Photorespiration

Last updated
Simplified photorespiration cycle Photorespiration allgemein.svg
Simplified photorespiration cycle
Simplified photorespiration and Calvin cycle Simplified photorespiration diagram.svg
Simplified photorespiration and Calvin cycle

Photorespiration (also known as the oxidative photosynthetic carbon cycle or C2 cycle) refers to a process in plant metabolism where the enzyme RuBisCO oxygenates RuBP, wasting some of the energy produced by photosynthesis. The desired reaction is the addition of carbon dioxide to RuBP (carboxylation), a key step in the Calvin–Benson cycle, but approximately 25% of reactions by RuBisCO instead add oxygen to RuBP (oxygenation), creating a product that cannot be used within the Calvin–Benson cycle. This process lowers the efficiency of photosynthesis, potentially lowering photosynthetic output by 25% in C3 plants. [1] Photorespiration involves a complex network of enzyme reactions that exchange metabolites between chloroplasts, leaf peroxisomes and mitochondria.

Contents

The oxygenation reaction of RuBisCO is a wasteful process because 3-phosphoglycerate is created at a lower rate and higher metabolic cost compared with RuBP carboxylase activity. While photorespiratory carbon cycling results in the formation of G3P eventually, around 25% of carbon fixed by photorespiration is re-released as CO2 [2] and nitrogen, as ammonia. Ammonia must then be detoxified at a substantial cost to the cell. Photorespiration also incurs a direct cost of one ATP and one NAD(P)H.

While it is common to refer to the entire process as photorespiration, technically the term refers only to the metabolic network which acts to rescue the products of the oxygenation reaction (phosphoglycolate).

Photorespiratory reactions

PhotorespirationFrom left to right: chloroplast, peroxisome, and mitochondrion Photorespiration.svg
PhotorespirationFrom left to right: chloroplast, peroxisome, and mitochondrion

Addition of molecular oxygen to ribulose-1,5-bisphosphate produces 3-phosphoglycerate (PGA) and 2-phosphoglycolate (2PG, or PG). PGA is the normal product of carboxylation, and productively enters the Calvin cycle. Phosphoglycolate, however, inhibits certain enzymes involved in photosynthetic carbon fixation (hence is often said to be an 'inhibitor of photosynthesis'). [3] It is also relatively difficult to recycle: in higher plants it is salvaged by a series of reactions in the peroxisome, mitochondria, and again in the peroxisome where it is converted into glycerate. Glycerate reenters the chloroplast and by the same transporter that exports glycolate. A cost of 1 ATP is associated with conversion to 3-phosphoglycerate (PGA) (Phosphorylation), within the chloroplast, which is then free to re-enter the Calvin cycle.

Several costs are associated with this metabolic pathway; the production of hydrogen peroxide in the peroxisome (associated with the conversion of glycolate to glyoxylate). Hydrogen peroxide is a dangerously strong oxidant which must be immediately split into water and oxygen by the enzyme catalase. The conversion of 2× 2Carbon glycine to 1× C3 serine in the mitochondria by the enzyme glycine-decarboxylase is a key step, which releases CO2, NH3, and reduces NAD to NADH. Thus, one CO
2
molecule is produced for every two molecules of O
2
(two deriving from RuBisCO and one from peroxisomal oxidations). The assimilation of NH3 occurs via the GS-GOGAT cycle, at a cost of one ATP and one NADPH.

Cyanobacteria have three possible pathways through which they can metabolise 2-phosphoglycolate. They are unable to grow if all three pathways are knocked out, despite having a carbon concentrating mechanism that should dramatically lower the rate of photorespiration (see below). [4]

Substrate specificity of RuBisCO

Oxygenase activity of RuBisCO RuBisCO reaction O2.svg
Oxygenase activity of RuBisCO

The oxidative photosynthetic carbon cycle reaction is catalyzed by RuBP oxygenase activity:

RuBP + O
2
→ Phosphoglycolate + 3-phosphoglycerate + 2 H+

During the catalysis by RuBisCO, an 'activated' intermediate is formed (an enediol intermediate) in the RuBisCO active site. This intermediate is able to react with either CO
2
or O
2
. It has been demonstrated that the specific shape of the RuBisCO active site acts to encourage reactions with CO
2
. Although there is a significant "failure" rate (~25% of reactions are oxygenation rather than carboxylation), this represents significant favouring of CO
2
, when the relative abundance of the two gases is taken into account: in the current atmosphere, O
2
is approximately 500 times more abundant, and in solution O
2
is 25 times more abundant than CO
2
. [5]

The ability of RuBisCO to specify between the two gases is known as its selectivity factor (or Srel), and it varies between species, [5] with angiosperms more efficient than other plants, but with little variation among the vascular plants. [6]

A suggested explanation of RuBisCO's inability to discriminate completely between CO
2
and O
2
is that it is an evolutionary relic:[ citation needed ] The early atmosphere in which primitive plants originated contained very little oxygen, the early evolution of RuBisCO was not influenced by its ability to discriminate between O
2
and CO
2
. [6]

Conditions which affect photorespiration

Photorespiration rates are affected by:

Altered substrate availability: lowered CO2 or increased O2

Factors which influence this include the atmospheric abundance of the two gases, the supply of the gases to the site of fixation (i.e. in land plants: whether the stomata are open or closed), the length of the liquid phase (how far these gases have to diffuse through water in order to reach the reaction site). For example, when the stomata are closed to prevent water loss during drought: this limits the CO2 supply, while O
2
production within the leaf will continue. In algae (and plants which photosynthesise underwater); gases have to diffuse significant distances through water, which results in a decrease in the availability of CO2 relative to O
2
. It has been predicted that the increase in ambient CO2 concentrations predicted over the next 100 years may lower the rate of photorespiration in most plants by around 50%[ citation needed ]. However, at temperatures higher than the photosynthetic thermal optimum, the increases in turnover rate are not translated into increased CO2 assimilation because of the decreased affinity of Rubisco for CO2. [7]

Increased temperature

At higher temperatures RuBisCO is less able to discriminate between CO2 and O
2
. This is because the enediol intermediate is less stable. Increasing temperatures also lower the solubility of CO2, thus lowering the concentration of CO2 relative to O
2
in the chloroplast.

Biological adaptation to minimize photorespiration

Maize uses the C4 pathway, minimizing photorespiration Field, corn, Liechtenstein, Mountains, Alps, Vaduz, sky, clouds, landscape.jpg
Maize uses the C4 pathway, minimizing photorespiration

The vast majority of plants are C3, meaning they photorespire when necessary. Certain species of plants or algae have mechanisms to lower the uptake of molecular oxygen by RuBisCO. These are commonly referred to as Carbon Concentrating Mechanisms (CCMs), as they increase the concentration of CO2 so that RuBisCO is less likely to produce glycolate through reaction with O
2
.

Biochemical carbon concentrating mechanisms

Biochemical CCMs concentrate carbon dioxide in one temporal or spatial region, through metabolite exchange. C4 and CAM photosynthesis both use the enzyme Phosphoenolpyruvate carboxylase (PEPC) to add CO
2
to a 4-carbon sugar. PEPC is faster than RuBisCO, and more selective for CO
2
.

C4

C4 plants capture carbon dioxide in their mesophyll cells (using an enzyme called phosphoenolpyruvate carboxylase which catalyzes the combination of carbon dioxide with a compound called phosphoenolpyruvate (PEP)), forming oxaloacetate. This oxaloacetate is then converted to malate and is transported into the bundle sheath cells (site of carbon dioxide fixation by RuBisCO) where oxygen concentration is low to avoid photorespiration. Here, carbon dioxide is removed from the malate and combined with RuBP by RuBisCO in the usual way, and the Calvin cycle proceeds as normal. The CO
2
concentrations in the Bundle Sheath are approximately 10–20 fold higher than the concentration in the mesophyll cells. [6]

This ability to avoid photorespiration makes these plants more hardy than other plants in dry and hot environments, wherein stomata are closed and internal carbon dioxide levels are low. Under these conditions, photorespiration does occur in C4 plants, but at a much lower level compared with C3 plants in the same conditions. C4 plants include sugar cane, corn (maize), and sorghum.

CAM (Crassulacean acid metabolism)

Overnight graph of CO2 absorbed by a CAM plant CAMplantgraph.jpg
Overnight graph of CO2 absorbed by a CAM plant

CAM plants, such as cacti and succulent plants, also use the enzyme PEP carboxylase to capture carbon dioxide, but only at night. Crassulacean acid metabolism allows plants to conduct most of their gas exchange in the cooler night-time air, sequestering carbon in 4-carbon sugars which can be released to the photosynthesizing cells during the day. This allows CAM plants to minimize water loss (transpiration) by maintaining closed stomata during the day. CAM plants usually display other water-saving characteristics, such as thick cuticles, stomata with small apertures, and typically lose around 1/3 of the amount of water per CO
2
fixed. [8]


C2

In C2 plants, the mitochondria of mesophyll cells have no glycine decarboxylase (GDC). C2 Photosynthesis.svg
In C2 plants, the mitochondria of mesophyll cells have no glycine decarboxylase (GDC).

C2 photosynthesis (also called glycine shuttle and photorespiratory CO2 pump) is a CCM that works by making use of as opposed to avoiding photorespiration. It performs carbon refixation by delaying the breakdown of photorespired glycine, so that the molecule is shuttled from the mesophyll into the bundle sheath. Once there, the glycine is decarboxylated in mitochondria as usual, releasing CO2 and concentrating it to triple the usual concentration. [9]

Although C2 photosynthesis is traditionally understood as an intermediate step between C3 and C4, a wide variety of plant lineages do end up in the C2 stage without further evolving, showing that it is an evolutionary steady state of its own. C2 may be easier to engineer into crops, as the phenotype requires fewer anatomical changes to produce. [9]

Algae

There have been some reports of algae operating a biochemical CCM: shuttling metabolites within single cells to concentrate CO2 in one area. This process is not fully understood. [10]

Biophysical carbon-concentrating mechanisms

This type of carbon-concentrating mechanism (CCM) relies on a contained compartment within the cell into which CO2 is shuttled, and where RuBisCO is highly expressed. In many species, biophysical CCMs are only induced under low carbon dioxide concentrations. Biophysical CCMs are more evolutionary ancient than biochemical CCMs. There is some debate as to when biophysical CCMs first evolved, but it is likely to have been during a period of low carbon dioxide, after the Great Oxygenation Event (2.4 billion years ago). Low CO
2
periods occurred around 750, 650, and 320–270 million years ago. [11]

Eukaryotic algae

In nearly all species of eukaryotic algae ( Chloromonas being one notable exception), upon induction of the CCM, ~95% of RuBisCO is densely packed into a single subcellular compartment: the pyrenoid. Carbon dioxide is concentrated in this compartment using a combination of CO2 pumps, bicarbonate pumps, and carbonic anhydrases. The pyrenoid is not a membrane-bound compartment but is found within the chloroplast, often surrounded by a starch sheath (which is not thought to serve a function in the CCM). [12]

Hornworts

Certain species of hornwort are the only land plants that are known to have a biophysical CCM involving concentration of carbon dioxide within pyrenoids in their chloroplasts.

Cyanobacteria

Cyanobacterial CCMs are similar in principle to those found in eukaryotic algae and hornworts, but the compartment into which carbon dioxide is concentrated has several structural differences. Instead of the pyrenoid, cyanobacteria contain carboxysomes, which have a protein shell, and linker proteins packing RuBisCO inside with a very regular structure. Cyanobacterial CCMs are much better understood than those found in eukaryotes, partly due to the ease of genetic manipulation of prokaryotes.

Possible purpose of photorespiration

Lowering photorespiration may not result in increased growth rates for plants. Photorespiration may be necessary for the assimilation of nitrate from soil. Thus, a lowering in photorespiration by genetic engineering or because of increasing atmospheric carbon dioxide may not benefit plants as has been proposed. [13] Several physiological processes may be responsible for linking photorespiration and nitrogen assimilation. Photorespiration increases availability of NADH, which is required for the conversion of nitrate to nitrite. Certain nitrite transporters also transport bicarbonate, and elevated CO2 has been shown to suppress nitrite transport into chloroplasts. [14] However, in an agricultural setting, replacing the native photorespiration pathway with an engineered synthetic pathway to metabolize glycolate in the chloroplast resulted in a 40 percent increase in crop growth. [15] [16] [17]

Although photorespiration is much lower in C4 species, it is still an essential pathway mutants without functioning 2-phosphoglycolate metabolism cannot grow in normal conditions. One mutant was shown to rapidly accumulate glycolate. [18]

Although the functions of photorespiration remain controversial, [19] it is widely accepted that this pathway influences a wide range of processes from bioenergetics, photosystem II function, and carbon metabolism to nitrogen assimilation and respiration. The oxygenase reaction of RuBisCO may prevent CO2 depletion near its active sites [20] and contributes to the regulation of CO2. concentration in the atmosphere [21] The photorespiratory pathway is a major source of hydrogen peroxide (H
2
O
2
) in photosynthetic cells. Through H
2
O
2
production and pyrimidine nucleotide interactions, photorespiration makes a key contribution to cellular redox homeostasis. In so doing, it influences multiple signalling pathways, in particular, those that govern plant hormonal responses controlling growth, environmental and defense responses, and programmed cell death. [19]

It has been postulated that photorespiration may function as a "safety valve", [22] preventing the excess of reductive potential coming from an overreduced NADPH-pool from reacting with oxygen and producing free radicals (oxidants), as these can damage the metabolic functions of the cell by subsequent oxidation of membrane lipids, proteins or nucleotides. The mutants deficient in photorespiratory enzymes are characterized by a high redox level in the cell, [23] impaired stomatal regulation, [24] and accumulation of formate. [25]

See also

Related Research Articles

<span class="mw-page-title-main">Photosynthesis</span> Biological process to convert light into chemical energy

Photosynthesis is a system of biological processes by which photosynthetic organisms, such as most plants, algae, and cyanobacteria, convert light energy, typically from sunlight, into the chemical energy necessary to fuel their activities. Photosynthetic organisms use intracellular organic compounds to store the chemical energy they produce in photosynthesis within organic compounds like sugars, glycogen, cellulose and starches. Photosynthesis is usually used to refer to oxygenic photosynthesis, a process that produces oxygen. To use this stored chemical energy, the organisms' cells metabolize the organic compounds through another process called cellular respiration. Photosynthesis plays a critical role in producing and maintaining the oxygen content of the Earth's atmosphere, and it supplies most of the biological energy necessary for complex life on Earth.

<span class="mw-page-title-main">RuBisCO</span> Key enzyme of photosynthesis involved in carbon fixation

Ribulose-1,5-bisphosphate carboxylase/oxygenase, commonly known by the abbreviations RuBisCo, rubisco, RuBPCase, or RuBPco, is an enzyme involved in the light-independent part of photosynthesis, including the carbon fixation by which atmospheric carbon dioxide is converted by plants and other photosynthetic organisms to energy-rich molecules such as glucose. It emerged approximately four billion years ago in primordial metabolism prior to the presence of oxygen on Earth. It is probably the most abundant enzyme on Earth. In chemical terms, it catalyzes the carboxylation of ribulose-1,5-bisphosphate.

<span class="mw-page-title-main">Pyrenoid</span> Organelle found within the chloroplasts of algae and hornworts

Pyrenoids are sub-cellular micro-compartments found in chloroplasts of many algae, and in a single group of land plants, the hornworts. Pyrenoids are associated with the operation of a carbon-concentrating mechanism (CCM). Their main function is to act as centres of carbon dioxide (CO2) fixation, by generating and maintaining a CO2 rich environment around the photosynthetic enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO). Pyrenoids therefore seem to have a role analogous to that of carboxysomes in cyanobacteria.

<span class="mw-page-title-main">Crassulacean acid metabolism</span> Metabolic process

Crassulacean acid metabolism, also known as CAM photosynthesis, is a carbon fixation pathway that evolved in some plants as an adaptation to arid conditions that allows a plant to photosynthesize during the day, but only exchange gases at night. In a plant using full CAM, the stomata in the leaves remain shut during the day to reduce evapotranspiration, but they open at night to collect carbon dioxide and allow it to diffuse into the mesophyll cells. The CO2 is stored as four-carbon malic acid in vacuoles at night, and then in the daytime, the malate is transported to chloroplasts where it is converted back to CO2, which is then used during photosynthesis. The pre-collected CO2 is concentrated around the enzyme RuBisCO, increasing photosynthetic efficiency. This mechanism of acid metabolism was first discovered in plants of the family Crassulaceae.

C<sub>4</sub> carbon fixation Photosynthetic process in some plants

C4 carbon fixation or the Hatch–Slack pathway is one of three known photosynthetic processes of carbon fixation in plants. It owes the names to the 1960s discovery by Marshall Davidson Hatch and Charles Roger Slack.

<span class="mw-page-title-main">Ribulose 1,5-bisphosphate</span> Chemical compound

Ribulose 1,5-bisphosphate (RuBP) is an organic substance that is involved in photosynthesis, notably as the principal CO2 acceptor in plants. It is a colourless anion, a double phosphate ester of the ketopentose called ribulose. Salts of RuBP can be isolated, but its crucial biological function happens in solution. RuBP occurs not only in plants but in all domains of life, including Archaea, Bacteria, and Eukarya.

C<sub>3</sub> carbon fixation Series of interconnected biochemical reactions

C3 carbon fixation is the most common of three metabolic pathways for carbon fixation in photosynthesis, the other two being C4 and CAM. This process converts carbon dioxide and ribulose bisphosphate (RuBP, a 5-carbon sugar) into two molecules of 3-phosphoglycerate through the following reaction:

<span class="mw-page-title-main">Calvin cycle</span> Light-independent reactions in photosynthesis

The Calvin cycle, light-independent reactions, bio synthetic phase, dark reactions, or photosynthetic carbon reduction (PCR) cycle of photosynthesis is a series of chemical reactions that convert carbon dioxide and hydrogen-carrier compounds into glucose. The Calvin cycle is present in all photosynthetic eukaryotes and also many photosynthetic bacteria. In plants, these reactions occur in the stroma, the fluid-filled region of a chloroplast outside the thylakoid membranes. These reactions take the products of light-dependent reactions and perform further chemical processes on them. The Calvin cycle uses the chemical energy of ATP and reducing power of NADPH from the light dependent reactions to produce sugars for the plant to use. These substrates are used in a series of reduction-oxidation (redox) reactions to produce sugars in a step-wise process; there is no direct reaction that converts several molecules of CO2 to a sugar. There are three phases to the light-independent reactions, collectively called the Calvin cycle: carboxylation, reduction reactions, and ribulose 1,5-bisphosphate (RuBP) regeneration.

<span class="mw-page-title-main">Carboxysome</span> Bacterial microcompartment containing the enzyme RuBisCo

Carboxysomes are bacterial microcompartments (BMCs) consisting of polyhedral protein shells filled with the enzymes ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO)—the predominant enzyme in carbon fixation and the rate limiting enzyme in the Calvin cycle—and carbonic anhydrase.

<span class="mw-page-title-main">Glyoxylic acid</span> Chemical compound

Glyoxylic acid or oxoacetic acid is an organic compound. Together with acetic acid, glycolic acid, and oxalic acid, glyoxylic acid is one of the C2 carboxylic acids. It is a colourless solid that occurs naturally and is useful industrially.

The light compensation point (Ic) is the light intensity on the light curve where the rate of photosynthesis exactly matches the rate of cellular respiration. At this point, the uptake of CO2 through photosynthetic pathways is equal to the respiratory release of carbon dioxide, and the uptake of O2 by respiration is equal to the photosynthetic release of oxygen. The concept of compensation points in general may be applied to other photosynthetic variables, the most important being that of CO2 concentration – CO2 compensation point (Γ).Interval of time in day time when light intensity is low due to which net gaseous exchange is zero is called as compensation point.

The photosynthetic efficiency is the fraction of light energy converted into chemical energy during photosynthesis in green plants and algae. Photosynthesis can be described by the simplified chemical reaction

<span class="mw-page-title-main">Phosphoenolpyruvate carboxylase</span> Class of enzymes

Phosphoenolpyruvate carboxylase (also known as PEP carboxylase, PEPCase, or PEPC; EC 4.1.1.31, PDB ID: 3ZGE) is an enzyme in the family of carboxy-lyases found in plants and some bacteria that catalyzes the addition of bicarbonate (HCO3) to phosphoenolpyruvate (PEP) to form the four-carbon compound oxaloacetate and inorganic phosphate:

<span class="mw-page-title-main">Glyoxylate reductase</span> Enzyme

Glyoxylate reductase, first isolated from spinach leaves, is an enzyme that catalyzes the reduction of glyoxylate to glycolate, using the cofactor NADH or NADPH.

Malate dehydrogenase (oxaloacetate-decarboxylating) (NADP<sup>+</sup>) Enzyme

Malate dehydrogenase (oxaloacetate-decarboxylating) (NADP+) (EC 1.1.1.40) or NADP-malic enzyme (NADP-ME) is an enzyme that catalyzes the chemical reaction in the presence of a bivalent metal ion:

<span class="mw-page-title-main">Phosphoglycolate phosphatase</span>

Phosphoglycolate phosphatase(EC 3.1.3.18; systematic name 2-phosphoglycolate phosphohydrolase), also commonly referred to as phosphoglycolate hydrolase, 2-phosphoglycolate phosphatase, P-glycolate phosphatase, and phosphoglycollate phosphatase, is an enzyme responsible for catalyzing the conversion of 2-phosphoglycolate into glycolate and phosphate:

The evolution of photosynthesis refers to the origin and subsequent evolution of photosynthesis, the process by which light energy is used to assemble sugars from carbon dioxide and a hydrogen and electron source such as water. It is believed that the pigments used for photosynthesis initially were used for protection from the harmful effects of light, particularly ultraviolet light. The process of photosynthesis was discovered by Jan Ingenhousz, a Dutch-born British physician and scientist, first publishing about it in 1779.

<span class="mw-page-title-main">Fractionation of carbon isotopes in oxygenic photosynthesis</span>

Photosynthesis converts carbon dioxide to carbohydrates via several metabolic pathways that provide energy to an organism and preferentially react with certain stable isotopes of carbon. The selective enrichment of one stable isotope over another creates distinct isotopic fractionations that can be measured and correlated among oxygenic phototrophs. The degree of carbon isotope fractionation is influenced by several factors, including the metabolism, anatomy, growth rate, and environmental conditions of the organism. Understanding these variations in carbon fractionation across species is useful for biogeochemical studies, including the reconstruction of paleoecology, plant evolution, and the characterization of food chains.

<span class="mw-page-title-main">Kinetic isotope effects of RuBisCO</span>

The kinetic isotope effect (KIE) of ribulose-1,5-bisphosphate carboxylase oxygenase (RuBisCO) is the isotopic fractionation associated solely with the step in the Calvin-Benson cycle where a molecule of carbon dioxide is attached to the 5-carbon sugar ribulose-1,5-bisphosphate (RuBP) to produce two 3-carbon sugars called 3-phosphoglycerate. This chemical reaction is catalyzed by the enzyme RuBisCO, and this enzyme-catalyzed reaction creates the primary kinetic isotope effect of photosynthesis. It is also largely responsible for the isotopic compositions of photosynthetic organisms and the heterotrophs that eat them. Understanding the intrinsic KIE of RuBisCO is of interest to earth scientists, botanists, and ecologists because this isotopic biosignature can be used to reconstruct the evolution of photosynthesis and the rise of oxygen in the geologic record, reconstruct past evolutionary relationships and environmental conditions, and infer plant relationships and productivity in modern environments.

<span class="mw-page-title-main">2-Phosphoglycolate</span> Chemical compound

2-Phosphoglycolate (chemical formula C2H2O6P3-; also known as phosphoglycolate, 2-PG, or PG) is a natural metabolic product of the oxygenase reaction mediated by the enzyme ribulose 1,5-bisphosphate carboxylase (RuBisCo).

References

  1. Sharkey T (1988). "Estimating the rate of photorespiration in leaves". Physiologia Plantarum. 73 (1): 147–152. doi:10.1111/j.1399-3054.1988.tb09205.x.
  2. Leegood RC (May 2007). "A welcome diversion from photorespiration". Nature Biotechnology. 25 (5): 539–40. doi:10.1038/nbt0507-539. PMID   17483837. S2CID   5015366.
  3. Peterhansel C, Krause K, Braun HP, Espie GS, Fernie AR, Hanson DT, Keech O, Maurino VG, Mielewczik M, Sage RF (July 2013). "Engineering photorespiration: current state and future possibilities". Plant Biology. 15 (4): 754–8. Bibcode:2013PlBio..15..754P. doi:10.1111/j.1438-8677.2012.00681.x. PMID   23121076.
  4. Eisenhut M, Ruth W, Haimovich M, Bauwe H, Kaplan A, Hagemann M (November 2008). "The photorespiratory glycolate metabolism is essential for cyanobacteria and might have been conveyed endosymbiontically to plants". Proceedings of the National Academy of Sciences of the United States of America. 105 (44): 17199–204. Bibcode:2008PNAS..10517199E. doi: 10.1073/pnas.0807043105 . PMC   2579401 . PMID   18957552.
  5. 1 2 Griffiths H (June 2006). "Plant biology: designs on Rubisco". Nature. 441 (7096): 940–1. Bibcode:2006Natur.441..940G. doi:10.1038/441940a. PMID   16791182. S2CID   31190084.
  6. 1 2 3 Ehleringer JR, Sage RF, Flanagan LB, Pearcy RW (March 1991). "Climate change and the evolution of C(4) photosynthesis". Trends in Ecology & Evolution. 6 (3): 95–9. doi:10.1016/0169-5347(91)90183-x. PMID   21232434.
  7. Hermida-Carrera, Carmen; Kapralov, Maxim V; Galmés, Jeroni (21 June 2016). "Rubisco catalytic properties and temperature response in crops". Plant Physiology. 171 (4): 2549–61. doi:10.1104/pp.16.01846. PMC   4972260 . PMID   27329223.
  8. Taiz L, Zeiger E (2010). "Chapter 8: Photosynthesis: The Carbon Reactions: Inorganic Carbon–Concentrating Mechanisms: Crassulacean Acid Metabolism (CAM)". Plant Physiology (Fifth ed.). Sinauer Associates, Inc. p. 222.
  9. 1 2 Lundgren, Marjorie R. (December 2020). "C 2 photosynthesis: a promising route towards crop improvement?". New Phytologist. 228 (6): 1734–1740. doi: 10.1111/nph.16494 . PMID   32080851.
  10. Giordano M, Beardall J, Raven JA (June 2005). "CO2 concentrating mechanisms in algae: mechanisms, environmental modulation, and evolution". Annual Review of Plant Biology. 56 (1): 99–131. doi:10.1146/annurev.arplant.56.032604.144052. PMID   15862091.
  11. Raven JA, Giordano M, Beardall J, Maberly SC (February 2012). "Algal evolution in relation to atmospheric CO2: carboxylases, carbon-concentrating mechanisms and carbon oxidation cycles". Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 367 (1588): 493–507. doi:10.1098/rstb.2011.0212. PMC   3248706 . PMID   22232762.
  12. Villarejo A, Martinez F, Pino Plumed M, Ramazanov Z (1996). "The induction of the CO2 concentrating mechanism in a starch-less mutant of Chlamydomonas reinhardtii". Physiologia Plantarum. 98 (4): 798–802. doi:10.1111/j.1399-3054.1996.tb06687.x.
  13. Rachmilevitch S, Cousins AB, Bloom AJ (August 2004). "Nitrate assimilation in plant shoots depends on photorespiration". Proceedings of the National Academy of Sciences of the United States of America. 101 (31): 11506–10. Bibcode:2004PNAS..10111506R. doi: 10.1073/pnas.0404388101 . PMC   509230 . PMID   15272076.
  14. Bloom AJ, Burger M, Rubio Asensio JS, Cousins AB (May 2010). "Carbon dioxide enrichment inhibits nitrate assimilation in wheat and Arabidopsis". Science. 328 (5980): 899–903. Bibcode:2010Sci...328..899B. doi:10.1126/science.1186440. PMID   20466933. S2CID   206525174.
  15. South PF, Cavanagh AP, Liu HW, Ort DR (January 2019). "Synthetic glycolate metabolism pathways stimulate crop growth and productivity in the field". Science. 363 (6422): eaat9077. doi: 10.1126/science.aat9077 . PMC   7745124 . PMID   30606819.
  16. Timmer J (7 December 2017). "We may now be able to engineer the most important lousy enzyme on the planet". Ars Technica. Retrieved 5 January 2019.
  17. Timmer J (3 January 2019). "Fixing photosynthesis by engineering it to recycle a toxic mistake". Ars Technica. Retrieved 5 January 2019.
  18. Zabaleta E, Martin MV, Braun HP (May 2012). "A basal carbon concentrating mechanism in plants?". Plant Science. 187: 97–104. doi:10.1016/j.plantsci.2012.02.001. hdl: 11336/13266 . PMID   22404837. S2CID   25371100.
  19. 1 2 Foyer CH, Bloom AJ, Queval G, Noctor G (2009). "Photorespiratory metabolism: genes, mutants, energetics, and redox signaling". Annual Review of Plant Biology. 60 (1): 455–84. doi:10.1146/annurev.arplant.043008.091948. PMID   19575589.
  20. Igamberdiev AU (2015). "Control of Rubisco function via homeostatic equilibration of CO2 supply". Frontiers in Plant Science. 6: 106. doi: 10.3389/fpls.2015.00106 . PMC   4341507 . PMID   25767475.
  21. Igamberdiev AU, Lea PJ (February 2006). "Land plants equilibrate O2 and CO2 concentrations in the atmosphere". Photosynthesis Research. 87 (2): 177–94. Bibcode:2006PhoRe..87..177I. doi:10.1007/s11120-005-8388-2. PMID   16432665. S2CID   10709679.
  22. Stuhlfauth T, Scheuermann R, Fock HP (April 1990). "Light Energy Dissipation under Water Stress Conditions: Contribution of Reassimilation and Evidence for Additional Processes". Plant Physiology. 92 (4): 1053–61. doi:10.1104/pp.92.4.1053. PMC   1062415 . PMID   16667370.
  23. Igamberdiev AU, Bykova NV, Lea PJ, Gardeström P (April 2001). "The role of photorespiration in redox and energy balance of photosynthetic plant cells: A study with a barley mutant deficient in glycine decarboxylase". Physiologia Plantarum. 111 (4): 427–438. doi:10.1034/j.1399-3054.2001.1110402.x. PMID   11299007.
  24. Igamberdiev AU, Mikkelsen TN, Ambus P, Bauwe H, Lea PJ, Gardeström P (2004). "Photorespiration Contributes to Stomatal Regulation and Carbon Isotope Fractionation: A Study with Barley, Potato and Arabidopsis Plants Deficient in Glycine Decarboxylase". Photosynthesis Research. 81 (2): 139–152. Bibcode:2004PhoRe..81..139I. doi:10.1023/B:PRES.0000035026.05237.ec. S2CID   9485316.
  25. Wingler A, Lea PJ, Leegood RC (1999). "Photorespiratory metabolism of glyoxylate and formate in glycine-accumulating mutants of barley and Amaranthus edulis 2". Planta. 207 (4): 518–526. Bibcode:1999Plant.207..518W. doi:10.1007/s004250050512. S2CID   34817815.

Further reading