photosynthesis

photosynthesis
photosynthetic /foh'teuh sin thet"ik/, adj.photosynthetically, adv.
/foh'teuh sin"theuh sis/, n. Biol., Biochem.
(esp. in plants) the synthesis of complex organic materials, esp. carbohydrates, from carbon dioxide, water, and inorganic salts, using sunlight as the source of energy and with the aid of chlorophyll and associated pigments.
[1895-1900; PHOTO- + SYNTHESIS]

* * *

Process by which green plants and certain other organisms transform light into chemical energy.

In green plants, light energy is captured by chlorophyll in the chloroplasts of the leaves and used to convert water, carbon dioxide, and minerals into oxygen and energy-rich organic compounds (simple and complex sugars) that are the basis of both plant and animal life. Photosynthesis consists of a number of photochemical and enzymatic reactions. It occurs in two stages. During the light-dependent stage (light reaction), chlorophyll absorbs light energy, which excites some electrons in the pigment molecules to higher energy levels; these leave the chlorophyll and pass along a series of molecules, generating formation of NADPH (an enzyme) and high-energy ATP molecules. Oxygen, released as a by-product, passes into the atmosphere through pores in the leaves. NADPH and ATP drive the second stage, the dark reaction (or Calvin cycle, discovered by Melvin Calvin), which does not require light. During this stage glucose is generated using atmospheric carbon dioxide. Photosynthesis is crucial for maintaining life on earth; if it ceased, there would soon be little food or other organic matter on the planet, and most organisms would disappear.

* * *

Introduction
 the process by which green plants (plant) and certain other organisms transform light energy into chemical energy. During photosynthesis in green plants, light energy is captured and used to convert water, carbon dioxide, and minerals into oxygen and energy-rich organic compounds.

      It would be impossible to overestimate the importance of photosynthesis in the maintenance of life on Earth. If photosynthesis ceased, there would soon be little food or other organic matter on Earth. Most organisms would disappear, and in time the Earth's atmosphere would become nearly devoid of gaseous oxygen. The only organisms able to exist under such conditions would be the chemosynthetic bacteria, which can utilize the chemical energy of certain inorganic compounds and thus are not dependent on the conversion of light energy.

      Photosynthesis also is responsible for the “fossil fuels (fossil fuel)” (i.e., coal, oil, and gas) that power industrial society. In past ages, green plants and small organisms that fed on plants increased faster than they were consumed, and their remains were deposited in the Earth's crust by sedimentation and other geological processes. There, protected from oxidation, these organic remains were slowly converted to fossil fuels. These fuels not only provide much of the energy used in factories, homes, and transportation, but they also serve as the raw material for plastics and other synthetic products. Unfortunately, modern civilization is using up in a few centuries the excess of photosynthetic production accumulated over millions of years.

      Requirements for food, materials, and energy in a world where human population is rapidly growing have created a need to increase both the amount of photosynthesis and the efficiency of converting photosynthetic output into products useful to people. One response to these needs—the so-called “Green Revolution”—has achieved enormous improvements in agricultural yield through the use of chemical fertilizers, pest and plant disease control, plant breeding, and mechanized tilling, harvesting, and crop processing. This effort has limited severe famines to a few areas of the world despite rapid population growth, but it has not eliminated widespread malnutrition.

      A second agricultural revolution, based on plant genetic engineering, may lead to increases in plant productivity and thereby partially alleviate malnutrition. Since the 1970s, molecular biologists have possessed the means to manipulate a plant's genetic material (DNA) to achieve improvements in disease and drought resistance, product yield and quality, frost hardiness, and other desirable properties. In the future, such genetic engineering may result in improvements in the process of photosynthesis.

General characteristics

Development of the idea
      The study of photosynthesis began in 1771, with observations made by the English chemist Joseph Priestley (Priestley, Joseph). Priestley had burned a candle in a closed container until the air within the container could no longer support combustion. He then placed a sprig of mint plant in the container and discovered that after several days the mint had produced some substance (later recognized as oxygen) that enabled the confined air to again support combustion. In 1779 the Dutch physician Jan Ingenhousz (Ingenhousz, Jan) expanded upon Priestley's work, showing that the plant must be exposed to light if the combustible substance (i.e., oxygen) was to be restored; he also demonstrated that this process required the presence of the green tissues of the plant.

      In 1782 it was demonstrated that the combustion-supporting gas (oxygen) was formed at the expense of another gas, or “fixed air,” which had been identified the year before as carbon dioxide. Gas-exchange experiments in 1804 showed that the gain in weight of a plant grown in a carefully weighed pot was the sum of carbon, which came entirely from absorbed carbon dioxide, and water taken up by plant roots. Almost half a century passed before the concept of chemical energy developed sufficiently to permit the discovery (in 1845) that light energy from the sun is stored as chemical energy in products formed during photosynthesis.

Overall reaction of photosynthesis
      In chemical terms, photosynthesis is a light-energized oxidation–reduction process (oxidation–reduction reaction). (Oxidation refers to the removal of electrons from a molecule; reduction refers to the gain of electrons by a molecule.) In plant photosynthesis, the energy of light is used to drive the oxidation of water (H2O), producing oxygen gas (O2), hydrogen ions (hydrogen ion) (H+), and electrons (electron). Most of the removed electrons and hydrogen ions ultimately are transferred to carbon dioxide (CO2), which is reduced to organic products. Other electrons and hydrogen ions are used to reduce nitrate and sulfate to amino and sulfhydryl groups in amino acids (amino acid), which are the building blocks of proteins. In most green cells, carbohydrates (carbohydrate)—especially starch and the sugar sucrose—are the major direct organic products of photosynthesis. The overall reaction in which carbohydrates—represented by the general formula (CH2O)—are formed during plant photosynthesis can be indicated by the following equation:

      This equation is merely a summary statement, for the process of photosynthesis actually involves numerous complex reactions. These reactions occur in two stages: the “light” stage, consisting of photochemical (photochemical reaction) (i.e., light-dependent) reactions; and the “dark” stage, comprising chemical reactions controlled by enzymes (enzyme) (organic catalysts). During the first stage, the energy of light is absorbed and used to drive a series of electron transfers, resulting in the synthesis of the energy-rich compound adenosine triphosphate (ATP) and the electron donor reduced nicotine adenine dinucleotide phosphate (NADPH). During the dark stage, the ATP and NADPH formed in the light reactions are used to reduce carbon dioxide to organic carbon compounds. This assimilation of inorganic carbon into organic compounds is called carbon fixation.

      During the 20th century, comparisons between photosynthetic processes in green plants and in certain photosynthetic sulfur bacteria (sulfur bacterium) provided important information about the photosynthetic mechanism. Sulfur bacteria use hydrogen sulfide (H2S) as a source of hydrogen atoms (hydrogen) and produce sulfur instead of oxygen during photosynthesis. The overall reaction is

      In the 1930s Dutch biologist Cornelis van Niel recognized that the utilization of carbon dioxide to form organic compounds was similar in the two types of photosynthetic organisms. Suggesting that differences existed in the light-dependent stage and in the nature of the compounds used as a source of hydrogen atoms, he proposed that hydrogen was transferred from hydrogen sulfide (in bacteria) or water (in green plants) to an unknown acceptor (called A), which was reduced to H2A. During the dark reactions, which are similar in both bacteria and green plants, the reduced acceptor (H2A) reacted with carbon dioxide (CO2) to form carbohydrate (CH2O) and to oxidize the unknown acceptor to A. This putative reaction can be represented as:

      Van Niel's proposal was important because the popular (but incorrect) theory had been that oxygen was removed from carbon dioxide (rather than hydrogen from water) and that carbon then combined with water to form carbohydrate (rather than the hydrogen from water combining with CO2 to form CH2O).

      By 1940 chemists were using heavy isotopes (radioactive isotope) to follow the reactions (isotopic tracer) of photosynthesis. Water marked with an isotope of oxygen (18O) was used in early experiments. Plants that photosynthesized in the presence of water containing H218O produced oxygen gas containing 18O; those that photosynthesized in the presence of normal water produced normal oxygen gas. These results provided strong support for van Niel's theory that the oxygen gas produced during photosynthesis is derived from water.

Basic products of photosynthesis
      As has been stated, carbohydrates are the most important direct organic product of photosynthesis in the majority of green plants. The formation of a simple carbohydrate, glucose, is indicated by a chemical equation,

      Little free glucose is produced in plants; instead, glucose units are linked together to form starch or are joined with fructose, another sugar, to form sucrose (see carbohydrate).

      Not only carbohydrates, as was once thought, but also amino acids, proteins, lipids (or fats), pigments, and other organic components of green tissues are synthesized during photosynthesis. Minerals supply the elements (e.g., nitrogen, N; phosphorus, P; sulfur, S) required to form these compounds. Chemical bonds are broken between oxygen (O) and carbon (C), hydrogen (H), nitrogen, and sulfur, and new bonds are formed in products that include gaseous oxygen (O2) and organic compounds. More energy is required to break the bonds between oxygen and other elements (e.g., in water, nitrate, and sulfate) than is released when new bonds form in the products. This difference in bond energy accounts for a large part of the light energy stored as chemical energy in the organic products formed during photosynthesis. Additional energy is stored in making complex molecules from simple ones.

Evolution of the process
      Although life and the quality of the atmosphere today depend on photosynthesis, it is likely that green plants evolved long after the first living cells. When the Earth was young, electrical storms and solar radiation probably provided the energy for the synthesis of complex molecules from abundant simpler ones, such as water, ammonia, and methane. The first living cells probably evolved from these complex molecules (see life: The primitive atmosphere (life)). For example, the accidental joining together (condensation) of the amino acid glycine and the fatty acid acetate may have formed complex organic molecules known as porphyrins (porphyrin); these molecules, in turn, may have evolved further into coloured molecules called pigments (coloration); e.g., chlorophylls (chlorophyll) of green plants, bacteriochlorophyll of photosynthetic bacteria, hemin (the red pigment of blood), and cytochromes (cytochrome), a group of pigment molecules essential in both photosynthesis and cellular respiration.

      Primitive coloured cells then had to evolve mechanisms for using the light energy absorbed by their pigments. At first, the energy may have been used immediately to initiate reactions useful to the cell. As the process for utilization of light energy continued to evolve, however, a larger part of the absorbed light energy probably was stored as chemical energy, to be used to maintain life. Green plants, with their ability to use light energy to convert carbon dioxide and water to carbohydrates and oxygen, are the culmination of this evolutionary process.

      The first oxygenic (oxygen-producing) cells probably were the cyanophytes, or “ blue-green algae,” which appeared about 2,000,000,000 to 3,000,000,000 years ago. These microscopic organisms are believed to have greatly increased the oxygen content of the atmosphere, making possible the development of aerobic (oxygen-using) organisms. Cyanophytes are prokaryotic cells (prokaryote); that is, they contain no distinct, membrane-enclosed subcellular particles (organelles), such as nuclei and chloroplasts. Green plants, by contrast, are composed of eukaryotic cells (eukaryote), in which the photosynthetic apparatus is contained within membrane-bound chloroplasts. There is a theory that the first photosynthetic eukaryotes were red algae that may have developed when nonphotosynthetic eukaryotic cells engulfed cyanophytes. Within the host cells, these cyanophytes are thought to have evolved into chloroplasts. Alternatively, the ancestors of chloroplasts in green plants may have been another oxygenic prokaryote like Prochloron, an organism that has been found only growing symbiotically inside ascidians.

      There are a number of photosynthetic bacteria that are not oxygenic (e.g., the sulfur bacteria previously discussed). The evolutionary pathway that led to these bacteria diverged from the one that resulted in oxygenic organisms. In addition to the absence of oxygen production, nonoxygenic photosynthesis differs from oxygenic photosynthesis in two other ways: light of longer wavelengths is absorbed and used by pigments called bacteriochlorophylls, and reduced compounds other than water (such as hydrogen sulfide or organic molecules) provide the electrons needed for the reduction of carbon dioxide.

Factors that influence the rate of photosynthesis
      The rate of photosynthesis is defined in terms of the rate of oxygen production either per unit mass (or area) of green plant tissues or per unit weight of total chlorophyll. The amount of light, the carbon dioxide supply, the temperature, the water supply, and the availability of minerals are the most important environmental factors that directly affect the rate of photosynthesis in land plants. The rate of photosynthesis also is determined by the plant species and its physiological state—e.g., its health, its maturity, and whether or not it is in flower.

Light intensity and temperature
      As has been mentioned, the complex mechanism of photosynthesis includes a photochemical, or light-dependent, stage and an enzymatic, or dark, stage that involves chemical reactions. These stages can be distinguished by studying the rates of photosynthesis at various degrees of light saturation (i.e., intensity) and at different temperatures (temperature). Over a range of moderate temperatures and at low to medium light (luminous intensity) intensities (relative to the normal range of the plant species), the rate of photosynthesis increases as the intensity increases and is independent of temperature. As the light intensity increases to higher levels, however, the rate becomes increasingly dependent on temperature and less dependent on intensity; light “saturation” is achieved at a specific light intensity, and the rate then is dependent only on temperature if all other factors are constant. In the light-dependent range before saturation, therefore, the rate of photosynthesis is determined by the rates of photochemical steps. At high light intensities, some of the chemical reactions of the dark stage become rate-limiting. At light saturation, rate increases with temperature until a point is reached beyond which no further rate increase can occur. In many land plants, moreover, a process called photorespiration occurs at high light intensities and temperatures. Photorespiration competes with photosynthesis and limits further increases in the rate of photosynthesis, especially if the supply of water is limited (see below Photorespiration (photosynthesis)).

Carbon dioxide
      Included among the rate-limiting steps of the dark stage of photosynthesis are the chemical reactions by which organic compounds are formed using carbon dioxide as a carbon source. The rates of these reactions can be increased somewhat by increasing the carbon dioxide concentration. During the past century, the level of carbon dioxide in the atmosphere has been rising due to the extensive combustion of fossil fuels. The atmospheric level of carbon dioxide climbed from about 0.028 percent in 1860 to 0.0315 percent by 1958 (when improved measurements began), and to 0.034 percent by 1981. This increase in carbon dioxide directly increases plant photosynthesis, but the size of the increase depends on the species and physiological condition of the plant. Furthermore, if increasing levels of atmospheric carbon dioxide result in climatic changes, including increased global temperatures as some meteorologists predict, these changes will affect photosynthesis rates.

Water
      For land plants, water availability can function as a limiting factor in photosynthesis and plant growth. Besides the requirement for water in the photosynthetic reaction itself, water is transpired (transpiration) from the leaves; that is, water evaporates from the leaves to the atmosphere via the stomates. These stomates (stomate) are small openings through the leaf epidermis, or outer skin; they permit the entry of carbon dioxide but also allow the exit of water vapour. The stomates open and close according to the physiological needs of the leaf. In hot and arid climates the stomates may close to conserve water, but this closure limits the entry of carbon dioxide and hence the rate of photosynthesis, while the wasteful process of photorespiration may increase. If the level of carbon dioxide in the atmosphere increases, more carbon dioxide could enter through a smaller opening of the stomates, so that more photosynthesis could occur with a given supply of water.

Minerals
      Several minerals are required for healthy plant growth and for maximum rates of photosynthesis. Nitrate or ammonia, sulfate, phosphate, iron, magnesium, and potassium are required in substantial amounts for the synthesis of amino acids, proteins, coenzymes, deoxyribonucleic acid (DNA) and ribonucleic acid (RNA), chlorophyll and other pigments, and other essential plant constituents. Smaller amounts of such elements as manganese, copper, and chlorine are required in photosynthesis. Some other trace elements are needed for various nonphotosynthetic functions in plants.

Internal factors
      Each plant species adapts to a range of environmental factors. Within this normal range of conditions, complex regulatory mechanisms in the plant's cells adjust the activities of enzymes (i.e., organic catalysts). These adjustments maintain a balance in the overall photosynthetic process and control it in accordance with the needs of the whole plant. With a given plant species, for example, doubling the carbon dioxide level might cause a temporary increase of nearly twofold in the rate of photosynthesis; a few hours later, however, the rate might fall to the original level because photosynthesis had made more sucrose than the rest of the plant could use. By contrast, another plant species provided with such carbon dioxide enrichment might be able to use more sucrose and would continue to photosynthesize and to grow faster throughout most of its life cycle.

Energy efficiency of photosynthesis (energy)
      The energy efficiency of photosynthesis is the ratio of the energy stored to the energy of light absorbed. The chemical energy stored is the difference between that contained in gaseous oxygen and organic compound products and the energy of water, carbon dioxide, and other reactants. The amount of energy stored can only be estimated because many products are formed, and these vary with the plant species and environmental conditions. If the equation for glucose formation given earlier is used to approximate the actual storage process, the production of one mole (i.e., 6.02 × 1023 molecules; abbreviated N) of oxygen and one-sixth mole of glucose results in the storage of about 117 kilocalories (kcal) of chemical energy. This amount must then be compared to the energy of light absorbed to produce one mole of oxygen in order to calculate the efficiency of photosynthesis.

      Light can be described as a wave of particles known as photons (photon); these are units of energy, or light quanta. The quantity N photons is called an einstein. The energy of light varies inversely with the length of the photon waves; that is, the shorter the wavelength, the greater the energy content. The energy (e) of a photon is given by the equation e = hc/λ, where c is the velocity of light, h is Planck's constant, and λ is the light wavelength. The energy (E) of an einstein is E = Ne = Nhc/λ = 28,600/λ, when E is in kilocalories and λ is given in nanometres (nm; 1 nm = 10−9 metres). An einstein of red light with a wavelength of 680 nm has an energy of about 42 kcal. Blue light has a shorter wavelength and therefore more energy than red light. Regardless of whether the light is blue or red, however, the same number of einsteins are required for photosynthesis per mole of oxygen formed. The part of the solar spectrum used by plants has an estimated mean wavelength of 570 nanometres; therefore, the energy of light used during photosynthesis is approximately 28,600/570, or 50 kilocalories per einstein.

      In order to compute the amount of light energy involved in photosynthesis, one other value is needed: the number of einsteins absorbed per mole of oxygen evolved. This is called the quantum requirement. The minimum quantum requirement for photosynthesis under optimal conditions is about nine. Thus the energy used is 9 × 50, or 450 kilocalories per mole of oxygen evolved. Therefore, the estimated maximum energy efficiency of photosynthesis is the energy stored per mole of oxygen evolved—117 kilocalories—divided by 450; that is, 117/450, or 26 percent.

      The actual percentage of solar energy stored by plants is much less than the maximum energy efficiency of photosynthesis. An agricultural crop in which the biomass (total dry weight) stores as much as 1 percent of total solar energy received on an annual area-wide basis is exceptional, although a few cases of higher yields (perhaps as much as 3.5 percent in sugarcane) are reported. There are several reasons for this difference between the predicted maximum efficiency of photosynthesis and the actual energy stored in biomass. First, more than half of the incident sunlight is composed of wavelengths too long to be absorbed, while some of the remainder is reflected or lost to the leaves. Consequently, plants can at best absorb only about 34 percent of the incident sunlight. Second, plants must carry out a variety of physiological processes in such nonphotosynthetic tissues as roots and stems; these processes, as well as cellular respiration in all parts of the plant, use up stored energy. Third, rates of photosynthesis in bright sunlight sometimes exceed the needs of the plants, resulting in the formation of excess sugars and starch. When this happens, the regulatory mechanisms of the plant slow down the process of photosynthesis, allowing more absorbed sunlight to go unused. Fourth, in many plants, energy is wasted by the process of photorespiration. Finally, the growing season may last only a few months of the year; sunlight received during other seasons is not used. Furthermore, it should be noted that if only agricultural products (e.g., seeds, fruits, and tubers, rather than total biomass) are considered as the end product of the energy conversion process of photosynthesis, the efficiency falls even further.

Chloroplasts, the photosynthetic units of green plants (chloroplast)
      The process of plant photosynthesis takes place entirely within the chloroplasts. Detailed studies of the role of these organelles date from the work of the British biochemist Robert Hill. About 1940 Hill discovered that green particles obtained from broken cells could produce oxygen from water in the presence of light and a chemical compound, such as ferric oxalate, able to serve as an electron acceptor. This process is known as the Hill reaction. During the 1950s Daniel Arnon and other American biochemists prepared plant cell fragments in which not only the Hill reaction but also the synthesis of the energy-storage compound ATP occurred. In addition, the coenzyme NADP was used as the final acceptor of electrons, replacing the nonphysiological electron acceptors used by Hill. His procedures were refined further so that individual small pieces of isolated chloroplast membranes, or lamellae, could perform the Hill reaction. These small pieces of lamellae were then fragmented into pieces so small that they performed only the light reactions of the photosynthetic process. It is now possible also to isolate the entire chloroplast so that it can carry out the complete process of photosynthesis, from light absorption, oxygen formation, and the reduction of carbon dioxide to the formation of glucose and other products.

Structural features
      The intricate structural organization of the photosynthetic apparatus is essential for the efficient performance of the complex process of photosynthesis. The chloroplast is enclosed in a double outer membrane, and its size approximates a spheroid about 2,500 nanometres thick and 5,000 nanometres long. Some single-celled algae have one chloroplast that occupies more than half the cell volume. Leaf cells of higher plants contain many chloroplasts, each approximately the size of the one in some algal cells.

      When thin sections of a chloroplast are examined under the electron microscope, several features are apparent. Chief among these are the intricate internal membranes (i.e., the lamellae) and the stroma, a colourless matrix in which the lamellae are embedded. Also visible are starch granules, which appear as dense bodies.

      The stroma is basically a solution of enzymes and small molecules. The dark reactions occur in the stroma, the soluble enzymes of which catalyze the conversion of carbon dioxide and minerals to carbohydrates and other organic compounds. The capacity for carbon fixation and reduction is lost if the outer membrane of the chloroplast is broken, allowing the stroma enzymes to leak out.

      A single lamella, which contains all the photosynthetic pigments, is approximately 10–15 nanometres thick. The lamellae exist in more-or-less flat sheets, a few of which extend through much of the length of the chloroplast. Examination of cross sections of lamellae under the electron microscope shows that their edges are joined to form closed hollow disks that are called thylakoids (“saclike”). The chloroplasts of most higher plants have regions, called grana, in which the thylakoids are very tightly stacked. When viewed by electron microscopy at an oblique angle, the grana appear as stacks of disks. When viewed in cross section, it is apparent that some thylakoids extend from one grana through the stroma into other grana. The thin aqueous spaces inside the thylakoids are believed to be connected with each other via these stroma thylakoids. These thylakoid spaces are isolated from the stroma spaces by the relatively impermeable lamellae.

      The light reactions occur exclusively in the thylakoids. The complex structural organization of lamellae is required for proper thylakoid function; intact thylakoids apparently are necessary for the formation of ATP. Thylakoids that have been broken down to smaller units can no longer form ATP, even when the conversion of light into chemical energy occurs during electron transport in these units. Such lamellar fragments can carry out the Hill reaction, with the transfer of electrons from water to NADP+.

Chemical composition of lamellae
Lipids (lipid)
      Lamellae consist of about equal amounts of lipids and proteins (protein). About one-fourth of the lipid portion of the lamellae consists of pigments and coenzymes; the remainder consists of various lipids, including polar compounds such as phospholipids (phospholipid) and galactolipids. These polar lipid molecules have “head” groups that attract water (i.e., are hydrophilic) and fatty acid “tails” that are oil soluble and repel water (i.e., are hydrophobic). When polar lipids are placed in an aqueous environment, they can line up with the fatty acid tails side by side. A second layer of phospholipids forms tail-to-tail with the first, establishing a lipid bilayer in which the hydrophilic heads are in contact with the aqueous solution on each side of the bilayer. Sandwiched between the heads are the hydrophobic tails, creating a hydrophobic environment from which water is excluded. This lipid bilayer is an essential feature of all biological membranes (see cell: The plasma membrane (cell)). The hydrophobic parts of proteins and lipid-soluble cofactors and pigments are dissolved or embedded in the lipid bilayer. Lamellar membranes can function as electrical insulating material and permit a charge, or potential difference, to develop across the membrane. Such a charge can be a source of chemical or electrical energy.

      Approximately one-fifth of the lamellar lipids are chlorophyll molecules; one type, chlorophyll a, is more abundant than the second type, chlorophyll b. The chlorophyll molecules are specifically bound to small protein molecules. Most of these chlorophyll-proteins are “light-harvesting” pigments. These absorb light and pass its energy on to special chlorophyll a molecules that are directly involved in the conversion of light energy to chemical energy. When one of these special chlorophyll a molecules is excited by light energy (as described later), it gives up an electron. There are two types of these special chlorophyll a molecules: one, called P680, has an absorption spectrum that peaks at 684 nanometres; the other, called P700, shows an absorption peak at 700 nanometres.

      Although chlorophylls are the main light-absorbing molecules in green plants, other pigments such as carotenes (carotene) and carotenoids (carotenoid) (which are responsible for the yellow-orange colour of carrots) also can absorb light and may supplement chlorophyll as the light-absorbing molecules in some plant cells. The light energy absorbed by these pigments must be passed to chlorophyll before conversion to chemical energy can occur.

Proteins
      Many of the lamellar proteins are components of the chlorophyll–protein complexes described above. Other proteins include enzymes and protein-containing coenzymes. Enzymes are required as organic catalysts for specific reactions within the lamellae. Protein coenzymes, also called cofactors (cofactor), include important electron carrier molecules called cytochromes (cytochrome), which are iron-containing pigments with the pigment portions attached to protein molecules. During electron transfer, an electron is accepted by an iron atom in the pigment portion of a cytochrome molecule, which thus is reduced; then the electron is transferred to the iron atom in the next cytochrome carrier in the electron transfer chain, thus oxidizing the first cytochrome and reducing the next one in the chain.

      In addition to the metal atoms found in the pigment portions of cytochrome molecules, metal atoms also are found in other protein molecules of the lamellae. In proteins with a total molecular weight of 900,000 (based on the weight of hydrogen as one), there are two atoms of manganese, 10 atoms of iron, and six atoms of copper. These metal atoms are required for the catalytic activity of some of the enzymes important in photosynthesis. The manganese atoms are involved in water-splitting and oxygen formation. Both copper- and iron-containing proteins function in electron transport between water and the final electron-acceptor molecule of the light stage of photosynthesis, an iron-containing protein called ferredoxin. Ferredoxin is a soluble component in the chloroplasts. In its reduced form, it gives electrons directly to the systems that reduce nitrate and sulfate and via NADPH to the system that reduces carbon dioxide. A copper-containing protein called plastocyanin (PC) carries electrons at one point in the electron transport chain. PC molecules are water soluble and can move through the inner space of the thylakoids, carrying electrons from one place to another.

Quinones
      Small molecules called plastoquinones are found in substantial numbers in the lamellae. Like the cytochromes, quinones have important roles in carrying electrons between the components of the light reactions. Since they are lipid soluble, they can diffuse through the membrane. They can carry one or two electrons and, in their reduced form (with added electrons), they carry hydrogen atoms that can be released as hydrogen ions when the added electrons are passed on, for example, to a cytochrome.

The process of photosynthesis: the light reactions

Light absorption and energy transfer
      The light energy absorbed by a chlorophyll molecule excites (excitation) some electrons within the structure of the molecule to higher energy levels, or excited states. Light of shorter wavelength (such as blue) has more energy than light of longer wavelength (such as red), so that absorption of blue light creates an excited state of higher energy. A molecule raised to this higher energy state quickly gives up the “extra” energy as heat and falls to its lowest excited state. This lowest excited state is similar to that of a molecule that has just absorbed the longest wavelength light capable of exciting it. In the case of chlorophyll a, this lowest excited state corresponds to that of a molecule that has absorbed red light of about 680 nanometres.

      The return of a chlorophyll a molecule from its lowest excited state to its original low-energy state (ground state) requires the release of the extra energy of the excited state. This can occur in one of several ways. In photosynthesis, most of this energy is conserved as chemical energy by the transfer of an electron from a special chlorophyll a molecule (P680 or P700) to an electron acceptor. When this electron transfer is blocked by inhibitors, such as the herbicide dichlorophenylmethylurea (DCMU), or by low temperature, the energy can be released as red light. Such re-emission of light is called fluorescence. The examination of fluorescence from photosynthetic material in which electron transfer has been blocked has proved to be a useful tool for scientists studying the light reactions.

The pathway of electrons
 The general features of a widely accepted mechanism for photoelectron transfer, in which two light reactions occur during the transfer of electrons from water to carbon dioxide, were proposed by Robert Hill and Fay Bendall in 1960. A modified scheme for this mechanism is shown in Figure 1—>. In this figure the vertical scale represents the relative potential (in volts) of various cofactors of the electron-transfer chain to be oxidized or reduced. Molecules that in their oxidized form have the strongest affinity for electrons (i.e., are strong oxidizing agents) are near the bottom of the scale. Molecules that in their oxidized form are difficult to reduce are near the top of the scale; once they have accepted electrons, these molecules are strong reducing agents.

      The actual photochemical steps are indicated by the two vertical arrows, which signify that the special pigments P680 and P700 receive light energy from the light-harvesting chlorophyll-protein molecules and are raised in energy from their ground state to excited states, symbolized as P*680 and P*700. In their excited state, these pigments are extremely strong reducing agents that quickly transfer electrons to the first acceptor. These first acceptors also are strong reducing agents and rapidly pass electrons to more stable carriers. In light reaction II the first acceptor may be pheophytin (Ph; a molecule similar to chlorophyll), which also has a strong reducing potential and quickly transfers electrons to the next acceptor. QA and QB are special quinones, similar to plastoquinone. They receive electrons from pheophytin and pass them to the intermediate electron carriers, which include the plastoquinone (PQ) pool and the cytochromes b and f (Cytb and Cytf) associated in a complex with an iron-sulfur protein (Fe-S).

      In light reaction I the identity of the first electron acceptor, X, is not known. It passes electrons on to iron-sulfur proteins (Fe-S-protein) in the lamellar membrane, after which the electrons flow to ferredoxin (Fd), a small, water-soluble iron-sulfur protein. When NADP+ and a suitable enzyme are present, two ferredoxin molecules, carrying one electron each, transfer two electrons to NADP+, which picks up a proton (i.e., a hydrogen ion) and becomes NADPH.

      Each time a P680 or P700 molecule gives up an electron, it returns to its ground (unexcited) state, but with a positive charge due to the loss of the electron. These positively charged ions are extremely strong oxidizing agents that remove an electron from a suitable donor. The P680+ of light reaction II is capable of taking electrons from water in the presence of appropriate catalysts. There is good evidence that two or more manganese atoms complexed with protein are involved in this catalysis, taking four electrons from two water molecules (with release of four hydrogen ions). The manganese-protein complex gives up these electrons one at a time via an unidentified carrier Z to P680+, reducing it to P680. When manganese is selectively removed by chemical treatment, the thylakoids lose the capacity to oxidize water, but all other parts of the electron pathway remain intact.

 In light reaction I, P700+ recovers electrons from plastocyanin (PC), which in turn receives them from intermediate carriers, including the plastoquinone pool and cytochrome b and cytochrome f molecules. The pool of intermediate carriers may receive electrons from water via light reaction II and QA and QB. Transfer of electrons from water to ferredoxin via the two light reactions and intermediate carriers is called noncyclic electron flow. Alternately, electrons may be transferred only by light reaction I, in which case they are recycled from ferredoxin back to the intermediate carriers (dashed line, Figure 1—>). This process is called cyclic electron flow.

Evidence of two light reactions
      Many lines of evidence support the concept of electron flow via two light reactions. An early study by the U.S. biochemist Robert Emerson employed the algae Chlorella, which was illuminated with red light alone, with blue light alone, and with red and blue light at the same time. Oxygen evolution was measured in each case. It was substantial with blue light alone but not with red light alone. With both red and blue light together, the amount of oxygen evolved far exceeded the sum of that seen with blue and red light alone. These experimental data pointed to the existence of two types of light reactions that, when operating in tandem, would yield the highest rate of oxygen evolution. It is now known that light reaction I can use light of a slightly longer wavelength than red (λ = 680 nanometres), while light reaction II requires light with a wavelength of 680 nanometres or shorter.

 Since those early studies, the two light reactions have been separated in many ways, including separation of the membrane particles in which each reaction occurs. As discussed previously, lamellae can be disrupted mechanically into fragments that absorb light energy and break the bonds of water molecules (i.e., oxidize water) to produce oxygen, hydrogen ions, and electrons. These electrons can be transferred to ferredoxin, the final electron acceptor of the light stage. No transfer of electrons from water to ferredoxin occurs if the herbicide DCMU is present. The subsequent addition of certain reduced dyes (i.e., electron donors) restores the light reduction of NADP+ but without oxygen production, suggesting that light reaction I but not light reaction II is functioning. It is now known that DCMU blocks the transfer of electrons from QA to the PQ pool (see Figure 1—>).

      When treated with certain detergents, lamellae can be broken down into smaller particles capable of carrying out single light reactions. One type of particle can absorb light energy, oxidize water, and produce oxygen (light reaction II), but a special dye molecule must be supplied to accept the electrons. In the presence of electron donors, such as a reduced dye, a second type of lamellar particle can absorb light and transfer electrons from the electron donor to ferredoxin (light reaction I).

Photosystems I and II
      The structural and photochemical properties of the minimum particles capable of performing light reactions I and II have received much study. Treatment of lamellar fragments with neutral detergents releases these particles, designated photosystem I and photosystem II, respectively. Subsequent harsher treatment (with charged detergents) and separation of the individual polypeptides with electrophoretic techniques has helped identify the components of the photosystems. Each photosystem consists of a light-harvesting complex and a core complex. Each core complex contains a reaction centre with the pigment (either P700 or P680) that can be photochemically oxidized, together with electron acceptors and electron donors. In addition, the core complex has some 40 to 60 chlorophyll molecules bound to proteins. In addition to the light absorbed by the chlorophyll molecules in the core complex, the reaction centres receive a major part of their excitation from the pigments (coloration) of the light-harvesting complex.

Quantum requirements
      The quantum requirements of the individual light reactions of photosynthesis are defined as the number of light photons absorbed for the transfer of one electron. The quantum requirement for each light reaction has been found to be approximately one photon. The total number of quanta required, therefore, to transfer the four electrons that result in the formation of one molecule of oxygen via the two light reactions should be four times two, or eight. It appears, however, that additional light is absorbed and used to form ATP by a cyclic photophosphorylation pathway (see next section). The actual quantum requirement, therefore, probably is nine to 10.

The process of photosynthesis: the conversion of light energy to ATP
      The electron transfers of the light reactions provide the energy for the synthesis of two compounds vital to the dark reactions: NADPH+ to NADPH. In this section, the synthesis of the energy-rich compound ATP is described.

      ATP is formed by the addition of a phosphate group to a molecule of adenosine diphosphate (ADP); or to state it in chemical terms, by the phosphorylation of ADP. This reaction requires a substantial input of energy, much of which is captured in the bond that links the added phosphate group to ADP. Because light energy powers this reaction in the chloroplasts, the production of ATP during photosynthesis is referred to as photophosphorylation.

 Unlike the production of NADPH, the photophosphorylation of ADP occurs in conjunction with both cyclic and noncyclic electron flow (see Figure 1—>). In fact researchers speculate that the sole purpose of cyclic electron flow may be for photophosphorylation, since this process involves no net transfer of electrons to reducing agents. The relative amounts of cyclic and noncyclic flow may be adjusted in accordance with changing physiological needs for ATP and reduced ferrodoxin and NADPH in chloroplasts. In contrast to electron transfer in light reactions I and II, which can occur in membrane fragments, intact thylakoids are required for efficient photophosphorylation. This requirement stems from the special nature of the mechanism linking photophosphorylation to electron flow in the lamellae.

      The theory relating the formation of ATP to electron flow in the membranes of both chloroplasts and mitochondria (mitochondrion) (the organelles responsible for ATP formation during cellular respiration) was first proposed by the English biochemist Peter Mitchell (Mitchell, Peter Dennis). This chemiosmotic theory has been somewhat modified to fit later experimental facts, and there is still debate over many of the details. The general features, however, are widely accepted. A central feature is the formation of a hydrogen ion (proton) concentration gradient and an electrical charge across intact lamellae. The potential energy stored by the proton gradient and electrical charge is then used to drive the energetically unfavourable conversion of ADP and inorganic phosphate (Pi) to ATP and water.

      The manganese-protein complex associated with light reaction II is exposed to the interior of the thylakoid. Consequently, the oxidation of water during light reaction II leads to release of hydrogen ions (protons) into the inner thylakoid space. Furthermore, it is likely that photoreaction II entails the transfer of electrons across the lamella toward its outer face, so that when plastoquinone molecules are reduced they can receive protons from the outside of the thylakoid. When these reduced plastoquinone molecules are oxidized, giving up electrons to the cytochrome-iron-sulfur complex, protons are released inside the thylakoid. Because the lamella is impermeable to them, the release of protons inside the thylakoid by oxidation of both water and plastoquinone leads to a higher concentration of protons inside the thylakoid than outside it. In other words, a proton gradient is established across the lamella. The movement of electrons (negatively charged particles) outward across the lamella during both light reactions results in the establishment of an electrical charge across the lamella. (Some scientists believe, however, that the proton gradient and electrical charge required for ATP formation need not be between inner and outer thylakoid space but only within the membrane.)

      An enzyme complex located partly in and on the lamellae catalyzes the reaction in which ATP is formed from ADP and inorganic phosphate. The reverse of this reaction is catalyzed by an enzyme called ATP-ase, hence the enzyme complex is sometimes called an ATP-ase complex. It is also called the coupling factor. It consists of hydrophilic polypeptides (F1), which project from the outer surface of the lamellae, and hydrophobic polypeptides (F0), which are embedded inside the lamellae. Researchers hypothesize that F0 forms a channel that permits protons to flow through the lamellar membrane to F1. The enzymes in F1 then catalyze ATP formation, using both the proton supply and the lamellar transmembrane charge.

      In summary, the use of light energy for ATP formation occurs indirectly: a proton gradient and electrical charge—built up in or across the lamellae as a consequence of electron flow in the light reactions—provide the energy to drive the synthesis of ATP from ADP and Pi.

The process of photosynthesis: carbon fixation and reduction
      The assimilation of carbon into organic compounds is the result of a complex series of enzymatically regulated chemical reactions—the dark reactions. This term is something of a misnomer, for these reactions can take place in either light or darkness. Furthermore, some of the enzymes involved in the so-called dark reactions become inactive in prolonged darkness.

Elucidation of the carbon pathway
      Radioactive isotopes of carbon (14C) and phosphorus (32P) have been valuable in identifying the intermediate compounds formed during carbon assimilation. A photosynthesizing plant does not strongly discriminate between the natural carbon isotopes and 14C. During photosynthesis in the presence of 14CO2, the compounds formed become labelled with the radioisotope. During very short exposures, only the first intermediates in the carbon-fixing pathway become labelled. Early investigations showed that some radioactive products were formed even when the light was turned off and the 14CO2 was added just afterward in the dark, confirming the nature of the carbon fixation as a “dark” reaction.

 The U.S. biochemist Melvin Calvin (Calvin, Melvin), a Nobel Prize recipient for his work on the carbon reduction cycle, allowed green plants to photosynthesize in the presence of radioactive carbon dioxide for a few seconds under various experimental conditions. Products that became labelled with radioactive carbon during Calvin's experiments included a three-carbon compound called 3-phosphoglycerate (abbreviated PGA, see Figure 2—>), sugar phosphates, amino acids, sucrose, and carboxylic acids. When photosynthesis was stopped after two seconds, the principal radioactive product was PGA, which therefore was identified as the first compound formed during carbon dioxide fixation in green plants.

      Further studies with 14C as well as with inorganic phosphate labelled with 32P led to the mapping of the carbon fixation and reduction pathway called the reductive pentose phosphate cycle (RPP cycle). An additional pathway for carbon transport in certain plants was later discovered in other laboratories (see below acids (photosynthesis)). All the steps in these pathways can be carried out in the laboratory by isolated enzymes in the dark. Several steps require the ATP or NADPH generated by the light reactions. In addition, some of the enzymes are fully active only when conditions simulate those in green cells exposed to light. In vivo, these enzymes are active during photosynthesis but not in the dark.

The reductive pentose phosphate cycle
 Overall reaction. The RPP cycle, in which carbon is fixed, reduced, and utilized, involves the formation of intermediate sugar phosphates in a cyclic sequence (Figure 2—>). One complete RPP cycle incorporates three molecules of carbon dioxide and produces one molecule of the three-carbon compound glyceraldehyde-3-phosphate (Gal3P). This three-carbon sugar phosphate usually is either exported from the chloroplasts or is converted to starch (dashed lines).

 Figure 2—> summarizes the main steps that take place during one complete turn of the RPP cycle. ATP and NADPH formed during the light reactions are utilized for key steps in this pathway and provide the energy and reducing equivalents (i.e., electrons) to drive the sequence in the direction shown. For each molecule of carbon dioxide that is fixed, two molecules of NADPH and three molecules of ATP from the light reactions are required. The overall reaction can be represented as follows:

      The cycle is composed of four stages: (1) carboxylation, (2) reduction, (3) isomerization/condensation/dismutation, and (4) phosphorylation.

 The initial incorporation of carbon dioxide, which is catalyzed by the enzyme ribulose 1,5-bisphosphate carboxylase, proceeds by the addition of carbon dioxide to the five-carbon compound ribulose 1,5-bisphosphate (RuBP) and the splitting of the resulting unstable six-carbon compound into two molecules of PGA, a three-carbon compound. As indicated in Figure 2—>, this reaction occurs three times during each complete turn of the cycle; thus, six molecules of PGA are produced.

      The six molecules of PGA are first phosphorylated with ATP by the enzyme PGA-kinase, yielding six molecules of 1,3-diphosphoglycerate (DPGA). These then are reduced with NADPH and the enzyme glyceraldehyde-3-phosphate dehydrogenase to give six molecules of Gal3P. These reactions are the reverse of two steps of the process glycolysis in cellular respiration (see Glycolysis (metabolism)).

Isomerization/condensation/dismutation
   For each complete RPP cycle, one of the Gal3P molecules, with its three carbon atoms, is the net product and may be transferred out of the chloroplast or converted to starch inside the chloroplast (dashed lines, Figure 2—>). For the cycle to regenerate, the other five Gal3P molecules (with a total of 15 carbon atoms) must be converted back to three molecules of five-carbon RuBP. The conversion of Gal3P to RuBP begins with a complex series of enzymatically regulated reactions (not shown in Figure 2—>) that lead to the synthesis of the five-carbon compound ribulose-5-phosphate (Ru5P). The occurrence of this complex series, which entails isomerization, condensation, and dismutation reactions, is indicated by the broken arrow in Figure 2—>.

      The three molecules of Ru5P are converted to the carboxylation substrate, RuBP, by the enzyme phosphoribulokinase, using ATP. This reaction, shown below, completes the cycle.

Regulation of the cycle
      Photosynthesis cannot occur at night, but the respiratory process of glycolysis—which uses some of the same reactions as the RPP cycle, except in the reverse—does take place. Thus, some steps in the RPP cycle would be wasteful if allowed to occur in the dark because they would counteract the reactions of glycolysis. For this reason, some enzymes of the RPP cycle are “turned off ” (i.e., become inactive) in the dark.

      Even in the presence of light, changes in physiological conditions frequently necessitate adjustments in the relative rates of reactions of the RPP cycle, so that enzymes for some reactions change in their catalytic activity. These alterations in enzyme activity typically are brought about by changes in levels of such chloroplast components as reduced ferredoxin, acids, and soluble components (e.g., Pi and magnesium ions).

Products of carbon reduction
      The most important use of Gal3P is its export from the chloroplasts to the cytoplasm of green cells, where it is used for biosynthesis (anabolism) of products needed by the plant. In land plants, a principal product is sucrose (sugar), which is translocated from the green cells of the leaves to other parts of the plant. Other key products include the carbon skeletons of certain primary amino acids, such as alanine, glutamate, and aspartate. To complete the synthesis of these compounds, amino groups are added to the appropriate carbon skeletons made from Gal3P. Sulfur amino acids such as cysteine are formed by adding sulfhydryl groups and amino groups. Other biosynthesis pathways lead from Gal3P to lipids, pigments, and most of the constituents of green cells.

      Starch synthesis (starch) and accumulation in the chloroplasts occurs particularly when photosynthetic carbon fixation exceeds the needs of the plant. Under such circumstances, sugar phosphates accumulate in the cytoplasm, binding cytoplasmic Pi. The export of Gal3P from the chloroplasts is tied to a one-for-one exchange of Pi for Gal3P, so less cytoplasmic Pi results in decreased export of Gal3P and decreased Pi in the chloroplast. These changes trigger alterations in the activities of regulated enzymes, leading in turn to increased starch synthesis. This starch can be broken down at night and used as a source of reduced carbon and energy for the physiological needs of the plant. Too much starch in the chloroplasts leads to diminished rates of photosynthesis, however. Thus, under what would seem to be the ideal photosynthetic conditions of a bright, warm day, many plants in fact have slower rates of photosynthesis in the afternoon.

Photorespiration
      Under conditions of high light intensity, hot weather, and water limitation, the productivity of the RPP cycle is limited in many plants by the occurrence of photorespiration. This process converts sugar phosphates back to carbon dioxide; it is initiated by the oxygenation of RuBP (i.e., the combination of gaseous oxygen [O2] with RuBP). This oxygenation reaction yields only one molecule of PGA and one molecule of a two-carbon acid, phosphoglycollate, which is subsequently converted in part to carbon dioxide. The reaction of oxygen with RuBP is in direct competition with the carboxylation reaction (CO2 + RuBP) that initiates the RPP cycle and is, in fact, catalyzed by the same protein, ribulose 1,5-bisphosphate carboxylase. The relative concentrations of oxygen and carbon dioxide within the chloroplasts determine whether oxygenation or carboxylation is favoured. The concentration of oxygen inside the chloroplasts may be higher than atmospheric (20 percent) owing to photosynthetic oxygen evolution, whereas the internal carbon dioxide concentration may be lower than atmospheric (0.035 percent) owing to photosynthetic uptake. Any increase in the internal carbon dioxide pressure tends to help the carboxylation reaction compete more effectively with oxygenation.

Carbon fixation via C4 acids
      Certain plants—including the important crops sugarcane and corn (maize), as well as other diverse species believed to have evolved in the drier tropical areas—have developed a special mechanism of carbon fixation that largely prevents photorespiration. The leaves of these plants have special anatomy and biochemistry. In particular, photosynthetic functions are divided between mesophyll and bundle sheath leaf cells. The carbon fixation pathway begins in the mesophyll cells, where carbon dioxide is added to the three-carbon acid phosphoenolpyruvate (PEPA) by an enzyme called phosphoenolpyruvate carboxylase. The product of this reaction is the four-carbon acid oxaloacetate, which is reduced to malate, another four-carbon acid, in one form of C4 pathway. Malate then is translocated to bundle sheath cells, which are located near the vascular system of the leaf. There, malate enters the chloroplasts and is oxidized and decarboxylated (i.e., loses CO2) by malic enzyme. This yields carbon dioxide, which is fed into the RPP cycle of the bundle sheath cells, and pyruvate, a three-carbon acid that is translocated back to the mesophyll cells. In the mesophyll chloroplasts, the enzyme pyruvate orthophosphate dikinase (PPDK) uses ATP (adenosine triphosphate) and Pi to convert pyruvate back to PEPA, completing the C4 cycle. There are several variations of this pathway in different species. For example, the amino acids aspartate and alanine can substitute for malate and pyruvate in some species.

      The C4 pathway acts as a shuttle for carrying carbon dioxide into the chloroplasts of the bundle sheath cells, where it is used in carbohydrate synthesis. The resulting higher level of internal carbon dioxide in these chloroplasts serves to increase the ratio of carboxylation to oxygenation, thus minimizing photorespiration. Although the plant must expend extra energy to drive this shuttle, the energy loss is more than compensated by the near elimination of photorespiration under conditions where it would otherwise occur. Sugarcane and certain other plants that employ this pathway have the highest annual yields of biomass of all species.

The molecular biology of photosynthesis
      Oxygenic photosynthesis occurs in both prokaryotic cells (prokaryote) (cyanophytes) and eukaryotic cells (algae and higher plants). In eukaryotic cells (eukaryote), which contain chloroplasts and a nucleus, the genetic information needed for the reproduction of the photosynthetic apparatus is contained partly in the chloroplast chromosome and partly in chromosomes of the nucleus. For example, the carboxylation enzyme ribulose 1,5-bisphosphate carboxylase is a large protein molecule comprising a complex of eight large polypeptide subunits and eight small polypeptide subunits. The gene for the large subunits is located in the chloroplast chromosome, while the gene for the small subunits is in the nucleus. Transcription of the DNA of the nuclear gene yields messenger RNA (mRNA) that encodes the information for the synthesis of the small polypeptides. During this synthesis, which occurs on the cytoplasmic ribosomes, some extra amino acid residues are added to form a recognition leader on the end of the polypeptide chain. This leader is recognized by special receptor sites on the outer chloroplast membrane; these receptor sites then allow the polypeptide to penetrate the membrane and enter the chloroplast. The leader is removed and the small subunits combine with the large subunits, which have been synthesized on chloroplast ribosomes according to mRNA transcribed from the chloroplast DNA. The expression of nuclear genes that code for proteins needed in the chloroplasts appears to be under control of events in the chloroplasts in some cases; for example, the synthesis of some nuclear-encoded chloroplast enzymes (enzyme) may occur only when light is absorbed by chloroplasts.

Additional Reading
Photosynthesis is discussed in Yash Pal Abrol, Prasanna Mohanty, and Govindjee (eds.), Photosynthesis: Photoreactions to Plant Productivity (1993); Frank B. Salisbury and Cleon W. Ross, Plant Physiology, 4th ed. (1992), chapters 10 and 11; R.P.F. Gregory, Biochemistry of Photosynthesis, 3rd ed. (1989); and D.O. Hall and K.K. Rao, Photosynthesis, 4th ed. (1987). Jerome A. Schiff (ed.), On the Origins of Chloroplasts (1982), studies the possible bacterial origins of chloroplasts; and Ralph A. Lewin (ed.), Origins of Plastids (1993), examines current research on chloroplast origins. Roderick K. Clayton, Photosynthesis: Physical Mechanisms and Chemical Patterns (1980), is a comprehensive review of research done during the 1970s, with ample discussion of background information. J. Kenneth Hoober, Chloroplasts (1984), is a concise, readable monograph on the structure and function of the photosynthetic organelle.James Alan Bassham

* * *


Universalium. 2010.

Игры ⚽ Поможем решить контрольную работу

Look at other dictionaries:

  • Photosynthesis — Pho to*syn the*sis, n. (Plant Physiol.) The process of constructive metabolism by which carbohydrates are formed from water vapor and the carbon dioxide of the air in the chlorophyll containing tissues of plants exposed to the action of light. It …   The Collaborative International Dictionary of English

  • photosynthesis — photosynthesis. См. фотосинтез. (Источник: «Англо русский толковый словарь генетических терминов». Арефьев В.А., Лисовенко Л.А., Москва: Изд во ВНИРО, 1995 г.) …   Молекулярная биология и генетика. Толковый словарь.

  • photosynthesis — 1898, loan translation of Ger. Photosynthese, from photo light + synthese synthesis. Another early word for it was photosyntax …   Etymology dictionary

  • photosynthesis —  Photosynthesis  Фотосинтез   Процесс образования органического вещества из углекислого газа и воды на свету при участии фотосинтетических пигментов (хлорофилл у растений, бактериохлорофилл и бактериородопсин у бактерий). В современной физиологии …   Толковый англо-русский словарь по нанотехнологии. - М.

  • photosynthesis — ► NOUN ▪ the process by which green plants use sunlight to synthesize nutrients from carbon dioxide and water. DERIVATIVES photosynthesize (also photosynthesise) verb photosynthetic adjective …   English terms dictionary

  • photosynthesis — [fōt΄ō sin′thə sis] n. [ModL: see PHOTO & SYNTHESIS] 1. the biological synthesis of chemical compounds in the presence of light 2. the production of organic substances, chiefly sugars, from carbon dioxide and water occurring in green plant cells… …   English World dictionary

  • Photosynthesis — Composite image showing the global distribution of photosynthesis, including both oceanic phytoplankton and vegetation …   Wikipedia

  • photosynthesis — Process by which green plants, algae, and some bacteria absorb light energy and use it to synthesize organic compounds (initially carbohydrates). In green plants, occurs in chloroplasts, that contain the photosynthetic pigments. Occurs by… …   Dictionary of molecular biology

  • photosynthesis — 1. The compounding or building up of chemical substances under the influence of light. 2. The process by which green plants, using chlorophyll and the energy of sunlight, produce carbohydrates from water and carbon dioxide, liberating mole …   Medical dictionary

  • photosynthesis — [[t]f ouθoʊsɪ̱nθəsɪs[/t]] N UNCOUNT Photosynthesis is the way that green plants make their food using sunlight. [TECHNICAL] …   English dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”