You are on page 1of 28

Carbon sequestration for mitigating hazardous effects of global warming

Adarsh K. Puri and T. Satyanarayana* Department of Microbiology, University of Delhi South Campus, New Delhi, India *Corresponding Author, E-mail: tsnarayana@gmail.com

Abstract The concentration of atmospheric CO2 has increased in an unprecedented manner after the industrial revolution. The whole scientific community, politicians and industrialists throughout the world are debating on the development of strategies to bring down the atmospheric carbon levels to the acceptable limits. Carbon sequestration has emerged as the most potential and effective way for mitigating global warming in the past few years. Carbon capture and storage in ocean and deep geological formations are the promising solutions. The current focus now, however, is on novel ideas that ensure a leakage-proof and cost-effective approach for long term and maximal storage of CO2. Capturing carbon by biological means is not only a mean of sequestering carbon, but may also lead to the production of useful products. 1. Introduction The thawing permafrost (Kolchugina and Vinson, 1993), melting glaciers (U.N. Environment Programme, 2008; Vaughan et al., 2003; Dyurgerov and Meier, 2000), raising sea levels (Meier and Wahr, 2002), changing hydrological cycle (Held and Soden, 2006; Mirza, 2002), increasing precipitation (Schnur, 2002), declining crop productivity (Tan and Shibasaki, 2003; Ortiz et al., 2008), early breeding of birds (Brown et al., 1999; Sydeman et al., 2001; Wormworth and Mallon, 2007), vanishing coral reefs (Donner et al., 2005; Crabbe, 2007), increasing health hazards (McMichael and Woodruff, 2004; Sutherst, 2004) are all predictable effects of the hottest debatable phenomenon on earth called global warming. Global warming, frequently referred to as climate change, is not just a theory or a distant threat. The 2007 Nobel peace prize to Intergovernmental Panel on Climate Change (IPCC) and Albert Arnold (Al) Gore Jr. has made every one realize the severity of the problem, and further, cleared all doubts being raised by naysayers over its reality. The overwhelming agreement among the worlds prominent scientists, governments and scientific bodies is that the Earth is heating up and that human activities are largely to blame [IPCC, 2007; National Research Council (NRC) 2001]. The global warming is expected to significantly disrupt the planets climate system. Minimization of greenhouse gas emissions to acceptable limits is the intrinsic environmental responsibility of the whole world.

Over the last 200 years since the Industrial Revolution, most of the worlds energy has been derived from burning finite resources of fossil fuels, mainly coal, oil and more recently, gas (Ansolabehere et al., 2007). Fossil fuels account for 80% of the global energy demand (Table 1). Table 1. Fuel sources and world energy demand [Source: Ansolabehere et al., 2007]
Sources Coal Natural gas Petroleum Nuclear Hydro Biomass and waste Geothermal, Solar and Wind Global energy demand met 25% 21% 34% 6.5% 2.2% 11% 0.4%

During the process, a billion tons of carbon dioxide and other green house gases (GHGs) have been spewed into the atmosphere. Energy sector accounts for the greatest share (36%) of carbon dioxide emissions. A large 1000 Megawatt coal power station releases around 5.5 million tons of CO2 annually (Evans and Furlong, 2003). Earths atmosphere is essentially transparent to incoming radiation from the Sun, as sunlight peaks in the visible part of the spectrum. On the other hand, thermal radiation from the Earth, in the form of long-wavelength infrared rays, lies in the absorption spectrum of carbon dioxide and other GHGs. These GHGs absorb radiation primarily in a very narrow frequency band (7-13m), while CO2 absorbs over a much larger (13-19m) spectral range (Halmann and Steinberg, 1999). This is why CO2 accounts for 21% of the greenhouse effect (after water vapour that accounts for 64%) than ozone (6%) and other trace gases (9%) (Barry and Charley, 1992; Marsh, 2001). Moreover, carbon dioxide makes up 68% of the total greenhouse gas emissions (Harrington and Foster, 1999). The atmospheric CO2 concentration has increased from 280 ppm in 1800, the beginning of industrial age, to 380 ppm today (IPCC, 2007; Takashaki, 2004; NOAA Climate Monitoring and Diagnostics Laboratory, 2003), and without any mitigation, it could reach levels of 700-900 ppm by the end of the 21st century, which could bring about severe climate change (Houghton et al., 2001; IPCC, 2001). The annual CO2 concentration growth rate was larger during the last 10 years (1995-2005 average: 1.9 ppm per year) than it has been since the beginning of direct atmospheric

measurements. In fact, eleven of the last twelve years (1995-2006) rank among the twelve warmest years since 1850 (IPCC, 2007; Hadley centre, 2006). This abrupt imbalance has disturbed the Earths carbon cycle that is normally kept in balance by the oceans, vegetation, soil and the forests. The most pressing technical and economic challenge of the present time is to supply energy demand for the world economic growth without affecting the Earths climate. That is why the current focus is on reducing fossil fuel usage and minimizing the emission of CO2 in atmosphere (Evans and Furlong, 2003). In spite of the great advances made in the field of renewable energy, it has not been possible to replace gas, coal and oil to meet the current energy needs (Collins, 1998). If fossil fuels, particularly coal, remain the dominant energy source of the 21st century, then stabilizing the concentration of atmospheric CO2 will require development of the capability to capture CO2 from the combustion of fossil fuels and store it safely away from the atmosphere (House et al., 2006). The hazards of global warming have reached to a magnitude that irreversible changes in the functioning of the planet are seriously feared. And mankind has been ruthless (harmful rather useless)! It is, therefore, implacable for the whole scientific community to restore permissible levels of CO2 by using the existing knowledge. Carbon sequestration or carbon capture and storage (CCS) has emerged as a potentially promising technology to deal with the problem of global warming. Several approaches are being considered, including geological, oceanic, and terrestrial sequestration, as well as CO2 conversion into useful materials. In this chapter, an attempt has been made to review the possible strategies for carbon sequestration. An emphasis has been laid on biological ways of carbon sequestration due to the drawbacks associated with ocean and geological sequestration approaches. This includes normal photosynthetic ways of carbon fixation, exploiting enzymatic machinery and metabolic pathways of microbes and use of various biomimetic approaches. 2. Gases contributing to global warming: Green House Gases (GHGs) Molecules of various greenhouse gases trap the heat that is expected to escape from earth. The extent of greenhouse effect contributed by different gases over a certain time frame is expressed in terms of their individual Global Warming Potential (GWP) taking CO2 as the reference gas (Table 2). The main greenhouse gases produced by human activity are carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O) and some halogenated compounds with high-GWP. Perfluorocarbons (PFCs), sulphur hexafluoride (SF6) and hydrofluorocarbons (HFCs) were added to the list of green house gases under the Kyoto Protocol to the United Nations Framework Convention on Climate Change (UNFCCC) in 1997. Non-CO2 greenhouse gases are also a matter of concern
3

owing to their significant contribution (30%) to the overall anthropogenic greenhouse effect since preindustrial times (IPCC, 2001). Table 2. Global warming potential of selected greenhouse gases.
Greenhouse Gases (GHGs) Source Average rate of increase (in % per year) Global warming potential (GWP) Estimated atmospheric lifetime (in years) NR

Fossil fuel combustion, 1 1.1-3.0 deforestation, changing land use, biomass burning, erosion Methane (CH4) Enteric fermentation in cattle 0.8-1.0* 21 and insects, biomass burning, waste burial, coal mines, gas leaks, rice fields, swamps and tundra Nitrous oxide (N2O) Aerosols, refrigeration and 0.25 310 air conditioning, plastic foams, solvents, computer industry, sterilants, medical supplies Sulphur hexafluoride (SF6) Electrical insulation, 7 23,900 semiconductor manufacture , magnesium foundries (cast) Perfluorocarbons (PFCs) Primary aluminum 1.3-3.2 6,500- 9,200 production and semiconductor manufacture Hydrofluorocarbons Man made alternatives to High variation 140-11,700 (HFCs) ozone depleting substances (ODSs) Source: USEPA, 2006; IPCC, 2001. Commonwealth Scientific and Industrial Research Organization, 2007. Million metric tons of CO2 equivalent over 100 year integration time (MMT CO2 Eq.). * Milich, 1999; Glantz and Krenz, 1992.

Carbon dioxide (CO2)

12

14

3,200

2,600- 50,000

1-260

Fenhann, J. (2000) NR = not reported because this values depends greatly on assumptions.

The amount of anthropogenic CO2 emitted to the atmosphere is much greater than any of other greenhouse gases. As a result, CO2 makes the highest contribution to the greenhouse effect despite its low GWP. 3. Carbon sequestration and its importance Carbon sequestration can be defined as the capture and secure storage of carbon that would otherwise be emitted to or remain in the atmosphere. The idea is to keep carbon emissions produced by human activities from reaching the atmosphere by capturing and diverting them to secure storage or to remove carbon from the atmosphere by various means and storing it (DOE, 1999).

Carbon sequestration could be a major tool for reducing carbon emissions from fossil fuels. Much work, however, remains to be done to understand the science and engineering aspects and potential of carbon sequestration options. Given the magnitude of carbon emission reductions needed to stabilize the atmospheric CO2 concentration, multiple approaches to carbon management will be needed. The natural carbon cycle is balanced over a long term, but dynamic over the short term. Historically, acceleration of natural processes that emit CO2 is eventually balanced by the acceleration of processes that sequester carbon, and vice versa. The current increase in atmospheric carbon is the result of anthropogenic mining and burning of fossil carbon, resulting in carbon emissions into the atmosphere. Developing new sequestration techniques and accelerating existing techniques would help in diminishing the net positive atmospheric carbon flux. 4. General methods of carbon sequestration 4.1. Ocean sequestration Oceans cover over 70% of the Earths surface with an average depth of about 3800 metres. Depending upon the oceanic equilibrium with the atmosphere, a significant amount of captured CO2 could be deliberately injected into the ocean at great depth, where it would remain isolated from the atmosphere for centuries. 4.1.1. Direct injection of CO2 Direct ocean CO2 disposal, first suggested by Marchetti in 1977, is now the biggest hope to use ocean as the largest sink for carbon sequestration purposes. A large literature is now available which has brought significant technological developments (Kobayashi, 1995; Aya et al., 1997; Brewer et al., 2000; Fer and Haugen, 2003) and improved our understanding the disposal of CO2 directly into the ocean. The research on ocean disposal options has mostly focused on predicting the behavior and the dissolution time scale of the released CO2. Different scenarios of CO2 disposal in the ocean have been proposed at various depths and in different forms in relation to the phase properties of CO2 (Ozaki et al., 2001; Shitashima et al., 2008). CO2 can be released directly in to the ocean in any of its physical forms- gas, liquid, solid or solid hydrate. But, it is important to study CO 2 induced density changes on the fluid dynamics of the ocean before its release. Dissolved CO2 increases the density of seawater (Fig.1) that affects its transport and mixing (Bradshaw, 1973; Song et al., 2005). Density of Injected CO2 is also controlled by geothermal gradient, which varies from 0.02oC/m to 0.04oC/m. The rate of CO2 dissolution in the seawater depends upon its physical form (gas, liquid, solid or solid hydrate), the depth and temperature of disposal, and the local water velocities.

Fig.1. Density profiles of liquid CO2 at 0oC and 10oC (solid curves), CO2-saturated seawater (dashed line), and seawater (dotted line). [Source: Fer and Haugen, 2003] CO2 disposal in gas/liquid/solid phase CO2 could potentially be released as a gas above 500m depth. However, due to lesser density of gas bubbles than surrounding seawater, these bubbles tend to rise up on the surface, dissolving at a radial speed of about 0.1 cm hr-1 (Teng et al., 1996). It is better to use CO2 diffusers to produce smaller CO2 bubbles, which can dissolve completely before reaching the surface. CO2 can exist as a liquid below roughly 500m to 2,500m depth. Due to lower temperature (<9 oC), CO2 hydrate tends to form on the droplet wall. Under these conditions, the radius of droplet would diminish at a speed of about 0.5 cm hr-1 (Brewer et al., 2002). Below roughly 3,000 m, liquid CO2 becomes denser than the surrounding seawater, and hence, tends to sink. Similarly CO2 released in solid form is also denser and would sink to sea floor dissolving at a radial speed of about 0.2 cm hr-1 (Aya et al., 1997). Large masses of solid CO2 reach sea floor before complete dissolution. CO2 Hydrate CO2 hydrates (5.75 H2O. CO2) are nonstoichiometric crystalline compounds that form at high pressure (greater than; 4.5 MPa) and low temperature (less than 9.85o C) by trapping CO2 molecules in hydrogen- bonded cages of H2O (Lee et al., 2002; Fer and Haugen, 2003). It can form in average ocean waters below 400m depth. Being 15% denser than seawater (Wannamaker and Adams, 2002), these hydrates tend to sink and dissolve rapidly into the relatively dilute ocean waters at a speed similar to that of solid CO2 (about 0.2 cm/hr-1) [Rehder et al., 2004; Teng et al., 1999]. It is, however, not feasible to release CO2 hydrates directly through the pipelines.

That is why a paste like composite of hydrate and seawater has been proposed to be used for sequestration purposes (Tsouris et al., 2004). Liquid CO2 Lakes When liquid CO2 is injected into a sea floor depression at a depth greater than 3,000 m (where it is denser than sea water), it accumulates as a stable large lake" of CO2. The dissolution of these liquid CO2 lakes is retarded by formation of a thin hydrate layer over it (Ohsumi, 1993; Fer and Haugen, 2003). While investigating different kinds of discharge pipes for CO2 lake creation on sea floor, Nakashiki (1997) proposed a floating discharge pipe that was simple and less likely to be damaged by wind and wave in storm conditions. Slurry of liquid CO2 mixed with dry ice in discharge pipe provides good conditions for lake formation (Aya et al., 1997). 4.1.2. Natural oceanic mineralization Oceans can sequester so much of CO2 not only because of their large volume but also because CO2 dissolves in water to form various ionic species that increases the total dissolved inorganic carbon (DIC) of seawater. Total dissolved inorganic carbon (DIC) is the sum of carbon contained in H2CO3, HCO3- and CO3-2. CO2 (g) + H2O H2CO3 (aq.) HCO3- + H+ CO3-2 + 2H+

Ocean surface water is supersaturated with respect to calcium carbonate, while the deeper ocean water would be with lower pH and remain under saturated. This makes organisms to produce calcium carbonate particles (e.g. corals) in the surface oceans, which settle and dissolve in under saturated regions of deep oceans. 5. Geological sequestration Since the first use of CO2 for large-scale recovery of residual oil from Texas reservoirs in 1972, the concept of using CO2 for beneficial purposes has got momentum. Long term operational experience with geological formations, its substantial capacity as a CO2 sink (Table 3) and its immediate availability has led to consideration of global warming problem through geological sequestration (Klara et al, 2003). Geological formations include depleted oil and gas fields, deep saline reservoirs and unminable coal seams.

Table 3. Estimates of storage capacities for different geological reservoirs Source: Gale, 2004.

Global capacity Gt CO2 % of emissions to 2050 Depleted oil and gas fields 920 45 Deep saline reservoirs 400- 10,000 20-500 Unminable coal seams 20 <2 CO2 can be trapped in geologic formations by three principal trapping mechanisms (DOE, 1993): (1) hydrodynamic trapping, where CO2 can be trapped under a low-permeability caprock like gas reservoirs or aquifers, (2) solubility trapping, where CO2 can be trapped in a dissolved phase in a liquid- like petroleum and (3) physical/mineral trapping, a relatively slower process which involves conversion of CO2 in the form of calcium, magnesium or iron carbonates. 5.1. Depleted oil and gas reservoirs A considerable amount of oil or gas is often present in depleted oil and gas reservoirs following rigorous primary recovery processes (Gentzis, 2000). CO2 is being injected into oil reservoirs as an established and successful technique (over 80 projects worldwide; Wells et al., 2007) for enhanced oil recovery (CO2-EOR). Injected CO2 reduces the interfacial tension between oil and the reservoir rock, expands the volume of oil (oil swelling) thereby reducing its viscosity and making its easier mobility towards the production well (oil recovery enhanced by 10-15%; Davison et al., 2001). Alternatively, in situations where CO2 is immiscible with oil, CO2 is injected to increase the reservoir pressure helping to push more oil towards the production well. Up to half of the injected CO2 is stored in the immobile oil remaining in the reservoir at the end of production. The rest is collected from the production well and get re-circulated. This improves the overall economics for sequestration projects (Holloway, 2005). Gas fields have much higher primary recovery rates (80-95%) than oil fields (Gale, 2004). This leaves a big void space in the reservoirs, which can be used for CO2 storage as a supercritical gas for thousand of years. Similarly, the void space that had previously been occupied by oil and natural gases is being used for large-scale sequestration of CO2. 5.2. Deep saline aquifers A large amount of underground water filled strata (aquifers) is too salty to be used for agriculture or human consumption. These aquifers can potentially be used as long-term CO2 reservoirs (IEA, 2001; IPCC, 2005). CO2 injected (with techniques similar to those for gas and oil fields) into these aquifers would displace brine and some of it would get partially dissolved (Nordbotten et al., 2005). A part of the injected CO2 is also reported to react with calcite and aluminosilicates to

Storage option

form permanent carbonates. The best example of CO2 storage in deep saline aquifer is the Sleipner project in the North Sea, which sequesters approximately 1 Mt CO2 annually (IPCC, 2005). 5.3. Deep unminable coal seams Storing CO2 deep into unminable coal seams appears to be a good approach due to its value added benefit of CO2 enhanced coal bed methane (ECBM) recovery. Coal beds typically contain large amounts of methane-rich gas (Holloway, 2005) that is adsorbed onto the surface of the coal. CO2 adsorbs more strongly on the micropores of coal than methane (CH4). However, the volumetric ratio of adsorbable CO2:CH4 depends on the type of coal. This ratio ranges from 1 for anthracite to about 10 for lignite coal (IPCC, 2005). This can be exploited to lock CO2 permanently on the micropores of coal provided the coal is never mined. Over 100,000 tons of CO2 has been successfully injected at Allison Unit in New Mexico, USA during a ECBM project (Davison et al., 2001). However, the scope of geological trapping is currently economically limited to point sources of CO2 emissions that are near geological formation of choice. Continuous monitoring along with exhaustive geophysical and geochemical study is needed to make sure the injected CO2 stays in ground. 6. Drawbacks associated with artificial approaches of carbon sequestration Permanence of the stored carbon through abiological sequestration methods is of great critical concern. Ocean and geological storage of carbon dioxide is associated with future risk of leakage from the site of injection (Oldenburg and Unger, 2003). Sequestered CO 2 may leak back into the atmosphere and impose future climate damages. If CO2 migrates out of the receiving geological formation and rises to the surface, it could cause local ecological damage, primarily by displacing soil gas and affecting plant roots. Moreover, upward migration of injected CO 2 could contaminate hydrocarbon reservoirs or surface drinking water supplies. In rare cases, rapid escape of CO2 may cause asphyxiation or toxicity risks to local animal and human populations. The limnic eruption of CO2 during 1986 at Lake Nyos, West of Cameroon is the most evident example, which killed more than 1700 people (Clarke, 2001). Deep-sea organisms are highly sensitive to any environmental disturbances. Increased partial pressure of CO2 (hypercapnia) and decreased pH of seawater caused by CO2 dissolution may affect the whole marine biodiversity (Shirayama, 1997). The scientific community is trying to get rid of leaky sequestration approaches. Novel concepts are being contemplated to find the most environment-friendly way to sequester CO 2. This includes the art of exploiting natural biological ways of capturing carbon and storing it in the most eco-compatible way.
9

7. Biological ways of carbon sequestration Biological systems have solutions to the most dreaded problems of all times. The photosynthetic fixation of atmospheric CO2 in plants and trees could be of great value in maintaining a CO2 balance in the atmosphere. Algal systems, on the other hand, being more efficient in photosynthetic capabilities are the choice of research for solving global warming problem. The biomass thus produced could be used as fuel for various heating and power purposes. Mankind is indebted to microbes for bringing and maintaining stable oxygenic conditions on Earth. A proper understanding of microbial systems and their processes will help in stabilizing atmospheric conditions in future too. Investigations are in progress to exploit carbonic anhydrase and other carboxylating enzymes to develop a promising CO2 mitigation strategy. Recent work on biomimetic approaches using immobilized carbonic anhydrase in bioreactors has a big hope for the safe future. 7.1. Exploiting Photosynthesis 7.1.1. Terrestrial carbon sequestration The process of carbon assimilation by photosynthesis has made forests, trees and crops as the major biological scrubbers of CO2. Terrestrial biomes are potential CO2 sinks (Table. 4). 7.1.1.1. Forest lands Afforestation (Ozawa et al., 1995) and reforestation (Yokoyama, 1997) leads to a net increase in plant carbon stocks. A young growing forest sequesters more carbon than a matured one. Forest management can contribute to carbon sequestration by promoting forest growth and biomass accumulation (Sedjo, 2006). 7.1.1.2. Agricultural lands Croplands: Improved cropland management (including agronomy, nutrient management, tillage/residue management and water management) has significant carbon sequestration potential (Dendoncker et al., 2004; Smith, 2004). Worldwide adoption of best management practices (BMPs) can sequester a considerable part of the lost carbon back into croplands (Lal et al., 1998).

Table 4. Global potential carbon sequestration (CS) rates of terrestrial carbon sinks Terrestrial carbon sinks Potential CS (GtC / year)

10

Agricultural lands Biomass croplands Grasslands Rangelands Forests Deserts and degraded lands Terrestrial sediments Boreal peatlands wetlands Total Source: DOE, 1999. and

0.85-0.90 0.5-0.8 0.5 1.2 1-3 0.8-1.3 0.7-1.7 other 0.1-0.7 5.65-10.1

Grasslands: Grasslands cover about 70% of the worlds agricultural area (Soussana and Luscher, 2007). Recent studies have suggested that tropical grasslands and savannas sequester approximately 0.5 Gt of carbon annually (Scurlock et al., 1998). There are reports of increased CO2 uptake in calcareous grasslands during day time at elevated CO2 levels (Amthor, 1995; Stocker et al., 1997; Niklaus et al., 2000). Range lands: Rangelands (including grasslands, shrub lands, deserts and tundra) occupy about half of the worlds land area, and contain more than a third of above- and below-ground C reserves (Allen-Diaz, 1996). Grazing and burning in rangelands have resulted in increased soil organic carbon storage (Schuman et al., 2002; Rice, 2000) 7.1.1.3. Biomass croplands This includes croplands, which apart from assimilating carbon and increasing soil organic matter produce value added products (e.g. biofuels).

7.1.1.4. Urban shade trees

11

Urban trees play a major role in sequestering CO2. One tree in urban area is equivalent to three to five forest trees. The average sequestration rate of an urban tree of 50m2 crown area has been estimated to be about 11-19 kg/year (Nowak, 1994; Akbari, 2002). 7.1.1.5. Soil sequestration Plants assimilate carbon through the process of photosynthesis and return some of it to the atmosphere through respiration. After the death and decomposition of plants, carbon in the form of plant tissue is either consumed by animals or added to the soil as litter. The primary way that carbon is stored in the soil is as soil organic matter (SOM). SOM is a complex mixture of carbon compounds, consisting of decomposing plant and animal tissue, microbes (protozoa, nematodes, fungi, and bacteria) and carbon associated with soil minerals. Soils contain three times more carbon than the amount stored in living plants and animals (Houghton et al., 1985). Increasing the soil organic carbon (SOC) by 0.01% would nullify the annual increase in atmospheric carbon due to anthropogenic CO2 emissions (Cole et al., 1996). 7.1.1.5.1. Role of microbial communities in soil Microbial community structure and various microbial processes have been shown to directly affect carbon sequestration in soil agro ecosystems. A thorough understanding of microbial community structure and processes is required for enhanced carbon sequestration in agricultural soils. A balance between microbial community dynamics and formation and degradation of microbial byproducts maintains the soil carbon content. Soil microbes also indirectly influence C cycling by improving soil aggregation, which physically protects SOM. Consequently, the microbial contribution to C sequestration is governed by the interactions between the amount of microbial biomass, microbial community structure, microbial byproducts, and soil properties such as texture, clay mineralogy, pore-size distribution, and aggregate dynamics (Six et al., 2006). Fungi and bacteria have been found to be responsible for most of the carbon transformations and long-term storage of carbon in soils. However, chances of persistent C storage are more in fungi due to their complex chemical composition (Holland and Coleman, 1987; Guggenberger et al., 1999) and higher carbon utilization efficiency. In fact, increased fungal to bacterial activity has been shown to be associated with increased carbon stored in soil (Bailey et al., 2002).

7.1.1.6. Carbon concentrating mechanisms (CCM) Photoautotropic organisms ranging from bacteria to higher plants have evolved with unique carbon concentrating mechanism (CCM) in response to the declining levels of CO 2 in their surrounding environment. It is proposed that ribulose-1,5-biphosphate carboxylase/oxygenase

12

(Rubisco) co-evolved during the process. The organization of the carboxysomes in prokaryotes and of the pyrenoids in eukaryotes, and the presence of membrane mechanisms for inorganic carbon (Ci) transport are central to the concentrating mechanisms (Kaplan and Reinhold, 1999). There can be different types of CCM based on the biochemical mechanisms in different photoautotropic organisms such as C4 photosynthesis and crassulacean acid metabolism (CAM) in terrestrial higher plants, active transport of inorganic carbon (Ci) primarily in cyanobacteria and CO2 concentration following acidification in a compartment adjacent to Rubisco found in some eukaryotic algae (Moroney and Ynalvez, 2007). Higher terrestrial plants having crassulacean acid metabolism (CAM) primarily capture CO 2 through PEP carboxylase located in the cytosol of their mesophyll cells. PEP carboxylase uses bicarbonate as its primary substrate for fixation of CO2 into oxaloacetate, thus CO2 entering from the external environment must be hydrated rapidly by a carbonic anhydrase (CA) and converted to bicarbonate. C4 carboxylic acids such as malate or aspartate formed in the mesophyll cell cytosol serve as the intermediate CO2 pool. CCM found in eukaryotic algae relies on the pH gradient set up across the chloroplast thylakoid membrane in the light. Light-driven photosynthetic electron transport sets up a pH around 8.0 in chloroplast stroma and a pH between 4 to 5 inside the thylakoid lumen. Under these conditions, bicarbonate is the predominant species of Ci in the chloroplast stroma, while CO2 is the most abundant form of Ci in the thylakoid lumen. Bicarbonate transporters on thylakoid membrane are proposed to help bicarbonate transport inside thylakoid lumen where it is converted into CO2 with the help of carbonic anhydrase (Pronina et al., 1981; Pronina and Semenenko, 1990; Moroney and Ynalvez, 2007). The CCM mechanisms make it possible for cells to enhance the delivery of CO 2 to ribulose-1,5biphosphate carboxylase/oxygenase (Rubisco) and limit the oxygenase activity of this enzyme (Raven and Falkowski, 1999; Omata et al., 2001; Miyachi et al., 2003; Ogawa and Kaplan, 2003). Cyanobacterial CCM is proposed to have developed in the Proterozoic era in response to falling pCO2 and rising pO2 levels. The CO2 taken up by the cyanobacterial cell is converted into HCO3- by extracellular carbonic anhydrase that diffuses into the carboxysome. Here HCO 3- is once again converted back into CO2 by intracellular carbonic anhydrase. This releases OH-, which raises the sheath pH and hence provides optimum conditions for maximal sheath calcification by cyanobacteria (Riding, 2006). CCM has significant role in capturing CO 2 by the process of biocalcification. 7.1.1.7. Algal photosynthesis Photosynthesis is much more efficient in microalgae than in terrestrial C3 and C4 plants (Kodama et al., 1993). This high efficiency is again due to the presence of both intracellular and

13

extracellular carbonic anhydrases and the CO2 concentrating mechanism (Miyashi et al., 1997). The present focus is on exploiting the ability of microalgae to convert solar energy and CO2 into O2 and carbohydrates. Considerable efforts have been made for CO2 fixation along with valuable material production by mass cultivation of algal cultures. To advance both the near- and longterm development and applications of microalgae for biofixation of CO2 and GHG mitigation, the U.S. Department of Energy (DOE) and EniTecnologie, the R&D arm of the Italian oil company Eni, with the assistance of the IEA Greenhouse Gas R&D Programme, in Cheltenham, Great Britain, organized the International Network for Biofixation of CO2 and Greenhouse Gas Abatement with Microalgae. Microalgal mass cultures can use CO2 from power plant flue gases for the production of biomass. The algal biomass thus produced can directly be used as health food for human consumption, as animal feed or in aquaculture, for biodiesel production or as fertilizer for agriculture. A fast growing marine green alga Cholococcum littorale is reported to tolerate high concentrations of CO2 (Kodama et al., 1993 ). Yun et al. (1996 and 1997) used waste water containing phosphate (46 g m-3) from a steel plant to raise cultures of the photosynthetic microalga Chlorella vulgaris. Flue gas containing 15% CO2 was supplemented further to get a CO2 fixation rate of 26 g CO2 m3 -1 h . Research is in progress on the development of a novel photobioreactors for enhanced CO 2 fixation and CaCO3 formation. CO2 fixation rate was increased from 80 to 260 mg l -1h-1 by using Chlorella vulgaris in a newly developed membrane-photobioreactor (Cheng et al., 2006). A novel multidisciplinary process has recently been proposed that uses algal biomass in a photobioreator to produce H2 apart from sequestering CO2 (Skjanes et al., 2007). Enhanced growth rate of marine macroalgae Gracilaria sp. and G. chilensis has been observed by increasing CO2 concentration from 650 ppm to 1250 ppm (Gao et al., 1993). Gao and McKinley (1994) proposed that this macroalgal culture could make important contributions to both biomass production for chemicals and fuel and CO2 remediation. 7.2. Non-Photosynthetic ways of capturing carbon 7.2.1. Methanogenic and acetogenic bacteria Nonphotosynthetic CO2 fixation occurs widely in nature by the methanogenic archaebacteria. These are obligate anaerobes that grow in freshwater and marine sediments, peats, swamps and wetlands, rice paddies, landfills, sewage sludge, manure piles, and the gut of animals. Methanogens are responsible for more than half of the methane released to the atmosphere. These methanogenic bacteria grow optimally at temperatures between 20 oC and 95 oC. Carbon monoxide dehydrogenase and/or acetyl-CoA synthase aid them to use carbon monoxide or carbon dioxide along with hydrogen as their sole energy source.

14

Waste gases from blast furnaces containing oxides of carbon were used for converting them into higher- Btu (more calorific value) methane using thermophilic methanogens (Bugante et al., 1989). A column bioreactor operated at 55 C and pH 7.4 was used for the process. A mixture of three culture of bacteria, viz. Rhodospirillum rubrum, Methanobacterium formidium and Methanosarcina barkeri was used for complete bioconversion of oxides of carbon to methane (Klassen et al., 1990). Acetogenesis, on the other hand, is involved in the recycling of 10 to 20% of the carbon on earth (Gollin et al., 1997). 7.2.2. Carbon sequestration using heterotrophic bacteria The concept of CO2 fixation in certain representatives of heterotrophic bacteria was first proposed by Wood and Werkman (1941). The idea, however, faced lot of criticism, as with many new findings in the scientific world. While working on propionic acid bacteria they proposed that CO2 and pyruvate combined to form oxaloacetate, which was later called as the WoodWerkman reaction. The same pathway can be exploited now for capturing carbon using heterotrophic bacteria. Carbonic anhydrases play a critical role in concentrating CO2 inside the cell. The capability of carbonic anhydrases to convert CO2 in bicarbonate may be utilized by carboxylases such as phosphoenolpyruvate (PEP) carboxylase and pyruvate carboxylase, to form oxaloacetate (Norici and Giordano, 2002). Such anapleurotic pathway exists in organisms to compensate for the loss of oxaloacetate siphoned off for the synthesis of amino acids of aspartate family (Fig. 2). Heterotrophic bacteria having maximal carbonic anhydrase and phosphoenolpyruvate carboxylase and/or pyruvate carboxylase titers may be raised in fermentors, and these can be flushed with flue gases with CO2 concentration to produce useful metabolites such as oxaloacetate and amino acids. Extensive research has been done on the production of glutamic acid and lysine by Corynebacterium glutamicum. The presence of both phosphoenolpyruvate carboxylase and pyruvate carboxylase and the PEPpyruvateoxaloacetate node (Sauer and Eikmanns, 2005) makes this bacterium suitable for fixing carbon in the form of amino acids. An increased bicarbonate supply by the action of carbonic anhydrases (in elevated CO 2 conditions) to phosphoenolpyruvate and pyruvate carboxylases may enhance their activity, thereby making the conditions favorable for enhanced lysine production. Previous studies on PEP and pyruvate carboxylase activity in relation to lysine production (Gubler et al., 1994; Jetten et al., 1994) supports the assumption. Work is in progress in our laboratory at South Campus of the University of Delhi to understand the effect of different levels of carbon dioxide on carbonic anhydrase, phosphoenolpyruvate carboxylase and pyruvate carboxylase titres and hence their overall effect on lysine production. Dual benefit of carbon sequestration along with useful byproduct formation makes this approach very attractive.

15

PEPC

PC

Fig. 2. Simplified TCA cycle showing the anapleurotic sequence involved in feeding of oxaloacetate via the activity of Carbonic Anhydrase (PEPC: Phosphoenolpyruvate Carboxylase; PC: Pyruvate Carboxylase)

7.2.3. The Biomimetic Approach/ Biomimetic remediation of CO2 Biomimetic approach involves identification of a biological process or structure and its application to solve a nonbiological problem (Bond et al., 2001). It has emerged as an environment friendly process, which can be operated at near ambient temperature and pressure with no costly CO2 concentration and compression steps. Microbes, being widespread in nature, play a major role in chemical cycles that influence atmosphere-hydrosphere composition and are extensively involved in the production and accumulation of various sediments deep inside the oceans. Containing about 150,000 Gt of CO2, carbonate minerals constitute the Earths largest reservoir of CO2 (Liu et al., 2005). Bacteria are the key organisms in the formation of microbial carbonates. Mineral carbonation has emerged as a new carbon capture and storage technology in the past few years. The idea of applying carbonation reactions for CO2 storage was proposed by Seifritz (1990). Carbonic anhydrases are the fastest enzymes known for their capability and efficiency for converting carbon dioxide into bicarbonates. Gillian M. Bond of New Mexico Tech, USA, started working on this enzyme for mineral CO2 sequestration since 2001. The process of carbon dioxide fixation can be carried out successfully with a stream of carbon dioxide (from flue gases) in a bioreactor. Various methods for carbonic anhydrase immobilization are being attempted for the development of an efficient biodegradable matrix that can ensure maximal activity along with its long term use in bioreactors for sequestration purposes. Carbonic anhydrase was recently immobilized in chitosan-coated alginate beads and the mineralization was studied by FT-IR (Simsek-Ege et al., 2002). A novel trickling spray

16

reactor employing immobilized carbonic anhydrase has been developed that enables concentration of CO2 from the emission stream (Bhattacharya et. al., 2004). Carbonic anhydrase is one of the fastest enzymes, which make mass transfer from the gas phase to aqueous phase. This biocatalytic fixation of carbon could be the answer to tackle atmospheric pollution. New cation sources are being identified for carbonate formation. Seawater is a good source for cations but its availability is limited to coastal locations only. Waste brines from industries and produced waters from the oil and gas industries have emerged as promising sources of cations for biomimetic carbonate formation. Development of an efficient process can sequester approximately 3.49 Mt (millions of metric tons) CO2 /year utilizing about 2.07 Mt calcium ions and 0.67 Mt magnesium ions from the waters produced from New Maxico and Permian Basin regions of texas (Liu et al., 2005). Most of the work on biomimetic sequestration uses carbonic anhydrase from animal sources. However, there is a need for a thermophilic carbonic anhydrase that sustains high pressure if we really want to use brine as the most favorable cationic source for mineral carbonation under in the deep-sea environment. A gene encoding a putative - type carbonic anhydrase in the methanoarchaeon Methanobacterium thermoautotrophicum has been expressed in E. coli and found to encode a thermostable (up to 75C) carbonic anhydrase (Smith and Ferry, 1999). Its activity at different hydrostatic pressures needs to be studied for its biomimetic applications in carbon sequestration. The biomimetic approach has now also been applied in relation to geological sequestration. A closed-loop fossil-fuel carbon cycle has been proposed to be developed, in which microbial consortium (comprising of methanogens) could be used to convert CO2 to methane at a commercially useful rate. This can be used either in a geological setting (following injection of CO2 into depleted oil and gas well, saline aquifer, etc.) or above ground in rapid-contact reactors (Beecy et al., 2000; Medina et al., 2001). Possibility of an on-site scrubber that would provide a plant-by-plant solution to CO 2 sequestration, apart from eliminating the concentration and transportation costs is the potential advantage of the biomimetic approach. Conclusions Several novel concepts and techniques are being attempted for a safe and permanent capture of CO2. Routine abiotic methods although appear promising at prima facie but costly concentration and transportation steps along with future leakage risks have led to focus on new biotic methods. Evolution has equipped plants and various domains of microbial life with different mechanisms for carbon fixation. The present need is to exploit these biological mechanisms along with existing biochemical engineering techniques for long term CO2 sequestration. Exhaustive study needs to be done on various metabolic pathways that employ carboxylases. Behavior of enzymes like carbonic anhydrase and Rubisco with gases other than CO2 in flue gas must be understood.
17

Despite finite sink capacity, biological approaches provide a natural and cost-effective method of carbon sequestration. Biotic and abiotic approaches have their own merits and demerits, and they are complementary and have the potential to mitigate the risks of climate change. The World Environment Day slogan for 2008 Kick the Habit! Towards a Low Carbon Economy shows a growing concern and recognition that climate change has become the defining issue of present era. Countries, companies and communities throughout the world are focusing attention on reducing the greenhouse gas emissions. Acknowledgements We gratefully acknowledge the financial assistance form the Department of Biotechnology (DBT), Government of India during the course of this manuscript preparation. References Akbari, H. (2002). Shade trees reduce building energy use and CO 2 emissions from power plants. Environmental Pollution, 116: S119-S126. Allen-Diaz, B. (1996). Rangelands in a changing climate: impacts, adaptations, and mitigation. In: Watson, R.T., Zinyowera, M.C. and Moss, R.H. (editors), Climate Change 1995. Impacts, Adaptations, and Mitigation of Climate Change: Scientific-Technical Analyses. Published for the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp. 131-158. Amthor, J.S. (1995). Terrestrial higher-plant response to increasing atmospheric [CO2] in relation to the global carbon cycle. Global Change Biology, 1: 243-274. Ansolabehere, S., Beer, J., Deutch, J., Ellerman, D., Herzog, H., Jacoby, H.D., Joskow, P.L., McRae, G., Lester, R., Moniz, E.J., Steinfold, E. and Katzer, J. (2007). MIT study on the future of coal, Massachusetts Institute of Technology, pp. ix. Aya, I., Kojima, R., Yamane, K., Brewer, P. G. and Pelter, E. T. (2003) In situ experiments of cold CO2 release in mid-depth. Proceedings of the International Conference on Greenhouse Gas Control Technologies, 30th September-4th October, Kyoto, Japan. Aya, I., Yamane, K. and Nariai, H. (1997). Solubility of CO2 and density of CO2 hydrate at 30 MPa. Energy, 22: 263-271. Bailey, V.L., Smith, J.L. and Bolton (Jr.). H. (2002). Fungal-to-bacterial ratios in soils investigated for the enhanced C sequestration. Soil Biology & Biochemistry, 34: 997-1007. Barry, R.G. and Charley, R.J. (1992). Atmosphere, weather and climate. Routledge, London, sixth edition.

18

Beecy, D. J., Ferrell, F. M. and Carey, J. K. (2000). Paper presented at the National Meeting of American Chemical Society, Washington, DC. Bhattacharya, S., Nayak, A., Schiavone, M. and Bhattacharya, S. K. (2004). Solubilization and Concentration of Carbon Dioxide: Novel spray reactors with Immobilized Carbonic Anhydrase. Biotechnology and Bioengineering, 86 : 37-46. Bond, G. M., Stringer, J., Donald, K., Brandvold, F., Simsek, A., Medina, M. G. and Egeland G. (2001). Development of integrated system for biomimetic CO2 sequestration using the enzyme Carbonic anhydrase. Energy & Fuels, 15: 309-316. Brewer, P.G., Peltzer, E.T., Friederich, G., Aya, I. and Yamane, K. (2000). Experiments on the ocean sequestration of fossil fuel CO2: pH measurements and hydrate formation. Marine Chemistry, 72: 83-93. Brown, J. L., Li, S. H. and Bhagabati, N. (1999) Long-Term Trend Toward Earlier Breeding in an American Bird: A Response to Global Warming? Proceedings of the National Academy of Sciences of the United States of America, 96(10): 5565-5569. Ceulesmans, R., Jach, M.E., Van de Velde, R., Lin J.X. and Stevens, M. (2002). Elevated atmospheric CO2 alters wood production, wood quality and wood strength of Scots pine (Pinus sylvestris L) after three years of enrichment. Global Change Biology, 8: 153-162. Cheng, L., Zhang, L., Chen. H. and Gao, C. (2006). Carbon dioxide removal from air by microalgae cultured in a membrane-photobioreactor. Separation and Purification Technology, 50: 324-329. Clarke, T. (2001). Taming Africas killer lake. Nature, 409: 54-55. Cole, C. V., Cerri, C., Minami, K., Mosier, A., Rosenberg, N., Sauerbeck, D., Dumanski, J., DuxburyJ., Freney, J., Gupta, R., Heinemeyer, O., Kolchugina, T., Lee, J., Paustian, K., Powlson, D., Sampson, N., Tiessen, H., van Noordwijk, M., and Zhao, Q. (1996) Agricultural Options for Mitigation of Greenhouse Gas Emissions. In: Watson, R. T., Zinyowera, M., and Moss, R. H. (editors.), Climate Change 1995. Impacts, Adaptations and Mitigation of Climate Change: Scientific-Technical Analyses, IPCC Working Group II, Cambridge University Press, Cambridge, U.K., pp. 745-771. Collins, L. (1998). Renewable energy from wood and paper: Technological and cultural implications. Technology in society, 20: 157-177. Commonwealth Scientific and Industrial Research Organization (2007). CO2 emissions increasing faster than expected, Media release. http://www.csiro.au/news/GlobalCarbonProject-PNAS.html (accessed on March 21, 2008)

19

Crabbe, M. J. C. (2007). Global warming and coral reefs: Modelling the effect of temperature on Acropora palmata colony growth. Computational Biology and Chemistry, 31: 294-297. Curtis, P.S. and Wang, X. (1998). A meta-analysis of elevated CO2 effects on woody plant mass, form and physiology. Oecologia, 113: 299-313. Davison J., Freund P. and Smith A. (2001). Putting carbon back into the ground; IEA Greenhouse Gas R&D Programme. Dendoncker, N., Wesemael, B. V., Rounsevell, D.A., Roelandt, C. and Lettens, S. (2004). Belgiums CO2 mitigation potential under improved cropland management. Agriculture, Ecosystems and Environment, 103: 101-116. Department of Energy (DOE). (1999). Report on carbon sequestration research and development. Washington, D.C., USA. DOE (Department of Energy). (1993). A research needs assessment for the capture, utilization and disposal of carbon dioxide from fossil fuel fired power plants. DOE/ER-30194, Washington, D.C., USA. Donner, S.D., Skirving,W.J., Little, C.M., Oppenheimer, M. and Hoegh-Guldberg, O. (2005). Global assessment of coral bleaching and required rates of adaptation under climate change. Global Change Biology, 11; 2251-2265. Dyurgerov, M.B. and Meier, M.F. (2000). Twentieth century climate change: Evidence from small glaciers. Proceedings of the National Academy of Sciences, 97(4): 1406-1411. Eamus, D. and Jarvis, P.G. (1989). The direct effects of increase in the global atmospheric CO 2 concentration on natural and commercial temperate trees and forests. Advances in Ecological Research, 19:1-55. Friedlingstein, P. and Solomon, S. (2005). Contributions of past and present human generations to committed warming caused by carbon dioxide. Proceedings of the National Acadamy of Sciences, USA, 102(31): 10832-10836. Fer, I. and Haugan, P.M. (2003). Dissolution from a liquid CO2 lake disposed in the deep ocean. Limnology and Oceanography, 48(2): 872-883. Gale, J. (2004). Geological storage of CO2: What do we know, where are the gaps and what more needs to be done? Energy, 29: 1329-1338. Gao, K. and McKinley, K.R. (1994). Use of macroalgae for marine biomass production and CO 2 remediation. Journal of Applied Phycology, 6: 45-60.

20

Gao, K., Aruga, Y., Asada, K., and Kiyohara, M. (1993). Influence of enhanced CO 2 on growth and photosynthesis of the red algae Gracilaria sp. and G. chilensis. Journal of Applied Phycology, 5: 563-571. Gentzis, T. (2000). Subsurface sequestration of carbon dioxide an overview from an Alberta (Canada) perspective. International Journal of Coal Geology, 43: 287-305. NOAA Climate Monitoring and Diagnostics Laboratory. (2003). Cooperative Atmospheric Data Integration Project: GLOBALVIEW-CO2, NOAA Climate Monitoring and Diagnostics Laboratory, Boulder, CO. ftp.cmdl.noaa.gov, path: ccg/co2/GLOBALVIEW (accessed on October 14, 2007). Gollin, D.J., Li, X.L., Liu, S.M. and Ljungdahl, L.G. (1997). Primary structure of the NADPdependent formate dehydrogenase of Clostridium thermoaceticum, a tungsten-selenium-ironsulfur containing enzyme, in Abstract, 4th International conference on Carbon Dioxide Utilization, Kyoto, Japan. O-31. Gubler, M., Park, S.M., Jetten, M., Stephanopoulos G. and Sinskey, A.J. (1994) Effects of phosphoenol pyruvate carboxylase deficiency on metabolism and lysine production in Corynebacterium glutamicum. Applied Microbiology Biotechnology, 40: 857-863. Guggenberger, G., Frey, S.D., Six, J., Paustian, K. and Elliot, E.T. (1999). Bacterial and fungal cell-wall residues in conventional and no-tillage agroecosystems. Soil Science Society of America Journal, 63: 1188-1198. Hadley Centre, United Kingdom (2006). Annual land, air and sea surface temperature anomalies,http//www.metoffice.gov.uk/research/hadleycentre/CR_data/Annual/land+sst_web.txt (accessed on May 15, 2006). Halmann, M.M. and Steinberg, M. (1999). Greenhouse gas carbon dioxide mitigation: Science and technology, Lewis publishers, Boca Raton, Florida. pp. 1-3. Hansen, J. and Sato, M. (2004). Greenhouse gas growth rates. Proceedings of the National Acadamy of Sciences, 101(46): 16109-16114. Harrington, L. and Foster, R. (1999). Australian residential building sector greenhouse gas emissions 1990-2010. Final Report, Energy Efficient Strategies. Australian Greenhouse Office. Held, I. M. and Soden, B. J. (2006). Robust responses of the hydrological cycle to global warming. Journal of Climate, 19: 5686-5699. Holland, E.A. and Coleman, D.C. (1987). Litter placement effects on microbial and organic matter dynamics in an agroecosystem. Ecology, 68: 425-433.

21

Holloway, S (2005). Underground sequestration of carbon dioxidea viable greenhouse gas mitigation option. Energy, 30: 2318-2333. Houghton, R.A.., R.D. Boone, J.M. Melillo, C.A. Palm, G.M. Woodwell and N. Myers. (1985). Net flux of carbon dioxide from terrestrial tropical forests in 1980. Nature, 316: 617-620. House, K.Z., Schrag D.P., Harvey,C.F., and Lackner K.S. (2006). Permanent carbon dioxide storage in deep-sea sediments, Proc. Natl. Acad. Sci. USA, vol. 103 no. 33, 12291-12295. Intergovernmental Panel on Climate Change. (2005). IPCC Special Report on Carbon Dioxide Capture and Storage. http://www.ipcc.ch/ (accessed on December 15th, 2007). Intergovernmental Panel on Climate Change. (2007). Fourth Assessment Report-Observations of Climate Change. Intergovernmental Panel on Climate Change, http//www.ipcc.ch (accessed on November 17, 2007). Jetten, M.S.M., Pitoc, G.A., Follettie, M.T. and Sinskey, A.J. (1994). Regulation of phospho(enol)-pyruvate and oxaloacetate-converting enzymes in Corynebacterium glutamicum, Applied Microbiology Biotechnology, 41: 47-52. Johnsen, K.H. (1993). Growth and ecophysiological responses of black spruce seedlings to elevated CO2 under varied water and nutrient additions. Canadian Journal of Forest Research, 23: 1033-1042. Kaplan, A. and Reinhold, L. (1999). CO2 concentrating mechanisms in photosynthetic microorganisms. Annual Review of Plant Physiology and Plant Molecular Biology, 50: 539570. Klara, S. M., Srivastava, R. D. and McIlvried, H. G. (2003). Integrated collaborative technology development program for CO2 sequestration in geologic formations - United States Department of Energy R&D. Energy Conversion and Management, 44: 2699-2712. Kleypas, J.A., Feely, R.A., Fabry, V.J., Langdon, C., Sabine, C.L. and Robbins, L.L. (eds). (2006). Impacts of Increasing Ocean Acidification on Coral Reefs and other Marine Calcifiers. Report from International Workshop on the Impacts of Increasing Atmospheric CO2 on Coral Reefs and Other Marine Calcifiers, 18-20 April 2005, St Petersburg, Florida sponsored by NSF/NOAA/USGS. Kobayashi, Y., (1995). Physical behavior of liquid CO2 in the ocean. In: Handa, N., Ohsumi, T. (editors), Direct Ocean Disposal of Carbon Dioxide. Terra Scientific Publishing Company (TERRAPUB), Tokyo. pp. 165-181. Kodama, M., Ikemoto, H., and Miyachi, S. (1993). A new species of highly CO2 tolerent fast growing marine microalga suitable for high-density culture. Journal of Marine Biotechnology, 1 : 21-25.
22

Kolchugina, T. P. and Vinson, T. S. (1993) Climate warming and the carbon cycle in the permafrost zone of the former Soviet Union. Permafrost and Periglacial Processes, 4(2): 149163. Lal, R., Kimble, J.M., Follett, R.F. and Cole, C.V. (1998). The Potential of U.S. Cropland to Sequester Carbon and Mitigate the Greenhouse Effect. Ann Arbor Press, Chelsea, MI, pp. 1. Liu, N., Bond, G.M., Abel, A., McPherson, B. J. and Stringer, J.(2005) Biomimetic sequestration of CO2 in carbonate form: Role of produced waters and other brines. Fuel Processing Technology, 86: 1615-1625. Marchetti, C. (1977). On geoengineering and the CO2 problem. Climate Change, 1: 59-68. Margolis, H.A. and Vzina, L.P. (1990). Atmospheric CO2 enrichment and the development of frost hardiness in containerized black spruce seedlings. Canadian Journal of Forest Research, 20: 1392-1398. Marsh, G.E. (2002). A global warming primer, The National Centre for Public Policy Research. USA, 420: 4-5. McMichael, A. and Woodruff, R. (2004). Climate change and risk to health: The risk is complex, and more than a sum of risks due to individual climatic factors. British Medical Journal, 329(7480): 1416-1417. Medina, M. G., Bond, G. M. and Stringer, J. (2001). An overview of carbon dioxide sequestration. The Electrochemical Society Interface, pp 26-30. Megonigal, J.P., Hines, M.E. and Visscher, P.T. (2004). Anaerobic metabolism: Linkages to trace gases and aerobic processes. In Schlesinger, W.H. (editor). Biogeochemistry. ElsevierPergamon, Oxford, UK., pp 317-424. Meier, M. F. and Wahr, J. M. (2002). Sea level is rising: Do we know why? Proceedings of the National Academy of Sciences, 99(10): 65246526. Mirza, M. M. Q. (2002). Global warming and changes in the probability of occurrence of floods in Bangladesh and implications. Global Environmental Change, 12: 127-138. Miyachi, S., Iwasaki, I. and Shiraiwa, Y. (2003). Historical perspective on microalgal and cyanobacterial acclimation to low- and extremely high-CO2 conditions. Photosynthesis Research, 77: 139-153. Miyashi, S., Kurano, N., Qiang, H., and Iwasaki, I. (1997). Carbon dioxide and microalgae, in Abstr. 4th International conference on carbon dioxide utilisation, Kyoto, Japan.
23

Moroney, J. V. and Ynalvez, R. A (2007) Proposed carbon dioxide concentrating mechanism in Chlamydomonas reinhardtii. Eukaryotic Cell, 6(8): 1251-1259. Murray, M.B., Smith, R.I., Leith, I.D., Fowler, D., Lee, H.S., Friend A.D. and Jarvis, P.G. (1994). Effects of elevated CO2, nutrition and climatic warming on bud phenology in Sitka spruce (Picea sitchensis) and their impact on the risk of frost damage. Tree Physiology, 14: 691706. Nakashiki, N. (1997). Lake-type storage concepts for CO2 disposal option. Waste Management, 17(5): 361-367. National Research Council (NRC). 2001. Climate Change Science: An Analysis of Some Key Questions. National Academy Press, Washington, D.C. Niklaus, P. A., Stocker, R., Krner, C. and Leadley, P. W. (2000). CO2 flux estimates tend to overestimate ecosystem C sequestration at elevated CO2. Functional Ecology, 14: 546-559. Nordbotten, J. M., Celia, M. A. and Bachu S. (2005). Injection and storage of CO 2 in deep saline aquifers: analytical solution for CO2 plume evolution during injection. Transport in Porous Media, 58: 339-360. Norici, A. and Giordano, M. (2002). Anaplerosis in microalgae. Recent Research and Development in Plant Physiology, 3: 153-164. Nowak, D.J. (1994). Atmospheric carbon dioxide reduction by Chicagos urban forest. In: McPherson, E.G., Nowak, D.J., Rowntree, R.A (editors), Chicagos Urban Forest Ecosystem: Results of the Chicago Urban Forest Climate Project (NE-186). Forest Service, US Dept. of Agriculture, Department of Agriculture, Radnor, PA., pp. 83-94. Ogawa, T., and Kaplan, A. (2003). Inorganic carbon acquisition systems in cyanobacteria. Photosynthesis Research, 77: 105-115. Ohsumi, T. (1993). Prediction of solute carbon dioxide behavior around a liquid carbon dioxide pool on deep ocean basin. Energy Conversion and Management, 34: 1059-1064. Omata, T., Gohta, S., Takahashi, Y., Harano, Y., and Maeda, S. (2001). Involvement of a CbbR homolog in low CO2-induced activation of the bicarbonate transporter operon in cyanobacteria. Journal of Bacteriology, 183: 1891-1898. O'Neill, B. C. and Oppenheimer, M. (2002). "Climate change - Dangerous climate impacts and the Kyoto protocol." Science 296 (5575): 1971-1972. Ortiz, R., Sayre, K. D., Govaerts, B., Gupta, R., Subbarao, G.V., Ban, T., Hodson, D., Dixon, J. M., Ortiz-Monasterio, J. I. and Reynolds, M. (2008). Climate change: Can wheat beat the heat? Agriculture, Ecosystems and Environment ( In Press).

24

Ozaki, M., Minamiura, J., Kitajima, Y., Mizokami, S., Takeuchi, K., and Hatakenaka, K. (2001). CO2 ocean sequestration by moving ships. Journal of Marine Science and Technology, 6: 51-58. Ozawa, H., Okabayashi, T., Komiyama, H. and Kaya, Y. (1995). Research of arid land afforestation technologies for carbon dioxide fixation. Energy Conversion and Management, 36: 911-914. Pronina, N. A. and Semenenko, V. E. (1990). Membrane-bound carbonic anhydrase takes part in CO2 concentration in algal cells. Current Research in Photosynthesis, 4:489-492. Pronina, N. A., Ramazanov, Z. M. and Semenenko, V. E. (1981). Carbonic anhydrase activity of Chlorella cells as a function of CO2 concentration. Soviet Plant Physiology, 28: 345-351. Raven, J. A. and Falkowski, P. G. (1999). Oceanic sinks for atmospheric CO 2. Plant, Cell and Environment, 22: 741-755. Rawaswamy, V., Boucher, O., Haigh, J., Hauglustaine, D., Haywood, J., Myhre, G., Nakajima, T., Shi, G. Y. and Solomon, S. (2001). In: Houghton, J. T., Ding, Y., Griggs, D. J., Noguer, M., van der Linden, P. J., Dai, X., Maskell, K. & Johnson, C. A. (editors), Climate Change 2001: The Scientific Basis: Contributions of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge Univ. Press, Cambridge, U.K., pp. 349416. Rehder, G., Kirby, S.H., Durham, W.B., Stern, L.A., Peltzer, E.T., Pinkston, J. and Brewer, P.G. (2004). Dissolution rates of pure methane hydrate and carbon dioxide hydrate in under-saturated sea water at 1000 m depth. Geochimica et Cosmochimica Acta, 68(2): 285-292. Riding , R. (2006). Cyanobacterial calcification, carbon dioxide concentrating mechanisms, and Proterozoic-Cambrian changes in atmospheric composition. Geobiology, 4: 299-316. Rice, C.W. (2000). Soil organic C and N in rangeland soils under elevated CO2 and land management. In: Proc., Advances in Terrestrial Ecosystem Carbon Inventory, Measurements, and Monitoring. 3-5 October 2000. USDA-ARS, USDA-FS, USDA-NRCS, US Dept. Energy, NASA, and National Council for Air and Stream Improvement, Raleigh, NC, pp. 83. Rogers, H.H., Ingham, G.E., Cure, J.D., Smith, J.M. and Surano, K.A. (1983). Response of selected plant species to elevated carbon dioxide in the field. Journal of Environmental Quality, 12: 569-574. Samuelson, L.J. and Seiler, J.R. (1994). Red spruce seedling gas exchange in response to elevated CO2, water stress and soil fertility treatments. Canadian Journal of Forest Research, 24: 954-959. Sauer, U. and Eikmanns, B. J. (2005).The PEPpyruvateoxaloacetate node as the switch point

25

for carbon flux distribution in bacteria. FEMS Microbiology Reviews, 29: 765-794. Schnur, R. (2002). The investment forecast. Nature, 415: 483-484. Schuman, G.E., Janzen, H.H. and Herrick J.E. (2002). Soil carbon dynamics and potential carbon sequestration by Rangelands. Environmental Pollution, 116: 391-396. Scurlock, J. M. O. and Hall, D. O. (1998). The global carbon sink: a grassland perspective. Global Change Biology, 4 (2): 229-233. Sedjo R. (2006). Forest and biological carbon sinks after Kyoto: Resources for the future, Washington, D.C., USA. www.rff.org (accessed on November 19, 2007). Seifritz, W. (1990). CO2 disposal by means of silicates. Nature, 345:486. Shirayama, Y. (1997). Biodiversity and biological impact of ocean disposal of carbon dioxide. Waste Management, 17: 381-384. Shitashima, K., Maeda, Y., Koike, Y., and Ohsumi, T. (2008). Natural analogue of the rise and dissolution of liquid CO2 in the ocean, International Journal of Greenhouse Gas, 2(1): 95-104. Simsek-Ege, F.A., Bond, G.M. and Stringer, J. (2002). Matrix molecular weight cut-off for encapsulation of carbonic anhydrase in polyelectrolyte beads. Journal of Biomaterials Science, 13: 1175-1187. Six, J., Frey, S. D., Thiet, R. K. and Batten, K. M. (2006). Bacterial and fungal contributions to carbon sequestration in agroecosystems. Soil Science Society of America, 70: 555-569. Skjanes, K., Lindblad, P. and Muller, J. (2007). BioCO2 - A multidisciplinary, biological approach using solar energy to capture CO2 while producing H2 and high value products. Biomolecular Engineering, 24: 405-413. Smith, K.S. and Ferry, J.G. (1999). A plant type class carbonic anhydrase from the thermophilic methanoarchaeon Methanobacterium thermoautotrophicum. Journal of Bacteriology, 181: 6247-6253. Smith, P. (2004). Carbon sequestration in croplands: the potential in Europe and the global context. Europian Journal of Agronomy, 20: 229-236. Soussana, J. F. and Luscher, A. (2007). Temperate grasslands and global atmospheric change: a review. Grass and Forage Science 62, 127-134. Stocker, R., Leadley, P.W. and Krner, C. (1997). Carbon and water fluxes in a calcareous grassland under elevated CO2. Functional Ecology, 11: 222-230.

26

Sutherst, R. W. (2004). Global change and human vulnerability to vector-borne diseases. Clinical Microbiology Reviews, 17(1): 136-173. Sydeman, W. J., Hester, M. M., Thayer, J. A., Gress, F., Martin, P. and Buffa, J. (2001). Climate change, reproductive performance and diet composition of marine birds in the southern California Current system, 19691997. Progress in Oceanography, 49; 309329. Takahashi, T. (2004). The fate of industrial carbon dioxide. Science, 305: 352-353. Tan, G. and Shibasaki, R. (2003). Global estimation of crop productivity and the impacts of global warming by GIS and EPIC integration. Ecological Modelling, 168: 357-370. Tsouris, C., Brewer, P.G., Peltzer, E., Walz, P., Riestenberg, D., Liang, L. and West O.R. (2004). Hydrate composite particles for ocean carbon sequestration: field verification. Environmental Science and Technology, 38(8): 2470-2475. Usui, N. and Ikenouchi, M. (1997). Biological CO2 fixation and utilization project by RITE.1. Highly-effective photobioreactor system. Energy Conversion and Management, 38: 487-492. U.N. Environment Programme (2008). Glaciers are melting faster than expected, UN Reports. ScienceDaily, http://www.sciencedaily.com /releases/2008/03/080317154235.htm (accessed on March 26, 2008). Vaughan, D. G., Marshall, G. J., Connolley, W. M., Parkinson, C. L., Mulvaney, R., Hodgson, D. A., King, J. C., Pudsey, C. J. and Turner, J. (2003). Recent rapid regional climate warming on the Antarctic Peninsula. Climatic Change, 60: 243274. Wannamaker, E.J. and Adams, E.E. (2002). Modelling descending carbon dioxide injections in the ocean. Proceedings of the 6th International Conference on Greenhouse Gas Control Technologies, 30th September-4th October, Kyoto, Japan. Wells, A.W., Diehl, J. R., Bromhal, G., Strazisar, B. R., Wilson, T. H. and White C. M. (2007). The use of tracers to assess leakage from the sequestration of CO 2 in a depleted oil reservoir, New Mexico, USA. Applied Geochemistry, 22: 996-1016. Wigley, T. M. L., R. Richels, and Edmonds, J. A. (1996). "Economic and environmental choices in the stabilization of atmospheric CO2 concentrations." Nature, 379 (6562): 240-243. Wood, H. G., Werkman, C. H., Hemingway, A., and Nier, A. O. (1941) Heavy Carbon as a tracer in heterotrophic carbon dioxide assimilation. Journal of Biological Chemistry, 139: 365 376. Wormworth, J. and Mallon, K. (2007). Bird Species and Climate Change: The Global Status Report. Climate Risk Pty Limited, Australia.

27

Yokoyama, S. (1997). Potential land area for reforestation and carbon dioxide mitigation effect through biomass energy conversion. Energy Conversion Management, 38: 569-573. Yun, Y.S., Lee, S.B., Park, J.M., Lee, C.I. and Yang., J.W. (1997). CO2 fixation by algal cultivation using wastewater nutrients. Journal of Chemical Technology and Biotechnology, 69: 451-455. Yun, Y.S., Park, J.M., and Yang, J.W. (1996). Enhancement of CO2 tolerance of Chlorella vulgaris by gradual increase of CO2 concentration. Biotechnology Techniques, 10: 713-716.

28

You might also like