Photosynthesis

Page semi-protected
Source: Wikipedia, the free encyclopedia.

Schematic of photosynthesis in plants. The carbohydrates produced are stored in or used by the plant.
Composite image showing the global distribution of photosynthesis, including both oceanic phytoplankton and terrestrial vegetation. Dark red and blue-green indicate regions of high photosynthetic activity in the ocean and on land, respectively.

Photosynthesis (

complex life on Earth.[2]

Some

carbon-fixing anoxygenic photosynthesis, where the simpler photopigment retinal and its microbial rhodopsin derivatives are used to absorb green light and power proton pumps to directly synthesize adenosine triphosphate (ATP), the "energy currency" of cells. Such archaeal photosynthesis might have been the earliest form of photosynthesis that evolved on Earth, going back as far as the Paleoarchean, preceding that of cyanobacteria (see Purple Earth hypothesis
).

While the details may differ between

plasma membrane. In these light-dependent reactions, some energy is used to strip electrons from suitable substances, such as water, producing oxygen gas. The hydrogen freed by the splitting of water is used in the creation of two important molecules that participate in energetic processes: reduced nicotinamide adenine dinucleotide phosphate
(NADPH) and ATP.

In plants, algae, and cyanobacteria, sugars are synthesized by a subsequent sequence of light-independent reactions called the

reduced and removed to form further carbohydrates, such as glucose. In other bacteria, different mechanisms like the reverse Krebs cycle
are used to achieve the same end.

The first photosynthetic organisms probably

power consumption of human civilization.[9] Photosynthetic organisms also convert around 100–115 billion tons (91–104 Pg petagrams, or a billion metric tons), of carbon into biomass per year.[10][11] Photosyntesis was first discovered in 1779 by Jan Ingenhousz
; he showed that plants need light, not just air, soil, and water.

Photosynthesis is vital for climate processes, as it captures carbon dioxide from the air and then binds it in plants, harvested products and soil.

petagrams) of carbon dioxide every year, i.e. 3.825 billion metric tons.[12]

Overview

Photosynthesis changes sunlight into chemical energy, splits water to liberate O2, and fixes CO2 into sugar.

Most photosynthetic organisms are

photoautotrophs, which means that they are able to synthesize food directly from carbon dioxide and water using energy from light. However, not all organisms use carbon dioxide as a source of carbon atoms to carry out photosynthesis; photoheterotrophs use organic compounds, rather than carbon dioxide, as a source of carbon.[2]

In plants, algae, and cyanobacteria, photosynthesis releases oxygen. This oxygenic photosynthesis is by far the most common type of photosynthesis used by living organisms. Some shade-loving plants (sciophytes) produce such low levels of oxygen during photosynthesis that they use all of it themselves instead of releasing it to the atmosphere.[13]

Although there are some differences between oxygenic photosynthesis in

plants, algae, and cyanobacteria, the overall process is quite similar in these organisms. There are also many varieties of anoxygenic photosynthesis, used mostly by bacteria, which consume carbon dioxide but do not release oxygen.[citation needed
]

Carbon dioxide is converted into sugars in a process called

endothermic redox reaction. In general outline, photosynthesis is the opposite of cellular respiration: while photosynthesis is a process of reduction of carbon dioxide to carbohydrates, cellular respiration is the oxidation of carbohydrates or other nutrients to carbon dioxide. Nutrients used in cellular respiration include carbohydrates, amino acids and fatty acids. These nutrients are oxidized to produce carbon dioxide and water, and to release chemical energy to drive the organism's metabolism
.

Photosynthesis and cellular respiration are distinct processes, as they take place through different sequences of chemical reactions and in different cellular compartments.[citation needed]

The general equation for photosynthesis as first proposed by Cornelis van Niel is:[14]

CO2carbon
dioxide
+ 2H2Aelectron donor +
photonslight energy[CH2O]carbohydrate
+ 2Aoxidized
electron
donor
+ H2Owater

Since water is used as the electron donor in oxygenic photosynthesis, the equation for this process is:

CO2carbon
dioxide
+ 2H2Owater + photonslight energy[CH2O]carbohydrate + O2oxygen + H2Owater

This equation emphasizes that water is both a reactant in the light-dependent reaction and a product of the light-independent reaction, but canceling n water molecules from each side gives the net equation:

CO2carbon
dioxide
+ H2O water + photonslight energy[CH2O]carbohydrate + O2 oxygen

Other processes substitute other compounds (such as arsenite) for water in the electron-supply role; for example some microbes use sunlight to oxidize arsenite to arsenate:[15] The equation for this reaction is:

CO2carbon
dioxide
+ (AsO3−
3
)

arsenite
+ photonslight energy(AsO3−
4
)

arsenate
+ COcarbon
monoxide
(used to build other compounds in subsequent reactions)[16]

Photosynthesis occurs in two stages. In the first stage, light-dependent reactions or light reactions capture the energy of light and use it to make the hydrogen carrier

NADPH and the energy-storage molecule ATP
. During the second stage, the light-independent reactions use these products to capture and reduce carbon dioxide.

Most organisms that use oxygenic photosynthesis use visible light for the light-dependent reactions, although at least three use shortwave infrared or, more specifically, far-red radiation.[17]

Some organisms employ even more radical variants of photosynthesis. Some archaea use a simpler method that employs a pigment similar to those used for vision in animals. The bacteriorhodopsin changes its configuration in response to sunlight, acting as a proton pump. This produces a proton gradient more directly, which is then converted to chemical energy. The process does not involve carbon dioxide fixation and does not release oxygen, and seems to have evolved separately from the more common types of photosynthesis.[18]

Photosynthetic membranes and organelles

Chloroplast ultrastructure:
  1. outer membrane
  2. intermembrane space
  3. inner membrane (1+2+3: envelope)
  4. stroma (aqueous fluid)
  5. thylakoid lumen (inside of thylakoid)
  6. thylakoid membrane
  7. granum (stack of thylakoids)
  8. thylakoid (lamella)
  9. starch
  10. ribosome
  11. plastidial DNA
  12. plastoglobule (drop of lipids)

In photosynthetic bacteria, the proteins that gather light for photosynthesis are embedded in

vesicles called intracytoplasmic membranes.[21] These structures can fill most of the interior of a cell, giving the membrane a very large surface area and therefore increasing the amount of light that the bacteria can absorb.[20]

In plants and algae, photosynthesis takes place in organelles called chloroplasts. A typical plant cell contains about 10 to 100 chloroplasts. The chloroplast is enclosed by a membrane. This membrane is composed of a phospholipid inner membrane, a phospholipid outer membrane, and an intermembrane space. Enclosed by the membrane is an aqueous fluid called the stroma. Embedded within the stroma are stacks of thylakoids (grana), which are the site of photosynthesis. The thylakoids appear as flattened disks. The thylakoid itself is enclosed by the thylakoid membrane, and within the enclosed volume is a lumen or thylakoid space. Embedded in the thylakoid membrane are integral and peripheral membrane protein complexes of the photosynthetic system.

Plants absorb light primarily using the

diatoms
resulting in a wide variety of colors.

These pigments are embedded in plants and algae in complexes called antenna proteins. In such proteins, the pigments are arranged to work together. Such a combination of proteins is also called a light-harvesting complex.[23]

Although all cells in the green parts of a plant have chloroplasts, the majority of those are found in specially adapted structures called

palisade
mesophyll cells where most of the photosynthesis takes place.

Light-dependent reactions

Light-dependent reactions of photosynthesis at the thylakoid membrane

In the light-dependent reactions, one molecule of the pigment chlorophyll absorbs one photon and loses one electron. This electron is taken up by a modified form of chlorophyll called pheophytin, which passes the electron to a quinone molecule, starting the flow of electrons down an electron transport chain that leads to the ultimate reduction of NADP to NADPH. In addition, this creates a proton gradient (energy gradient) across the chloroplast membrane, which is used by ATP synthase in the synthesis of ATP. The chlorophyll molecule ultimately regains the electron it lost when a water molecule is split in a process called photolysis, which releases oxygen.

The overall equation for the light-dependent reactions under the conditions of non-cyclic electron flow in green plants is:[24]

2 H2O + 2 NADP+ + 3 ADP + 3 Pi + light → 2 NADPH + 2 H+ + 3 ATP + O2

Not all

photosynthetic organisms their color (e.g., green plants, red algae, purple bacteria) and is the least effective for photosynthesis in the respective organisms
.

Z scheme

The "Z scheme"

In

NADPH. The light-dependent reactions are of two forms: cyclic and non-cyclic
.

In the non-cyclic reaction, the photons are captured in the light-harvesting

NADPH is a product of the terminal redox reaction in the Z-scheme. The electron enters a chlorophyll molecule in Photosystem I. There it is further excited by the light absorbed by that photosystem. The electron is then passed along a chain of electron acceptors to which it transfers some of its energy. The energy delivered to the electron acceptors is used to move hydrogen ions across the thylakoid membrane into the lumen. The electron is eventually used to reduce the coenzyme NADP with an H+
to NADPH (which has functions in the light-independent reaction); at that point, the path of that electron ends.

The cyclic reaction is similar to that of the non-cyclic but differs in that it generates only ATP, and no reduced NADP (NADPH) is created. The cyclic reaction takes place only at photosystem I. Once the electron is displaced from the photosystem, the electron is passed down the electron acceptor molecules and returns to photosystem I, from where it was emitted, hence the name cyclic reaction.

Water photolysis

photosynthetic organisms.[25][26]

Light-independent reactions

Calvin cycle

In the

: 128 

3 CO2 + 9 ATP + 6 NADPH + 6 H+ → C3H6O3-phosphate + 9 ADP + 8 Pi + 6 NADP+ + 3 H2O
Overview of the Calvin cycle and carbon fixation

Carbon fixation produces the three-carbon sugar intermediate, which is then converted into the final carbohydrate products. The simple carbon sugars photosynthesis produces are then used to form other organic compounds, such as the building material cellulose, the precursors for lipid and amino acid biosynthesis, or as a fuel in cellular respiration. The latter occurs not only in plants but also in animals when the carbon and energy from plants is passed through a food chain
.

The

lipids
.

Carbon concentrating mechanisms

On land

pyruvate, and all the species ending in "-ate" are shown as unionized acids, such as malic acid
and so on).

In

ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) and decrease in carbon fixation. Some plants have evolved mechanisms to increase the CO2 concentration in the leaves under these conditions.[27]

Plants that use the

C2 photosynthesis, which involves carbon-concentration by selective breakdown of photorespiratory glycine, is both an evolutionary precursor to C4 and a useful carbon-concentrating mechanism in its own right.[30]

oxaloacetate, which is then reduced to malate. Decarboxylation of malate during the day releases CO2 inside the leaves, thus allowing carbon fixation to 3-phosphoglycerate by RuBisCO. CAM is used by 16,000 species of plants.[31]

Calcium-oxalate-accumulating plants, such as Amaranthus hybridus and Colobanthus quitensis, show a variation of photosynthesis where calcium oxalate crystals function as dynamic carbon pools, supplying carbon dioxide (CO2) to photosynthetic cells when stomata are partially or totally closed. This process was named alarm photosynthesis. Under stress conditions (e.g., water deficit), oxalate released from calcium oxalate crystals is converted to CO2 by an oxalate oxidase enzyme, and the produced CO2 can support the Calvin cycle reactions. Reactive hydrogen peroxide (H2O2), the byproduct of oxalate oxidase reaction, can be neutralized by catalase. Alarm photosynthesis represents a photosynthetic variant to be added to the well-known C4 and CAM pathways. However, alarm photosynthesis, in contrast to these pathways, operates as a biochemical pump that collects carbon from the organ interior (or from the soil) and not from the atmosphere.[32][33]

In water

hydrocarbonate ions (HCO
3
). Before the CO2 can diffuse out, RuBisCO concentrated within the carboxysome quickly sponges it up. HCO
3
ions are made from CO2 outside the cell by another carbonic anhydrase and are actively pumped into the cell by a membrane protein. They cannot cross the membrane as they are charged, and within the cytosol they turn back into CO2 very slowly without the help of carbonic anhydrase. This causes the HCO
3
ions to accumulate within the cell from where they diffuse into the carboxysomes.[34] Pyrenoids in algae and hornworts also act to concentrate CO2 around RuBisCO.[35]

Order and kinetics

The overall process of photosynthesis takes place in four stages:[11]

Stage Event Site Time scale
1
Energy transfer in antenna chlorophyll
Thylakoid membranes in the chloroplasts
Femtosecond to picosecond
2
photochemical reactions
Picosecond to nanosecond
3
ATP synthesis
Microsecond to millisecond
4 Carbon fixation and export of stable products Stroma of the chloroplasts and the cell cytosol Millisecond to second

Efficiency

Plants usually convert light into chemical energy with a photosynthetic efficiency of 3–6%.[36][37] Absorbed light that is unconverted is

light reaction of photosynthesis by using chlorophyll fluorometers.[38]

Actual plants' photosynthetic efficiency varies with the

electric energy at an efficiency of approximately 6–20% for mass-produced panels, and above 40% in laboratory
devices. Scientists are studying photosynthesis in hopes of developing plants with increased yield.[37]

The efficiency of both light and dark reactions can be measured, but the relationship between the two can be complex. For example, the

C3 plants can use for carbon fixation or photorespiration.[40] Electrons may also flow to other electron sinks.[41][42][43] For this reason, it is not uncommon for authors to differentiate between work done under non-photorespiratory conditions and under photorespiratory conditions.[44][45][46]

barometric pressure, leaf area, and photosynthetically active radiation (PAR), it becomes possible to estimate, "A" or carbon assimilation, "E" or transpiration, "gs" or stomatal conductance, and "Ci" or intracellular CO2.[49] However, it is more common to use chlorophyll fluorescence for plant stress measurement, where appropriate, because the most commonly used parameters FV/FM and Y(II) or F/FM' can be measured in a few seconds, allowing the investigation of larger plant populations.[46]

Gas exchange systems that offer control of CO2 levels, above and below ambient, allow the common practice of measurement of A/Ci curves, at different CO2 levels, to characterize a plant's photosynthetic response.[49]

Integrated chlorophyll fluorometer – gas exchange systems allow a more precise measure of photosynthetic response and mechanisms.[47][48] While standard gas exchange photosynthesis systems can measure Ci, or substomatal CO2 levels, the addition of integrated chlorophyll fluorescence measurements allows a more precise measurement of CC, the estimation of CO2 concentration at the site of carboxylation in the chloroplast, to replace Ci.[48][50] CO2 concentration in the chloroplast becomes possible to estimate with the measurement of mesophyll conductance or gm using an integrated system.[47][48][51]

Photosynthesis measurement systems are not designed to directly measure the amount of light the leaf absorbs, but analysis of chlorophyll fluorescence, P700- and P515-absorbance, and gas exchange measurements reveal detailed information about, e.g., the photosystems, quantum efficiency and the CO2 assimilation rates. With some instruments, even wavelength dependency of the photosynthetic efficiency can be analyzed.[52]

A phenomenon known as quantum walk increases the efficiency of the energy transport of light significantly. In the photosynthetic cell of an alga, bacterium, or plant, there are light-sensitive molecules called chromophores arranged in an antenna-shaped structure called a photocomplex. When a photon is absorbed by a chromophore, it is converted into a quasiparticle referred to as an exciton, which jumps from chromophore to chromophore towards the reaction center of the photocomplex, a collection of molecules that traps its energy in a chemical form accessible to the cell's metabolism. The exciton's wave properties enable it to cover a wider area and try out several possible paths simultaneously, allowing it to instantaneously "choose" the most efficient route, where it will have the highest probability of arriving at its destination in the minimum possible time.

Because that quantum walking takes place at temperatures far higher than quantum phenomena usually occur, it is only possible over very short distances. Obstacles in the form of destructive interference cause the particle to lose its wave properties for an instant before it regains them once again after it is freed from its locked position through a classic "hop". The movement of the electron towards the photo center is therefore covered in a series of conventional hops and quantum walks.[53][54][55]

Evolution

thylakoid membranes preserved in 1.75-billion-year-old cherts.[60]

photosynthetic reaction center
.

Symbiosis and the origin of chloroplasts

Plant cells with visible chloroplasts (from a moss, Plagiomnium affine)

Several groups of

mollusks, such as Elysia viridis and Elysia chlorotica, also maintain a symbiotic relationship with chloroplasts they capture from the algae in their diet and then store in their bodies (see Kleptoplasty). This allows the mollusks to survive solely by photosynthesis for several months at a time.[62][63] Some of the genes from the plant cell nucleus have even been transferred to the slugs, so that the chloroplasts can be supplied with proteins they need to survive.[64]

An even closer form of symbiosis may explain the origin of chloroplasts. Chloroplasts have many similarities with photosynthetic

CoRR Hypothesis proposes that this co-location of genes with their gene products is required for redox regulation of gene expression, and accounts for the persistence of DNA in bioenergetic organelles.[68]

Photosynthetic eukaryotic lineages

Symbiotic and kleptoplastic organisms excluded:

Except for the euglenids, which are found within the Excavata, all of these belong to the Diaphoretickes. Archaeplastida and the photosynthetic Paulinella got their plastids, which are surrounded by two membranes, through primary endosymbiosis in two separate events, by engulfing a cyanobacterium. The plastids in all the other groups have either a red or green algal origin, and are referred to as the "red lineages" and the "green lineages". The only known exception is the ciliate Pseudoblepharisma tenue, which in addition to its plastids that originated from green algae also has a purple sulfur bacterium as symbiont. In dinoflagellates and euglenids the plastids are surrounded by three membranes, and in the remaining lines by four. A nucleomorph, remnants of the original algal nucleus located between the inner and outer membranes of the plastid, is present in the cryptophytes (from a red alga) and chlorarachniophytes (from a green alga).[69] Some dinoflagellates that lost their photosynthetic ability later regained it again through new endosymbiotic events with different algae. While able to perform photosynthesis, many of these eukaryotic groups are mixotrophs and practice heterotrophy to various degrees.

Photosynthetic prokaryotic lineages

Early photosynthetic systems, such as those in

reducing at that time.[70]

With a possible exception of Heimdallarchaeota, photosynthesis is not found in archaea.[71] Haloarchaea are phototrophic and can absorb energy from the sun, but do not harvest carbon from the atmosphere and are therefore not photosynthetic.[72] Instead of chlorophyll they use rhodopsins, which convert light-energy to ion gradients but cannot mediate electron transfer reactions.[73][74]

In bacteria eight photosynthetic lineages are currently known:[75][76][77][78]

Cyanobacteria and the evolution of photosynthesis

The biochemical capacity to use water as the source for electrons in photosynthesis evolved once, in a

continental shelves near the end of the Proterozoic, but only with the Mesozoic (251–66 Ma) radiations of dinoflagellates, coccolithophorids, and diatoms did the primary production of oxygen in marine shelf waters take modern form. Cyanobacteria remain critical to marine ecosystems as primary producers of oxygen in oceanic gyres, as agents of biological nitrogen fixation, and, in modified form, as the plastids of marine algae.[82]

Experimental history

Discovery

Although some of the steps in photosynthesis are still not completely understood, the overall photosynthetic equation has been known since the 19th century.

Portrait of Jan Baptist van Helmont by Mary Beale, c. 1674

Jan van Helmont began the research of the process in the mid-17th century when he carefully measured the mass of the soil a plant was using and the mass of the plant as it grew. After noticing that the soil mass changed very little, he hypothesized that the mass of the growing plant must come from the water, the only substance he added to the potted plant. His hypothesis was partially accurate – much of the gained mass comes from carbon dioxide as well as water. However, this was a signaling point to the idea that the bulk of a plant's biomass comes from the inputs of photosynthesis, not the soil itself.

Joseph Priestley, a chemist and minister, discovered that when he isolated a volume of air under an inverted jar and burned a candle in it (which gave off CO2), the candle would burn out very quickly, much before it ran out of wax. He further discovered that a mouse could similarly "injure" air. He then showed that a plant could restore the air the candle and the mouse had "injured."[83]

In 1779, Jan Ingenhousz repeated Priestley's experiments. He discovered that it was the influence of sunlight on the plant that could cause it to revive a mouse in a matter of hours.[83][84]

In 1796,

Nicolas-Théodore de Saussure showed that the increase in mass of the plant as it grows could not be due only to uptake of CO2 but also to the incorporation of water. Thus, the basic reaction by which organisms use photosynthesis to produce food (such as glucose) was outlined.[85]

Refinements

green bacteria, he was the first to demonstrate that photosynthesis is a light-dependent redox reaction in which hydrogen reduces (donates its atoms as electrons and protons
to) carbon dioxide.

Robert Emerson discovered two light reactions by testing plant productivity using different wavelengths of light. With the red alone, the light reactions were suppressed. When blue and red were combined, the output was much more substantial. Thus, there were two photosystems, one absorbing up to 600 nm wavelengths, the other up to 700 nm. The former is known as PSII, the latter is PSI. PSI contains only chlorophyll "a", PSII contains primarily chlorophyll "a" with most of the available chlorophyll "b", among other pigments. These include phycobilins, which are the red and blue pigments of red and blue algae, respectively, and fucoxanthol for brown algae and diatoms. The process is most productive when the absorption of quanta is equal in both PSII and PSI, assuring that input energy from the antenna complex is divided between the PSI and PSII systems, which in turn powers the photochemistry.[11]

Robert Hill thought that a complex of reactions consisted of an intermediate to cytochrome b6 (now a plastoquinone), and that another was from cytochrome f to a step in the carbohydrate-generating mechanisms. These are linked by plastoquinone, which does require energy to reduce cytochrome f. Further experiments to prove that the oxygen developed during the photosynthesis of green plants came from water were performed by Hill in 1937 and 1939. He showed that isolated chloroplasts give off oxygen in the presence of unnatural reducing agents like iron oxalate, ferricyanide or benzoquinone after exposure to light. In the Hill reaction:[86]

2 H2O + 2 A + (light, chloroplasts) → 2 AH2 + O2

A is the electron acceptor. Therefore, in light, the electron acceptor is reduced and oxygen is evolved. Samuel Ruben and Martin Kamen used radioactive isotopes to determine that the oxygen liberated in photosynthesis came from the water.

Melvin Calvin works in his photosynthesis laboratory.

Melvin Calvin and Andrew Benson, along with James Bassham, elucidated the path of carbon assimilation (the photosynthetic carbon reduction cycle) in plants. The carbon reduction cycle is known as the Calvin cycle, but many scientists refer to it as the Calvin-Benson, Benson-Calvin, or even Calvin-Benson-Bassham (or CBB) Cycle.

Nobel Prize–winning scientist Rudolph A. Marcus was later able to discover the function and significance of the electron transport chain.

Otto Heinrich Warburg and Dean Burk discovered the I-quantum photosynthesis reaction that splits CO2, activated by the respiration.[87]

In 1950, first experimental evidence for the existence of photophosphorylation in vivo was presented by Otto Kandler using intact Chlorella cells and interpreting his findings as light-dependent ATP formation.[88] In 1954, Daniel I. Arnon et al. discovered photophosphorylation in vitro in isolated chloroplasts with the help of P32.[89][90]

Louis N. M. Duysens and Jan Amesz discovered that chlorophyll "a" will absorb one light, oxidize cytochrome f, while chlorophyll "a" (and other pigments) will absorb another light but will reduce this same oxidized cytochrome, stating the two light reactions are in series.

Development of the concept

In 1893, the American botanist Charles Reid Barnes proposed two terms, photosyntax and photosynthesis, for the biological process of synthesis of complex carbon compounds out of carbonic acid, in the presence of chlorophyll, under the influence of light. The term photosynthesis is derived from the Greek phōs (φῶς, gleam) and sýnthesis (σύνθεσις, arranging together),[91][92][93] while another word that he designated was photosyntax, from sýntaxis (σύνταξις, configuration). Over time, the term photosynthesis came into common usage. Later discovery of anoxygenic photosynthetic bacteria and photophosphorylation necessitated redefinition of the term.[94]

C3 : C4 photosynthesis research

In the late 1940s at the University of California, Berkeley, the details of photosynthetic carbon metabolism were sorted out by the chemists Melvin Calvin, Andrew Benson, James Bassham and a score of students and researchers utilizing the carbon-14 isotope and paper chromatography techniques.[95] The pathway of CO2 fixation by the algae Chlorella in a fraction of a second in light resulted in a three carbon molecule called phosphoglyceric acid (PGA). For that original and ground-breaking work, a Nobel Prize in Chemistry was awarded to Melvin Calvin in 1961. In parallel, plant physiologists studied leaf gas exchanges using the new method of infrared gas analysis and a leaf chamber where the net photosynthetic rates ranged from 10 to 13 μmol CO2·m−2·s−1, with the conclusion that all terrestrial plants have the same photosynthetic capacities, that are light saturated at less than 50% of sunlight.[96][97]

Later in 1958–1963 at Cornell University, field grown maize was reported to have much greater leaf photosynthetic rates of 40 μmol CO2·m−2·s−1 and not be saturated at near full sunlight.[98][99] This higher rate in maize was almost double of those observed in other species such as wheat and soybean, indicating that large differences in photosynthesis exist among higher plants. At the University of Arizona, detailed gas exchange research on more than 15 species of monocots and dicots uncovered for the first time that differences in leaf anatomy are crucial factors in differentiating photosynthetic capacities among species.[100][101] In tropical grasses, including maize, sorghum, sugarcane, Bermuda grass and in the dicot amaranthus, leaf photosynthetic rates were around 38−40 μmol CO2·m−2·s−1, and the leaves have two types of green cells, i.e. outer layer of mesophyll cells surrounding a tightly packed cholorophyllous vascular bundle sheath cells. This type of anatomy was termed Kranz anatomy in the 19th century by the botanist Gottlieb Haberlandt while studying leaf anatomy of sugarcane.[102] Plant species with the greatest photosynthetic rates and Kranz anatomy showed no apparent photorespiration, very low CO2 compensation point, high optimum temperature, high stomatal resistances and lower mesophyll resistances for gas diffusion and rates never saturated at full sun light.[103] The research at Arizona was designated a Citation Classic in 1986.[101] These species were later termed C4 plants as the first stable compound of CO2 fixation in light has four carbons as malate and aspartate.[104][105][106] Other species that lack Kranz anatomy were termed C3 type such as cotton and sunflower, as the first stable carbon compound is the three-carbon PGA. At 1000 ppm CO2 in measuring air, both the C3 and C4 plants had similar leaf photosynthetic rates around 60 μmol CO2·m−2·s−1 indicating the suppression of photorespiration in C3 plants.[100][101]

Factors

The leaf is the primary site of photosynthesis in plants.

There are four main factors influencing photosynthesis[clarification needed] and several corollary factors. The four main are:[107]

Total photosynthesis is limited by a range of environmental factors. These include the amount of light available, the amount of leaf area a plant has to capture light (shading by other plants is a major limitation of photosynthesis), the rate at which carbon dioxide can be supplied to the chloroplasts to support photosynthesis, the availability of water, and the availability of suitable temperatures for carrying out photosynthesis.[108]

Light intensity (irradiance), wavelength and temperature

Absorbance spectra of free chlorophyll a (blue) and b (red) in a solvent. The action spectra of chlorophyll molecules are slightly modified in vivo depending on specific pigment–protein interactions.

The process of photosynthesis provides the main input of free energy into the biosphere, and is one of four main ways in which radiation is important for plant life.[109]

The radiation climate within plant communities is extremely variable, in both time and space.

In the early 20th century,

Gabrielle Matthaei investigated the effects of light intensity (irradiance
) and temperature on the rate of carbon assimilation.

  • At constant temperature, the rate of carbon assimilation varies with irradiance, increasing as the irradiance increases, but reaching a plateau at higher irradiance.
  • At low irradiance, increasing the temperature has little influence on the rate of carbon assimilation. At constant high irradiance, the rate of carbon assimilation increases as the temperature is increased.

These two experiments illustrate several important points: First, it is known that, in general,

photochemical reactions are not affected by temperature. However, these experiments clearly show that temperature affects the rate of carbon assimilation, so there must be two sets of reactions in the full process of carbon assimilation. These are the light-dependent 'photochemical' temperature-independent stage, and the light-independent, temperature-dependent stage. Second, Blackman's experiments illustrate the concept of limiting factors. Another limiting factor is the wavelength of light. Cyanobacteria, which reside several meters underwater, cannot receive the correct wavelengths required to cause photoinduced charge separation in conventional photosynthetic pigments. To combat this problem, Cyanobacteria have a light-harvesting complex called Phycobilisome.[110]
This complex is made up of a series of proteins with different pigments which surround the reaction center.

Carbon dioxide levels and photorespiration

Photorespiration

As carbon dioxide concentrations rise, the rate at which sugars are made by the light-independent reactions increases until limited by other factors. RuBisCO, the enzyme that captures carbon dioxide in the light-independent reactions, has a binding affinity for both carbon dioxide and oxygen. When the concentration of carbon dioxide is high, RuBisCO will fix carbon dioxide. However, if the carbon dioxide concentration is low, RuBisCO will bind oxygen instead of carbon dioxide. This process, called photorespiration, uses energy, but does not produce sugars.

RuBisCO oxygenase activity is disadvantageous to plants for several reasons:

  1. One product of oxygenase activity is phosphoglycolate (2 carbon) instead of
    Calvin-Benson cycle
    .
  2. Phosphoglycolate is quickly metabolized to glycolate that is toxic to a plant at a high concentration; it inhibits photosynthesis.
  3. Salvaging glycolate is an energetically expensive process that uses the glycolate pathway, and only 75% of the carbon is returned to the Calvin-Benson cycle as 3-phosphoglycerate. The reactions also produce ammonia (NH3), which is able to diffuse out of the plant, leading to a loss of nitrogen.
A highly simplified summary is:
2 glycolate + ATP → 3-phosphoglycerate + carbon dioxide + ADP + NH3

The salvaging pathway for the products of RuBisCO oxygenase activity is more commonly known as photorespiration, since it is characterized by light-dependent oxygen consumption and the release of carbon dioxide.

See also

References

  1. ^ "Photosynthesis". Lexico UK English Dictionary. Oxford University Press. Archived from the original on 2022-08-11. Retrieved 2023-07-15.
  2. ^
    PMID 16997562
    .
  3. . This initial incorporation of carbon into organic compounds is known as carbon fixation.
  4. .
  5. .
  6. .
  7. from the original on 2010-08-14. Retrieved 2012-07-07. 100×1015 grams of carbon/year fixed by photosynthetic organisms, which is equivalent to 4×1018 kJ/yr = 4×1021 J/yr of free energy stored as reduced carbon.
  8. from the original on 2016-09-02. Retrieved 2016-02-21. The average global rate of photosynthesis is 130 TW.
  9. ^ "World Consumption of Primary Energy by Energy Type and Selected Country Groups, 1980–2004". Energy Information Administration. July 31, 2006. Archived from the original (XLS) on November 9, 2006. Retrieved 2007-01-20.
  10. from the original on 2018-09-25. Retrieved 2018-04-20.
  11. ^ .
  12. .
  13. ^ Plants: Diversity and Evolution
  14. from the original on 2023-01-19. Retrieved 2019-04-17.
  15. ^ Anaerobic Photosynthesis, Chemical & Engineering News, 86, 33, August 18, 2008, p. 36
  16. from the original on 2020-07-28. Retrieved 2020-01-17.
  17. ^ "Scientists discover unique microbe in California's largest lake". Archived from the original on 2009-07-12. Retrieved 2009-07-20.
  18. .
  19. .
  20. ^ .
  21. .
  22. on 2014-11-02. Retrieved 2009-02-03.
  23. .
  24. ^ .
  25. ^ "Yachandra/Yano Group". Lawrence Berkeley National Laboratory. Archived from the original on 2019-07-22. Retrieved 2019-07-22.
  26. PMID 18250316
    .
  27. ^ .
  28. .
  29. from the original on 2023-01-19. Retrieved 2019-04-17.
  30. .
  31. .
  32. .
  33. .
  34. .
  35. .
  36. ^ Miyamoto K. "Chapter 1 – Biological energy production". Renewable biological systems for alternative sustainable energy production (FAO Agricultural Services Bulletin – 128). Food and Agriculture Organization of the United Nations. Archived from the original on 7 September 2013. Retrieved 4 January 2009.
  37. ^ from the original on 2022-04-07. Retrieved 2018-04-03.
  38. ^ .
  39. ^ Govindjee R. "What is Photosynthesis?". Biology at Illinois. Archived from the original on 27 May 2014. Retrieved 17 April 2014.
  40. from the original on 2023-01-19. Retrieved 2019-04-17.
  41. ^ Baker NR, Oxborough K (2004). "Chapter 3: Chlorophyll fluorescence as a probe of photosynthetic productivity". In Papaqeorgiou G, Govindjee (eds.). Chlorophylla Fluorescence a Signature of Photosynthesis. Dordrecht, The Netherlands: Springer. pp. 66–79. Archived from the original on 2023-01-19. Retrieved 2019-04-17.
  42. .
  43. .
  44. .
  45. .
  46. ^ .
  47. ^ .
  48. ^ a b c d Ribas-Carbo M, Flexas J, Robinson SA, Tcherkez GG (2010). "In vivo measurement of plant respiration". University of Wollongong Research Online.
  49. ^
    PMID 14512377
    .
  50. .
  51. .
  52. .
  53. ^ Palmer J (21 June 2013). "Plants 'seen doing quantum physics'". BBC News. Archived from the original on 3 October 2018. Retrieved 21 June 2018.
  54. ^ Lloyd S (10 March 2014). "Quantum Biology: Better living through quantum mechanics". The Nature of Reality. Nova: PBS Online; WGBH Boston. Archived from the original on 3 July 2017. Retrieved 8 September 2017.
  55. S2CID 25760719
    .
  56. ^ Davis K (2 October 2004). "Photosynthesis got a really early start". New Scientist. Archived from the original on 1 May 2015. Retrieved 8 September 2017.
  57. ^ Hooper R (19 August 2006). "Revealing the dawn of photosynthesis". New Scientist. Archived from the original on 24 May 2015. Retrieved 8 September 2017.
  58. PMID 29560463. Archived from the original
    on 1 April 2019. Retrieved 23 March 2018.
  59. ^ Howard V (7 March 2018). "Photosynthesis Originated A Billion Years Earlier Than We Thought, Study Shows". Astrobiology Magazine. Archived from the original on October 1, 2020. Retrieved 23 March 2018.{{cite web}}: CS1 maint: unfit URL (link)
  60. S2CID 266752333
    .
  61. .
  62. .
  63. .
  64. .
  65. .
  66. ]
  67. .
  68. .
  69. .
  70. from the original on 2023-01-19. Retrieved 2019-04-17.
  71. .
  72. . CBC/Radio-Canada. 26 October 2018.
  73. .
  74. .
  75. .
  76. .
  77. .
  78. .
  79. .
  80. ^ "Cyanobacteria: Fossil Record". Ucmp.berkeley.edu. Archived from the original on 2010-08-24. Retrieved 2010-08-26.
  81. from the original on 2023-01-19. Retrieved 2019-04-17.
  82. from the original on 2023-01-19. Retrieved 2019-04-17.
  83. ^ .
  84. .
  85. ^ Rabinowitch EI (1945). Photosynthesis and Related Processes. Vol. 1. Archived from the original on 2020-08-06. Retrieved 2019-12-14 – via Biodiversity Heritage Library.
  86. S2CID 21567780. Archived from the original
    (PDF) on 2008-03-09. Retrieved 2015-08-27.
  87. ^ Otto Warburg – Biography Archived 2010-12-15 at the Wayback Machine. Nobelprize.org (1970-08-01). Retrieved on 2011-11-03.
  88. (PDF) from the original on 2018-06-24. Retrieved 2018-06-26.
  89. .
  90. .
  91. ^ "Photosynthesis". Online Etymology Dictionary. Archived from the original on 2013-03-07. Retrieved 2013-05-23.
  92. Perseus Project
    .
  93. Perseus Project
    .
  94. S2CID 11265932
    .
  95. .
  96. .
  97. .
  98. .
  99. .
  100. ^ .
  101. ^ a b c El-Sharkawy MA, Hesketh JD (1986). "Citation Classic-Photosynthesis among species in relation to characteristics of leaf anatomy and CO2 diffusion resistances" (PDF). Curr. Cont./Agr.Biol.Environ. 27: 14.
  102. ^ Haberlandt G (1904). Physiologische Pflanzanatomie. Leipzig: Engelmann. Archived from the original on 2023-01-19. Retrieved 2019-04-17.
  103. ^ El-Sharkawy MA (1965). Factors Limiting Photosynthetic Rates of Different Plant Species (Ph.D. thesis). The University of Arizona, Tucson.
  104. ^ Karpilov YS (1960). "The distribution of radioactvity in carbon-14 among the products of photosynthesis in maize". Proc. Kazan Agric. Inst. 14: 15–24.
  105. PMID 16656075
    .
  106. .
  107. . Retrieved 2023-02-09.
  108. from the original on 2023-01-19. Retrieved 2019-04-17.
  109. from the original on 2023-01-19. Retrieved 2019-04-17.
  110. .

Further reading

Books

Papers

External links