Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

Carbon fixation

Carbon fixation or сarbon assimilation is


the process by which inorganic carbon
(particularly in the form of carbon
dioxide) is converted to organic
compounds by living organisms. The
compounds are then used to store
energy and as structure for other
biomolecules. The most prominent
example of carbon fixation is
photosynthesis; another form known as
chemosynthesis can take place in the
absence of sunlight.
Cyanobacteria such as these carry out
photosynthesis. Their emergence foreshadowed the
evolution of many photosynthetic plants, which
oxygenated Earth's atmosphere.

Organisms that grow by fixing carbon are


called autotrophs, which include
photoautotrophs (which use sunlight),
and lithoautotrophs (which use inorganic
oxidation). Heterotrophs are not
themselves capable of carbon fixation
but are able to grow by consuming the
carbon fixed by autotrophs or other
heterotrophs. "Fixed carbon", "reduced
carbon", and "organic carbon" may all be
used interchangeably to refer to various
organic compounds.[1]

Net vs. gross CO2 fixation

Graphic showing net annual amounts of CO2


fixation by land and sea-based organisms.

It is estimated that approximately 258


billion tons of carbon dioxide are
converted by photosynthesis annually.
The majority of the fixation occurs in
terrestrial environments, especially the
tropics. The gross amount of carbon
dioxide fixed is much larger since
approximately 40% is consumed by
respiration following photosynthesis.[1]
Given the scale of this process, it is
understandable that RuBisCO is the most
abundant protein on Earth.

Overview of pathways
Six autotrophic carbon fixation pathways
are known as of 2011. The Calvin cycle
fixes carbon in the chloroplasts of plants
and algae, and in the cyanobacteria. It
also fixes carbon in the anoxygenic
photosynthesis in one type of
proteobacteria called purple bacteria,
and in some non-phototrophic
proteobacteria.[2]

Of the five other autotrophic pathways,


two are known only in bacteria (the
reductive citric acid cycle and the 3-
hydroxypropionate cycle), two only in
archaea (two variants of the 3-
hydroxypropionate cycle), and one in
both bacteria and archaea (the reductive
acetyl CoA pathway).

Oxygenic photosynthesis
In photosynthesis, energy from sunlight
drives the carbon fixation pathway.
Oxygenic photosynthesis is used by the
primary producers—plants, algae, and
cyanobacteria. They contain the pigment
chlorophyll, and use the Calvin cycle to
fix carbon autotrophically. The process
works like this:

2H2O → 4e− + 4H+ + O2


CO2 + 4e− + 4H+ → CH2O + H2O

In the first step, water is dissociated into


electrons, protons, and free oxygen. This
allows the use of water, one of the most
abundant substances on Earth, as an
electron donor—as a source of reducing
power. The release of free oxygen is a
side-effect of enormous consequence.
The first step uses the energy of sunlight
to oxidize water to O2, and, ultimately, to
produce ATP

ADP + Pi ⇌ ATP + H2O

and the reductant, NADPH

NADP+ + 2e− + 2H+ ⇌ NADPH + H+

In the second step, called the Calvin


cycle, the actual fixation of carbon
dioxide is carried out. This process
consumes ATP and NADPH. The Calvin
cycle in plants accounts for the
preponderance of carbon fixation on
land. In algae and cyanobacteria, it
accounts for the preponderance of
carbon fixation in the oceans. The Calvin
cycle converts carbon dioxide into sugar,
as triose phosphate (TP), which is
glyceraldehyde 3-phosphate (GAP)
together with dihydroxyacetone
phosphate (DHAP):

3 CO2 + 12 e− + 12 H+ + Pi → TP + 4
H2O

An alternative perspective accounts for


NADPH (source of e−) and ATP:

3 CO2 + 6 NADPH + 6 H+ + 9 ATP + 5


H2O → TP + 6 NADP+ + 9 ADP + 8 Pi

The formula for inorganic phosphate (Pi)


is HOPO32− + 2H+. Formulas for triose
and TP are C2H3O2-CH2OH and C2H3O2-
CH2OPO32− + 2H+
Evolutionary considerations …

Somewhere between 3.8 and 2.3 billion


years ago, the ancestors of
cyanobacteria evolved oxygenic
photosynthesis,[3][4] enabling the use of
the abundant yet relatively oxidized
molecule H2O as an electron donor to the
electron transport chain of light-
catalyzed proton-pumping responsible
for efficient ATP synthesis.[5][6] When this
evolutionary breakthrough occurred,
autotrophy (growth using inorganic
carbon as the sole carbon source) is
believed to have already been developed.
However, the proliferation of
cyanobacteria, due to their novel ability to
exploit water as a source of electrons,
radically altered the global environment
by oxygenating the atmosphere and by
achieving large fluxes of CO2
consumption.[7]

CO2 concentrating mechanisms …

Many photosynthetic organisms have not


acquired CO2 concentrating mechanisms
(CCMs), which increase the
concentration of CO2 available to the
initial carboxylase of the Calvin cycle, the
enzyme RuBisCO. The benefits of a CCM
include increased tolerance to low
external concentrations of inorganic
carbon, and reduced losses to
photorespiration. CCMs can make plants
more tolerant of heat and water stress.

CO2 concentrating mechanisms use the


enzyme carbonic anhydrase (CA), which
catalyze both the dehydration of
bicarbonate to CO2 and the hydration of
CO2 to bicarbonate

HCO3− + H+ ⇌ CO2 + H2O

Lipid membranes are much less


permeable to bicarbonate than to CO2.
To capture inorganic carbon more
effectively, some plants have adapted the
anaplerotic reactions

HCO3− + H+ + PEP → OAA + Pi


catalyzed by PEP carboxylase (PEPC), to
carboxylate phosphoenolpyruvate (PEP)
to oxaloacetate (OAA) which is a C4
dicarboxylic acid.

CAM plants …

CAM plants that use Crassulacean acid


metabolism as an adaptation for arid
conditions. CO2 enters through the
stomata during the night and is
converted into the 4-carbon compound,
malic acid, which releases CO2 for use in
the Calvin cycle during the day, when the
stomata are closed. The dung jade plant
(Crassula ovata) and cacti are typical of
CAM plants. Sixteen thousand species of
plants use CAM.[8] These plants have a
carbon isotope signature of −20 to −10
‰.[9]

C4 plants …

C4 plants preface the Calvin cycle with


reactions that incorporate CO2 into one
of the 4-carbon compounds, malic acid
or aspartic acid. C4 plants have a
distinctive internal leaf anatomy. Tropical
grasses, such as sugar cane and maize
are C4 plants, but there are many
broadleaf plants that are C4. Overall,
7600 species of terrestrial plants use C4
carbon fixation, representing around 3%
of all species.[10] These plants have a
carbon isotope signature of −16 to −10
‰.[9]

C3 plants …

The large majority of plants are C3


plants. They are so-called to distinguish
them from the CAM and C4 plants, and
because the carboxylation products of
the Calvin cycle are 3-carbon
compounds. They lack C4 dicarboxylic
acid cycles, and therefore have higher
CO2 compensation points than CAM or
C4 plants. C3 plants have a carbon
isotope signature of −24 to −33‰.[9]

Bacteria and cyanobacteria …


Almost all cyanobacteria and some
bacteria utilize carboxysomes to
concentrate carbon dioxide.
Carboxysomes are protein shells filled
with the enzyme RuBisCO and a carbonic
anhydrase. The carbonic anhydrase
produces CO2 from the bicarbonate that
diffuses into the carboxysome. The
surrounding shell provides a barrier to
carbon dioxide loss, helping to increase
its concentration around RuBisCO.

Other autotrophic pathways

Reverse Krebs cycle …


The reverse Krebs cycle, also known as
reverse TCA cycle (rTCA) or reductive
citric acid cycle, is an alternative to the
standard Calvin-Benson cycle for carbon
fixation. It has been found in strict
anaerobic or microaerobic bacteria (as
Aquificales)[11] and anaerobic archea. It
was discovered by Evans, Buchanan and
Arnon in 1966 working with the
photosynthetic green sulfur bacterium
Chlorobium limicola.[12] The cycle
involves the biosynthesis of acetyl-CoA
from two molecules of CO2.[13] The key
steps of the reverse Krebs cycle are:

Oxaloacetate to malate, using NADH +


H+
Fumarate to succinate, catalyzed by an
oxidoreductase, Fumarate reductase

Succinate to succinyl-CoA, an ATP


dependent step

Succinyl-CoA to alpha-ketoglutarate,
using one molecule of CO2

Alpha-ketoglutarate to isocitrate, using


NADPH + H+ and another molecule of
CO2
Citrate converted into oxaloacetate
and acetyl-CoA, this is an ATP
dependent step and the key enzyme is
the ATP citrate lyase

This pathway is cyclic due to the


regeneration of the oxaloacetate.[14]

The reverse Krebs cycle is used by


microorganisms in anaerobic
environments. In particular, it is one of
the most used pathways in hydrothermal
vents by the Epsilonproteobacteria.[15]
This feature is very important in oceans.
Without it, there would be no primary
production in aphotic environments,
which would lead to habitats without life.
So this kind of primary production is
called "dark primary production".[16]

One other important aspect is the


symbiosis between
Gammaproteobacteria and Riftia
pachyptila. These bacteria can switch
from the Calvin-Benson cycle to the rTCA
cycle and vice versa in response to
different concentrations of H2S in the
environment.[17]

Reductive acetyl CoA pathway …


The reductive acetyl CoA pathway (CoA)
pathway, also known as the Wood-
Ljungdahl pathway, was discovered by
Harland G. Wood and Lars G. Ljungdahl
in 1965, thanks to their studies on
Clostridium thermoaceticum, a Gram
positive bacterium now named Moorella
thermoacetica.[18] It is an acetogen, an
anaerobic bacteria that uses CO2 as
electron acceptor and carbon source,
and H2 as an electron donor to form
acetic acid.[19][20][21][22] This metabolism
is wide spread within the phylum
Firmicutes, especially in the Clostridia.[19]

The pathway is also used by


methanogens, which are mainly
Euryarchaeota, and several anaerobic
chemolithoautotrophs, such as sulfate-
reducing bacteria and archaea. It is
probably performed also by the
Brocadiales, an order of Planctomycetes
that oxidize ammonia in anaerobic
condition.[13][23][24][25][26][27][28]
Hydrogenotrophic methanogenesis,
which is only found in certain archaea
and accounts for 80% of global
methanogenesis, is also based on the
reductive acetyl CoA pathway.

The Carbon Monoxide


Dehydrogenase/Acetyl-CoA Synthase is
the oxygen-sensitive enzyme that
permits the reduction of CO2 to CO and
the synthesis of acetyl-CoA in several
reactions.[29]

One branch of this pathway, the methyl


branch, is similar but non-homologous
between bacteria and archaea. In this
branch happens the reduction of CO2 to a
methyl residue bound to a cofactor. The
intermediates are formate for bacteria
and formyl-methanofuran for archaea,
and also the carriers, tetrahydrofolate
and tetrahydropterins respectively in
bacteria and archaea, are different, such
as the enzymes forming the cofactor-
bound methyl group.[13]

Otherwise, the carbonyl branch is


homologous between the two domains
and consists of the reduction of another
molecule of CO2 to a carbonyl residue
bound to an enzyme, catalyzed by the CO
dehydrogenase/acetyl-CoA synthase.
This key enzyme is also the catalyst for
the formation of acetyl-CoA starting from
the products of the previous reactions,
the methyl and the carbonyl
residues.[29][30]

This carbon fixation pathway requires


only one molecule of ATP for the
production of one molecule of pyruvate,
which makes this process one of the
main choice for chemolithoautotrophs
limited in energy and living in anaerobic
conditions.[13]
3-Hydroxypropionate bicycle …

The 3-Hydroxypropionate bicycle, also


known as 3-HP/malyl-CoA cycle, was
discovered by Helge Holo in 1989. It's a
pathway of carbon fixation and is utilized
by green non-sulfur phototrophs of
Chloroflexaceae family, including the
maximum exponent of this family
Chloroflexus auranticus by which this way
was discovered and demonstrated.[31]

The 3-Hydroxipropionate bicycle is


composed of two cycles and the name of
this way comes from the 3-
Hydroxyporopionate which corresponds
to an intermediate characteristic of it.
The first cycle is a way of synthesis of
glycoxilate. During this cycle two
bicarbonate molecules are fixed thanks
to the action of two enzymes: the Acetyl-
CoA carboxylase catalyzes the
carboxylation of the Acetyl-CoA to
Malonyl-CoA and Propionyl-CoA
carboxylase catalyses the carboxylation
of Propionyl-CoA to Methylamalonyl-CoA.
From this point a series of reactions lead
to the formation of glycoxylate which will
thus become part of the second
cycle.[32][33]

In the second cycle, glycoxilate is


approximately one molecule of Propionyl-
CoA forming Methylamalonyl-CoA. This,
in turn, is then converted through a series
of reactions into Citramalyl-CoA. The
Citramalyl-CoA is split into pyruvate and
Acetyl-CoA thanks to the enzyme MMC
lyase. At this point the pyruvate is
released, while the Acetyl-CoA is reused
and carboxylated again at Malonyl-coa
thus reconstituting the cycle.[34]

19 are the total reactions involved in 3-


Hydroxypropionate bicycle and 13 are the
multifunctional enzymes used. The
multifunctionality of these enzymes is an
important feature of this pathway which
thus allows the fixation of 3 bicarbonate
molecules.[34]
It is a very expensive way: 7 ATP
molecules are used for the synthesis of
the new pyruvate and 3 ATP for the
phosphate triose.[33]

An important characteristic of this cycle


is that it allows the co-assimilation of
numerous compounds making it suitable
for the mixotrophic organisms.[33]

Two other cycles related to the 3-


hydroxypropionate cycle

A variant of the 3-hydroxypropionate


cycle was found to operate in the aerobic
extreme thermoacidophile archaeon
Metallosphaera sedula. This pathway is
called the 3-hydroxypropionate/4-
hydroxybutyrate cycle.[35]

Yet another variant of the 3-


hydroxypropionate cycle is the
dicarboxylate/4-hydroxybutyrate cycle. It
was discovered in anaerobic archaea. It
was proposed in 2008 for the
hyperthermophile archeon Ignicoccus
hospitalis.[36]

Chemosynthesis
Chemosynthesis is carbon fixation driven
by energy obtained by oxidating
inorganic substances (e.g., hydrogen gas
or hydrogen sulfide), rather than from
sunlight. Sulfur- and hydrogen-oxidizing
bacteria often use the Calvin cycle or the
reductive citric acid cycle.[37]

Non-autotrophic pathways
Although almost all heterotrophs cannot
synthesize complete organic molecules
from carbon dioxide, some carbon
dioxide is incorporated in their
metabolism.[38] Notably pyruvate
carboxylase consumes carbon dioxide
(as bicarbonate ions) as part of
gluconeogenesis, and carbon dioxide is
consumed in various anaplerotic
reactions. Recently, also 6-
phosphogluconate dehydrogenase was
shown to catalyze the reductive
carboxylation of ribulose 5-phosphate to
6-phosphogluconate in E. coli under
elevated CO2 concentrations.[39]
Considering the CO2 concentration in the
habitat of E. coli (e.g. the mammalian
gut[40]), this reaction might happen also
naturally. In the future, this property could
be exploited for the design of synthetic
carbon fixation routes.

Carbon isotope
discrimination
Some carboxylases, particularly
RuBisCO, preferentially bind the lighter
carbon stable isotope carbon-12 over the
heavier carbon-13. This is known as
carbon isotope discrimination and
results in carbon-12 to carbon-13 ratios
in the plant that are higher than in the
free air. Measurement of this ratio is
important in the evaluation of water use
efficiency in plants,[41][42][43] and also in
assessing the possible or likely sources
of carbon in global carbon cycle studies.

References
1. Geider RJ, et al. (2001). "Primary
productivity of planet earth:
biological determinants and physical
constraints in terrestrial and aquatic
habitats". Global Change Biol. 7 (8):
849–882.
Bibcode:2001GCBio...7..849G .
doi:10.1046/j.1365-
2486.2001.00448.x .
2. Swan BK, Martinez-Garcia M,
Preston CM, Sczyrba A, Woyke T,
Lamy D, et al. (September 2011).
"Potential for chemolithoautotrophy
among ubiquitous bacteria lineages
in the dark ocean". Science. 333
(6047): 1296–300.
Bibcode:2011Sci...333.1296S .
doi:10.1126/science.1203690 .
PMID 21885783 .
S2CID 206533092 .
3. Cardona T, Sánchez-Baracaldo P,
Rutherford AW, Larkum AW (March
2019). "Early Archean origin of
Photosystem II" . Geobiology. 17 (2):
127–150. doi:10.1111/gbi.12322 .
PMC 6492235 . PMID 30411862 .
4. Cardona T, Murray JW, Rutherford
AW (May 2015). "Origin and
Evolution of Water Oxidation before
the Last Common Ancestor of the
Cyanobacteria" . Molecular Biology
and Evolution. 32 (5): 1310–28.
doi:10.1093/molbev/msv024 .
PMC 4408414 . PMID 25657330 .
5. Brasier M, McLoughlin N, Green O,
Wacey D (June 2006). "A fresh look
at the fossil evidence for early
Archaean cellular life" .
Philosophical Transactions of the
Royal Society of London. Series B,
Biological Sciences. 361 (1470):
887–902.
doi:10.1098/rstb.2006.1835 .
PMC 1578727 . PMID 16754605 .
6. Tomitani A, Knoll AH, Cavanaugh
CM, Ohno T (April 2006). "The
evolutionary diversification of
cyanobacteria: molecular-
phylogenetic and paleontological
perspectives" . Proceedings of the
National Academy of Sciences of the
United States of America. 103 (14):
5442–7.
Bibcode:2006PNAS..103.5442T .
doi:10.1073/pnas.0600999103 .
PMC 1459374 . PMID 16569695 .
7. Kopp RE, Kirschvink JL, Hilburn IA,
Nash CZ (August 2005). "The
Paleoproterozoic snowball Earth: a
climate disaster triggered by the
evolution of oxygenic
photosynthesis" . Proceedings of the
National Academy of Sciences of the
United States of America. 102 (32):
11131–6.
Bibcode:2005PNAS..10211131K .
doi:10.1073/pnas.0504878102 .
PMC 1183582 . PMID 16061801 .
8. Dodd AN, Borland AM, Haslam RP,
Griffiths H, Maxwell K (April 2002).
"Crassulacean acid metabolism:
plastic, fantastic" . Journal of
Experimental Botany. 53 (369): 569–
80. doi:10.1093/jexbot/53.369.569 .
PMID 11886877 .
9. O'Leary MH (1988). "Carbon
isotopes in photosynthesis".
BioScience. 38 (5): 328–336.
doi:10.2307/1310735 .
JSTOR 1310735 . S2CID 29110460 .
10. Sage RF, Meirong L, Monson RK
(1999). "16. The Taxonomic
Distribution of C4 Photosynthesis".
In Sage RF, Monson RK (eds.). C4
Plant Biology. pp. 551–580. ISBN 0-
12-614440-0.
11. Wächtershäuser G. Before enzymes
and templates: theory of surface
metabolism. OCLC 680443998 .
12. Fuchs G (13 October 2011).
"Alternative pathways of carbon
dioxide fixation: insights into the
early evolution of life?". Annual
Review of Microbiology. 65 (1): 631–
58. doi:10.1146/annurev-micro-
090110-102801 . PMID 21740227 .
13. Hügler M, Sievert SM (15 January
2011). "Beyond the Calvin cycle:
autotrophic carbon fixation in the
ocean". Annual Review of Marine
Science. 3 (1): 261–89.
Bibcode:2011ARMS....3..261H .
doi:10.1146/annurev-marine-
120709-142712 . PMID 21329206 .
S2CID 44800487 .
14. Buchanan BB, Arnon DI (April 1990).
"A reverse KREBS cycle in
photosynthesis: consensus at last".
Photosynthesis Research. 24 (1):
47–53. doi:10.1007/bf00032643 .
PMID 24419764 . S2CID 2753977 .
15. Grzymski JJ, Murray AE, Campbell
BJ, Kaplarevic M, Gao GR, Lee C,
et al. (November 2008).
"Metagenome analysis of an
extreme microbial symbiosis reveals
eurythermal adaptation and
metabolic flexibility" . Proceedings
of the National Academy of Sciences
of the United States of America. 105
(45): 17516–21.
Bibcode:2008PNAS..10517516G .
doi:10.1073/pnas.0802782105 .
PMC 2579889 . PMID 18987310 .
16. Baltar F, Herndl GJ (11 June 2019).
"Is dark carbon fixation relevant for
oceanic primary production
estimates?" (PDF). doi:10.5194/bg-
2019-223 .
17. Markert S, Arndt C, Felbeck H,
Becher D, Sievert SM, Hügler M, et al.
(January 2007). "Physiological
proteomics of the uncultured
endosymbiont of Riftia pachyptila".
Science. 315 (5809): 247–50.
Bibcode:2007Sci...315..247M .
doi:10.1126/science.1132913 .
hdl:1912/1514 . OCLC 655249163 .
PMID 17218528 . S2CID 45745396 .
18. Ljungdahl L, Wood HG (April 1965).
"Incorporation of C14 from Carbon
Dioxide into Sugar Phosphates,
Carboxylic Acids, and Amino Acids
by Clostridium thermoaceticum" .
Journal of Bacteriology. 89 (4):
1055–64. doi:10.1128/jb.89.4.1055-
1064.1965 . PMC 277595 .
PMID 14276095 .
19. Drake HL, Gössner AS, Daniel SL
(March 2008). "Old acetogens, new
light". Annals of the New York
Academy of Sciences. 1125 (1):
100–28.
Bibcode:2008NYASA1125..100D .
doi:10.1196/annals.1419.016 .
PMID 18378590 . S2CID 24050060 .
20. Ljungdahl LG (1969). "Total
synthesis of acetate from CO2 by
heterotrophic bacteria". Annual
Review of Microbiology. 23 (1): 515–
38.
doi:10.1146/annurev.mi.23.100169.0
02503 . PMID 4899080 .
21. Ljungdahl LG (1 January 1986). "The
autotrophic pathway of acetate
synthesis in acetogenic bacteria".
Annual Review of Microbiology. 40
(1): 415–50.
doi:10.1146/annurev.micro.40.1.41
5 . PMID 3096193 .
22. Ljungdahl LG (2009). "A life with
acetogens, thermophiles, and
cellulolytic anaerobes". Annual
Review of Microbiology. 63 (1): 1–
25.
doi:10.1146/annurev.micro.091208.0
73617 . PMID 19575555 .
23. Jansen K, Thauer RK, Widdel F,
Fuchs G (1984). "Carbon
assimilation pathways in sulfate
reducing bacteria. Formate, carbon
dioxide, carbon monoxide, and
acetate assimilation by Desulfovibrio
baarsii". Archives of Microbiology.
138 (3): 257–262.
doi:10.1007/bf00402132 .
ISSN 0302-8933 . S2CID 8587232 .
24. Zeikus JG, Kerby R, Krzycki JA
(March 1985). "Single-carbon
chemistry of acetogenic and
methanogenic bacteria". Science.
227 (4691): 1167–73.
Bibcode:1985Sci...227.1167Z .
doi:10.1126/science.3919443 .
PMID 3919443 .
25. Schauder R, Preuß A, Jetten M,
Fuchs G (1989). "Oxidative and
reductive acetyl CoA/carbon
monoxide dehydrogenase pathway
in Desulfobacterium autotrophicum".
Archives of Microbiology. 151 (1):
84–89. doi:10.1007/bf00444674 .
26. Fuchs G (1994). "Variations of the
Acetyl-CoA Pathway in Diversely
Related Microorganisms That Are
Not Acetogens". Acetogenesis.
Springer US. pp. 507–520.
doi:10.1007/978-1-4615-1777-1_19 .
ISBN 978-1-4613-5716-2.
27. Vornolt J, Kunow J, Stetter KO,
Thauer RK (1 February 1995).
"Enzymes and coenzymes of the
carbon monoxide dehydrogenase
pathway for autotrophic CO2 fixation
in Archaeoglobus lithotrophicus and
the lack of carbon monoxide
dehydrogenase in the heterotrophic
A. profundus". Archives of
Microbiology. 163 (2): 112–118.
doi:10.1007/s002030050179 .
ISSN 0302-8933 .
28. Strous M, Pelletier E, Mangenot S,
Rattei T, Lehner A, Taylor MW, et al.
(April 2006). "Deciphering the
evolution and metabolism of an
anammox bacterium from a
community genome". Nature. 440
(7085): 790–4.
Bibcode:2006Natur.440..790S .
doi:10.1038/nature04647 .
PMID 16598256 . S2CID 4402553 .
29. Pezacka E, Wood HG (October
1984). "Role of carbon monoxide
dehydrogenase in the autotrophic
pathway used by acetogenic
bacteria" . Proceedings of the
National Academy of Sciences of the
United States of America. 81 (20):
6261–5.
Bibcode:1984PNAS...81.6261P .
doi:10.1073/pnas.81.20.6261 .
PMC 391903 . PMID 6436811 .
30. Ragsdale SW, Wood HG (April 1985).
"Acetate biosynthesis by acetogenic
bacteria. Evidence that carbon
monoxide dehydrogenase is the
condensing enzyme that catalyzes
the final steps of the synthesis" . The
Journal of Biological Chemistry. 260
(7): 3970–7. doi:10.1016/S0021-
9258(18)89217-1 . PMID 2984190 .
31. Strauss G, Fuchs G (August 1993).
"Enzymes of a novel autotrophic CO2
fixation pathway in the phototrophic
bacterium Chloroflexus aurantiacus,
the 3-hydroxypropionate cycle".
European Journal of Biochemistry.
215 (3): 633–43. doi:10.1111/j.1432-
1033.1993.tb18074.x .
PMID 8354269 .
32. Herter S, Busch A, Fuchs G
(November 2002). "L-Malyl-
coenzyme A lyase/beta-methylmalyl-
coenzyme A lyase from Chloroflexus
aurantiacus, a bifunctional enzyme
involved in autotrophic CO(2)
fixation" . Journal of Bacteriology.
184 (21): 5999–6006.
doi:10.1128/jb.184.21.5999-
6006.2002 . PMC 135395 .
PMID 12374834 .
33. Berg IA (March 2011). "Ecological
aspects of the distribution of
different autotrophic CO2 fixation
pathways" . Applied and
Environmental Microbiology. 77 (6):
1925–36. doi:10.1128/aem.02473-
10 . PMC 3067309 .
PMID 21216907 .
34. Zarzycki J, Brecht V, Müller M, Fuchs
G (December 2009). "Identifying the
missing steps of the autotrophic 3-
hydroxypropionate CO2 fixation
cycle in Chloroflexus aurantiacus" .
Proceedings of the National
Academy of Sciences of the United
States of America. 106 (50): 21317–
22. doi:10.1073/pnas.0908356106 .
PMC 2795484 . PMID 19955419 .
35. Berg IA, Kockelkorn D, Buckel W,
Fuchs G (December 2007). "A 3-
hydroxypropionate/4-
hydroxybutyrate autotrophic carbon
dioxide assimilation pathway in
Archaea". Science. 318 (5857):
1782–6.
Bibcode:2007Sci...318.1782B .
doi:10.1126/science.1149976 .
PMID 18079405 . S2CID 13218676 .
36. Huber H, Gallenberger M, Jahn U,
Eylert E, Berg IA, Kockelkorn D, et al.
(June 2008). "A dicarboxylate/4-
hydroxybutyrate autotrophic carbon
assimilation cycle in the
hyperthermophilic Archaeum
Ignicoccus hospitalis" . Proceedings
of the National Academy of Sciences
of the United States of America. 105
(22): 7851–6.
Bibcode:2008PNAS..105.7851H .
doi:10.1073/pnas.0801043105 .
PMC 2409403 . PMID 18511565 .
37. Encyclopedia of Microbiology .
Academic Press. 2009. pp. 83–84.
ISBN 9780123739445.
38. Nicole Kresge; Robert D. Simoni;
Robert L. Hill (2005). "The Discovery
of Heterotrophic Carbon Dioxide
Fixation by Harland G. Wood" . The
Journal of Biological Chemistry. 280
(18): e15.
39. Satanowski A, Dronsella B, Noor E,
Vögeli B, He H, Wichmann P, et al.
(November 2020). "Awakening a
latent carbon fixation cycle in
Escherichia coli" . Nature
Communications. 11 (1): 5812.
Bibcode:2020NatCo..11.5812S .
doi:10.1038/s41467-020-19564-5 .
PMC 7669889 . PMID 33199707 .
40. Levitt MD (June 1971). "Volume and
composition of human intestinal gas
determined by means of an intestinal
washout technic" . The New England
Journal of Medicine. 284 (25):
1394–8.
doi:10.1056/nejm19710624284250
2 . PMID 5578321 .
41. Adiredjo AL, Navaud O, Muños S,
Langlade NB, Lamaze T, Grieu P (3
July 2014). "Genetic control of water
use efficiency and leaf carbon
isotope discrimination in sunflower
(Helianthus annuus L.) subjected to
two drought scenarios" . PLOS ONE.
9 (7): e101218.
Bibcode:2014PLoSO...9j1218A .
doi:10.1371/journal.pone.0101218 .
PMC 4081578 . PMID 24992022 .
42. Farquhar GD, Ehleringer JR, Hubick
KT (June 1989). "Carbon Isotope
Discrimination and Photosynthesis".
Annual Review of Plant Physiology
and Plant Molecular Biology. 40 (1):
503–537.
doi:10.1146/annurev.pp.40.060189.0
02443 . S2CID 12988287 .
43. Seibt U, Rajabi A, Griffiths H, Berry JA
(March 2008). "Carbon isotopes and
water use efficiency: sense and
sensitivity". Oecologia. 155 (3): 441–
54. Bibcode:2008Oecol.155..441S .
doi:10.1007/s00442-007-0932-7 .
PMID 18224341 . S2CID 451126 .

Further reading
Berg IA (March 2011). "Ecological aspects
of the distribution of different autotrophic
CO2 fixation pathways" . Applied and
Environmental Microbiology. 77 (6): 1925–
36. doi:10.1128/AEM.02473-10 .
PMC 3067309 . PMID 21216907 .
Keeling PJ (October 2004). "Diversity and
evolutionary history of plastids and their
hosts" . American Journal of Botany. 91
(10): 1481–93.
doi:10.3732/ajb.91.10.1481 .
PMID 21652304 . S2CID 17522125 .
Keeling PJ (2009). "Chromalveolates and
the evolution of plastids by secondary
endosymbiosis" (PDF). The Journal of
Eukaryotic Microbiology. 56 (1): 1–8.
doi:10.1111/j.1550-7408.2008.00371.x .
PMID 19335769 . S2CID 34259721 .
Archived from the original (PDF) on 9 July
2009.
Keeling PJ (March 2010). "The
endosymbiotic origin, diversification and
fate of plastids" . Philosophical
Transactions of the Royal Society of
London. Series B, Biological Sciences. 365
(1541): 729–48.
doi:10.1098/rstb.2009.0103 .
PMC 2817223 . PMID 20124341 .
Timme RE, Bachvaroff TR, Delwiche CF
(2012). "Broad phylogenomic sampling and
the sister lineage of land plants" . PLOS
ONE. 7 (1): e29696.
Bibcode:2012PLoSO...7E9696T .
doi:10.1371/journal.pone.0029696 .
PMC 3258253 . PMID 22253761 .
Spiegel FW (February 2012). "Evolution.
Contemplating the first Plantae". Science.
335 (6070): 809–10.
Bibcode:2012Sci...335..809S .
doi:10.1126/science.1218515 .
PMID 22344435 . S2CID 36584136 .
Price DC, Chan CX, Yoon HS, Yang EC, Qiu
H, Weber AP, et al. (February 2012).
"Cyanophora paradoxa genome elucidates
origin of photosynthesis in algae and
plants" (PDF). Science. 335 (6070): 843–7.
Bibcode:2012Sci...335..843P .
doi:10.1126/science.1213561 .
PMID 22344442 . S2CID 17190180 .
Archived from the original (PDF) on 14 May
2013.
Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Carbon_fixation&oldid=1004928918"

Last edited 30 days ago by OAbot

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like