Carbon dioxide in Earth's atmosphere

Last updated

Atmospheric CO2 concentration measured at Mauna Loa Observatory in Hawaii from 1958 to 2023 (also called the Keeling Curve). The rise in CO2 over that time period is clearly visible. The concentration is expressed as mmole per mole, or ppm. Mauna Loa CO2 monthly mean concentration.svg
Atmospheric CO2 concentration measured at Mauna Loa Observatory in Hawaii from 1958 to 2023 (also called the Keeling Curve). The rise in CO2 over that time period is clearly visible. The concentration is expressed as μmole per mole, or ppm.

In Earth's atmosphere, carbon dioxide is a trace gas that plays an integral part in the greenhouse effect, carbon cycle, photosynthesis and oceanic carbon cycle. It is one of three main greenhouse gases in the atmosphere of Earth. Water vapor is the primary greenhouse gas, as of 2010, contributing 50% of the greenhouse effect, followed by carbon dioxide at 20%. [1] The current global average concentration of carbon dioxide (CO2) in the atmosphere is 421 ppm (0.04%) as of May 2022. [2] This is an increase of 50% since the start of the Industrial Revolution, up from 280 ppm during the 10,000 years prior to the mid-18th century. [3] [2] [4] The increase is due to human activity. [5]

Contents

As of March 2024, the monthly average concentration of CO2 reached a new record high of 425.22 parts per million (ppm), marking an increase of 4.7 ppm over March 2023. By the latest measurement, levels had further escalated to 427.48 ppm. [6] This continuous increase in CO2 concentrations is a clear indicator of ongoing global environmental stress, primarily driven by the burning of fossil fuels, which is the principal cause of this rise and also a major contributor to climate change. [7] Other significant human activities that emit CO2 include cement production, deforestation, and biomass burning.

Carbon dioxide is a greenhouse gas. It absorbs and emits infrared radiation at its two infrared-active vibrational frequencies. The two wavelengths are 4.26  μm (2,347 cm−1) (asymmetric stretching vibrational mode) and 14.99 μm (667 cm−1) (bending vibrational mode). CO2 plays a significant role in influencing Earth's surface temperature through the greenhouse effect. [8] Light emission from the Earth's surface is most intense in the infrared region between 200 and 2500 cm−1, [9] as opposed to light emission from the much hotter Sun which is most intense in the visible region. Absorption of infrared light at the vibrational frequencies of atmospheric CO2 traps energy near the surface, warming the surface of Earth and its lower atmosphere. Less energy reaches the upper atmosphere, which is therefore cooler because of this absorption. [10]

The increase in atmospheric concentrations of CO2 and other long-lived greenhouse gases such as methane increase the absorption and emission of infrared radiation by the atmosphere. This has led to a rise in average global temperature and ocean acidification. Another direct effect is the CO2 fertilization effect. The increase in atmospheric concentrations of CO2 causes a range of further effects of climate change on the environment and human living conditions.

The present atmospheric concentration of CO2 is the highest for 14 million years. [11] Concentrations of CO2 in the atmosphere were as high as 4,000 ppm during the Cambrian period about 500 million years ago, and as low as 180 ppm during the Quaternary glaciation of the last two million years. [3] Reconstructed temperature records for the last 420 million years indicate that atmospheric CO2 concentrations peaked at approximately 2,000 ppm. This peak happened during the Devonian period (400 million years ago). Another peak occurred in the Triassic period (220–200 million years ago). [12]

Between 1850 and 2019 the Global Carbon Project estimates that about 2/3rds of excess carbon dioxide emissions have been caused by burning fossil fuels, and a little less than half of that has stayed in the atmosphere. Carbon Sources and Sinks.svg
Between 1850 and 2019 the Global Carbon Project estimates that about 2/3rds of excess carbon dioxide emissions have been caused by burning fossil fuels, and a little less than half of that has stayed in the atmosphere.

Current situation

Since the start of the Industrial Revolution, atmospheric CO2 concentration have been increasing, causing global warming and ocean acidification. [13] In October 2023 the average level of CO2 in Earth's atmosphere, adjusted for seasonal variation, was 422.17 parts per million by volume (ppm). [14] Figures are published monthly by the National Oceanic & Atmospheric Administration (NOAA). [15] [16] The value had been about 280 ppm during the 10,000 years up to the mid-18th century. [3] [2] [4]

Each part per million of CO2 in the atmosphere represents approximately 2.13 gigatonnes of carbon, or 7.82 gigatonnes of CO2. [17]

It was pointed out in 2021 that "the current rates of increase of the concentration of the major greenhouse gases (carbon dioxide, methane and nitrous oxide) are unprecedented over at least the last 800,000 years". [18] :515

It has been estimated that 2,400 gigatons of CO₂ have been emitted by human activity since 1850, with some absorbed by oceans and land, and about 950 gigatons remaining in the atmosphere. Around 2020 the emission rate was over 40 gigatons per year. [19]

Some fraction (a projected 20–35%) of the fossil carbon transferred thus far will persist in the atmosphere as elevated CO2 levels for many thousands of years after these carbon transfer activities begin to subside. [20] [21]

Annual and regional fluctuations

Atmospheric CO2 concentrations fluctuate slightly with the seasons, falling during the Northern Hemisphere spring and summer as plants consume the gas and rising during northern autumn and winter as plants go dormant or die and decay. The level drops by about 6 or 7 ppm (about 50 Gt) from May to September during the Northern Hemisphere's growing season, and then goes up by about 8 or 9 ppm. The Northern Hemisphere dominates the annual cycle of CO2 concentration because it has much greater land area and plant biomass in mid-latitudes (30-60 degrees) than the Southern Hemisphere. Concentrations reach a peak in May as the Northern Hemisphere spring greenup begins, and decline to a minimum in October, near the end of the growing season. [22] [23]

Concentrations also vary on a regional basis, most strongly near the ground with much smaller variations aloft. In urban areas concentrations are generally higher [24] and indoors they can reach 10 times background levels.

Measurements and predictions made in the recent past

Measurement techniques

Carbon dioxide observations from 2008 to 2017 showing the seasonal variations and the difference between northern and southern hemispheres Global distribution of Carbon Dioxide.jpg
Carbon dioxide observations from 2008 to 2017 showing the seasonal variations and the difference between northern and southern hemispheres

The concentrations of carbon dioxide in the atmosphere are expressed as parts per million by volume (abbreviated as ppmv, or ppm(v), or just ppm). To convert from the usual ppmv units to ppm mass (abbreviated as ppmm, or ppm(m)), multiply by the ratio of the molar mass of CO2 to that of air, i.e. times 1.52 (44.01 divided by 28.96).

The first reproducibly accurate measurements of atmospheric CO2 were from flask sample measurements made by Dave Keeling at Caltech in the 1950s. [32] Measurements at Mauna Loa have been ongoing since 1958. Additionally, measurements are also made at many other sites around the world. Many measurement sites are part of larger global networks. Global network data are often made publicly available.

Data networks

There are several surface measurement (including flasks and continuous in situ) networks including NOAA/ERSL, [33] WDCGG, [34] and RAMCES. [35] The NOAA/ESRL Baseline Observatory Network, and the Scripps Institution of Oceanography Network [36] data are hosted at the CDIAC at ORNL. The World Data Centre for Greenhouse Gases (WDCGG), part of GAW, data are hosted by the JMA. The Reseau Atmospherique de Mesure des Composes an Effet de Serre database (RAMCES) is part of IPSL.

From these measurements, further products are made which integrate data from the various sources. These products also address issues such as data discontinuity and sparseness. GLOBALVIEW-CO2 is one of these products. [37]

Analytical methods to investigate sources of CO2

Causes of the current increase

Anthropogenic CO2 emissions

The US, China and Russia have cumulatively contributed the greatest amounts of CO2 since 1850. 20211026 Cumulative carbon dioxide CO2 emissions by country - bar chart.svg
The US, China and Russia have cumulatively contributed the greatest amounts of CO2 since 1850.

While CO2 absorption and release is always happening as a result of natural processes, the recent rise in CO2 levels in the atmosphere is known to be mainly due to human (anthropogenic) activity. [18] Anthropogenic carbon emissions exceed the amount that can be taken up or balanced out by natural sinks. [42] Thus carbon dioxide has gradually accumulated in the atmosphere and, as of May 2022, its concentration is 50% above pre-industrial levels. [2]

The extraction and burning of fossil fuels, releasing carbon that has been underground for many millions of years, has increased the atmospheric concentration of CO2. [4] [13] As of year 2019 the extraction and burning of geologic fossil carbon by humans releases over 30 gigatonnes of CO2 (9 billion tonnes carbon) each year. [43] This larger disruption to the natural balance is responsible for recent growth in the atmospheric CO2 concentration. [31] [44] Currently about half of the carbon dioxide released from the burning of fossil fuels is not absorbed by vegetation and the oceans and remains in the atmosphere. [45]

Burning fossil fuels such as coal, petroleum, and natural gas is the leading cause of increased anthropogenic CO2; deforestation is the second major cause. In 2010, 9.14 gigatonnes of carbon (GtC, equivalent to 33.5 gigatonnes of CO2 or about 4.3 ppm in Earth's atmosphere) were released from fossil fuels and cement production worldwide, compared to 6.15 GtC in 1990. [46] In addition, land use change contributed 0.87 GtC in 2010, compared to 1.45 GtC in 1990. [46] In the period 1751 to 1900, about 12 GtC were released as CO2 to the atmosphere from burning of fossil fuels, whereas from 1901 to 2013 the figure was about 380 GtC. [47]

The International Energy Agency estimates that the top 1% of emitters globally each had carbon footprints of over 50 tonnes of CO2 in 2021, more than 1,000 times greater than those of the bottom 1% of emitters. The global average energy-related carbon footprint is around 4.7 tonnes of CO2 per person. [48]

Roles in natural processes on Earth

Greenhouse effect

Greenhouse gases allow sunlight to pass through the atmosphere, heating the planet, but then absorb and redirect the infrared radiation (heat) the planet emits Climate Change Schematic.svg
Greenhouse gases allow sunlight to pass through the atmosphere, heating the planet, but then absorb and redirect the infrared radiation (heat) the planet emits
CO2 reduces the flux of thermal radiation emitted to space (causing the large dip near 667 cm ), thereby contributing to the greenhouse effect. Spectral Greenhouse Effect.png
CO2 reduces the flux of thermal radiation emitted to space (causing the large dip near 667 cm ), thereby contributing to the greenhouse effect.
Longwave-infrared absorption coefficients of water vapor and carbon dioxide. For wavelengths near 15-microns, CO2 is a much stronger absorber than water vapor. Longwave Absorption Coefficients of H2O and CO2.svg
Longwave-infrared absorption coefficients of water vapor and carbon dioxide. For wavelengths near 15-microns, CO2 is a much stronger absorber than water vapor.

On Earth, carbon dioxide is the most relevant, direct greenhouse gas that is influenced by human activities. Water is responsible for most (about 36–70%) of the total greenhouse effect, and the role of water vapor as a greenhouse gas depends on temperature. Carbon dioxide is often mentioned in the context of its increased influence as a greenhouse gas since the pre-industrial (1750) era. In 2013, the increase in CO2 was estimated to be responsible for 1.82 W m−2 of the 2.63 W m−2 change in radiative forcing on Earth (about 70%). [49]

Earth's natural greenhouse effect makes life as we know it possible, and carbon dioxide in the atmosphere plays a significant role in providing for the relatively high temperature on Earth. The greenhouse effect is a process by which thermal radiation from a planetary atmosphere warms the planet's surface beyond the temperature it would have in the absence of its atmosphere. [50] [51] [52]

The concept of more atmospheric CO2 increasing ground temperature was first published by Svante Arrhenius in 1896. [53] The increased radiative forcing due to increased CO2 in the Earth's atmosphere is based on the physical properties of CO2 and the non-saturated absorption windows where CO2 absorbs outgoing long-wave energy. The increased forcing drives further changes in Earth's energy balance and, over the longer term, in Earth's climate. [18]

Carbon cycle

This diagram of the carbon cycle shows the movement of carbon between land, atmosphere, and oceans in billions of metric tons of carbon per year. Yellow numbers are natural fluxes, red are human contributions, white are stored carbon. Carbon cycle.jpg
This diagram of the carbon cycle shows the movement of carbon between land, atmosphere, and oceans in billions of metric tons of carbon per year. Yellow numbers are natural fluxes, red are human contributions, white are stored carbon.

Atmospheric carbon dioxide plays an integral role in the Earth's carbon cycle whereby CO2 is removed from the atmosphere by some natural processes such as photosynthesis and deposition of carbonates, to form limestones for example, and added back to the atmosphere by other natural processes such as respiration and the acid dissolution of carbonate deposits. There are two broad carbon cycles on Earth: the fast carbon cycle and the slow carbon cycle. The fast carbon cycle refers to movements of carbon between the environment and living things in the biosphere whereas the slow carbon cycle involves the movement of carbon between the atmosphere, oceans, soil, rocks, and volcanism. Both cycles are intrinsically interconnected and atmospheric CO2 facilitates the linkage.

Natural sources of atmospheric CO2 include volcanic outgassing, the combustion of organic matter, wildfires and the respiration processes of living aerobic organisms. Man-made sources of CO2 include the burning of fossil fuels, as well as some industrial processes such as cement making.

Annual CO2 flows from anthropogenic sources (left) into Earth's atmosphere, land, and ocean sinks (right) since year 1960. Units in equivalent gigatonnes carbon per year. Global carbon budget components.png
Annual CO2 flows from anthropogenic sources (left) into Earth's atmosphere, land, and ocean sinks (right) since year 1960. Units in equivalent gigatonnes carbon per year.

Natural sources of CO2 are more or less balanced by natural carbon sinks, in the form of chemical and biological processes which remove CO2 from the atmosphere. For example, the decay of organic material in forests, grasslands, and other land vegetation - including forest fires - results in the release of about 436  gigatonnes of CO2 (containing 119 gigatonnes carbon) every year, while CO2 uptake by new growth on land counteracts these releases, absorbing 451 Gt (123 Gt C). [55] Although much CO2 in the early atmosphere of the young Earth was produced by volcanic activity, modern volcanic activity releases only 130 to 230  megatonnes of CO2 each year. [56]

From the human pre-industrial era to 1940, the terrestrial biosphere represented a net source of atmospheric CO2 (driven largely by land-use changes), but subsequently switched to a net sink with growing fossil carbon emissions. [57]

Oceanic carbon cycle

Air-sea exchange of CO2 CO2 pump hg.svg
Air-sea exchange of CO2

The Earth's oceans contain a large amount of CO2 in the form of bicarbonate and carbonate ions—much more than the amount in the atmosphere. The bicarbonate is produced in reactions between rock, water, and carbon dioxide.

From 1850 until 2022, the ocean has absorbed 26% of total anthropogenic emissions. [13] However, the rate at which the ocean will take it up in the future is less certain. Even if equilibrium is reached, including dissolution of carbonate minerals, the increased concentration of bicarbonate and decreased or unchanged concentration of carbonate ion will give rise to a higher concentration of un-ionized carbonic acid and dissolved CO2. This higher concentration in the seas, along with higher temperatures, would mean a higher equilibrium concentration of CO2 in the air. [58] [59]

Effects of current increase

Direct effects

Physical drivers of global warming that has happened so far. Future global warming potential for long lived drivers like carbon dioxide emissions is not represented. Whiskers on each bar show the possible error range. Physical Drivers of climate change.svg
Physical drivers of global warming that has happened so far. Future global warming potential for long lived drivers like carbon dioxide emissions is not represented. Whiskers on each bar show the possible error range.

Direct effects of increasing CO2 concentrations in the atmosphere include increasing global temperatures, ocean acidification and a CO2 fertilization effect on plants and crops. [60]

Temperature rise on land

Changes in global temperatures over the past century provide evidence for the effects of increasing greenhouse gases. When the climate system reacts to such changes, climate change follows. Measurement of the GST is one of the many lines of evidence supporting the scientific consensus on climate change, which is that humans are causing warming of Earth's climate system.

The global average and combined land and ocean surface temperature, show a warming of 1.09 °C (range: 0.95 to 1.20 °C) from 1850–1900 to 2011–2020, based on multiple independently produced datasets. [61] :5 The trend is faster since the 1970s than in any other 50-year period over at least the last 2000 years. [61] :8

Temperature rise in oceans

It is clear that the ocean is warming as a result of climate change, and this rate of warming is increasing. [62] :9 The global ocean was the warmest it had ever been recorded by humans in 2022. [63] This is determined by the ocean heat content, which exceeded the previous 2021 maximum in 2022. [63] The steady rise in ocean temperatures is an unavoidable result of the Earth's energy imbalance, which is primarily caused by rising levels of greenhouse gases. [63] Between pre-industrial times and the 2011–2020 decade, the ocean's surface has heated between 0.68 and 1.01 °C. [64] :1214

The majority of ocean heat gain occurs in the Southern Ocean. For example, between the 1950s and the 1980s, the temperature of the Antarctic Southern Ocean rose by 0.17 °C (0.31 °F), nearly twice the rate of the global ocean. [65]

Ocean acidification

Ocean acidification means that the average seawater pH value is dropping over time. Mean-seawater-ph.png
Ocean acidification means that the average seawater pH value is dropping over time.

Ocean acidification is the ongoing decrease in the pH of the Earth's ocean. Between 1950 and 2020, the average pH of the ocean surface fell from approximately 8.15 to 8.05. [67] Carbon dioxide emissions from human activities are the primary cause of ocean acidification, with atmospheric carbon dioxide (CO2) levels exceeding 410 ppm (in 2020). CO2 from the atmosphere is absorbed by the oceans. This chemical reaction produces carbonic acid (H2CO3) which dissociates into a bicarbonate ion (HCO3) and a hydrogen ion (H+). The presence of free hydrogen ions (H+) lowers the pH of the ocean, increasing acidity (this does not mean that seawater is acidic yet; it is still alkaline, with a pH higher than 8). Marine calcifying organisms, such as mollusks and corals, are especially vulnerable because they rely on calcium carbonate to build shells and skeletons. [68]

A change in pH by 0.1 represents a 26% increase in hydrogen ion concentration in the world's oceans (the pH scale is logarithmic, so a change of one in pH units is equivalent to a tenfold change in hydrogen ion concentration). Sea-surface pH and carbonate saturation states vary depending on ocean depth and location. Colder and higher latitude waters are capable of absorbing more CO2. This can cause acidity to rise, lowering the pH and carbonate saturation levels in these areas. There are several other factors that influence the atmosphere-ocean CO2 exchange, and thus local ocean acidification. These include ocean currents and upwelling zones, proximity to large continental rivers, sea ice coverage, and atmospheric exchange with nitrogen and sulfur from fossil fuel burning and agriculture. [69] [70] [71]

CO2 fertilization effect

The CO2 fertilization effect or carbon fertilization effect causes an increased rate of photosynthesis while limiting leaf transpiration in plants. Both processes result from increased levels of atmospheric carbon dioxide (CO2). [72] [73] The carbon fertilization effect varies depending on plant species, air and soil temperature, and availability of water and nutrients. [74] [75] Net primary productivity (NPP) might positively respond to the carbon fertilization effect. [76] Although, evidence shows that enhanced rates of photosynthesis in plants due to CO2 fertilization do not directly enhance all plant growth, and thus carbon storage. [74] The carbon fertilization effect has been reported to be the cause of 44% of gross primary productivity (GPP) increase since the 2000s. [77] Earth System Models, Land System Models and Dynamic Global Vegetation Models are used to investigate and interpret vegetation trends related to increasing levels of atmospheric CO2. [74] [78] However, the ecosystem processes associated with the CO2 fertilization effect remain uncertain and therefore are challenging to model. [79] [80]

Terrestrial ecosystems have reduced atmospheric CO2 concentrations and have partially mitigated climate change effects. [81] The response by plants to the carbon fertilization effect is unlikely to significantly reduce atmospheric CO2 concentration over the next century due to the increasing anthropogenic influences on atmospheric CO2. [73] [74] [82] [83] Earth's vegetated lands have shown significant greening since the early 1980s [84] largely due to rising levels of atmospheric CO2. [85] [86] [87] [88]

Theory predicts the tropics to have the largest uptake due to the carbon fertilization effect, but this has not been observed. The amount of CO2 uptake from CO2 fertilization also depends on how forests respond to climate change, and if they are protected from deforestation. [89]

Other direct effects

CO2 emissions have also led to the stratosphere contracting by 400 meters since 1980, which could affect satellite operations, GPS systems and radio communications. [90]

Indirect effects and impacts

062821Yreka Fire CalFire -2wiki.jpg
Bleachedcoral.jpg
Village Telly in Mali.jpg
US Navy 071120-M-8966H-005 An aerial view over southern Bangladesh reveals extensive flooding as a result of Cyclone Sidr.jpg
Some climate change effects, clockwise from top left: Wildfire caused by heat and dryness, bleached coral caused by ocean acidification and heating, coastal flooding caused by storms and sea level rise, and environmental migration caused by desertification
Effects of climate change are well documented and growing for Earth's natural environment and human societies. Changes to the climate system include an overall warming trend, changes to precipitation patterns, and more extreme weather. As the climate changes it impacts the natural environment with effects such as more intense forest fires, thawing permafrost, and desertification. These changes impact ecosystems and societies, and can become irreversible once tipping points are crossed. Climate activists are engaged in a range of activities around the world that seek to ameliorate these issues or prevent them from happening. [91]
Overview of climatic changes and their effects on the ocean. Regional effects are displayed in italics. Overview of climatic changes and their effects on the ocean.png
Overview of climatic changes and their effects on the ocean. Regional effects are displayed in italics.
There are many effects of climate change on oceans. One of the main ones is an increase in ocean temperatures. More frequent marine heatwaves are linked to this. The rising temperature contributes to a rise in sea levels due to melting ice sheets. Other effects on oceans include sea ice decline, reducing pH values and oxygen levels, as well as increased ocean stratification. All this can lead to changes of ocean currents, for example a weakening of the Atlantic meridional overturning circulation (AMOC). [62] The main root cause of these changes are the emissions of greenhouse gases from human activities, mainly burning of fossil fuels. Carbon dioxide and methane are examples of greenhouse gases. The additional greenhouse effect leads to ocean warming because the ocean takes up most of the additional heat in the climate system. [93] The ocean also absorbs some of the extra carbon dioxide that is in the atmosphere. This causes the pH value of the seawater to drop. [94] Scientists estimate that the ocean absorbs about 25% of all human-caused CO2 emissions. [94]

Approaches for reducing CO2 concentrations

A model of the behavior of carbon in the atmosphere from 1 September 2014 to 31 August 2015. The height of Earth's atmosphere and topography have been vertically exaggerated and appear approximately 40 times higher than normal to show the complexity of the atmospheric flow.

Carbon dioxide has unique long-term effects on climate change that are nearly "irreversible" for a thousand years after emissions stop (zero further emissions). The greenhouse gases methane and nitrous oxide do not persist over time in the same way as carbon dioxide. Even if human carbon dioxide emissions were to completely cease, atmospheric temperatures are not expected to decrease significantly in the short term. This is because the air temperature is determined by a balance between heating, due to greenhouse gases, and cooling due to heat transfer to the ocean. If emissions were to stop, CO2 levels and the heating effect would slowly decrease, but simultaneously the cooling due to heat transfer would diminish (because sea temperatures would get closer to the air temperature), with the result that the air temperature would decrease only slowly. Sea temperatures would continue to rise, causing thermal expansion and some sea level rise. [58] Lowering global temperatures more rapidly would require carbon sequestration or geoengineering.

Various techniques have been proposed for removing excess carbon dioxide from the atmosphere.

Carbon dioxide removal (CDR) is a process in which carbon dioxide (CO2) is removed from the atmosphere by deliberate human activities and durably stored in geological, terrestrial, or ocean reservoirs, or in products. [95] :2221 This process is also known as carbon removal, greenhouse gas removal or negative emissions. CDR is more and more often integrated into climate policy, as an element of climate change mitigation strategies. [96] [97] Achieving net zero emissions will require first and foremost deep and sustained cuts in emissions, and then—in addition—the use of CDR ("CDR is what puts the net into net zero emissions" [98] ). In the future, CDR may be able to counterbalance emissions that are technically difficult to eliminate, such as some agricultural and industrial emissions. [99] :114

Concentrations in the geologic past

CO2 concentrations over the last 500 Million years Phanerozoic Carbon Dioxide.png
CO2 concentrations over the last 500 Million years
Concentration of atmospheric CO2 over the last 40,000 years, from the Last Glacial Maximum to the present day. The current rate of increase is much higher than at any point during the last deglaciation. CO2 40k.png
Concentration of atmospheric CO2 over the last 40,000 years, from the Last Glacial Maximum to the present day. The current rate of increase is much higher than at any point during the last deglaciation.

Estimates in 2023 found that the current carbon dioxide concentration in the atmosphere may be the highest it has been in the last 14 million years. [11] However the IPCC Sixth Assessment Report estimated similar levels 3 to 3.3 million years ago in the mid-Pliocene warm period. This period can be a proxy for likely climate outcomes with current levels of CO2. [100] :Figure 2.34

Carbon dioxide is believed to have played an important effect in regulating Earth's temperature throughout its 4.54 billion year history. Early in the Earth's life, scientists have found evidence of liquid water indicating a warm world even though the Sun's output is believed to have only been 70% of what it is today. Higher carbon dioxide concentrations in the early Earth's atmosphere might help explain this faint young sun paradox. When Earth first formed, Earth's atmosphere may have contained more greenhouse gases and CO2 concentrations may have been higher, with estimated partial pressure as large as 1,000  kPa (10  bar ), because there was no bacterial photosynthesis to reduce the gas to carbon compounds and oxygen. Methane, a very active greenhouse gas, may have been more prevalent as well. [101] [102]

Carbon dioxide concentrations have shown several cycles of variation from about 180 parts per million during the deep glaciations of the Holocene and Pleistocene to 280 parts per million during the interglacial periods. Carbon dioxide concentrations have varied widely over the Earth's history. It is believed to have been present in Earth's first atmosphere, shortly after Earth's formation. The second atmosphere, consisting largely of nitrogen and CO
2
was produced by outgassing from volcanism, supplemented by gases produced during the late heavy bombardment of Earth by huge asteroids. [103] A major part of carbon dioxide emissions were soon dissolved in water and incorporated in carbonate sediments.

The production of free oxygen by cyanobacterial photosynthesis eventually led to the oxygen catastrophe that ended Earth's second atmosphere and brought about the Earth's third atmosphere (the modern atmosphere) 2.4 billion years ago. Carbon dioxide concentrations dropped from 4,000 parts per million during the Cambrian period about 500 million years ago to as low as 180 parts per million 20,000 years ago . [3]

Drivers of ancient-Earth CO2 concentration

On long timescales, atmospheric CO2 concentration is determined by the balance among geochemical processes including organic carbon burial in sediments, silicate rock weathering, and volcanic degassing. The net effect of slight imbalances in the carbon cycle over tens to hundreds of millions of years has been to reduce atmospheric CO2. On a timescale of billions of years, such downward trend appears bound to continue indefinitely as occasional massive historical releases of buried carbon due to volcanism will become less frequent (as earth mantle cooling and progressive exhaustion of internal radioactive heat proceed further). The rates of these processes are extremely slow; hence they are of no relevance to the atmospheric CO2 concentration over the next hundreds or thousands of years.

Photosynthesis in the geologic past

Over the course of Earth's geologic history CO2 concentrations have played a role in biological evolution. The first photosynthetic organisms probably evolved early in the evolutionary history of life and most likely used reducing agents such as hydrogen or hydrogen sulfide as sources of electrons, rather than water. [104] Cyanobacteria appeared later, and the excess oxygen they produced contributed to the oxygen catastrophe, [105] which rendered the evolution of complex life possible. In recent geologic times, low CO2 concentrations below 600 parts per million might have been the stimulus that favored the evolution of C4 plants which increased greatly in abundance between 7 and 5 million years ago over plants that use the less efficient C3 metabolic pathway. [106] At current atmospheric pressures photosynthesis shuts down when atmospheric CO2 concentrations fall below 150 ppm and 200 ppm although some microbes can extract carbon from the air at much lower concentrations. [107] [108]

Measuring ancient-Earth CO2 concentration

Over 400,000 years of ice core data: Graph of CO2 (green), reconstructed temperature (blue) and dust (red) from the Vostok ice core Vostok Petit data.svg
Over 400,000 years of ice core data: Graph of CO2 (green), reconstructed temperature (blue) and dust (red) from the Vostok ice core
Correspondence between temperature and atmospheric CO2 during the last 800,000 years Temperature-change-and-carbon-dioxide-change-measured-from-the-EPICA-Dome-C-ice-core-in-Antarctica-v2.jpg
Correspondence between temperature and atmospheric CO2 during the last 800,000 years

The most direct method for measuring atmospheric carbon dioxide concentrations for periods before instrumental sampling is to measure bubbles of air (fluid or gas inclusions) trapped in the Antarctic or Greenland ice sheets. The most widely accepted of such studies come from a variety of Antarctic cores and indicate that atmospheric CO2 concentrations were about 260–280 ppm immediately before industrial emissions began and did not vary much from this level during the preceding 10,000 years. [109] [110] The longest ice core record comes from East Antarctica, where ice has been sampled to an age of 800,000 years. [111] During this time, the atmospheric carbon dioxide concentration has varied between 180 and 210 ppm during ice ages, increasing to 280–300 ppm during warmer interglacials. [112] [113]

CO2 mole fractions in the atmosphere have gone up by around 35 percent since the 1900s, rising from 280 parts per million by volume to 387 parts per million in 2009. One study using evidence from stomata of fossilized leaves suggests greater variability, with CO2 mole fractions above 300 ppm during the period ten to seven thousand years ago, [114] though others have argued that these findings more likely reflect calibration or contamination problems rather than actual CO2 variability. [115] [116] Because of the way air is trapped in ice (pores in the ice close off slowly to form bubbles deep within the firn) and the time period represented in each ice sample analyzed, these figures represent averages of atmospheric concentrations of up to a few centuries rather than annual or decadal levels.

Ice cores provide evidence for greenhouse gas concentration variations over the past 800,000 years. Both CO2 and CH
4
concentrations vary between glacial and interglacial phases, and these variations correlate strongly with temperature. Direct data does not exist for periods earlier than those represented in the ice core record, a record that indicates that CO2 mole fractions stayed within a range of 180 ppm to 280 ppm throughout the last 800,000 years, until the increase of the last 250 years. However, various proxy measurements and models suggest larger variations in past epochs: 500 million years ago CO2 levels were likely 10 times higher than now. [117]

Various proxy measurements have been used to try to determine atmospheric CO2 concentrations millions of years in the past. These include boron and carbon isotope ratios in certain types of marine sediments, and the numbers of stomata observed on fossil plant leaves. [106]

Phytane is a type of diterpenoid alkane. It is a breakdown product of chlorophyll, and is now used to estimate ancient CO2 levels. [118] Phytane gives both a continuous record of CO2 concentrations but it also can overlap a break in the CO2 record of over 500 million years. [118]

600 to 400 million years ago

There is evidence for high CO2 concentrations of over 6,000 ppm between 600 and 400 million years ago, and of over 3,000 ppm between 200 and 150 million years ago. [119] [ failed verification ]

Indeed, higher CO2 concentrations are thought to have prevailed throughout most of the Phanerozoic Eon, with concentrations four to six times current concentrations during the Mesozoic era, and ten to fifteen times current concentrations during the early Palaeozoic era until the middle of the Devonian period, about 400 million years ago. [120] [121] [122] The spread of land plants is thought to have reduced CO2 concentrations during the late Devonian, and plant activities as both sources and sinks of CO2 have since been important in providing stabilizing feedbacks. [123]

Earlier still, a 200-million year period of intermittent, widespread glaciation extending close to the equator (Snowball Earth) appears to have been ended suddenly, about 550 Ma, by a colossal volcanic outgassing that raised the CO2 concentration of the atmosphere abruptly to 12%, about 350 times modern levels, causing extreme greenhouse conditions and carbonate deposition as limestone at the rate of about 1 mm per day. [124] This episode marked the close of the Precambrian Eon, and was succeeded by the generally warmer conditions of the Phanerozoic, during which multicellular animal and plant life evolved. No volcanic CO2 emission of comparable scale has occurred since. In the modern era, emissions to the atmosphere from volcanoes are approximately 0.645 billion tons of CO2 per year, whereas humans contribute 29 billion tons of CO2 each year. [125] [124] [126] [127]

60 to 5 million years ago

Atmospheric CO2 concentration continued to fall after about 60 million years ago. About 34 million years ago, the time of the Eocene–Oligocene extinction event and when the Antarctic ice sheet started to take its current form, CO2 was about 760 ppm, [128] and there is geochemical evidence that concentrations were less than 300 ppm by about 20 million years ago. Decreasing CO2 concentration, with a tipping point of 600 ppm, was the primary agent forcing Antarctic glaciation. [129] Low CO2 concentrations may have been the stimulus that favored the evolution of C4 plants, which increased greatly in abundance between 7 and 5 million years ago. [106]

See also

Related Research Articles

<span class="mw-page-title-main">Causes of climate change</span> Effort to scientifically ascertain mechanisms responsible for recent global warming

The scientific community has been investigating the causes of climate change for decades. After thousands of studies, it came to a consensus, where it is "unequivocal that human influence has warmed the atmosphere, ocean and land since pre-industrial times." This consensus is supported by around 200 scientific organizations worldwide, The dominant role in this climate change has been played by the direct emissions of carbon dioxide from the burning of fossil fuels. Indirect CO2 emissions from land use change, and the emissions of methane, nitrous oxide and other greenhouse gases play major supporting roles.

<span class="mw-page-title-main">Carbon dioxide</span> Chemical compound with formula CO₂

Carbon dioxide is a chemical compound with the chemical formula CO2. It is made up of molecules that each have one carbon atom covalently double bonded to two oxygen atoms. It is found in the gas state at room temperature, and as the source of available carbon in the carbon cycle, atmospheric CO2 is the primary carbon source for life on Earth. In the air, carbon dioxide is transparent to visible light but absorbs infrared radiation, acting as a greenhouse gas. Carbon dioxide is soluble in water and is found in groundwater, lakes, ice caps, and seawater. When carbon dioxide dissolves in water, it forms carbonate and mainly bicarbonate, which causes ocean acidification as atmospheric CO2 levels increase.

<span class="mw-page-title-main">Greenhouse effect</span> Atmospheric phenomenon causing planetary warming

The greenhouse effect occurs when greenhouse gases in a planet's atmosphere insulate the planet from losing heat to space, raising its surface temperature. Surface heating can happen from an internal heat source as in the case of Jupiter, or from its host star as in the case of the Earth. In the case of Earth, the Sun emits shortwave radiation (sunlight) that passes through greenhouse gases to heat the Earth's surface. In response, the Earth's surface emits longwave radiation that is mostly absorbed by greenhouse gases. The absorption of longwave radiation prevents it from reaching space, reducing the rate at which the Earth can cool off.

<span class="mw-page-title-main">Global warming potential</span> Potential heat absorbed by a greenhouse gas

Global warming potential (GWP) is an index to measure how much infrared thermal radiation a greenhouse gas would absorb over a given time frame after it has been added to the atmosphere. The GWP makes different greenhouse gases comparable with regard to their "effectiveness in causing radiative forcing". It is expressed as a multiple of the radiation that would be absorbed by the same mass of added carbon dioxide, which is taken as a reference gas. Therefore, the GWP has a value of 1 for CO2. For other gases it depends on how strongly the gas absorbs infrared thermal radiation, how quickly the gas leaves the atmosphere, and the time frame being considered.

<span class="mw-page-title-main">Carbon cycle</span> Natural processes of carbon exchange

The carbon cycle is that part of the biogeochemical cycle by which carbon is exchanged among the biosphere, pedosphere, geosphere, hydrosphere, and atmosphere of Earth. Other major biogeochemical cycles include the nitrogen cycle and the water cycle. Carbon is the main component of biological compounds as well as a major component of many minerals such as limestone. The carbon cycle comprises a sequence of events that are key to making Earth capable of sustaining life. It describes the movement of carbon as it is recycled and reused throughout the biosphere, as well as long-term processes of carbon sequestration (storage) to and release from carbon sinks.

<span class="mw-page-title-main">Radiative forcing</span> Concept in climate science on solar irradiance

Radiative forcing is a concept used in climate science to quantify the change in energy balance in Earth's atmosphere. Various factors contribute to this change in energy balance, such as concentrations of greenhouse gases and aerosols, and changes in surface albedo and solar irradiance. In more technical terms, it is defined as "the change in the net, downward minus upward, radiative flux due to a change in an external driver of climate change." These external drivers are distinguished from feedbacks and variability that are internal to the climate system, and that further influence the direction and magnitude of imbalance. Radiative forcing on Earth is meaningfully evaluated at the tropopause and at the top of the stratosphere. It is quantified in units of watts per square meter, and often summarized as an average over the total surface area of the globe.

<span class="mw-page-title-main">Keeling Curve</span> Graph of atmospheric CO2 from 1958 to the present

The Keeling Curve is a graph of the accumulation of carbon dioxide in the Earth's atmosphere based on continuous measurements taken at the Mauna Loa Observatory on the island of Hawaii from 1958 to the present day. The curve is named for the scientist Charles David Keeling, who started the monitoring program and supervised it until his death in 2005.

Trace gases are gases that are present in small amounts within an environment such as a planet's atmosphere. Trace gases in Earth's atmosphere are gases other than nitrogen (78.1%), oxygen (20.9%), and argon (0.934%) which, in combination, make up 99.934% of its atmosphere.

<span class="mw-page-title-main">Climate sensitivity</span> Concept in climate science

Climate sensitivity is a key measure in climate science and describes how much Earth's surface will warm for a doubling in the atmospheric carbon dioxide (CO2) concentration. Its formal definition is: "The change in the surface temperature in response to a change in the atmospheric carbon dioxide (CO2) concentration or other radiative forcing." This concept helps scientists understand the extent and magnitude of the effects of climate change.

<span class="mw-page-title-main">Charles David Keeling</span> American scientist (1928-2005)

Charles David Keeling was an American scientist whose recording of carbon dioxide at the Mauna Loa Observatory confirmed Svante Arrhenius's proposition (1896) of the possibility of anthropogenic contribution to the greenhouse effect and global warming, by documenting the steadily rising carbon dioxide levels. The Keeling Curve measures the progressive buildup of carbon dioxide, a greenhouse gas, in the atmosphere.

Throughout Earth's climate history (Paleoclimate) its climate has fluctuated between two primary states: greenhouse and icehouse Earth. Both climate states last for millions of years and should not be confused with the much smaller glacial and interglacial periods, which occur as alternating phases within an icehouse period and tend to last less than 1 million years. There are five known icehouse periods in Earth's climate history, namely the Huronian, Cryogenian, Andean-Saharan, Late Paleozoic and Late Cenozoic glaciations.

<span class="mw-page-title-main">Carbonate–silicate cycle</span> Geochemical transformation of silicate rocks

The carbonate–silicate geochemical cycle, also known as the inorganic carbon cycle, describes the long-term transformation of silicate rocks to carbonate rocks by weathering and sedimentation, and the transformation of carbonate rocks back into silicate rocks by metamorphism and volcanism. Carbon dioxide is removed from the atmosphere during burial of weathered minerals and returned to the atmosphere through volcanism. On million-year time scales, the carbonate-silicate cycle is a key factor in controlling Earth's climate because it regulates carbon dioxide levels and therefore global temperature.

<span class="mw-page-title-main">Carbon dioxide removal</span> Removal of atmospheric carbon dioxide through human activity

Carbon dioxide removal (CDR) is a process in which carbon dioxide is removed from the atmosphere by deliberate human activities and durably stored in geological, terrestrial, or ocean reservoirs, or in products. This process is also known as carbon removal, greenhouse gas removal or negative emissions. CDR is more and more often integrated into climate policy, as an element of climate change mitigation strategies. Achieving net zero emissions will require first and foremost deep and sustained cuts in emissions, and then—in addition—the use of CDR. In the future, CDR may be able to counterbalance emissions that are technically difficult to eliminate, such as some agricultural and industrial emissions.

<span class="mw-page-title-main">Greenhouse gas</span> Gas in an atmosphere with certain absorption characteristics

Greenhouse gases (GHGs) are the gases in the atmosphere that raise the surface temperature of planets such as the Earth. What distinguishes them from other gases is that they absorb the wavelengths of radiation that a planet emits, resulting in the greenhouse effect. The Earth is warmed by sunlight, causing its surface to radiate heat, which is then mostly absorbed by greenhouse gases. Without greenhouse gases in the atmosphere, the average temperature of Earth's surface would be about −18 °C (0 °F), rather than the present average of 15 °C (59 °F).

<span class="mw-page-title-main">Atmospheric methane</span> Methane in Earths atmosphere

Atmospheric methane is the methane present in Earth's atmosphere. The concentration of atmospheric methane is increasing due to methane emissions, and is causing climate change. Methane is one of the most potent greenhouse gases. Methane's radiative forcing (RF) of climate is direct, and it is the second largest contributor to human-caused climate forcing in the historical period. Methane is a major source of water vapour in the stratosphere through oxidation; and water vapour adds about 15% to methane's radiative forcing effect. The global warming potential (GWP) for methane is about 84 in terms of its impact over a 20-year timeframe, and 28 in terms of its impact over a 100-year timeframe.

<span class="mw-page-title-main">Climate change feedbacks</span> Feedback related to climate change

Climate change feedbacks are natural processes that impact how much global temperatures will increase for a given amount of greenhouse gas emissions. Positive feedbacks amplify global warming while negative feedbacks diminish it. Feedbacks influence both the amount of greenhouse gases in the atmosphere and the amount of temperature change that happens in response. While emissions are the forcing that causes climate change, feedbacks combine to control climate sensitivity to that forcing.

<span class="mw-page-title-main">Atmospheric carbon cycle</span> Transformation of atmospheric carbon between various forms

The atmospheric carbon cycle accounts for the exchange of gaseous carbon compounds, primarily carbon dioxide, between Earth's atmosphere, the oceans, and the terrestrial biosphere. It is one of the faster components of the planet's overall carbon cycle, supporting the exchange of more than 200 billion tons of carbon in and out of the atmosphere throughout the course of each year. Atmospheric concentrations of CO2 remain stable over longer timescales only when there exists a balance between these two flows. Methane, Carbon monoxide (CO), and other human-made compounds are present in smaller concentrations and are also part of the atmospheric carbon cycle.

<span class="mw-page-title-main">Oceanic carbon cycle</span> Ocean/atmosphere carbon exchange process

The oceanic carbon cycle is composed of processes that exchange carbon between various pools within the ocean as well as between the atmosphere, Earth interior, and the seafloor. The carbon cycle is a result of many interacting forces across multiple time and space scales that circulates carbon around the planet, ensuring that carbon is available globally. The Oceanic carbon cycle is a central process to the global carbon cycle and contains both inorganic carbon and organic carbon. Part of the marine carbon cycle transforms carbon between non-living and living matter.

CO<sub>2</sub> fertilization effect Fertilization from increased levels of atmospheric carbon dioxide

The CO2 fertilization effect or carbon fertilization effect causes an increased rate of photosynthesis while limiting leaf transpiration in plants. Both processes result from increased levels of atmospheric carbon dioxide (CO2). The carbon fertilization effect varies depending on plant species, air and soil temperature, and availability of water and nutrients. Net primary productivity (NPP) might positively respond to the carbon fertilization effect. Although, evidence shows that enhanced rates of photosynthesis in plants due to CO2 fertilization do not directly enhance all plant growth, and thus carbon storage. The carbon fertilization effect has been reported to be the cause of 44% of gross primary productivity (GPP) increase since the 2000s. Earth System Models, Land System Models and Dynamic Global Vegetation Models are used to investigate and interpret vegetation trends related to increasing levels of atmospheric CO2. However, the ecosystem processes associated with the CO2 fertilization effect remain uncertain and therefore are challenging to model.

<span class="mw-page-title-main">Climate restoration</span>

Climate restoration is the climate change goal and associated actions to restore CO2 to levels humans have actually survived long-term, below 300 ppm. This would restore the Earth system generally to a safe state, for the well-being of future generations of humanity and nature. Actions include carbon dioxide removal from the Carbon dioxide in Earth's atmosphere, which, in combination with emissions reductions, would reduce the level of CO2 in the atmosphere and thereby reduce the global warming produced by the greenhouse effect of an excess of CO2 over its pre-industrial level. Actions also include restoring pre-industrial atmospheric methane levels by accelerating natural methane oxidation.

References

  1. Gavin, Schmidt (2010), Taking the Measure of the Greenhouse Effect , retrieved 24 August 2024
  2. 1 2 3 4 "Carbon dioxide now more than 50% higher than pre-industrial levels". National Oceanic and Atmospheric Administration. 3 June 2022. Archived from the original on 5 June 2022. Retrieved 14 June 2022.
  3. 1 2 3 4 Eggleton, Tony (2013). A Short Introduction to Climate Change. Cambridge University Press. p. 52. ISBN   9781107618763. Archived from the original on 14 March 2023. Retrieved 14 March 2023.
  4. 1 2 3 "The NOAA Annual Greenhouse Gas Index (AGGI) – An Introduction". NOAA Global Monitoring Laboratory/Earth System Research Laboratories. Archived from the original on 27 November 2020. Retrieved 18 December 2020.
  5. Etheridge, D.M.; L.P. Steele; R.L. Langenfelds; R.J. Francey; J.-M. Barnola; V.I. Morgan (1996). "Natural and anthropogenic changes in atmospheric CO2 over the last 1000 years from air in Antarctic ice and firn". Journal of Geophysical Research. 101 (D2): 4115–28. Bibcode:1996JGR...101.4115E. doi:10.1029/95JD03410. ISSN   0148-0227. S2CID   19674607.
  6. Pierre-Louis, Kendra (10 May 2024). "Carbon Dioxide Just Took an Ominous, Record-Breaking Jump". www.bloomberg.com. Retrieved 13 May 2024.
  7. IPCC (2022) Summary for policy makers Archived 12 March 2023 at the Wayback Machine in Climate Change 2022: Mitigation of Climate Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Archived 2 August 2022 at the Wayback Machine , Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA
  8. Petty, G.W. (2004). "A First Course in Atmospheric Radiation". Eos Transactions. 85 (36): 229–51. Bibcode:2004EOSTr..85..341P. doi: 10.1029/2004EO360007 .
  9. Atkins, P.; de Paula, J. (2006). Atkins' Physical Chemistry (8th ed.). W.H. Freeman. p.  462. ISBN   978-0-7167-8759-4.
  10. "Carbon Dioxide Absorbs and Re-emits Infrared Radiation". UCAR Center for Science Education. 2012. Archived from the original on 21 September 2017. Retrieved 9 September 2017.
  11. 1 2 Ahmed, Issam. "Current carbon dioxide levels last seen 14 million years ago". phys.org. Retrieved 8 February 2024.
  12. "Climate and CO2 in the Atmosphere". Archived from the original on 6 October 2018. Retrieved 10 October 2007.
  13. 1 2 3 Friedlingstein, Pierre; O'Sullivan, Michael; Jones, Matthew W.; Andrew, Robbie M.; Gregor, Luke; Hauck, Judith; Le Quéré, Corinne; Luijkx, Ingrid T.; Olsen, Are; Peters, Glen P.; Peters, Wouter; Pongratz, Julia; Schwingshackl, Clemens; Sitch, Stephen; Canadell, Josep G. (11 November 2022). "Global Carbon Budget 2022". Earth System Science Data. 14 (11): 4811–4900. Bibcode:2022ESSD...14.4811F. doi: 10.5194/essd-14-4811-2022 . hdl: 20.500.11850/594889 . Creative Commons by small.svg  This article incorporates textfrom this source, which is available under the CC BY 4.0 license.
  14. "Parts per million" refers to the number of carbon dioxide molecules per million molecules of dry air. "Carbon Dioxide LATEST MEASUREMENT". Climate Change: Vital Signs of the Planet. NASA Global Climate Change. Archived from the original on 17 April 2022. Updated monthly.
  15. "Global Monitoring Laboratory - Trends in Atmospheric Carbon Dioxide". National Oceanic & Atmospheric Administration. Latest figure, and graphs of trend; frequently updated
  16. "Table of atmospheric CO₂ since 1958, updated monthly". National Oceanic & Atmospheric Administration. The actual figures fluctuate month-by-month throughout the year, so figures for the same month of different years should be compared, or a seasonally corrected figure used.
  17. "Conversion Tables". Carbon Dioxide Information Analysis Center. Oak Ridge National Laboratory. 18 July 2020. Archived from the original on 27 September 2017. Retrieved 18 July 2020. Alt URL Archived 23 February 2016 at the Wayback Machine
  18. 1 2 3 Eyring, V., N.P. Gillett, K.M. Achuta Rao, R. Barimalala, M. Barreiro Parrillo, N. Bellouin, C. Cassou, P.J. Durack, Y. Kosaka, S.  McGregor, S. Min, O. Morgenstern, and Y. Sun, 2021: Chapter 3: Human Influence on the Climate System Archived 7 March 2023 at the Wayback Machine . In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Archived 9 August 2021 at the Wayback Machine [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L.  Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 423–552, doi : 10.1017/9781009157896.005
  19. "The World Counts". The World Counts. Retrieved 4 December 2023.
  20. Archer D (2009). "Atmospheric lifetime of fossil fuel carbon dioxide". Annual Review of Earth and Planetary Sciences. 37 (1): 117–34. Bibcode:2009AREPS..37..117A. doi:10.1146/annurev.earth.031208.100206. hdl:2268/12933. Archived from the original on 24 February 2021. Retrieved 7 March 2021.
  21. Joos F, Roth R, Fuglestvedt JS, Peters GP, Enting IG, Von Bloh W, et al. (2013). "Carbon dioxide and climate impulse response functions for the computation of greenhouse gas metrics: A multi-model analysis". Atmospheric Chemistry and Physics. 13 (5): 2793–2825. doi: 10.5194/acpd-12-19799-2012 . hdl: 20.500.11850/58316 . Archived from the original on 22 July 2020. Retrieved 7 March 2021.
  22. Rasmussen, Carl Edward. "Atmospheric Carbon Dioxide Growth Rate". Archived from the original on 14 March 2023. Retrieved 14 March 2023.
  23. "Frequently Asked Questions". Carbon Dioxide Information Analysis Center (CDIAC). Archived from the original on 17 August 2011. Retrieved 13 June 2007.
  24. George K, Ziska LH, Bunce JA, Quebedeaux B (2007). "Elevated atmospheric CO2 concentration and temperature across an urban–rural transect". Atmospheric Environment. 41 (35): 7654–7665. Bibcode:2007AtmEn..41.7654G. doi:10.1016/j.atmosenv.2007.08.018. Archived from the original on 15 October 2019. Retrieved 12 September 2019.
  25. "Carbon Budget 2009 Highlights". globalcarbonproject.org. Archived from the original on 16 December 2011. Retrieved 2 November 2012.
  26. "Carbon dioxide passes symbolic mark". BBC. 10 May 2013. Archived from the original on 23 May 2019. Retrieved 10 May 2013.
  27. "Up-to-date weekly average CO2 at Mauna Loa". NOAA. Archived from the original on 24 May 2019. Retrieved 1 June 2019.
  28. "Greenhouse gas levels pass symbolic 400ppm CO2 milestone". The Guardian . Associated Press. 1 June 2012. Archived from the original on 22 January 2014. Retrieved 11 May 2013.
  29. Kunzig, Robert (9 May 2013). "Climate Milestone: Earth's CO2 Level Passes 400 ppm". National Geographic . Archived from the original on 15 December 2013. Retrieved 12 May 2013.
  30. 1 2 "Trends in Atmospheric Carbon Dioxide". Earth System Research Laboratories. NOAA. Archived from the original on 25 January 2013. Retrieved 14 March 2023.
  31. "The Early Keeling Curve | Scripps CO2 Program". scrippsco2.ucsd.edu. Archived from the original on 8 October 2022. Retrieved 14 March 2023.
  32. "NOAA CCGG page Retrieved 2 March 2016". Archived from the original on 11 August 2011. Retrieved 14 March 2023.
  33. WDCGG webpage Archived 6 April 2016 at the Wayback Machine Retrieved 2 March 2016
  34. RAMCES webpage [ permanent dead link ] Retrieved 2 March 2016
  35. "CDIAC CO2 page Retrieved 9 February 2016". Archived from the original on 13 August 2011. Retrieved 14 March 2023.
  36. "GLOBALVIEW-CO2 information page. Retrieved 9 February 2016". Archived from the original on 31 January 2020. Retrieved 14 March 2023.
  37. e.g. Gosh, Prosenjit; Brand, Willi A. (2003). "Stable isotope ratio mass spectrometry in global climate change research" (PDF). International Journal of Mass Spectrometry . 228 (1): 1–33. Bibcode:2003IJMSp.228....1G. CiteSeerX   10.1.1.173.2083 . doi:10.1016/S1387-3806(03)00289-6. Archived (PDF) from the original on 11 August 2017. Retrieved 2 July 2012. Global change issues have become significant due to the sustained rise in atmospheric trace gas concentrations (CO2, N
    2
    O
    , CH
    4
    ) over recent years, attributable to the increased per capita energy consumption of a growing global population.
  38. Keeling, Charles D.; Piper, Stephen C.; Whorf, Timothy P.; Keeling, Ralph F. (2011). "Evolution of natural and anthropogenic fluxes of atmospheric CO2 from 1957 to 2003". Tellus B. 63 (1): 1–22. Bibcode:2011TellB..63....1K. doi: 10.1111/j.1600-0889.2010.00507.x . ISSN   0280-6509.
  39. Bender, Michael L.; Ho, David T.; Hendricks, Melissa B.; Mika, Robert; Battle, Mark O.; Tans, Pieter P.; Conway, Thomas J.; Sturtevant, Blake; Cassar, Nicolas (2005). "Atmospheric O2/N2changes, 1993–2002: Implications for the partitioning of fossil fuel CO2sequestration". Global Biogeochemical Cycles. 19 (4): n/a. Bibcode:2005GBioC..19.4017B. doi: 10.1029/2004GB002410 . ISSN   0886-6236.
  40. Evans, Simon (5 October 2021). "Analysis: Which countries are historically responsible for climate change? / Historical responsibility for climate change is at the heart of debates over climate justice". CarbonBrief.org. Carbon Brief. Archived from the original on 26 October 2021. Source: Carbon Brief analysis of figures from the Global Carbon Project, CDIAC, Our World in Data, Carbon Monitor, Houghton and Nassikas (2017) and Hansis et al (2015).
  41. Ballantyne, A.P.; Alden, C.B.; Miller, J.B.; Tans, P.P.; White, J.W.C. (2012). "Increase in observed net carbon dioxide uptake by land and oceans during the past 50 years". Nature. 488 (7409): 70–72. Bibcode:2012Natur.488...70B. doi:10.1038/nature11299. ISSN   0028-0836. PMID   22859203. S2CID   4335259.
  42. 1 2 Friedlingstein, P., Jones, M., O'Sullivan, M., Andrew, R., Hauck, J., Peters, G., Peters, W., Pongratz, J., Sitch, S., Le Quéré, C. and 66 others (2019) "Global carbon budget 2019". Earth System Science Data, 11(4): 1783–1838. doi : 10.5194/essd-11-1783-2019. CC-BY icon.svg Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  43. Dlugokencky, E. (5 February 2016). "Annual Mean Carbon Dioxide Data". Earth System Research Laboratories. NOAA. Archived from the original on 14 March 2023. Retrieved 12 February 2016.
  44. A.P. Ballantyne; C.B. Alden; J.B. Miller; P.P. Tans; J.W. C. White (2012). "Increase in observed net carbon dioxide uptake by land and oceans during the past 50 years". Nature . 488 (7409): 70–72. Bibcode:2012Natur.488...70B. doi:10.1038/nature11299. PMID   22859203. S2CID   4335259.
  45. 1 2 "Global carbon budget 2010 (summary)". Tyndall Centre for Climate Change Research. Archived from the original on 23 July 2012.
  46. Calculated from file global.1751_2013.csv in Archived 22 October 2011 at the Wayback Machine from the Carbon Dioxide Information Analysis Center.
  47. IEA (2023), The world’s top 1% of emitters produce over 1000 times more CO2 than the bottom 1%, IEA, Paris https://www.iea.org/commentaries/the-world-s-top-1-of-emitters-produce-over-1000-times-more-co2-than-the-bottom-1 , License: CC BY 4.0
  48. "IPCC Fifth Assessment Report – Chapter 8: Anthropogenic and Natural Radiative Forcing" (PDF). Archived (PDF) from the original on 22 October 2018. Retrieved 14 March 2023.
  49. "Annex II Glossary". Intergovernmental Panel on Climate Change. Archived from the original on 3 November 2018. Retrieved 15 October 2010.
  50. A concise description of the greenhouse effect is given in the Intergovernmental Panel on Climate Change Fourth Assessment Report, "What is the Greenhouse Effect?" FAQ 1.3 – AR4 WGI Chapter 1: Historical Overview of Climate Change Science Archived 30 November 2018 at the Wayback Machine , IPCC Fourth Assessment Report, Chapter 1, p. 115: "To balance the absorbed incoming [solar] energy, the Earth must, on average, radiate the same amount of energy back to space. Because the Earth is much colder than the Sun, it radiates at much longer wavelengths, primarily in the infrared part of the spectrum (see Figure 1). Much of this thermal radiation emitted by the land and ocean is absorbed by the atmosphere, including clouds, and reradiated back to Earth. This is called the greenhouse effect."
    Stephen H. Schneider, in Geosphere-biosphere Interactions and Climate, Lennart O. Bengtsson and Claus U. Hammer, eds., Cambridge University Press, 2001, ISBN   0-521-78238-4, pp. 90–91.
    E. Claussen, V.A. Cochran, and D.P. Davis, Climate Change: Science, Strategies, & Solutions, University of Michigan, 2001. p. 373.
    A. Allaby and M. Allaby, A Dictionary of Earth Sciences, Oxford University Press, 1999, ISBN   0-19-280079-5, p. 244.
  51. Smil, Vaclav (2003). The Earth's Biosphere: Evolution, Dynamics, and Change. MIT Press. p. 107. ISBN   978-0-262-69298-4. Archived from the original on 14 March 2023. Retrieved 14 March 2023.{{cite book}}: CS1 maint: extra punctuation (link)
  52. Arrhenius, Svante (1896). "On the influence of carbonic acid in the air upon the temperature of the ground" (PDF). Philosophical Magazine and Journal of Science: 237–76. Archived (PDF) from the original on 18 November 2020. Retrieved 14 March 2023.
  53. Riebeek, Holli (16 June 2011). "The Carbon Cycle". Earth Observatory. NASA. Archived from the original on 5 March 2016. Retrieved 5 April 2018.
  54. Kayler, Z.; Janowiak, M.; Swanston, C. (2017). "The Global Carbon Cycle". Considering Forest and Grassland Carbon in Land Management (PDF). United States Department of Agriculture, Forest Service. pp. 3–9. Archived (PDF) from the original on 7 July 2022. Retrieved 14 March 2023.{{cite book}}: |journal= ignored (help)
  55. Gerlach, T.M. (4 June 1991). "Present-day CO2 emissions from volcanoes". Eos, Transactions, American Geophysical Union . 72 (23): 249, 254–55. Bibcode:1991EOSTr..72..249.. doi:10.1029/90EO10192.
  56. Junling Huang; Michael B. McElroy (2012). "The Contemporary and Historical Budget of Atmospheric CO2" (PDF). Canadian Journal of Physics. 90 (8): 707–16. Bibcode:2012CaJPh..90..707H. doi:10.1139/p2012-033. Archived (PDF) from the original on 3 August 2017. Retrieved 14 March 2023.
  57. 1 2 Susan Solomon; Gian-Kasper Plattner; Reto Knutti; Pierre Friedlingstein (February 2009). "Irreversible climate change due to carbon dioxide emissions". Proc. Natl. Acad. Sci. USA. 106 (6): 1704–09. Bibcode:2009PNAS..106.1704S. doi: 10.1073/pnas.0812721106 . PMC   2632717 . PMID   19179281.
  58. Archer, David; Eby, Michael; Brovkin, Victor; Ridgwell, Andy; Cao, Long; Mikolajewicz, Uwe; Caldeira, Ken; Matsumoto, Katsumi; Munhoven, Guy; Montenegro, Alvaro; Tokos, Kathy (2009). "Atmospheric Lifetime of Fossil Fuel Carbon Dioxide". Annual Review of Earth and Planetary Sciences. 37 (1): 117–34. Bibcode:2009AREPS..37..117A. doi:10.1146/annurev.earth.031208.100206. hdl: 2268/12933 . ISSN   0084-6597. Archived from the original on 14 March 2023. Retrieved 14 March 2023.
  59. Keeling, Charles D. (5 August 1997). "Climate change and carbon dioxide: An introduction". Proceedings of the National Academy of Sciences. 94 (16): 8273–8274. Bibcode:1997PNAS...94.8273K. doi: 10.1073/pnas.94.16.8273 . ISSN   0027-8424. PMC   33714 . PMID   11607732.
  60. 1 2 IPCC (2021). "Summary for Policymakers" (PDF). The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. ISBN   978-92-9169-158-6.
  61. 1 2 "Summary for Policymakers". The Ocean and Cryosphere in a Changing Climate (PDF). 2019. pp. 3–36. doi:10.1017/9781009157964.001. ISBN   978-1-00-915796-4. Archived (PDF) from the original on 29 March 2023. Retrieved 26 March 2023.
  62. 1 2 3 Cheng, Lijing; Abraham, John; Trenberth, Kevin E.; Fasullo, John; Boyer, Tim; Mann, Michael E.; Zhu, Jiang; Wang, Fan; Locarnini, Ricardo; Li, Yuanlong; Zhang, Bin; Yu, Fujiang; Wan, Liying; Chen, Xingrong; Feng, Licheng (2023). "Another Year of Record Heat for the Oceans". Advances in Atmospheric Sciences. 40 (6): 963–974. Bibcode:2023AdAtS..40..963C. doi: 10.1007/s00376-023-2385-2 . ISSN   0256-1530. PMC   9832248 . PMID   36643611. Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
  63. Fox-Kemper, B., H.T. Hewitt, C. Xiao, G. Aðalgeirsdóttir, S.S. Drijfhout, T.L. Edwards, N.R. Golledge, M. Hemer, R.E. Kopp, G. Krinner, A. Mix, D. Notz, S. Nowicki, I.S. Nurhati, L. Ruiz, J.-B. Sallée, A.B.A. Slangen, and Y. Yu, 2021: Chapter 9: Ocean, Cryosphere and Sea Level Change Archived 2022-10-24 at the Wayback Machine . In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Archived 2021-08-09 at the Wayback Machine [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1211–1362
  64. Gille, Sarah T. (15 February 2002). "Warming of the Southern Ocean Since the 1950s". Science . 295 (5558): 1275–1277. Bibcode:2002Sci...295.1275G. doi:10.1126/science.1065863. PMID   11847337. S2CID   31434936.
  65. Ritchie, Roser, Mispy, Ortiz-Ospina. "SDG 14 – Measuring progress towards the Sustainable Development Goals Archived 22 January 2022 at the Wayback Machine ." SDG-Tracker.org, website (2018).
  66. Terhaar, Jens; Frölicher, Thomas L.; Joos, Fortunat (2023). "Ocean acidification in emission-driven temperature stabilization scenarios: the role of TCRE and non-CO2 greenhouse gases". Environmental Research Letters. 18 (2): 024033. Bibcode:2023ERL....18b4033T. doi:10.1088/1748-9326/acaf91. ISSN   1748-9326. S2CID   255431338. Figure 1f
  67. Ocean acidification due to increasing atmospheric carbon dioxide (PDF). Royal Society. 2005. ISBN   0-85403-617-2.
  68. Jiang, Li-Qing; Carter, Brendan R.; Feely, Richard A.; Lauvset, Siv K.; Olsen, Are (2019). "Surface ocean pH and buffer capacity: past, present and future". Scientific Reports. 9 (1): 18624. Bibcode:2019NatSR...918624J. doi: 10.1038/s41598-019-55039-4 . PMC   6901524 . PMID   31819102. CC-BY icon.svg Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 16 October 2017 at the Wayback Machine
  69. Zhang, Y.; Yamamoto-Kawai, M.; Williams, W.J. (16 February 2020). "Two Decades of Ocean Acidification in the Surface Waters of the Beaufort Gyre, Arctic Ocean: Effects of Sea Ice Melt and Retreat From 1997–2016". Geophysical Research Letters. 47 (3). doi: 10.1029/2019GL086421 . S2CID   214271838.
  70. Beaupré-Laperrière, Alexis; Mucci, Alfonso; Thomas, Helmuth (31 July 2020). "The recent state and variability of the carbonate system of the Canadian Arctic Archipelago and adjacent basins in the context of ocean acidification". Biogeosciences. 17 (14): 3923–3942. Bibcode:2020BGeo...17.3923B. doi: 10.5194/bg-17-3923-2020 . S2CID   221369828.
  71. Ueyama M, Ichii K, Kobayashi H, Kumagai TO, Beringer J, Merbold L, et al. (17 July 2020). "Inferring CO2 fertilization effect based on global monitoring land-atmosphere exchange with a theoretical model". Environmental Research Letters. 15 (8): 084009. Bibcode:2020ERL....15h4009U. doi: 10.1088/1748-9326/ab79e5 . ISSN   1748-9326.
  72. 1 2 Tharammal T, Bala G, Narayanappa D, Nemani R (April 2019). "Potential roles of CO2 fertilization, nitrogen deposition, climate change, and land use and land cover change on the global terrestrial carbon uptake in the twenty-first century". Climate Dynamics. 52 (7–8): 4393–4406. Bibcode:2019ClDy...52.4393T. doi:10.1007/s00382-018-4388-8. ISSN   0930-7575. S2CID   134286531.
  73. 1 2 3 4 Hararuk O, Campbell EM, Antos JA, Parish R (December 2018). "Tree rings provide no evidence of a CO2 fertilization effect in old-growth subalpine forests of western Canada". Global Change Biology. 25 (4): 1222–1234. Bibcode:2019GCBio..25.1222H. doi: 10.1111/gcb.14561 . PMID   30588740.
  74. Cartwright J (16 August 2013). "How does carbon fertilization affect crop yield?". environmentalresearchweb. Environmental Research Letters. Archived from the original on 27 June 2018. Retrieved 3 October 2016.
  75. Smith WK, Reed SC, Cleveland CC, Ballantyne AP, Anderegg WR, Wieder WR, et al. (March 2016). "Large divergence of satellite and Earth system model estimates of global terrestrial CO2 fertilization". Nature Climate Change. 6 (3): 306–310. Bibcode:2016NatCC...6..306K. doi:10.1038/nclimate2879. ISSN   1758-678X.
  76. Chen C, Riley WJ, Prentice IC, Keenan TF (March 2022). "CO2 fertilization of terrestrial photosynthesis inferred from site to global scales". Proceedings of the National Academy of Sciences of the United States of America. 119 (10): e2115627119. Bibcode:2022PNAS..11915627C. doi: 10.1073/pnas.2115627119 . PMC   8915860 . PMID   35238668.
  77. Bastos A, Ciais P, Chevallier F, Rödenbeck C, Ballantyne AP, Maignan F, Yin Y, Fernández-Martínez M, Friedlingstein P, Peñuelas J, Piao SL (7 October 2019). "Contrasting effects of CO2 fertilization, land-use change and warming on seasonal amplitude of Northern Hemisphere CO2 exchange". Atmospheric Chemistry and Physics. 19 (19): 12361–12375. Bibcode:2019ACP....1912361B. doi: 10.5194/acp-19-12361-2019 . ISSN   1680-7324.
  78. Li Q, Lu X, Wang Y, Huang X, Cox PM, Luo Y (November 2018). "Leaf Area Index identified as a major source of variability in modelled CO2 fertilization". Biogeosciences. 15 (22): 6909–6925. doi: 10.5194/bg-2018-213 .
  79. Albani M, Medvigy D, Hurtt GC, Moorcroft PR (December 2006). "The contributions of land-use change, CO2 fertilization, and climate variability to the Eastern US carbon sink: Partitioning of the Eastern US Carbon Sink". Global Change Biology. 12 (12): 2370–2390. doi:10.1111/j.1365-2486.2006.01254.x. S2CID   2861520.
  80. Wang S, Zhang Y, Ju W, Chen JM, Ciais P, Cescatti A, et al. (December 2020). "Recent global decline of CO2 fertilization effects on vegetation photosynthesis". Science. 370 (6522): 1295–1300. Bibcode:2020Sci...370.1295W. doi:10.1126/science.abb7772. hdl: 10067/1754050151162165141 . PMID   33303610. S2CID   228084631.
  81. Sugden AM (11 December 2020). Funk M (ed.). "A decline in the carbon fertilization effect". Science. 370 (6522): 1286.5–1287. Bibcode:2020Sci...370S1286S. doi:10.1126/science.370.6522.1286-e. S2CID   230526366.
  82. Kirschbaum MU (January 2011). "Does enhanced photosynthesis enhance growth? Lessons learned from CO2 enrichment studies". Plant Physiology. 155 (1): 117–24. doi:10.1104/pp.110.166819. PMC   3075783 . PMID   21088226.
  83. "Global Green Up Slows Warming". earthobservatory.nasa.gov. 18 February 2020. Retrieved 27 December 2020.
  84. Tabor A (8 February 2019). "Human Activity in China and India Dominates the Greening of Earth". NASA. Retrieved 27 December 2020.
  85. Zhu Z, Piao S, Myneni RB, Huang M, Zeng Z, Canadell JG, et al. (1 August 2016). "Greening of the Earth and its drivers". Nature Climate Change. 6 (8): 791–795. Bibcode:2016NatCC...6..791Z. doi:10.1038/nclimate3004. S2CID   7980894.
  86. Hille K (25 April 2016). "Carbon Dioxide Fertilization Greening Earth, Study Finds". NASA. Retrieved 27 December 2020.
  87. "If you're looking for good news about climate change, this is about the best there is right now". Washington Post. Retrieved 11 November 2016.
  88. Schimel D, Stephens BB, Fisher JB (January 2015). "Effect of increasing CO2 on the terrestrial carbon cycle". Proceedings of the National Academy of Sciences of the United States of America. 112 (2): 436–41. Bibcode:2015PNAS..112..436S. doi: 10.1073/pnas.1407302112 . PMC   4299228 . PMID   25548156.
  89. Pisoft, Petr (25 May 2021). "Stratospheric contraction caused by increasing greenhouse gases". Environmental Research Letters. 16 (6): 064038. Bibcode:2021ERL....16f4038P. doi: 10.1088/1748-9326/abfe2b .
  90. CounterAct; Women's Climate Justice Collective (4 May 2020). "Climate Justice and Feminism Resource Collection". The Commons Social Change Library. Retrieved 8 July 2024.
  91. Käse, Laura; Geuer, Jana K. (2018). "Phytoplankton Responses to Marine Climate Change – an Introduction". YOUMARES 8 – Oceans Across Boundaries: Learning from each other. pp. 55–71. doi:10.1007/978-3-319-93284-2_5. ISBN   978-3-319-93283-5. S2CID   134263396.
  92. Cheng, Lijing; Abraham, John; Hausfather, Zeke; Trenberth, Kevin E. (11 January 2019). "How fast are the oceans warming?". Science. 363 (6423): 128–129. Bibcode:2019Sci...363..128C. doi:10.1126/science.aav7619. PMID   30630919. S2CID   57825894.
  93. 1 2 Doney, Scott C.; Busch, D. Shallin; Cooley, Sarah R.; Kroeker, Kristy J. (17 October 2020). "The Impacts of Ocean Acidification on Marine Ecosystems and Reliant Human Communities". Annual Review of Environment and Resources. 45 (1): 83–112. doi: 10.1146/annurev-environ-012320-083019 . CC-BY icon.svg Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License Archived 2017-10-16 at the Wayback Machine
  94. IPCC, 2021: "Annex VII: Glossary". Matthews, J.B.R., V. Möller, R. van Diemen, J.S. Fuglestvedt, V. Masson-Delmotte, C. Méndez, S. Semenov, A. Reisinger (eds.). In "Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change". Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.). Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 2215–2256, doi:10.1017/9781009157896.022
  95. Schenuit, Felix; Colvin, Rebecca; Fridahl, Mathias; McMullin, Barry; Reisinger, Andy; Sanchez, Daniel L.; Smith, Stephen M.; Torvanger, Asbjørn; Wreford, Anita; Geden, Oliver (4 March 2021). "Carbon Dioxide Removal Policy in the Making: Assessing Developments in 9 OECD Cases". Frontiers in Climate. 3: 638805. doi: 10.3389/fclim.2021.638805 . hdl: 1885/270309 . ISSN   2624-9553.
  96. Geden, Oliver (May 2016). "An actionable climate target". Nature Geoscience. 9 (5): 340–342. Bibcode:2016NatGe...9..340G. doi:10.1038/ngeo2699. ISSN   1752-0908. Archived from the original on 25 May 2021. Retrieved 7 March 2021.
  97. Ho, David T. (4 April 2023). "Carbon dioxide removal is not a current climate solution — we need to change the narrative". Nature. 616 (7955): 9. Bibcode:2023Natur.616....9H. doi:10.1038/d41586-023-00953-x. ISSN   0028-0836. PMID   37016122. S2CID   257915220.
  98. M. Pathak, R. Slade, P.R. Shukla, J. Skea, R. Pichs-Madruga, D. Ürge-Vorsatz,2022: Technical Summary. In: Climate Change 2022: Mitigation of Climate Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [P.R. Shukla, J. Skea, R. Slade, A. Al Khourdajie, R. van Diemen, D. McCollum, M. Pathak, S. Some, P. Vyas, R. Fradera, M. Belkacemi, A. Hasija, G. Lisboa, S. Luz, J. Malley, (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA. doi: 10.1017/9781009157926.002.
  99. Gulev, S.K., P.W. Thorne, J. Ahn, F.J. Dentener, C.M. Domingues, S. Gerland, D. Gong, D.S. Kaufman, H.C. Nnamchi, J.  Quaas, J.A. Rivera, S. Sathyendranath, S.L. Smith, B. Trewin, K. von Schuckmann, and R.S. Vose, 2021: Chapter 2: Changing State of the Climate System. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R.  Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 287–422, doi:10.1017/9781009157896.004.
  100. Walker, James C.G. (June 1985). "Carbon dioxide on the early earth" (PDF). Origins of Life and Evolution of the Biosphere. 16 (2): 117–27. Bibcode:1985OrLi...16..117W. doi:10.1007/BF01809466. hdl: 2027.42/43349 . PMID   11542014. S2CID   206804461. Archived (PDF) from the original on 14 September 2012. Retrieved 30 January 2010.
  101. Pavlov, Alexander A.; Kasting, James F.; Brown, Lisa L.; Rages, Kathy A.; Freedman, Richard (May 2000). "Greenhouse warming by CH4 in the atmosphere of early Earth". Journal of Geophysical Research. 105 (E5): 11981–90. Bibcode:2000JGR...10511981P. doi: 10.1029/1999JE001134 . PMID   11543544.
  102. Zahnle, K.; Schaefer, L.; Fegley, B. (2010). "Earth's Earliest Atmospheres". Cold Spring Harbor Perspectives in Biology. 2 (10): a004895. doi:10.1101/cshperspect.a004895. PMC   2944365 . PMID   20573713.
  103. Olson JM (May 2006). "Photosynthesis in the Archean era". Photosynth. Res. 88 (2): 109–17. Bibcode:2006PhoRe..88..109O. doi:10.1007/s11120-006-9040-5. PMID   16453059. S2CID   20364747.
  104. Buick R (August 2008). "When did oxygenic photosynthesis evolve?". Philos. Trans. R. Soc. Lond. B Biol. Sci. 363 (1504): 2731–43. Bibcode:2008RSPTB.363.2731B. doi:10.1098/rstb.2008.0041. PMC   2606769 . PMID   18468984.
  105. 1 2 3 Osborne, C.P.; Beerling, D.J. (2006). "Nature's green revolution: the remarkable evolutionary rise of C4 plants". Philosophical Transactions of the Royal Society B: Biological Sciences. 361 (1465): 173–94. doi:10.1098/rstb.2005.1737. PMC   1626541 . PMID   16553316.
  106. Lovelock, J. E. (1972). "Gaia as seen through the atmosphere". Atmospheric Environment. 6 (8): 579–580. Bibcode:1972AtmEn...6..579L. doi:10.1016/0004-6981(72)90076-5. Archived from the original on 3 November 2011. Retrieved 22 March 2014.
  107. Li, K.-F. (30 May 2009). "Atmospheric pressure as a natural climate regulator for a terrestrial planet with a biosphere". Proceedings of the National Academy of Sciences. 106 (24): 9576–9579. Bibcode:2009PNAS..106.9576L. doi: 10.1073/pnas.0809436106 . PMC   2701016 . PMID   19487662. Archived from the original on 12 February 2013. Retrieved 22 March 2014.
  108. Etheridge, D.M.; Steele, L.P.; Langenfelds, R.L.; Francey, R.J.; Barnola, JM; Morgan, VI (June 1998). "Historical CO2 record derived from a spline fit (20-year cutoff) of the Law Dome DE08 and DE08-2 ice cores". Carbon Dioxide Information Analysis Center. Oak Ridge National Laboratory. Archived from the original on 5 March 2012. Retrieved 12 June 2007.
  109. Flückiger, Jacqueline (2002). "High-resolution Holocene N
    2
    O
    ice core record and its relationship with CH
    4
    and CO2"
    . Global Biogeochemical Cycles. 16 (1): 1010. Bibcode:2002GBioC..16.1010F. doi: 10.1029/2001GB001417 .
  110. Amos, J. (4 September 2006). "Deep ice tells long climate story". BBC News. Archived from the original on 23 January 2013. Retrieved 28 April 2010.
  111. Hileman B. (November 2005). "Ice Core Record Extended: Analyses of trapped air show current CO2 at highest level in 650,000 years". Chemical & Engineering News . 83 (48): 7. doi:10.1021/cen-v083n048.p007. ISSN   0009-2347. Archived from the original on 15 May 2019. Retrieved 28 January 2010.
  112. Vostok Ice Core Data Archived 27 February 2015 at the Wayback Machine , ncdc.noaa.gov Archived 22 April 2021 at the Wayback Machine
  113. Friederike Wagner; Bent Aaby; Henk Visscher (2002). "Rapid atmospheric CO2 changes associated with the 8,200-years-B.P. cooling event". Proc. Natl. Acad. Sci. USA. 99 (19): 12011–14. Bibcode:2002PNAS...9912011W. doi: 10.1073/pnas.182420699 . PMC   129389 . PMID   12202744.
  114. Andreas Indermühle; Bernhard Stauffer; Thomas F. Stocker (1999). "Early Holocene Atmospheric CO2 Concentrations". Science. 286 (5446): 1815. doi: 10.1126/science.286.5446.1815a .IndermÜhle, A (1999). "Early Holocene atmospheric CO2concentrations". Science. 286 (5446): 1815a–15. doi: 10.1126/science.286.5446.1815a .
  115. H. J. Smith; M. Wahlen; D. Mastroianni (1997). "The CO2 concentration of air trapped in GISP2 ice from the Last Glacial Maximum-Holocene transition". Geophysical Research Letters. 24 (1): 1–4. Bibcode:1997GeoRL..24....1S. doi:10.1029/96GL03700. S2CID   129667062.
  116. File:Phanerozoic Carbon Dioxide.png
  117. 1 2 Witkowski, Caitlyn (28 November 2018). "Molecular fossils from phytoplankton reveal secular pCO2 trend over the Phanerozoic". Science Advances. 2 (11): eaat4556. Bibcode:2018SciA....4.4556W. doi:10.1126/sciadv.aat4556. PMC   6261654 . PMID   30498776.
  118. "IPCC: Climate Change 2001: The Scientific Basis" (PDF). Archived (PDF) from the original on 29 August 2022. Retrieved 14 March 2023.
  119. Berner, Robert A. (January 1994). "GEOCARB II: a revised model of atmospheric CO2 over Phanerozoic time". American Journal of Science. 294 (1): 56–91. Bibcode:1994AmJS..294...56B. doi: 10.2475/ajs.294.1.56 .
  120. Royer, D.L.; R.A. Berner; D.J. Beerling (2001). "Phanerozoic atmospheric CO2 change: evaluating geochemical and paleobiological approaches". Earth-Science Reviews. 54 (4): 349–92. Bibcode:2001ESRv...54..349R. doi:10.1016/S0012-8252(00)00042-8.
  121. Berner, Robert A.; Kothavala, Zavareth (2001). "GEOCARB III: a revised model of atmospheric CO2 over Phanerozoic time" (PDF). American Journal of Science. 301 (2): 182–204. Bibcode:2001AmJS..301..182B. CiteSeerX   10.1.1.393.582 . doi:10.2475/ajs.301.2.182. Archived (PDF) from the original on 25 April 2006.
  122. Beerling, D.J.; Berner, R.A. (2005). "Feedbacks and the co-evolution of plants and atmospheric CO2". Proc. Natl. Acad. Sci. USA. 102 (5): 1302–05. Bibcode:2005PNAS..102.1302B. doi: 10.1073/pnas.0408724102 . PMC   547859 . PMID   15668402.
  123. 1 2 Hoffmann, PF; AJ Kaufman; GP Halverson; DP Schrag (1998). "A neoproterozoic snowball earth". Science. 281 (5381): 1342–46. Bibcode:1998Sci...281.1342H. doi:10.1126/science.281.5381.1342. PMID   9721097. S2CID   13046760.
  124. Siegel, Ethan. "How Much CO2 Does A Single Volcano Emit?". Forbes. Archived from the original on 6 June 2017. Retrieved 6 September 2018.
  125. Gerlach, TM (1991). "Present-day CO2 emissions from volcanoes". Transactions of the American Geophysical Union. 72 (23): 249–55. Bibcode:1991EOSTr..72..249.. doi:10.1029/90EO10192.
  126. See also: "U.S. Geological Survey". 14 June 2011. Archived from the original on 25 September 2012. Retrieved 15 October 2012.
  127. "New CO2 data helps unlock the secrets of Antarctic formation". Physorg.com. 13 September 2009. Archived from the original on 15 July 2011. Retrieved 28 January 2010.
  128. Pagani, Mark; Huber, Matthew; Liu, Zhonghui; Bohaty, Steven M.; Henderiks, Jorijntje; Sijp, Willem; Krishnan, Srinath; Deconto, Robert M. (2 December 2011). "Drop in carbon dioxide levels led to polar ice sheet, study finds". Science. 334 (6060): 1261–4. Bibcode:2011Sci...334.1261P. doi:10.1126/science.1203909. PMID   22144622. S2CID   206533232. Archived from the original on 22 May 2013. Retrieved 14 May 2013.