Introduction
The last about 1000 years constitute the best opportunity previous to the instrumental period to study the transient interaction of external forcing and internal variability in climate, atmospheric CO, and the carbon cycle on interannual to multi-decadal timescales. In fact, the instrumental record is often too short to draw strong conclusions on multi-decadal variability. The relatively stable climate together with the abundance of high-resolution climate proxy and ice core data makes the last millennium an interesting target and test bed for modeling studies. However, the large and sometimes controversial body of literature on the magnitude and impact of solar and volcanic forcing on interannual to multi-decadal climate variability illustrates the challenges inherent in establishing a robust understanding of a period that is characterized by a small signal-to-noise ratio in many quantities and for which uncertainties in the external forcing remain
Compared to glacial–interglacial climate change, the last millennium experienced little climate variability, yet there is evidence for distinct climate states during that period
Further, the last millennium offers the possibility to study the natural variability in the carbon cycle and its response to external forcing. Models with a carbon cycle module are extensively tested against present-day observations and widely used for emission-driven future projections
As for physical climate quantities, explosive volcanic eruptions constitute an important forcing for the carbon cycle. The sensitivity of the carbon cycle to such eruptions has been investigated by , , , and using different Earth system models. For this short-lived forcing, the land response appears to be the driver of most post-eruption carbon cycle changes, with a range of magnitudes and time horizons associated with the different models. Further, pointed out that the magnitude of the carbon cycle response to volcanoes depends critically on the climate state during the eruption.
The third Paleoclimate Modelling Intercomparison Project
This paper is structured as follows: a description of the model and experimental setup is presented in Sect. 2. In Sects. 3 and 4 we address the general simulated climate and carbon cycle evolution and investigate forced and unforced variability in the simulated climate by comparing models to reconstructions and models to models. Section 5 focuses on the response of the climate and carbon cycle to volcanic forcing. Section 6 deals with estimating the climate–carbon-cycle sensitivity in the Community Earth System Model (CESM). A discussion and conclusions follow in Sect. 7.
Data and methods
Model description
The Community Earth System Model
The atmospheric component of CESM 1.0.1 is the Community Atmosphere Model version 4
The ocean component is the Parallel Ocean Program version 2
List of simulations conducted for this study. See text for details regarding the forcing. TSI: total solar irradiance; GHGs: greenhouse gases; : anthropogenic CO emissions from fossil fuel burning and cement production; LULUC; land use and land use change.
Control simulation (CTRL) | Transient simulation (CESM) | ||
---|---|---|---|
Forcing | 850 CE (500 years) | 850–2099 CE | |
TSI | 1360.228 W m | adjusted and | |
Volcanic | none | ||
GHGs | CO (279.3 ppm) | ||
CH (674.5 ppb) | |||
NO (266.9 ppb) | |||
none | and | ||
Aerosol | 1850f CE from | ||
Orbital | 1990 CE after | 1990 CE after | |
LULUC | 850 CE from | and |
Experimental setup
Table provides an overview of the simulations conducted for this study. First, a 500-year control simulation with perpetual 850 CE forcing (hereafter CTRL) was branched off from an 1850 CE control simulation with CCSM4 . However, restart files for the land component were taken from an 850 CE control simulation, kindly provided by the NCAR, in which the land use maps by were applied. This procedure has the advantage that the slowly reacting soil and ecosystem carbon stocks are closer to 850 CE conditions than in the 1850 CE control simulation. A transient simulation covering the period 850–2099 CE was then branched off from year 258 of CTRL. Despite the shortness of CTRL leading up to the start of the transient simulation, most quantities of the surface climate, such as air temperature, sea ice, or upper-ocean temperature, can be considered reasonably equilibrated at the start of the transient simulation, as the forcing levels due to total solar irradiance (TSI) and most greenhouse gases are similar between 1850 and 850 CE . However, weak trends in CTRL are still detectable in slowly reacting quantities, such as deep-ocean temperature (below 2000 m; 0.04 C 100 yr ), Atlantic meridional overturning circulation (0.22 Sv 100 yr ), Antarctic Circumpolar Current ( 0.70 Sv 100 yr ), dissolved inorganic carbon in the ocean ( 0.01 % 100 yr ), or soil carbon storage ( 0.2 % C 100 yr ). The Antarctic Bottom Water formation rate shows no drift.
Forcings used in the last millennium simulation with CESM. (a) TSI in comparison with the different TSI reconstructions proposed by PMIP3. (b) Volcanic forcing as total volcanic aerosol mass. (c) Radiative forcing
[Figure omitted. See PDF]
The applied transient forcing largely follows the PMIP3 protocols and the Coupled Model Intercomparison Project 5
The volcanic forcing follows from 850 to 2001 CE. It provides estimates of the stratospheric sulfate aerosol loadings from volcanic eruptions as a function of latitude, altitude, and month and is implemented in CESM as a fixed single-size distribution in the three layers in the lower stratosphere . Post-2001 CE volcanic forcing remains zero.
Land use and land use changes (LULUC) are based on from 850 to 1500 CE, when this data set is splined into , a synthesis data set that extends into the future. The two data sets do not join smoothly but exhibit a small stepwise change in the distribution of crop land and pasture at the year 1500 CE. Up until about 1850 CE global anthropogenic LULUC are small; however, they can be significant regionally . Approaching the industrial era, LULUC accelerate, dominated by the expansion of crop land and pasture. Here, only net changes in land use area are considered. The impact of shifting cultivation and wood harvest on carbon emissions from land use is neglected; these processes are estimated to have contributed around 30 % to the total carbon emissions from land use .
The temporal evolution of long-lived greenhouse gases (GHGs; CO, CH, and NO) is prescribed based on estimates from high-resolution Antarctic ice cores that are joined with measurements in the mid-twentieth century
Aerosols such as sulfate, black and organic carbon, dust, and sea salt are implemented as non-time-varying up to 1850 CE, perpetually inducing the spatial distributions of the 1850 CE control simulation during this time . Post-1850 CE, the time-varying aerosol data sets provided by are used, although CESM only includes a representation of direct aerosol effects. Similarly, nitrogen (NH and NO) input is held constant until it starts to be time-varying from 1850 CE onwards, also following . Iron fluxes from sediments are held fixed .
Other model simulations
In addition to comparing CESM results to output from current Model Intercomparison Projects, we also compare them to results from a similar simulation with CCSM4 and IPSL-CM5A-LR (Institut Pierre Simon Laplace–climate model 5A–low resolution) , two simulations without interactive carbon cycle. The solar and volcanic forcing reconstructions applied to CCSM4 and IPSL-CM5A-LR are identical to ours with the exception of the scaling of TSI that we applied to CESM. The goal here is to investigate the question of whether different solar forcing amplitudes applied to the same physical model (CESM vs. CCSM4) have a larger effect than applying the same solar forcing to two different physical models (IPSL-CM5A-LR vs. CCSM4).
Selected forcing details of simulations used in comparisons with CESM. TSI: total solar irradiance; LULUC: land use and land use change.
CESM | CCSM4 | IPSL-CM5A-LR | MPI-ESM | |
---|---|---|---|---|
Forcing | ||||
TSI | adjusted | |||
and | and | and | ||
Volcanic | ||||
Orbital | 1990 CE, | transient; | transient; | transient; |
LULUC | non-transient | |||
and | and |
Further, we compare CESM to MPI-ESM (Max-Planck-Institut Earth System Model)
(a) Northern Hemisphere and (b) Southern Hemisphere temperature anomalies in model simulations and reconstructions. The anomalies are with reference to 1500–1899 CE (left panels) and 1850–1899 CE (right panels). Gray shading in (a) indicates the reconstruction overlap ; in (b), it indicates the reconstruction by . The 5–95 % range of the simulations from the third Paleoclimate Modelling Intercomparison Project (PMIP3) and the fifth Coupled Model Intercomparison Project (CMIP5; applying the RCP8.5) are given in green and red shading, respectively. Note that MPI-ESM applies the A1B scenario , which has a weaker forcing than RCP8.5. Hemispheric means from observations are shown as thick black line . All time series have been smoothed by a local regression filter which suppresses variability higher than 30 years. The Medieval Climate Anomaly (MCA) and the Little Ice Age (LIA) are indicated as defined in . (c) Evolution of atmospheric CO in CESM (black), MPI-ESM (grey; ensemble range), from ice cores (red), from measurements (orange), and from RCP8.5 used to force the radiative code in CESM (magenta). The small inset in the middle panel shows the observed annual cycle at Mauna Loa, Hawaii, and a 2 2 average over Hawaii from CESM, both derived from the period 1958–2012.
[Figure omitted. See PDF]
(a) Mean June–August (JJA) Arctic ( 60 N land) solar insolation in CCSM4 with time-varying orbital parameters and CESM with fixed orbital parameters. (b) Arctic JJA temperature difference between CCSM4 and CESM. The least-squares linear trend of this temperature difference is given in red. (c) Arctic JJA temperature anomalies (from their 850 to 1850 AD mean) versus solar insolation as 100- and 200-year averages (10 and 5 circles, respectively) from CCSM4 and CESM (red and blue, respectively). The least-squares linear trend for each cloud of 100- and 200-year averages is given in the respective color. The shading envelops the range of temperature versus solar insolation for each cloud of means.
[Figure omitted. See PDF]
General climate and carbon cycle evolution
Temperature
The simulated annual mean Northern Hemisphere (NH) surface air temperature (SAT) follows the general evolution of proxy reconstructions: a warm Medieval Climate Anomaly (MCA, 950–1250 CE), a transition into the colder Little Ice Age (LIA, 1400–1700 CE), followed by the anthropogenically driven warming of the nineteenth and twentieth centuries (Fig. ). The NH MCA-to-LIA cooling amounts to 0.26 0.18 C
In CESM, the inception of the NH LIA occurs in concert with decreasing TSI and a sequence of strong volcanic eruptions during the thirteenth century. Reconstructions differ substantially in this matter and start to cool as early as 1100 CE or as late as 1400 CE. Further, new regional multi-proxy reconstructions of temperature provide no support for a hemispherical or globally synchronous MCA or LIA but show a clear tendency towards colder temperatures and exceptionally cold decades over most continents in the second half of the millennium .
The last millennium simulation with CCSM4 shows a largely coherent behavior with CESM in terms of amplitude and decadal variability in NH SAT (850–1850 CE correlation of 5-year filtered annual means , ). The difference in NH SAT due to the different TSI amplitudes in CESM and CCSM4 scales roughly with the regression slope of NH SAT vs. TSI of both CESM and CCSM4 (0.13 C per W m), although internal variability can easily mask this effect at times. For example, the Maunder Minimum (1675–1704 CE), the 30-year period with the lowest TSI values and – when using TSI scaling – with the largest difference between CESM and CCSM4 (1.5 W m), is only 0.14 C cooler than in CCSM4 and not 0.20 C as expected from the regression.
The NH temperature evolution of additional PMIP3 and CMIP5 simulations shows that the multi-model range is within the range of the reconstructions and encompasses the instrument-based observations (Fig. ). Disagreement between models and reconstructions exists in particular on the magnitude of response to the eruptions at 1258 CE and around 1350 CE. The 1258 CE eruption is the largest volcanic event recorded for the last millennium, and its climatic impact was likely enhanced through the cumulative effect of three smaller eruptions following shortly after . However, the pronounced cooling that is simulated by the models for this cluster of eruptions is largely absent in temperature reconstructions. Conversely, around 1350 CE temperature reconstructions show a decadal-scale cooling presumably due to volcanoes that is absent in the models, as the reconstructed volcanic forcing shows only two relatively small eruptions around that time. Part of this incoherent picture may arise from the unknown aerosol size distribution , the geographic location of past volcanic eruptions , and differences in reconstruction methods. As many proxy reconstructions of temperature rely heavily on tree ring data, it is worth noting that the dendrochronology community is currently debating whether the trees' response to volcanic eruptions resembles the true magnitude of the eruption .
Disagreement among the models exists on the relative amplitude of the MCA, where most models show colder conditions than CESM and CCSM4. Remarkably, the simulation by IPSL-CM5A-LR applied the same TSI and volcanic forcing as CCSM4, yet it comes to lie at the lower end of the PMIP3 model range during the MCA. In other words, the way in which models respond to variations in TSI and other forcings can still make a larger difference in the simulated amplitude than the scaling of TSI by a factor of 2, which in turn complicates a proper detection and attribution of solar forcing during the last millennium . Further disagreement among the models exists on the response to volcanic eruptions, where CESM and CCSM4 are among the more sensitive models
The simulated mean SAT of the Southern Hemisphere (SH) generally shows a similar evolution as in the NH, with the signature of the MCA and LIA superimposed on a weak millennial cooling trend. Models and reconstructions disagree to a larger extent in the SH than in the NH, in particular regarding cold excursions due to large volcanic eruptions, which are largely absent in the reconstructions. Similar results have been reported in a recent study on interhemispheric temperature variations that finds much less phasing of the two hemispheres in reconstructions than in models, potentially related to underestimated internal variability on the SH in models . A lingering question of climate modeling in general is whether models are too global in their response to external forcing. That is, they might show too little regional variability that is independent of the global mean response during a forced period. However, the uncertainties in the early period of the reconstructions make it impossible to robustly answer this question. Similar to the NH, the industrial warming in the SH from 1851–1880 to 1981–2010 CE (0.53 0.07 C) is overestimated by CESM (0.71 0.13 C).
The differential warming between the hemispheres in CESM is among the smallest among CMIP5 models (not shown). This is mainly due to the underestimated deep-water formation in the Southern Ocean, leading to a comparably strong warming of the SH and likely an underestimation of the oceanic uptake of anthropogenic carbon . With a transient climate response of 1.73 C and an equilibrium climate sensitivity of 3.20 C , CESM lies in the middle of recent estimates of 1.0 to 2.5 C and 1.5 to 4.5 C .
Orbital forcing
To detect and attribute the influence of orbital forcing on SAT trends during the last millennium, we compare our simulation with fixed orbital parameters to the CCSM4 simulation with time-varying orbital parameters (Fig. ). While both models experience a negative long-term trend in global TSI until about 1850 CE (Fig. ), the difference arising from the different orbital setup can be best seen in Arctic summer land insolation (Fig. ). Hence, Arctic summer land SAT has been proposed as a quantity that, on timescales of centuries to millennia, may be affected by orbital forcing . However, we find no detectable difference between the two simulations in the trend of Arctic summer land SAT (Fig. b). In fact, the Arctic multi-decadal to centennial summer land SAT anomalies in CESM span a very similar range as in CCSM4, despite CESM not accounting for time-varying orbital parameters: Fig. c shows non-overlapping 100- and 200-year mean SAT anomalies plotted against the corresponding mean solar insolation. The results from CCSM4 suggest a clear relationship between of the two quantities; however, the results of CESM show that nearly identical SAT anomalies are possible without orbital forcing. In other words, while we detect a long-term cooling trend in Arctic summer SAT in both CESM and CCSM4, we fail to attribute this trend to orbital forcing alone, as suggested by . This is confirmed in new simulations with decomposed forcing, again comparing simulations with fixed and time-varying orbital parameters (B. Otto-Bliesner, personal communication, 2014).
Cumulative carbon emissions in Pg C by different components over different time periods in CESM. Positive (negative) values indicate emission to (uptake from) the atmosphere.
850–1500 CE | 1501–1750 CE | 1751–2011 CE | 2012–2100 CE | |
---|---|---|---|---|
Ocean | 26.0 | 4.0 | 151.3 | 413.0 |
Land | 15.0 | 10.3 | 82.5 | 139.4 |
Land (without LULUC) | 24.4 | 9.3 | 94.7 | 436.3 |
Fossil Fuels | 0.0 | 0.0 | 358.0 | 1851.5 |
Carbon cycle
The prognostic carbon cycle module in CESM allows us to study the response of the carbon cycle to transient external forcing. The land biosphere is a carbon sink during most of the first half of the last millennium, but becomes a source as anthropogenic land cover changes start to have a large-scale impact on the carbon cycle (Table ). The ocean is a carbon source at the beginning and becomes a sink in the second half of the last millennium. The residual of these fluxes represents changes in the atmospheric reservoir of carbon, illustrated in Fig. c by the prognostic CO concentration. The amplitude of the simulated concentration does not resemble the one reconstructed from ice cores (i.e., imposed on the radiative code of CESM); in particular, the prominent CO drop in the seventeenth century is not captured by CESM. This raises the question of whether the sensitivity of the carbon cycle to external forcing is too weak in CESM, whether the imposed land use changes are too modest , whether major changes in ocean circulation are not captured by models , and whether the ice core records are affected by uncertainties due to in situ production of CO . Ensemble simulations with MPI-ESM also do not reproduce the reconstructed amplitudes or the drop . Further, Earth system models of intermediate complexity or vegetation models driven by GCM (general circulation model) output do not reproduce the uptake of carbon by either ocean or land needed to explain the reconstructed amplitudes .
Annual mean net carbon flux from the atmosphere to (a) land and (b) ocean. Green bars given the full and 10–90 % range from the preindustrial part of the simulation. Observational estimates are from .
[Figure omitted. See PDF]
The rise in atmospheric CO due to fossil-fuel combustion is in good agreement with ice cores until about the 1940s. After that, a growing offset exists, leading to an overestimation of about 20 ppm by 2005 in CESM, qualitatively similar to the CMIP5 multi-model mean . From the observational estimates one can diagnose that the discrepancy arises primarily from overestimated carbon release from land
The twenty-first century sees substantial emissions from fossil-fuel burning under RCP8.5 (Fig. c). In addition, LULUC is associated with a positive flux into the atmosphere, particularly until around 2050 CE (Table ). After accounting for LULUC (which constitutes a carbon loss for land) the net land sink increases to about 7 Pg C yr at the end of the twenty-first century (Fig. a). The rate of ocean uptake, on the other hand, peaks around 2070 at about 5 Pg C yr, despite the fact that atmospheric CO continues to rise (Fig. c). This decoupling of the trends in atmospheric CO growth and ocean uptake flux is linked to nonlinearities in the carbon chemistry . The change in dissolved inorganic carbon per unit change in the partial pressure of CO decreases with increasing CO, and thus so does the uptake capacity of the ocean. Additionally, differences in the ventilation timescales of the upper and the deep ocean likely play a role. While the surface ocean and the thermocline exchanges carbon on annual-to-multi-decadal timescales with the atmosphere, it takes a century to ventilate the deep ocean, as evidenced by chlorofluorocarbon and radiocarbon data . CESM has a documented low bias in Southern Ocean ventilation due to too shallow mixed layer depths, contributing to the underestimated carbon uptake of the ocean .
The prognostic atmospheric CO increases to 1156 ppm by 2100 CE. This would imply a forcing of 7.6 W m from CO relative to 850 CE, significantly more than the approximately 6.5 W m that are imposed by the radiative code
Figure places the current and projected changes within the context of preindustrial variability. Estimated interannual variability prior to 1750 CE is 0.94 Pg C yr (1 standard deviation) for the net atmosphere–land and 0.42 Pg C yr for the net atmosphere–ocean flux. The much larger interannual variability in land than ocean flux is consistent with independent estimates and results from other models
Five-year filtered zonal mean anomalies of surface air temperature (SAT), relative to the 850–1849 CE mean from (a) CESM and (b) MPI-ESM. (c) 100-year running-window correlation of zonal mean SAT from CESM and MPI-ESM. A 0.75 Tukey window has been applied to the data before correlation to weaken sharp transitions. Stippling indicates significance (5 % level), taking into account autocorrelation estimated from the entire time period. Panel (d): as (c) but for the correlation of CESM with CCSM4. Panel (e): as (d) but for global mean SAT. Small inset on top shows volcanic and solar forcing of CESM and MPI-ESM. Volcanic forcing of CESM scaled to have the same radiative forcing as MPI-ESM for Pinatubo in 1991 CE. Solar forcing relative to 1850 CE.
[Figure omitted. See PDF]
Model–model coherence
A classical approach to assess the robustness of model results is to rely on the multi-model mean response to a given forcing . However, as there are only very few last millennium simulations with comprehensive Earth system models to date, this approach is not feasible for investigating the decadal-scale climate–carbon-cycle responses to external forcing in the period before 1850 CE. Instead, we estimate periods of forced variability with a 100-year running-window correlation of CESM and MPI-ESM, indicating phasing of the two models. The time series are anomalies from their 850–1849 mean and are smoothed with a 5-year local regression filter before calculating the correlation. Thereby, we focus on the preindustrial period, as the twentieth and twenty-first centuries are dominated by anthropogenic trends, which are nontrivial to remove for a proper correlation analysis. In addition, regression analysis is used.
Temperature
Figure a and b show anomalies of zonal mean annual SAT from CESM and MPI-ESM. In both models the northern high latitudes show the strongest trend, from positive anomalies during the MCA to negative anomalies during the LIA. This is consistent with the current understanding of polar amplification during either warm or cold phases . The twentieth and twenty-first centuries then see the strong anthropogenic warming, although this occurs earlier in CESM due to missing negative forcings from indirect aerosol effects (Sect. ). Superimposed on the preindustrial long-term negative trend are volcanic cooling events. In CESM many of these are global and are able to considerably cool the SH extra-tropics around 60 S, while in MPI-ESM the SH extra-tropics are only weakly affected. These differences are likely related to the Southern Ocean heat uptake rates in the two models
Regression of total solar irradiance (TSI) on surface air temperature (SAT) for the period 850–1850 CE in (a) CESM and (b) MPI-ESM. Time series at each grid point have been 5-year-filtered. Only significant regression coefficients at the 5 % level are shown. The small panel shows zonal means.
[Figure omitted. See PDF]
The phasing on interannual to decadal scales between the two models is largely restricted to periods of volcanic activity and, within those, mainly to land-dominated latitudes (except Antarctica, which shows no forced variability on these timescales; Fig. c). Despite the largest absolute temperature anomalies occurring in the Arctic, the correlations are highest in the subtropics, due to the smaller interannual variability there. Periods of centennial trends, such as the MCA or the Arctic cooling during the Maunder Minimum around 1700 CE, do not show up in the correlation analysis that focuses on 100-year windows, suggesting that multi-decadal low-frequency forcing, such as centennial TSI trends, or internal feedback mechanisms are responsible for the missing correlation. A regression analysis between the 5-year filtered annual TSI and SAT at each grid point (different filter lengths of up to 50 years have been tested as well without the results changing) reveals a clear link between the two quantities at high latitudes. In CESM this link seems to be driven primarily by a displacement of the sea ice edge (Arctic) and Southern Ocean heat uptake (Fig. a). As the sea ice response has not been detected in an earlier model version
Five-year-filtered zonal mean anomalies of horizontally averaged ocean temperature, relative to 850–1849 CE from (a) CESM and (b) MPI-ESM. Panel (c): 100-year running-window correlation of zonal mean SAT from CESM and MPI-ESM. A 0.75 Tukey window has been applied to the data before correlation to weaken sharp transitions. Stippling indicates significance at the 5 % level, taking into account autocorrelation estimated from the entire time period. Panel (d): 100-year running-window correlation of the Atlantic meridional overturning circulation (AMOC) in CESM and MPI-ESM.
[Figure omitted. See PDF]
In addition to the comparison with MPI-ESM, Fig. d shows results from the correlation analysis between CESM and CCSM4, two simulations that in terms of physics differ only in their applied TSI amplitude and orbital parameters. Not unexpectedly, there are generally more robust signals of forced variability as compared to CESM vs. MPI-ESM (Fig. c), very likely due to the identical physical model components in CESM and CCSM4. Similarly, global mean SAT shows generally stronger phasing between CESM and CCSM4 (Fig. e). However, the latitudinal and temporal pattern of the CESM vs. CCSM4 analysis agrees well with the one arising from CESM vs. MPI-ESM (Fig. c; with exception of the much stronger phasing in CESM and CCSM4 during the volcanic eruptions in the 1450s) and suggests that the physical mechanism behind periods of phasing is robust across the two models.
Five-year filtered zonal mean anomalies of horizontally integrated dissolved inorganic carbon (DIC), relative to 850–1849 CE, from (a) CESM and (b) MPI-ESM. Panel (c): 100-year running-window correlation of zonal mean SAT from CESM and MPI-ESM. A 0.75 Tukey window has been applied to the data before correlation to weaken sharp transitions. Stippling indicates significance at the 5 % level, taking into account autocorrelation estimated from the entire time period.
[Figure omitted. See PDF]
Applied to ocean temperature, the above approach enables us to investigate the penetration depth of a forced signal seen at the surface (Fig. ). Indeed, most of the surface signals also show up as significant correlations down to depths of about 150–200 m, and their timing again suggests volcanic forcing as the origin. Reduced heat loss from the tropical equatorial Pacific together with reduced heat uptake at high latitudes are responsible for ocean cooling after volcanoes (not shown). The Atlantic Meridional Overturning Circulations (AMOC) in the CESM and MPI-ESM shows no significant correlation; however, the highest correlation occurs during the thirteenth century and coincides with a phasing of the upper-ocean temperatures due to strong volcanic forcing (Fig. d). The correlation between CESM and CCSM4 at that time is even higher and points to a significant imprint of the volcanic forcing on ocean circulation . However, during the remainder of the millennium, no phasing of the AMOC is found.
Carbon cycle
We apply the same correlation analysis to zonally integrated land and ocean carbon fluxes from the two models to detect forced variability in the carbon cycle. Compared to SAT hardly any phasing can be found between the models in atmosphere-to-land carbon fluxes (not shown), which is due to its large interannual variability and to distinctly different responses to external forcing in the two models, as will be illustrated in Sect. . Similarly, there is little model phasing in net atmosphere-to-ocean carbon fluxes (not shown). Results become somewhat clearer when considering globally integrated upper-ocean dissolved inorganic carbon (DIC; Fig. ). There appear to exist spurious trends in CESM, likely related to model drift. We repeated the analysis, but with the CESM output detrended in each grid cell by subtracting the CTRL over the corresponding period 850–1372 CE. Due to the shortness of CTRL, we cannot apply this method to the whole simulation. However, these tests showed that the correlation between the two simulations is largely insensitive to the drift in CESM. In Fig. c there are periods of coherent carbon drawdown coinciding with volcanic eruptions around 1450 and 1815 CE in response to temperature-driven solubility changes. Interestingly, MPI-ESM shows a distinct behavior for the strong eruption of 1258 CE, with a prolonged ocean carbon loss after a weak initial uptake. CESM shows a stronger and more sustained carbon uptake, leading to no correlation between the two models for this eruption. The reasons for this discrepancy are discussed in Sect. .
Superposed epoch analysis of the strongest three (top3) and subsequent strongest seven eruptions (top10) of the period 850–1850 CE in (a–e) CESM and (f–j) MPI-ESM for (a, f) global mean surface air temperature, (b, g) global mean precipitation, (c, h) atmospheric carbon given in Pg C on the left axis and in ppm CO on the right axis, (d, i) ocean carbon, and (e, j) land carbon. Time series are deseasonalized and calculated as anomalies to the mean of the preceding 5 years. The shading shows the 10–90 % range.
[Figure omitted. See PDF]
Generally, the largest changes in upper-ocean carbon storage occur in response to volcanoes and take place in the tropical Pacific , with other significant changes occurring in the North and South Pacific, the subtropical Atlantic, and the Arctic (Sect. ). Within the tropical oceans, the models show different characteristics: CESM shows a larger variability in DIC than MPI-ESM and, when influenced by anthropogenic emissions in the twentieth and twenty-first centuries, takes up a larger portion of the total ocean carbon uptake than in MPI-ESM (not shown). In MPI-ESM, the Southern Ocean shows stronger variability and larger carbon uptake in the twenty-first century, illustrating the different behavior of the two models in terms of ocean carbon cycle variability and trend magnitude, closely related to the different mixed layer depths in the Southern Ocean region.
Volcanic forcing
To further isolate the response of the climate system and carbon cycle to volcanic eruptions, a superposed epoch analysis is applied to both simulations. Thereby, composite time series for the strongest three (top3) and subsequent strongest seven eruptions (top10), according to optical depth anomaly, over the period 850–1850 CE are calculated for the CESM and MPI-ESM (Fig. ). The time series are calculated as deseasonalized monthly anomalies from the 5 years preceding an eruption.
The physical parameters' global mean surface air temperature and global mean precipitation decrease in both models after volcanic eruptions, although the response of CESM is stronger by roughly a factor of 2–2.5 (Fig. a, b, f, g). Consequently, CESM temperature and precipitation take longer ( 15 years) to relax back to pre-eruption values than MPI-ESM ( 9 years).
The atmospheric carbon inventory, on the other hand, shows a remarkably different response in the two models. In CESM the atmosphere initially looses about 2–3 Pg C, irrespectively of the eruption strength, with the minimum occurring after about 1–2 years. In the top10 case, values return to normal after about 16 years, while in the top3 case, they tend to return already after about 6 years and overshoot. This overshoot is not straightforward to understand and did not seem to occur in earlier versions of the model . In MPI-ESM the response is a priori more straightforward and slower: in the top10 case the atmosphere looses about 2.5 Pg C, reaches a minimum after 2–4 years, and returns to pre-eruption values after 10–16 years. The top3 case reaches its minimum (6 Pg C) a bit faster, but then takes about 20 years to return to pre-eruption values .
Partitioning these atmospheric carbon changes into land and ocean changes indicates that land is primarily responsible for the differing response behavior of the two models, confirming the findings in the previous section. While in both models, land drives the atmospheric change by taking up carbon initially, it is released back to the atmosphere within about 3 years in CESM but retained in land areas for at least 15 years in MPI-ESM
Superposed epoch analysis of the strongest three (top3) and subsequent strongest seven eruptions (top10) for tropical land (25 S to 25 N) in CESM during the period 850–1850 CE. Land carbon inventory changes split up in (a) vegetation, (b) dead biomass (litter and wooden debris), and (c) soil. Furthermore, changes in (d) solar radiation, (e) net primary production (NPP), and (f) loss of carbon through fire. Time series are deseasonalized and calculated as anomalies from the mean of the preceding 5 years. The shading shows the 10–90 % range.
[Figure omitted. See PDF]
A closer look at CESM reveals a distinct response to the top3 and the top10 volcanoes. The response to top3 must be understood as the interaction of a number of processes: the initial global cooling triggers a La Niña-like response and a corresponding cloud and precipitation reduction that is particularly pronounced over tropical land, where large changes in carbon storage also occur (see Fig. a–c for the spatial pattern). Figure and the following analysis therefore focuses on tropical land. Direct solar radiation decreases, and indirect radiation increases, with a net decrease (Fig. d). These unfavorable conditions cause a reduction in net primary productivity and a strong decrease in vegetation (8 Pg C; Fig. a and e). At the same time, decomposition of dead biomass becomes less efficient due to reduced temperature
Composites of top10 post-volcanic eruption years as anomalies from the preceding 5 years, averaged over (left) the first 2 years starting with the year of the eruption and (right) the following 3 years. (a) Surface air temperature, (b) precipitation, (c) total land carbon, (d) dissolved inorganic carbon (DIC) integrated over the top 200 m. Shading or stippling indicates significance at the 5 % level. Note that for land carbon in an individual grid cell hardly any significant changes are detected due to the large interannual variability.
[Figure omitted. See PDF]
The ocean, on the other hand, shows a qualitatively similar response in CESM and MPI-ESM with an uptake of carbon and a gradual relaxation back to pre-eruption values over 20 or more years. In CESM the radiative cooling leads to increased uptake in the Western Pacific, while in the Eastern Pacific, cooling is less as this region is more controlled by upwelling rather than direct radiative forcing, as suggested by (Fig. d). Two or more years after the volcano, a La Niña-like pattern settles in both surface temperature as well as carbon uptake. Some model differences exist; e.g., in the top3 case of MPI-ESM, the ocean starts to release carbon, compensating for the persistent positive anomaly in the land inventory
Global mean changes in response to Pinatubo. (a) Global mean surface air temperature and (b) atmospheric carbon, as (left y axis) Pg C and (right y axis) ppm CO equivalent, both deseasonalized and linearly detrended over 30 years centered on June 1991; temperature observations were corrected for El Niño–Southern Oscillation and other dynamical components ; CO observations were corrected for El Niño–Southern Oscillation and anthropogenic emissions .
[Figure omitted. See PDF]
In an attempt to validate the two models, one is restrained to the well-observed eruption of Pinatubo in 1991 CE, as the CO records from ice cores do not adequately resolve short-term variations induced by volcanoes over the last millennium. Figure shows the global temperature and atmospheric carbon response to Pinatubo as extracted from observations, CESM, and the three-member ensemble of MPI-ESM. Note that the effects of El Niño–Southern Oscillation and anthropogenic emissions have been removed from the CO observations to obtain a tentative estimate of the actual CO response to the Pinatubo eruption . The initial cooling of about 0.5 C and the relaxation back to initial temperatures around 1998 CE is captured well by both models. The MPI-ESM ensemble, however, shows a large and consistent variation around 1995 CE, seemingly related to a phasing of El Niño–Southern Oscillation (ENSO) variability in response to the eruption
Climate–carbon-cycle sensitivity
Due to the absence of large anthropogenic disturbances of the carbon cycle, the last millennium represents a test bed to estimate the climate–carbon-cycle sensitivity, expressed as ppm C, and can thus potentially help to constrain this quantity
Here, we estimate the climate–carbon-cycle sensitivity for CESM as follows. We focus on the period before significant LULUC (850–1500 CE) and apply different low-pass filters of 20 to 120 years, taking 5-year increments, to the time series of NH SAT and global CO. The filtering aims at minimizing the influence of short-lived forcings such as volcanic eruptions that have a relatively direct impact on temperature and CO (as seen above) and thus may hinder the detection of a low-frequency influence of temperature on CO. For each filter length, we determine the highest lag correlation of the two time series, considering lags of up to 100 years. Due to the design of our simulation, we expect NH SAT to lead CO, which is confirmed by all lag correlations indicating positive lags for NH SAT (peak of lag correlation at 80.5 3.4 years). We regress the lagged time series and find a median estimate of 1.3 ppm C with a range from 1.0 to 1.8 ppm C, depending on the filter length. About 1 ppm C is explained by the land carbon cycle, while the ocean shows smaller sensitivities of about 0.4 ppm C. Note that we use NH SAT in order to be comparable with existing studies . Using global SAT instead of NH SAT can influence the sensitivity estimate, especially for the forced simulation: including the vast ocean area of the SH tends to dampen temperature variability induced by volcanoes and TSI variations. With temperature variability dampened, the sensitivity increases to 1.7 ppm C (1.4–2.1).
Temporal dependence of the climate–carbon-cycle sensitivity in CESM. Normalized probability density functions (PDFs) of climate–carbon-cycle sensitivity for 200-year windows overlapping by 50 years (color-filled) for the full period 850–1500 CE (black solid) and for the CTRL (black dashed). The spread of each PDF arises from the range of low-pass filters applied (20 to 120 years).
[Figure omitted. See PDF]
This estimate is barely within the reconstruction-constrained range of 1.7–21.4 ppm C and suggests a comparably low sensitivity of the carbon cycle in CESM. This low sensitivity is in agreement with, e.g., . Note that found different sensitivities for the early and late part of the last millennium with the mean for the period 1050–1549 CE being 4.3 ppm C. Indeed, a strong temporal dependence of the climate–carbon-cycle sensitivity is also found in CESM when looking at individual 200-year windows (Fig. ). The period 1300–1500 CE even shows negative sensitivity, which seems to be related to the different timescales with which SAT and CO relax back to the pre-eruption conditions after perturbations from large volcanic eruptions (Fig. a and c): atmospheric CO decreases after having overshot, while SAT increases after the initial cooling, leading to a negative correlation of the two quantities.
This illustrates the time-variant character of the climate–carbon-cycle sensitivity, which substantially complicates any attempt to constrain it by last millennium data and warrants caution when making inferences from past to future sensitivities. Besides, and found the sensitivity to vary greatly in a coupled model with the timescale and magnitude of volcanic forcing considered. This issue is further highlighted by the larger sensitivity derived for idealized +1 % CO year simulations with CESM (11.9 ppm C), for which a dependence on the background state, the scenario, and even the method is reported . Further, it is worth stressing that such sensitivity estimates cannot be extrapolated easily across timescales, as different processes might be at play .
Applying the identical analysis to CTRL reveals other timescales of climate–carbon-cycle feedback, suggesting maximum lags of less than 10 years and a sensitivity of 2.3 (1.4–2.9) ppm C. Using global SAT instead of NH SAT has no discernible effect (2.3 ppm C), as the CTRL does not see volcanoes or TSI variations. A later peak in the lag correlation of NH SAT and CO is found at 73.3 1.1 years in CTRL, i.e., close to where the forced simulation shows its highest lag correlation, but these lag correlations are much weaker ( 0.4 compared to 0.7 in the forced simulation). This is generally consistent with the finding by that a forced simulation exhibits increased power on lower frequencies compared to a control simulation.
Discussion and conclusions
This study presents a simulation from 850 to 2100 CE with the fully coupled CESM, including the carbon cycle, and provides an overview of the imprint of external forcing on different climate and carbon cycle diagnostics in the simulation. For comparison we draw on a number of PMIP3 simulations, particularly simulations with CCSM4 and MPI-ESM. The evolution of NH SAT during the preindustrial era in CESM is in reasonable agreement with both reconstructions and other models, albeit the uncertainties in reconstructions and forcing still being considerable. Compared to more reliable data in the twentieth century, the anthropogenic warming in CESM is overestimated due to a lack of negative forcing from indirect aerosol effects. In the SH, CESM and most other models do not capture the evolution of the mean SAT as well. The discrepancies could be explained by (i) significant model biases in SH and also interhemispheric SAT variability , (ii) spectral biases in proxies used in the reconstructions , (iii) uncertainties in the external forcing , or (iv) natural internal variability . Unfortunately, these potential explanations are neither exclusive nor independent. Arguments for model bias come from the fact that reconstructed interhemispheric SAT variability lies outside the models' range over 40 % of the time ; but these arguments are weakened by the uncertainty in external forcing. We show here that implementing the same TSI forcing in two different models results in a larger difference in simulated SAT than implementing two different TSI forcings in the same model. Hence, model structural uncertainty remains an issue in determining the role of external forcing over the last millennium.
Albeit beyond the scope of this study, detecting structural and spatial dependencies such as those illustrated here offers an opportunity to reconcile the discrepancies (e.g., regarding SH volcanic signals) between reconstructions and simulations, which might originate from sampling bias, model deficiencies, a combination of these, or the fact that reality may be the one realization that, by chance, is not encompassed by a multi-model ensemble .
Further, we compare simulations with and without orbital forcing and were not able to attribute northern high-latitude SAT trends over the last millennium to orbital forcing. This hampers, if not challenges, the validation of recent findings based on proxy archives that claim a distinct low-frequency orbital component in millennial trends . Instead, the decreasing trend in annual TSI – as opposed to seasonal and regional insolation – together with local feedbacks is able to account for a similar magnitude of trend.
When forced with emissions from LULUC, TSI variations, and volcanic eruptions over the last millennium, both CESM and MPI-ESM do not reproduce atmospheric CO variability, as suggested by ice cores. Notably, the large drop of CO in the seventeenth century is not reproduced, similar to the case in earlier studies . hypothesized that the unique, globally synchronous cooling during the LIA (which might be related to ocean dynamics) can serve as an explanation for this drop. While both CESM and MPI-ESM show a global cooling during the LIA, they develop no apparent phasing of ocean dynamics or carbon uptake and do not show any marked CO reduction around that time, leaving this issue unresolved. The strong volcanic forcing during the thirteenth century, on the other hand, is able to synchronize the AMOC on decadal scales, confirming similar results from the Bergen Climate Model and IPSL-CM5A-LR . With anthropogenic emissions, land and ocean carbon uptake rates emerge from the envelope of natural variability as simulated for the last millennium by about 1947 and 1877 CE, respectively. Atmospheric CO and global temperature emerge by 1755 and 1966 CE, suggesting that changes in carbon-cycle-related variables would be easier to detect than temperature, given sufficient observational data .
We find forced decadal-scale variability in CESM and MPI-ESM in response to major volcanic eruptions in both SAT and upper-ocean temperature, while the response in carbon cycle quantities is less coherent among models
Volcanoes trigger a coherent global response in SAT and precipitation that is qualitatively in line with earlier studies on the volcanic influence on climate and the carbon cycle
The climate–carbon-cycle sensitivity of CESM as estimated from the anthropogenically unperturbed first part of the last millennium is between 1.0 and 2.1 ppm C, depending on the filtering and the exact time period considered. Generally, the sensitivity of the carbon cycle to temperature variations in CESM is comparably small and reveals a strong component of unforced natural variability. In a transient last-millennium simulation with small temperature variations, the proper detection of a lead–lag relation between temperature and the carbon cycle is complicated by the superposition of perturbations and responses. In addition to the classic climate–carbon-cycle sensitivity experiments
Despite the challenges that paleoclimate modeling faces, a number of lessons regarding forcing and structural uncertainties can be learned from these experiments. In order to better understand the role of internal versus externally forced variability – which remains particularly critical for a period of relatively weak external forcing, such as the last millennium – larger simulation ensembles as well as ensembles with decomposed forcing should become standard in paleoclimate modeling. Since these are computationally expensive simulations, this calls for an informed discussion on the optimal usage of computing resources, to which studies like the one here can contribute valuable information. At the same time, uncertainties in forcings and reconstructions need to be further reduced to be able to better validate models in the past with the goal of constraining their future response. Key targets for such constraints are the sensitivity of temperature to solar and volcanic forcing and the climate–carbon-cycle sensitivity.
Acknowledgements
We gratefully acknowledge Axel Timmermann, Bette Otto-Bliesner, Peter Lawrence, and Rosie Fisher for valuable discussions as well as four anonymous reviewers for very helpful comments. We are grateful to the NCAR in Boulder, USA, for providing the code of the CESM, to the World Climate Research Programme's Working Group on Coupled Modelling, which is responsible for CMIP, and to the climate modeling groups for producing and making available their model output. This study is supported by the Swiss National Science Foundation (grant no. 200020 147174), and the European Commission through Seventh Framework Program (FP7) projects CARBOCHANGE (grant no. 264879) and Past4Future (grant no. 243908). J. Mignot has benefited from the support of the French Agence Nationale de la Recherche (HAMOC: ANR 13-BLAN-06-0003). The simulations for this study were performed on a CRAY XT5 and XE6 at the Swiss National Supercomputing Centre (CSCS) in Lugano. Edited by: G. Bala
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2015. This work is published under http://creativecommons.org/licenses/by/3.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Under the protocols of phase 3 of the Paleoclimate Modelling Intercomparison Project, a number of simulations were produced that provide a range of potential climate evolutions from the last millennium to the end of the current century. Here, we present the first simulation with the Community Earth System Model (CESM), which includes an interactive carbon cycle, that covers the last millennium. The simulation is continued to the end of the twenty-first century. Besides state-of-the-art forcing reconstructions, we apply a modified reconstruction of total solar irradiance to shed light on the issue of forcing uncertainty in the context of the last millennium. Nevertheless, we find that structural uncertainties between different models can still dominate over forcing uncertainty for quantities such as hemispheric temperatures or the land and ocean carbon cycle response. Compared to other model simulations, we find forced decadal-scale variability to occur mainly after volcanic eruptions, while during other periods internal variability masks potentially forced signals and calls for larger ensembles in paleoclimate modeling studies. At the same time, we were not able to attribute millennial temperature trends to orbital forcing, as has been suggested recently. The climate–carbon-cycle sensitivity in CESM during the last millennium is estimated to be between 1.0 and 2.1 ppm
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Climate and Environmental Physics, University of Bern, Bern, Switzerland; Oeschger Centre for Climate Change Research, University of Bern, Bern, Switzerland; now at: National Center for Atmospheric Research, Boulder, USA
2 Climate and Environmental Physics, University of Bern, Bern, Switzerland; Oeschger Centre for Climate Change Research, University of Bern, Bern, Switzerland
3 Climate and Environmental Physics, University of Bern, Bern, Switzerland; Oeschger Centre for Climate Change Research, University of Bern, Bern, Switzerland; LOCEAN Laboratory, Sorbonne Universités, Paris, France