GLORIA

GEOMAR Library Ocean Research Information Access

Your email was sent successfully. Check your inbox.

An error occurred while sending the email. Please try again.

Proceed reservation?

Export
  • 1
    Online Resource
    Online Resource
    Elsevier BV ; 2018
    In:  Joule Vol. 2, No. 8 ( 2018-08), p. 1405-1407
    In: Joule, Elsevier BV, Vol. 2, No. 8 ( 2018-08), p. 1405-1407
    Type of Medium: Online Resource
    ISSN: 2542-4351
    Language: English
    Publisher: Elsevier BV
    Publication Date: 2018
    detail.hit.zdb_id: 2952490-8
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 2
    Online Resource
    Online Resource
    American Chemical Society (ACS) ; 2013
    In:  Environmental Science & Technology Vol. 47, No. 23 ( 2013-12-03), p. 13900-13901
    In: Environmental Science & Technology, American Chemical Society (ACS), Vol. 47, No. 23 ( 2013-12-03), p. 13900-13901
    Type of Medium: Online Resource
    ISSN: 0013-936X , 1520-5851
    RVK:
    Language: English
    Publisher: American Chemical Society (ACS)
    Publication Date: 2013
    detail.hit.zdb_id: 280653-8
    detail.hit.zdb_id: 1465132-4
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 3
    Online Resource
    Online Resource
    American Chemical Society (ACS) ; 2013
    In:  Environmental Science & Technology Vol. 47, No. 12 ( 2013-06-18), p. 6718-6719
    In: Environmental Science & Technology, American Chemical Society (ACS), Vol. 47, No. 12 ( 2013-06-18), p. 6718-6719
    Type of Medium: Online Resource
    ISSN: 0013-936X , 1520-5851
    RVK:
    Language: English
    Publisher: American Chemical Society (ACS)
    Publication Date: 2013
    detail.hit.zdb_id: 280653-8
    detail.hit.zdb_id: 1465132-4
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 4
    In: Earth System Dynamics, Copernicus GmbH, Vol. 8, No. 3 ( 2017-07-18), p. 577-616
    Abstract: Abstract. Global temperature is a fundamental climate metric highly correlated with sea level, which implies that keeping shorelines near their present location requires keeping global temperature within or close to its preindustrial Holocene range. However, global temperature excluding short-term variability now exceeds +1 °C relative to the 1880–1920 mean and annual 2016 global temperature was almost +1.3 °C. We show that global temperature has risen well out of the Holocene range and Earth is now as warm as it was during the prior (Eemian) interglacial period, when sea level reached 6–9 m higher than today. Further, Earth is out of energy balance with present atmospheric composition, implying that more warming is in the pipeline, and we show that the growth rate of greenhouse gas climate forcing has accelerated markedly in the past decade. The rapidity of ice sheet and sea level response to global temperature is difficult to predict, but is dependent on the magnitude of warming. Targets for limiting global warming thus, at minimum, should aim to avoid leaving global temperature at Eemian or higher levels for centuries. Such targets now require negative emissions, i.e., extraction of CO2 from the air. If phasedown of fossil fuel emissions begins soon, improved agricultural and forestry practices, including reforestation and steps to improve soil fertility and increase its carbon content, may provide much of the necessary CO2 extraction. In that case, the magnitude and duration of global temperature excursion above the natural range of the current interglacial (Holocene) could be limited and irreversible climate impacts could be minimized. In contrast, continued high fossil fuel emissions today place a burden on young people to undertake massive technological CO2 extraction if they are to limit climate change and its consequences. Proposed methods of extraction such as bioenergy with carbon capture and storage (BECCS) or air capture of CO2 have minimal estimated costs of USD 89–535 trillion this century and also have large risks and uncertain feasibility. Continued high fossil fuel emissions unarguably sentences young people to either a massive, implausible cleanup or growing deleterious climate impacts or both.
    Type of Medium: Online Resource
    ISSN: 2190-4987
    Language: English
    Publisher: Copernicus GmbH
    Publication Date: 2017
    detail.hit.zdb_id: 2578793-7
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 5
  • 6
    Online Resource
    Online Resource
    American Chemical Society (ACS) ; 2013
    In:  Environmental Science & Technology Vol. 47, No. 9 ( 2013-05-07), p. 4889-4895
    In: Environmental Science & Technology, American Chemical Society (ACS), Vol. 47, No. 9 ( 2013-05-07), p. 4889-4895
    Type of Medium: Online Resource
    ISSN: 0013-936X , 1520-5851
    RVK:
    Language: English
    Publisher: American Chemical Society (ACS)
    Publication Date: 2013
    detail.hit.zdb_id: 280653-8
    detail.hit.zdb_id: 1465132-4
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 7
    In: Atmospheric Chemistry and Physics, Copernicus GmbH, Vol. 16, No. 6 ( 2016-03-22), p. 3761-3812
    Abstract: Abstract. We use numerical climate simulations, paleoclimate data, and modern observations to study the effect of growing ice melt from Antarctica and Greenland. Meltwater tends to stabilize the ocean column, inducing amplifying feedbacks that increase subsurface ocean warming and ice shelf melting. Cold meltwater and induced dynamical effects cause ocean surface cooling in the Southern Ocean and North Atlantic, thus increasing Earth's energy imbalance and heat flux into most of the global ocean's surface. Southern Ocean surface cooling, while lower latitudes are warming, increases precipitation on the Southern Ocean, increasing ocean stratification, slowing deepwater formation, and increasing ice sheet mass loss. These feedbacks make ice sheets in contact with the ocean vulnerable to accelerating disintegration. We hypothesize that ice mass loss from the most vulnerable ice, sufficient to raise sea level several meters, is better approximated as exponential than by a more linear response. Doubling times of 10, 20 or 40 years yield multi-meter sea level rise in about 50, 100 or 200 years. Recent ice melt doubling times are near the lower end of the 10–40-year range, but the record is too short to confirm the nature of the response. The feedbacks, including subsurface ocean warming, help explain paleoclimate data and point to a dominant Southern Ocean role in controlling atmospheric CO2, which in turn exercised tight control on global temperature and sea level. The millennial (500–2000-year) timescale of deep-ocean ventilation affects the timescale for natural CO2 change and thus the timescale for paleo-global climate, ice sheet, and sea level changes, but this paleo-millennial timescale should not be misinterpreted as the timescale for ice sheet response to a rapid, large, human-made climate forcing. These climate feedbacks aid interpretation of events late in the prior interglacial, when sea level rose to +6–9 m with evidence of extreme storms while Earth was less than 1 °C warmer than today. Ice melt cooling of the North Atlantic and Southern oceans increases atmospheric temperature gradients, eddy kinetic energy and baroclinicity, thus driving more powerful storms. The modeling, paleoclimate evidence, and ongoing observations together imply that 2 °C global warming above the preindustrial level could be dangerous. Continued high fossil fuel emissions this century are predicted to yield (1) cooling of the Southern Ocean, especially in the Western Hemisphere; (2) slowing of the Southern Ocean overturning circulation, warming of the ice shelves, and growing ice sheet mass loss; (3) slowdown and eventual shutdown of the Atlantic overturning circulation with cooling of the North Atlantic region; (4) increasingly powerful storms; and (5) nonlinearly growing sea level rise, reaching several meters over a timescale of 50–150 years. These predictions, especially the cooling in the Southern Ocean and North Atlantic with markedly reduced warming or even cooling in Europe, differ fundamentally from existing climate change assessments. We discuss observations and modeling studies needed to refute or clarify these assertions.
    Type of Medium: Online Resource
    ISSN: 1680-7324
    Language: English
    Publisher: Copernicus GmbH
    Publication Date: 2016
    detail.hit.zdb_id: 2092549-9
    detail.hit.zdb_id: 2069847-1
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 8
    Online Resource
    Online Resource
    The Royal Society ; 2007
    In:  Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences Vol. 365, No. 1856 ( 2007-07-15), p. 1925-1954
    In: Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, The Royal Society, Vol. 365, No. 1856 ( 2007-07-15), p. 1925-1954
    Abstract: Palaeoclimate data show that the Earth's climate is remarkably sensitive to global forcings. Positive feedbacks predominate. This allows the entire planet to be whipsawed between climate states. One feedback, the ‘albedo flip’ property of ice/water, provides a powerful trigger mechanism. A climate forcing that ‘flips’ the albedo of a sufficient portion of an ice sheet can spark a cataclysm. Inertia of ice sheet and ocean provides only moderate delay to ice sheet disintegration and a burst of added global warming. Recent greenhouse gas (GHG) emissions place the Earth perilously close to dramatic climate change that could run out of our control, with great dangers for humans and other creatures. Carbon dioxide (CO 2 ) is the largest human-made climate forcing, but other trace constituents are also important. Only intense simultaneous efforts to slow CO 2 emissions and reduce non-CO 2 forcings can keep climate within or near the range of the past million years. The most important of the non-CO 2 forcings is methane (CH 4 ), as it causes the second largest human-made GHG climate forcing and is the principal cause of increased tropospheric ozone (O 3 ), which is the third largest GHG forcing. Nitrous oxide (N 2 O) should also be a focus of climate mitigation efforts. Black carbon (‘black soot’) has a high global warming potential (approx. 2000, 500 and 200 for 20, 100 and 500 years, respectively) and deserves greater attention. Some forcings are especially effective at high latitudes, so concerted efforts to reduce their emissions could preserve Arctic ice, while also having major benefits for human health, agricultural productivity and the global environment.
    Type of Medium: Online Resource
    ISSN: 1364-503X , 1471-2962
    RVK:
    Language: English
    Publisher: The Royal Society
    Publication Date: 2007
    detail.hit.zdb_id: 208381-4
    detail.hit.zdb_id: 1462626-3
    SSG: 11
    SSG: 5,1
    SSG: 5,21
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 9
    In: Oxford Open Climate Change, Oxford University Press (OUP), Vol. 3, No. 1 ( 2023-02-14)
    Abstract: Improved knowledge of glacial-to-interglacial global temperature change yields Charney (fast-feedback) equilibrium climate sensitivity 1.2 ± 0.3°C (2σ) per W/m2, which is 4.8°C ± 1.2°C for doubled CO2. Consistent analysis of temperature over the full Cenozoic era—including ‘slow’ feedbacks by ice sheets and trace gases—supports this sensitivity and implies that CO2 was 300–350 ppm in the Pliocene and about 450 ppm at transition to a nearly ice-free planet, exposing unrealistic lethargy of ice sheet models. Equilibrium global warming for today’s GHG amount is 10°C, which is reduced to 8°C by today’s human-made aerosols. Equilibrium warming is not ‘committed’ warming; rapid phaseout of GHG emissions would prevent most equilibrium warming from occurring. However, decline of aerosol emissions since 2010 should increase the 1970–2010 global warming rate of 0.18°C per decade to a post-2010 rate of at least 0.27°C per decade. Thus, under the present geopolitical approach to GHG emissions, global warming will exceed 1.5°C in the 2020s and 2°C before 2050. Impacts on people and nature will accelerate as global warming increases hydrologic (weather) extremes. The enormity of consequences demands a return to Holocene-level global temperature. Required actions include: (1) a global increasing price on GHG emissions accompanied by development of abundant, affordable, dispatchable clean energy, (2) East-West cooperation in a way that accommodates developing world needs, and (3) intervention with Earth’s radiation imbalance to phase down today’s massive human-made ‘geo-transformation’ of Earth’s climate. Current political crises present an opportunity for reset, especially if young people can grasp their situation.
    Type of Medium: Online Resource
    ISSN: 2634-4068
    Language: English
    Publisher: Oxford University Press (OUP)
    Publication Date: 2023
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
  • 10
    Online Resource
    Online Resource
    IOP Publishing ; 2013
    In:  Environmental Research Letters Vol. 8, No. 1 ( 2013-03-01), p. 011006-
    In: Environmental Research Letters, IOP Publishing, Vol. 8, No. 1 ( 2013-03-01), p. 011006-
    Abstract: Rahmstorf et al ’s (2012) conclusion that observed climate change is comparable to projections, and in some cases exceeds projections, allows further inferences if we can quantify changing climate forcings and compare those with projections. The largest climate forcing is caused by well-mixed long-lived greenhouse gases. Here we illustrate trends of these gases and their climate forcings, and we discuss implications. We focus on quantities that are accurately measured, and we include comparison with fixed scenarios, which helps reduce common misimpressions about how climate forcings are changing. Annual fossil fuel CO 2 emissions have shot up in the past decade at about 3% yr -1 , double the rate of the prior three decades (figure 1). The growth rate falls above the range of the IPCC (2001) ‘Marker’ scenarios, although emissions are still within the entire range considered by the IPCC SRES (2000). The surge in emissions is due to increased coal use (blue curve in figure 1), which now accounts for more than 40% of fossil fuel CO 2 emissions. Figure 1. CO 2 annual emissions from fossil fuel use and cement manufacture, an update of figure 16 of Hansen (2003) using data of British Petroleum (BP 2012) concatenated with data of Boden et al (2012). The resulting annual increase of atmospheric CO 2 (12-month running mean) has grown from less than 1 ppm yr -1 in the early 1960s to an average ~2 ppm yr -1 in the past decade (figure 2). Although CO 2 measurements were not made at sufficient locations prior to the early 1980s to calculate the global mean change, the close match of global and Mauna Loa data for later years suggests that Mauna Loa data provide a good approximation of global change (figure 2), thus allowing a useful estimate of annual global change beginning with the initiation of Mauna Loa measurements in 1958 by Keeling et al (1973). Figure 2. Annual increase of CO 2 based on data from the NOAA Earth System Research Laboratory (ESRL 2012). CO 2 change and global temperature change are 12-month running means of differences for the same month of consecutive years. Nino index (Nino3.4 area) is 12-month running mean. Both temperature indices use data from Hansen et al (2010). Annual mean CO 2 amount in 1958 was 315 ppm (Mauna Loa) and in 2012 was 394 ppm (Mauna Loa) and 393 ppm (Global). Interannual variability of CO 2 growth is correlated with ENSO (El Nino Southern Oscillation) variations of tropical temperatures (figure 2). Ocean–atmosphere CO 2 exchange is affected by ENSO (Chavez et al 1999), but ENSO seems to have a greater impact on atmospheric CO 2 via the terrestrial carbon cycle through effects on the water cycle, temperature, and fire, as discussed in a large body of literature (referenced, e.g., by Schwalm et al 2011). In addition, volcanoes, such as the 1991 Mount Pinatubo eruption, slow the increase of atmospheric CO 2 (Rothenberg et al 2012), at least in part because photosynthesis is enhanced by the increased proportion of diffuse sunlight (Gu et al 2003, Mercado et al 2009). Watson (1997) suggests that volcanic dust deposited on the ocean surface may also contribute to CO 2 uptake by increasing ocean productivity. An important question is whether ocean and terrestrial carbon sinks will tend to saturate as human-made CO 2 emissions continue. Piao et al (2008) and Zhao and Running (2010) suggest that there already may be a reduction of terrestrial carbon uptake, while Le Quéré et al (2007) and Schuster and Watson (2007) find evidence of decreased carbon uptake in the Southern Ocean and North Atlantic Ocean, respectively. However, others (Knorr 2009, Sarmiento et al 2010, Ballantyne et al 2012) either cast doubt on the reality of a reduced uptake strength or find evidence for increased uptake. An informative presentation of CO 2 observations is the ratio of annual CO 2 increase in the air divided by annual fossil fuel CO 2 emissions (Keeling et al 1973), the ‘airborne fraction’ (figure 3, right scale). An alternative definition of airborne fraction includes in the denominator of this ratio an estimated net anthropogenic CO 2 source from changes in land use, but this latter term is much more uncertain than the two terms involved in the Keeling et al (1973) definition. For example, analysis by Harris et al (2012) reveals a range as high as a factor of 2–4 in estimates of recent land use emissions; see also the discussion by Sarmiento et al (2010). However, note that the airborne fraction becomes smaller when estimated land use emissions are included, with the uptake fraction (one minus airborne fraction) typically greater than 0.5. Figure 3. Fossil fuel CO 2 emissions (left scale) and airborne fraction, i.e., the ratio of observed atmospheric CO 2 increase to fossil fuel CO 2 emissions. Final three points are 5-, 3- and 1-year means. The simple Keeling airborne fraction, clearly, is not increasing (figure 3). Thus the net ocean plus terrestrial sink for carbon emissions has increased by a factor of 3–4 since 1958, accommodating the emissions increase by that factor. Remarkably, and we will argue importantly, the airborne fraction has declined since 2000 (figure 3) during a period without any large volcanic eruptions. The 7-year running mean of the airborne fraction had remained close to 60% up to 2000, except for the period affected by Pinatubo. The airborne fraction is affected by factors other than the efficiency of carbon sinks, most notably by changes in the rate of fossil fuel emissions (Gloor et al 2010). However, it is the dependence of the airborne fraction on fossil fuel emission rate that makes the post-2000 downturn of the airborne fraction particularly striking. The change of emission rate in 2000 from 1.5% yr -1 to 3.1% yr -1 (figure 1), other things being equal, would have caused a sharp increase of the airborne fraction (the simple reason being that a rapid source increase provides less time for carbon to be moved downward out of the ocean’s upper layers). A decrease in land use emissions during the past decade (Harris et al 2012) could contribute to the decreasing airborne fraction in figure 3, although Malhi (2010) presents evidence that tropical forest deforestation and regrowth are approximately in balance, within uncertainties. Land use change can be only a partial explanation for the decrease of the airborne fraction; something more than land use change seems to be occurring. We suggest that the huge post-2000 increase of uptake by the carbon sinks implied by figure 3 is related to the simultaneous sharp increase in coal use (figure 1). Increased coal use occurred primarily in China and India (Boden et al 2012; BP 2012; see graphs at www.columbia.edu/~mhs119/Emissions/Emis_moreFigs/ ). Satellite radiance measurements for July–December, months when desert dust does not dominate aerosol amount, yield an increase of aerosol optical depth in East Asia of about 4% yr -1 during 2000–2006 (van Donkelaar et al 2008). Associated gaseous and particulate emissions increased rapidly after 2000 in China and India (Lu et al 2011, Tian et al 2010). Some decrease of the sulfur component of emissions occurred in China after 2006 as wide application of flue-gas desulfurization began to be initiated (Lu et al 2010), but this was largely offset by continuing emission increases from India (Lu et al 2011). We suggest that the surge of fossil fuel use, mainly coal, since 2000 is a basic cause of the large increase of carbon uptake by the combined terrestrial and ocean carbon sinks. One mechanism by which fossil fuel emissions increase carbon uptake is by fertilizing the biosphere via provision of nutrients essential for tissue building, especially nitrogen, which plays a critical role in controlling net primary productivity and is limited in many ecosystems (Gruber and Galloway 2008). Modeling (e.g., Thornton et al 2009) and field studies (Magnani et al 2007) confirm a major role of nitrogen deposition, working in concert with CO 2 fertilization, in causing a large increase in net primary productivity of temperate and boreal forests. Sulfate aerosols from coal burning also might increase carbon uptake by increasing the proportion of diffuse insolation, as noted above for Pinatubo aerosols, even though the total solar radiation reaching the surface is reduced. Thus we see the decreased CO 2 airborne fraction since 2000 as sharing some of the same causes as the decreased airborne fraction after the Pinatubo eruption (figure 3). CO 2 fertilization is likely the major effect, as a plausible addition of 5 TgN yr -1 from fossil fuels and net ecosystem productivity of 200 kgC kgN -1 (Magnani et al 2007, 2008) yields an annual carbon drawdown of 1 GtC yr -1 , which is of the order of what is needed to explain the post-2000 anomaly in airborne CO 2 . However, an aerosol-induced increase of diffuse radiation might also contribute. Although tropospheric aerosol properties are not accurately monitored, there are suggestions of an upward trend of stratospheric background aerosols since 2000 (Hofmann et al 2009, Solomon et al 2011), which could be a consequence of more tropospheric aerosols at low latitudes where injection of tropospheric air into the stratosphere occurs (Holton et al 1995). We discuss climate implications of the reduced CO 2 airborne fraction after presenting data for other greenhouse gases. Atmospheric CH 4 is increasing more slowly than in IPCC scenarios (figure 4), which were defined more than a decade ago (IPCC 2001). However, after remaining nearly constant for several years, CH 4 has increased during the past five years, pushing slightly above the level that was envisaged in the Alternative Scenario of Hansen et al (2000). Reduction of CH 4 , besides slowdown in CO 2 growth in the twenty first century and a decline of CO 2 in the twenty second century, is a principal requirement to achieve a low climate forcing that stabilizes climate, in part because CH 4 also affects tropospheric ozone and stratospheric water vapor. The Alternative Scenario, defined in detail by Hansen and Sato (2004), keeps maximum global warming at ~1.5 °C relative to 1880–1920, under the assumption that fast-feedback climate sensitivity is ~3 °C for doubled CO 2 (Hansen et al 2007). The Alternative Scenario allows CO 2 to reach 475 ppm in 2100 before declining slowly; this scenario assumes that reductions of non-CO 2 greenhouse gases and black carbon aerosols can be achieved sufficient to balance the warming effect of likely future decreases of reflective aerosols. Figure 4. Observed atmospheric CH 4 amount and scenarios for twenty first century. Alternative scenario (Hansen et al 2000, Hansen and Sato 2004) yields maximum global warming ~1.5 °C above 1880–1920. Other scenarios are from IPCC (2001). Forcing on right hand scale is adjusted forcing, Fa, relative to values in 2000 (Hansen et al 2007). There are anthropogenic sources of CH 4 that potentially could be reduced, indeed, the leveling off of CH 4 amount during the past 20 years seems to have been caused by decreased venting in oil fields (Simpson et al 2012), but the feasibility of overall CH 4 reduction also depends on limiting global warming itself, because of the potential for amplifying climate-CH 4 feedbacks (Archer et al 2009, Koven et al 2011). Furthermore, reduction of atmospheric CH 4 might become problematic if unconventional mining of gas, such as ‘hydro-fracking’, expands widely (Cipolla 2009), as discussed further below. The growth rate for the total climate forcing by well-mixed greenhouse gases has remained below the peak values reached in the 1970s and early 1980s, has been relatively stable for about 20 years, and is falling below IPCC (2001) scenarios (figure 5). However, the greenhouse gas forcing is growing faster than in the Alternative Scenario. MPTGs and OTGs in figure 5 are Montreal Protocol Trace Gases and Other Trace Gases (Hansen and Sato 2004). Figure 5. Five-year mean of the growth rate of climate forcing by well-mixed greenhouse gases, an update of figure 4 of Hansen and Sato (2004). Forcing calculations use equations of Hansen et al (2000). The moderate uncertainties in radiative calculations affect the scenarios and actual greenhouse gas results equally and thus do not alter the conclusion that the actual forcing falls below that of the IPCC scenarios. If greenhouse gases were the only climate forcing, we would be tempted to infer from Rahmstorf’s conclusion (that actual climate change has exceeded IPCC projections) and our conclusion (that actual greenhouse gas forcings are slightly smaller than IPCC scenarios) that actual climate sensitivity is on the high side of what has generally been assumed. Although that may be a valid inference, the evidence is weakened by the fact that other climate forcings are not negligible in comparison to the greenhouse gases and must be accounted for. Natural forcings, by changing solar irradiance and volcanic aerosols, are well-measured since the late 1970s and included in most IPCC (2007) climate simulations. The difficulty is human-made aerosols. Aerosols are readily detected in satellite observations, but determination of their climate forcing requires accurate knowledge of changes in aerosol amount, size distribution, absorption and vertical distribution on a global basis—as well as simultaneous data on changes in cloud properties to allow inference of the indirect aerosol forcing via induced cloud changes. Unfortunately, the first satellite mission capable of measuring the needed aerosol characteristics (Aerosol Polarimetry Sensor on the Glory satellite, (Mishchenko et al 2007)) suffered a launch failure and as yet there are no concrete plans for a replacement mission. The human-made aerosol climate forcing thus remains uncertain. IPCC (2007) concludes that aerosols are a negative (cooling) forcing, probably between -0.5 and -2.5 W m -2 . Hansen et al (2011), based mainly on analysis of Earth’s energy imbalance, derive an aerosol forcing -1.6 ± 0.3 W m -2 , consistent with an analysis of Murphy et al (2009) that suggests an aerosol forcing about -1.5 W m -2 (see discussion in Hansen et al (2011)). This large negative aerosol forcing reduces the net climate forcing of the past century by about half (IPCC 2007; figure 1 of Hansen et al 2011). Coincidentally, this leaves net climate forcing comparable to the CO 2 forcing alone. Reduction of the net human-made climate forcing by aerosols has been described as a ‘Faustian bargain’ (Hansen and Lacis 1990, Hansen 2009), because the aerosols constitute deleterious particulate air pollution. Reduction of the net climate forcing by half will continue only if we allow air pollution to build up to greater and greater amounts. More likely, humanity will demand and achieve a reduction of particulate air pollution, whereupon, because the CO 2 from fossil fuel burning remains in the surface climate system for millennia, the ‘devil’s payment’ will be extracted from humanity via increased global warming. So is the new data we present here good news or bad news, and how does it alter the ‘Faustian bargain’? At first glance there seems to be some good news. First, if our interpretation of the data is correct, the surge of fossil fuel emissions, especially from coal burning, along with the increasing atmospheric CO 2 level is ‘fertilizing’ the biosphere, and thus limiting the growth of atmospheric CO 2 . Also, despite the absence of accurate global aerosol measurements, it seems that the aerosol cooling effect is probably increasing based on evidence of aerosol increases in the Far East and increasing ‘background’ stratospheric aerosols. Both effects work to limit global warming and thus help explain why the rate of global warming seems to be less this decade than it has been during the prior quarter century. This data interpretation also helps explain why multiple warnings that some carbon sinks are ‘drying up’ and could even become carbon sources, e.g., boreal forests infested by pine bark beetles (Kurz et al 2008) and the Amazon rain forest suffering from drought (Lewis et al 2011), have not produced an obvious impact on atmospheric CO 2 . However, increased CO 2 uptake does not necessarily mean that the biosphere is healthier or that the increased carbon uptake will continue indefinitely (Matson et al 2002, Galloway et al 2002, Heimann and Reichstein 2008, Gruber and Galloway 2008). Nor does it change the basic facts about the potential magnitude of the fossil fuel carbon source (figure 6) and the long lifetime of the CO 2 in the surface carbon reservoirs (atmosphere, ocean, soil, biosphere) once the fossil fuels are burned (Archer 2005). Fertilization of the biosphere affects the distribution of the fossil fuel carbon among these reservoirs, at least on the short run, but it does not alter the fact that the fossil carbon will remain in these reservoirs for millennia. Figure 6. Fossil fuel CO 2 emissions and carbon content (1 ppm atmospheric CO 2 ~2.12 GtC). Historical emissions are from Boden et al (2012). Estimated reserves and potentially recoverable resources are based on energy content values of Energy Information Administration (EIA 2011), German Advisory Council (GAC 2011), and Global Energy Assessment (GEA 2012). We convert energy content to carbon content using emission factors of Table 4.2 of IPCC (2007) for coal, gas, and conventional oil, and, following IPCC, we use an emission factor of unconventional oil the same as that for coal. Humanity, so far, has burned only a small portion (purple area in figure 6) of total fossil fuel reserves and resources. Yet deleterious effects of warming are apparent (IPCC 2007), even though only about half of the warming due to gases now in the air has appeared, the remainder still ‘in the pipeline’ due to the inertia of the climate system (Hansen et al 2011). Already it seems difficult to avoid passing the ‘guardrail’ of no more than 2 °C global warming that was agreed in the Copenhagen Accord of the United Nations Framework Convention on Climate Change (UNFCCC 2010). And Hansen et al (2008), based primarily on paleoclimate data and evidence of deleterious climate impacts already at 385 ppm CO 2 , concluded that an appropriate initial target for CO 2 was 350 ppm, which implied a global temperature limit, relative to 1880–1920 of about 1 °C. What is clear is that most of the remaining fossil fuels must be left in the ground if we are to avoid dangerous human-made interference with climate. The principal implication of our present analysis probably relates to the Faustian bargain. Increased short-term masking of greenhouse gas warming by fossil fuel particulate and nitrogen pollution represents a ‘doubling down’ of the Faustian bargain, an increase in the stakes. The more we allow the Faustian debt to build, the more unmanageable the eventual consequences will be. Yet globally there are plans to build more than 1000 coal-fired power plants (Yang and Cui 2012) and plans to develop some of the dirtiest oil sources on the planet (EIA 2011). These plans should be vigorously resisted. We are already in a deep hole—it is time to stop digging. Acknowledgments We thank ClimateWorks, Energy Foundation, Gerry Lenfest (Lenfest Foundation), Lee Wasserman (Rockefeller Family Foundation), and Stephen Toben (Flora Family Foundation) for research and communications support. References Archer D 2005 Fate of fossil fuel CO2 in geologic time J. Geophys. Res. 110 C09505 Archer D, Buffett B and Brovkin V 2009 Ocean methane hydrates as a slow tipping point in the global carbon cycle Proc. Natl Acad. Sci. 106 20596–601 Ballantyne A P, Alden C B, Miller J B, Tans P P and White J W C 2012 Increase in observed net carbon dioxide uptake by land and oceans during the past 50 years Nature 488 70-– Boden T A, Marland G and Andres R J 2012 Global, Regional, and National Fossil-Fuel CO 2 Emissions (Oak Ridge, TN: Carbon Dioxide Information and Analysis Center, Oak Ridge National Laboratory, US Department of Energy) doi: 10.3334/CDIAC/00001_V2012 BP (British Petroleum) 2012 Statistical Review of World Energy 2012 ( www.bp.com/sectionbodycopy.do?categoryId=7500 & contentId=7068481 ) Chavez F P et al 1999 Biological and chemical response of the Equatorial Pacific Ocean to the 1997–1998 El Nino Science 286 2126–31 Cipolla C L 2009 Modeling production and evaluating fracture performance in unconventional gas reservoirs J. Petrol. Technol. 61 84–90 Earth System Research Laboratory (ESRL) 2012 Trends in Atmospheric Carbon Dioxide ( www.esrl.noaa.gov/gmd/ccgg/trends/ ) Energy Information Administration (EIA) 2011 International Energy Outlook ( www.eia.gov/forecasts/ieo/pdf/0484(2011).pdf , accessed Sep. 2011) Galloway J N, Cowling E B, Seitzinger S P and Socowlow R H 2002 Reactive nitrogen: too much of a good thing? AMBIO 31 60–3 German Advisory Council on Global Change (GAC) 2011 World in Transition—A Social Contract for Sustainability ( www.wbgu.de/en/flagship-reports/fr-2011-a-social-contract/ , accessed Oct. 2011) Global Energy Assessment (GEA) 2012 Toward a Sustainable Future ed T B Johanson et al (Luxemburg: International Institute for Applied Systems Analysis) p 118 Gloor M, Sarmiento J L and Gruber N 2010 What can be learned about carbon cycle climate feedbacks from the CO2 airborne fraction? Atmos. Chem. Phys. 10 7739–51 Gruber N and Galloway J N 2008 An Earth-system perspective of the global nitrogen cycle Nature 451 293–6 Gu L et al 2003 Response of a deciduous forest to the Mount Pinatubo eruption: enhanced photosynthesis Science 299 2035–8 Hansen J 2003 Can we defuse the global warming time bomb? Natural Science posted 1 Aug 2003 Hansen J 2009 Storms of My Grandchildren: The Truth About the Coming Climate Catastrophe and Our Last Chance to Save Humanity (New York: Bloomsbury) p 304 Hansen J E and Lacis A A 1990 Sun and dust versus greenhouse gases: an assessment of their relative roles in global climate change Nature 346 713–9 Hansen J, Ruedy R, Sato M and Lo K 2010 Global surface temperature change Rev. Geophys. 48 RG4004 Hansen J and Sato M 2004 Greenhouse gas growth rates Proc. Natl Acad. Sci. 101 16109–14 Hansen J, Sato M, Kharecha P and von Schuckmann K 2011 Earth’s energy imbalance and implications Atmos. Chem. Phys. 11 13421–49 Hansen J, Sato M, Ruedy R, Lacis A and Oinas V 2000 Global warming in the twenty-first century: an alternative scenario Proc. Natl Acad. Sci. 97 9875–80 Hansen J et al 2007 Dangerous human-made interference with climate: a GISS modelE study Atmos. Chem. Phys. 7 2287–312 Hansen J et al 2008 Target atmospheric CO2: where should humanity aim? Open Atmos. Sci. 2 217–31 Harris N L et al 2012 Baseline map of carbon emissions from deforestation in tropical regions Science 336 1573–6 Heimann M and Reichstein M 2008 Terrestrial ecosystem carbon dynamics and climate feedbacks Nature 451 289–92 Hofmann D, Barnes J, O’Niel M, Trudeau M and Neely R 2009 Increase in background stratospheric aerosol observed with lidar at Mauna Loa Observatory and Boulder, Colorado Geophys. Res. Lett. 36 L15808 Holton J R et al 1995 Stratosphere–troposphere exchange Rev. Geophys. 33 403–39 IPCC (Intergovernmental Panel on Climate Change) 2000 Special Report on Emission Scenarios (SRES) ed N Nakicenovic et al (Cambridge: Cambridge University Press) p 599 IPCC (Intergovernmental Panel on Climate Change) 2001 Climate Change 2001: The Scientific Basis ed J T Houghton et al (Cambridge: Cambridge University Press) p 881 IPCC (Intergovernmental Panel on Climate Change) 2007 Climate Change 2007: The Physical Science Basis ed S Solomon et al (Cambridge: Cambridge University Press) p 996 Keeling C D, Whorf T P, Wahlen M and van der Plicht J 1973 Interannual extremes in the rate of rise of atmospheric carbon dioxide since 1980 Nature 375 666–70 Knorr W 2009 Is the airborne fraction of anthropogenic CO2 emissions increasing? Geophys. Res. Lett. 36 L21710 Koven C D et al 2011 Permafrost carbon-climate feedbacks accelerte global warming Proc. Natl Acad. Sci. 108 14769–74 Kurz W A et al 2008 Mountain pine beetle and forest carbon feedback to climate change Nature 452 987–90 Le Quéré C et al 2007 Saturation of the Southern Ocean CO2 sink due to recent climate change Science 316 1735–8 Lewis S L, Brando P M, Phillips O L, van der Heijden G M F and Nepstad D 2011 The 2010 Amazon drought Science 331 554 Lu Z, Zhang Q and Streets D G 2011 Sulfur dioxide and primary carbonaceous aerosol emissions in China and India, 1996–2010 Atmos. Chem. Phys. 11 9839–64 Lu Z et al 2010 Sulfur dioxide emissions in China and sulfur trends in East Asia since 2000 Atmos. Chem Phys. 10 6311–31 Magnani F et al 2007 The human footprint in the carbon cycle of temperate and boreal forests Nature 447 848–50 Magnani F et al 2008 Magnani et al. reply Nature 451 E28–9 Malhi Y 2010 The carbon balance of tropical forest regions, 1990–2005 Curr. Opin. Environ. Sustain. 2 237–44 Matson P, Lohse K A and Hall S J 2002 The globalization of nitrogen deposition: consequences for terrestrial ecosystems AMBIO 31 113–9 Mercado L M et al 2009 Impact of changes in diffuse radiation on the global land carbon sink Nature 458 1014–7 Mishchenko M I et al 2007 Accurate monitoring of terrestrial aerosols and total solar irradiance: introducing the Glory mission Bull. Am. Meteorol. Soc. 88 677–91 Murphy D M et al 2009 An observationally based energy balance for the Earth since 1950 J. Geophys. Res. 114 D17107 Piao S et al 2008 Net carbon dioxide losses of northern ecosystems in response to autumn warming Nature 451 49–52 Rahmstorf S, Foster G and Cazenave A 2012 Comparing climate projections to observations up to 2011 Environ. Res. Lett. 7 044035 Rothenberg D et al 2012 Volcano impacts on climate and biogeochemistry in a coupled carbon-climate model Earth Syst. Dyn. Discuss. 3 279–323 Sarmiento J L et al 2010 Trends and regional distributions of land and ocean carbon sinks Biogeosciences 7 2351–67 Schuster U and Watson A J 2007 A variable and decreasing sink for atmospheric CO2 in the North Atlantic J. Geophys. Res. 112 C11006 Schwalm C R et al 2011 Does terrestrial drought explain global CO2 flux anomalies induced by El Nino? Biogeosciences 8 2493–506 Simpson I J et al 2012 Long-term decline of global atmospheric ethane concentrations and implications for methane Nature 488 490–4 Solomon S et al 2011 The persistently variable ‘background’ stratospheric aerosol layer and global climate change Science 333 866–70 Thornton P E et al 2009 Carbon–nitrogen interactions regulate climate-carbon cycle feedbacks: results from an atmosphere-ocean general circulation model Biogeosciences 6 2099–120 Tian H Z et al 2010 Trend and characteristics of atmospheric emissions of Hg, As, and Se from coal combustion in China, 1980–2007 Atmos. Chem. Phys. 10 11905–19 UNFCCC (United Nations Framework Convention on Climate Change) 2010 Copenhagen Accord ( http://unfccc.int/resource/docs/2009/cop15/eng/11a01.pdf , accessed 25 Nov. 2012) van Donkelaar A et al 2008 Analysis of aircraft and satellite measurements from Interconinental Chemical Transport Experiment (INTEX-B) to quantify long-range transport of East Asian sulfur to Canada Atmos. Chem. Phys. 8 2999–3014 Watson A J 1997 Volcanic iron, CO2, ocean productivity and climate Nature 385 587–8 Yang A and Cui Y 2012 Global coal risk assessment: data analysis and market research WRI Working Paper (Washington, DC: World Resources Institute) ( www.wri.org/publication/global-coal-risk-assessment ) Zhao M and Running S W 2010 Drought-induced reduction in global terrestrial net primary production from 2000 through 2009 Science 329 940–3
    Type of Medium: Online Resource
    ISSN: 1748-9326
    Language: Unknown
    Publisher: IOP Publishing
    Publication Date: 2013
    detail.hit.zdb_id: 2255379-4
    Location Call Number Limitation Availability
    BibTip Others were also interested in ...
Close ⊗
This website uses cookies and the analysis tool Matomo. More information can be found here...