai approximation. SIAM Journal on Scientific Computing, 37(1):A295–A318, 2015. doi: 10.1137/130942024.
- Malhi et al. [2004] Y. Malhi, O. L. Phillips, W. Cramer, A. Bondeau, S. Schaphoff, W. Lucht, B. Smith, and S. Sitch. Tropical forests and the global carbon cycle: impacts of atmospheric carbon dioxide, climate change and rate of deforestation. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences, 359(1443):331–343, 2004. doi: 10.1098/rstb.2003.1428.
- Baccini et al. [2012] A. Baccini, S. J. Goetz, W. S. Walker, N. T. Laporte, M. Sun, D. Sulla-Menashe, J. Hackler, P. S. A. Beck, R. Dubayah, M. A. Friedl, et al. Estimated carbon dioxide emissions from tropical deforestation improved by carbon-density maps. Nature climate change, 2(3):182–185, 2012. doi: 10.1038/nclimate1354.
- Brunton et al. [2016] S. L. Brunton, J. L. Proctor, and J. N. Kutz. Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proceedings of the National Academy of Sciences, 113(15):3932–3937, 2016. doi: 10.1073/pnas.1517384113.
- Horváth [2016] L. Horváth. Sharp gronwall–bellman type integral inequalities with delay. Electronic Journal of Qualitative Theory of Differential Equations, 2016:1–25, 11 2016.
- Harr and Suomi [1971] T. H. Vonder Harr and V. Suomi. Measurements of the earth’s radiation budget from satellites during a 5 year period. i. extended time and space means. Journal of Atmospheric Science, (28):305 – 314, 1971.
- Stephens et al. [2012] G. Stephens, J. Li, M. Wild, C. A. Clayson, N. Loeb, S Kato, T. L’Ecuyer, P. W. Stackhouse Jr, Matthew Lebsock, and T. Andrews. An update on earth’s energy balance in light of the latest global observations. Nature Geoscience, 5:691 – 696, 2012.
- Goode et al. [2001] P. R. Goode, J. Qiu, V. Yurchyshyn, J. Hickey, M-C. Chu, E. Kolbe, C. T. Brown, and S. E. Koonin. Earthshine observations of the earth’s reflectance. Geophys. Res. Lett., 28(9):1671 – 1674, 2001.