Fact-checked by Grok 2 weeks ago

Photosynthesis

Photosynthesis is the physico-chemical process by which plants, algae, and photosynthetic bacteria use light energy to drive the synthesis of organic compounds, primarily carbohydrates, from carbon dioxide and water, releasing oxygen as a byproduct. The overall chemical equation for this oxygenic process is 6CO₂ + 6H₂O + light energy → C₆H₁₂O₆ + 6O₂, where glucose (C₆H₁₂O₆) serves as the primary energy-storing molecule. This fundamental reaction occurs in specialized organelles called chloroplasts in eukaryotic organisms like plants and algae, or in dedicated membranes in prokaryotes such as cyanobacteria. The process unfolds in two main stages: the and the light-independent reactions (also known as the Calvin-Benson cycle). In the light-dependent phase, which takes place in the membranes of , is absorbed by pigments like , exciting electrons that split water molecules (H₂O) into oxygen (O₂), protons, and electrons; this generates energy carriers ATP and NADPH. The light-independent phase, occurring in the , utilizes ATP and NADPH to fix atmospheric CO₂ into three-carbon sugars, which are then assembled into glucose and other carbohydrates. Photosynthesis is indispensable to life on , converting into that sustains nearly all ecosystems and producing the oxygen that enables aerobic . Annually, it processes approximately 200 billion tonnes of CO₂ and generates 140 billion tonnes of O₂, forming the basis of global food chains and contributing to the planet's oxygen-rich atmosphere, which originated from ancient cyanobacterial activity around 2.4 billion years ago. Without photosynthesis, complex life as we know it would not exist, as it provides both the energy for primary producers and the fuels derived from ancient photosynthetic organisms that modern society.

Overview and Importance

Definition and Basic Process

Photosynthesis is a fundamental autotrophic process by which photoautotrophic organisms, such as , , and certain , convert light energy into to synthesize organic compounds from inorganic precursors, primarily (CO₂) and (H₂O). This process enables these organisms to produce their own food, distinguishing them from heterotrophs that rely on external organic sources./08%3A_Photosynthesis/8.01%3A_Photosynthesis_-_An_Overview/8.1A%3A_Overview_of_Photosynthesis) It occurs in specialized structures and involves the capture of to drive endergonic reactions that build complex molecules. The overall reaction of photosynthesis can be summarized by the simplified balanced equation: $6 \mathrm{CO_2} + 6 \mathrm{H_2O} + \text{light energy} \rightarrow \mathrm{C_6H_{12}O_6} + 6 \mathrm{O_2} This equation represents the net stoichiometry for the production of one glucose molecule (C₆H₁₂O₆) from six molecules each of CO₂ and H₂O, with oxygen (O₂) released as a byproduct, though it abstracts the multi-step nature of the actual biochemical pathway. In more generalized terms, the process can be expressed as H₂O + CO₂ + light → O₂ + CH₂O, where CH₂O symbolizes carbohydrates. The primary inputs for photosynthesis are light energy (typically from ), CO₂ absorbed from the atmosphere, and H₂O taken up from the environment, while the key outputs are carbohydrates such as glucose for and O₂ released into the atmosphere. These outputs form the foundation of global flow, as photosynthetic serve as primary producers that sustain food chains by providing and oxygen essential for nearly all aerobic on . The process consists of that capture and light-independent reactions that fix carbon, though detailed mechanisms are covered elsewhere.

Ecological and Biological Significance

Photosynthesis has profoundly shaped Earth's , most notably through the of oxygenic photosynthesis in ancient , which initiated the around 2.4 billion years ago. This event marked a pivotal shift, as photosynthetic oxygen production accumulated in the atmosphere, rising from trace levels to enable the development of aerobic respiration and complex multicellular life forms. The oxygenation fundamentally altered planetary geochemistry, paving the way for diverse ecosystems dependent on oxygen. The cumulative impact of ancient photosynthesis persists today, with the Earth's atmosphere maintaining approximately 21% oxygen, a direct legacy of cyanobacterial activity that continues to be replenished by modern photosynthetic organisms. This oxygen-rich environment supports aerobic life across the planet, underscoring photosynthesis's role in sustaining biological diversity and metabolic processes essential for higher organisms. In contemporary ecosystems, photosynthesis drives the global by fixing 100–120 billion tons of carbon annually into , primarily through net primary productivity in terrestrial and environments. This vast production forms the basis for approximately 99% of Earth's , providing the foundational energy and carbon resources that sustain heterotrophic food webs, from microbes to apex predators./02:_Unit_II-_The_Cell/2.05:_Photosynthesis/2.5.02:_Overview_of_Photosynthesis) Furthermore, photosynthetic sinks absorb about 50% of annual CO₂ emissions, buffering atmospheric CO₂ accumulation and stabilizing dynamics.

Cellular Sites and Structures

Photosynthetic Apparatus in Prokaryotes

In prokaryotes, the photosynthetic apparatus is integrated directly into cellular membranes rather than being housed in specialized organelles like chloroplasts found in eukaryotes. This setup allows for a more streamlined organization, where light-harvesting and components are embedded within bilayers of the membrane or derived internal structures. Prokaryotic phototrophs encompass both oxygenic and anoxygenic types, with such as performing oxygenic photosynthesis and conducting anoxygenic versions. In oxygenic photosynthetic prokaryotes, particularly , the apparatus is localized to membranes, which form extensive networks of flattened sacs distinct from the plasma membrane and . These thylakoids contain photosystems I and II, large protein-pigment complexes embedded in the , where molecules are organized into reaction centers that initiate upon light absorption. Unlike eukaryotic systems, cyanobacterial thylakoids lack an envelope and are synthesized within the , enabling dynamic assembly and repair of the photosynthetic machinery. In certain primitive like Gloeobacter violaceus, the apparatus is instead incorporated into the plasma membrane itself, without dedicated thylakoids, highlighting evolutionary diversity in membrane specialization. Anoxygenic photosynthetic prokaryotes, such as (e.g., Rhodobacter sphaeroides and Chromatium vinosum), house their apparatus in intracytoplasmic membranes (ICMs), which are invaginations of the plasma membrane that expand under light conditions to increase surface area for energy capture. These ICMs embed reaction centers containing a or b, along with light-harvesting complexes (LH1 and LH2) that organize pigments into ring-like structures surrounding the core reaction center for efficient energy funneling. A key distinction is the use of alternative electron donors like (H2S) instead of , allowing these bacteria to thrive in , sulfide-rich environments without producing oxygen. This -based simplicity in prokaryotes contrasts with the compartmentalized chloroplasts of eukaryotes, where thylakoids are enclosed within a double for enhanced regulation.

Chloroplasts and Membranes in Eukaryotes

In eukaryotic photosynthetic organisms, including , , and certain protists, chloroplasts serve as the primary organelles for photosynthesis, characterized by a double envelope that separates the internal contents from the . The outer is highly permeable due to abundant porin proteins, allowing free passage of ions and small molecules, while the inner is selectively permeable, featuring specific transporters for metabolites and ions. This envelope encloses an aqueous matrix known as the and a system of internal membranes called thylakoids, enabling compartmentalized reactions that enhance efficiency compared to the integrated plasma membranes of prokaryotic precursors. The thylakoids form a highly organized network of flattened, discoid vesicles that are stacked into cylindrical structures termed grana, which are interconnected by unstacked regions called stroma thylakoids or stroma lamellae. These grana stacks, typically consisting of 10–20 thylakoids per granum, create a large surface area for embedding protein complexes and are a hallmark of eukaryotic chloroplast architecture, optimizing light capture and electron transport. The thylakoid membranes house the light-dependent reactions of photosynthesis, with (PSII) predominantly located in the appressed grana regions and (PSI) enriched in the exposed stroma thylakoids, facilitating spatial separation that supports cyclic and noncyclic electron flow. The , a dense, enzyme-rich fluid surrounding the thylakoids, functions as the site for light-independent reactions, containing the soluble enzymes of the Calvin-Benson-Bassham cycle, such as ribulose-1,5-bisphosphate carboxylase/oxygenase (), which catalyzes fixation into organic compounds. Additionally, the stroma includes chloroplast ribosomes, circular DNA, and for protein synthesis and maintenance of the organelle's genetic autonomy. This compartment also stores granules and ions, contributing to osmotic balance and metabolic regulation within the . Chlorophyll pigments are distributed across the thylakoid membranes within associated with the . , the primary pigment, is bound at the reaction centers of both and PSII, while is enriched in the peripheral antenna proteins, particularly the major II (LHCII), which forms trimers and binds approximately 60% of PSII-associated chlorophylls in the grana thylakoids. These chlorophyll a/b-binding proteins, including Lhcb1–3 in LHCII, enhance light absorption across a broader and transfer excitation energy to the reaction centers, with LHCII supercomplexes stabilizing PSII in the stacked membrane regions.

Light-Dependent Reactions

Absorption of Light and Pigments

Photosynthesis begins with the absorption of light by specialized pigments embedded in the membranes of chloroplasts. These pigments capture photons from sunlight, converting their energy into excited electrons that drive the . The primary pigment, chlorophyll a, is essential for this process, as it directly participates in the photochemical reactions at the reaction centers. Accessory pigments, such as chlorophyll b and , broaden the range of wavelengths that can be utilized, enhancing overall light capture efficiency. Chlorophyll a exhibits characteristic absorption peaks in the blue-violet region at approximately 430 and in the red region at approximately 680 , corresponding to its Soret and bands, respectively. Chlorophyll b absorbs similarly but with a slight shift, peaking around 450 and 640 , allowing it to transfer to chlorophyll a. Carotenoids, including and xanthophylls, primarily absorb in the blue-green spectrum ( ), which chlorophylls absorb less efficiently, and also serve protective roles by dissipating excess to prevent damage from high . The photosynthetic , which measures the rate of photosynthesis across wavelengths, closely aligns with the combined absorption of these pigments, peaking in the blue and red regions where chlorophyll a dominates. Light energy is captured not only by isolated pigments but through organized antenna systems known as light-harvesting complexes (LHCs). These protein-pigment assemblies, such as LHCII in , contain hundreds of chlorophyll a, chlorophyll b, and molecules arranged to maximize photon interception. Excitation energy migrates rapidly among pigments within the LHC via (), a non-radiative process where energy is transferred from donor to acceptor molecules based on spectral overlap and proximity. This efficient funneling directs the energy to the core reaction centers, where it is used for charge separation. At the reaction centers of photosystems II and I, specialized chlorophyll a pairs—P680 and P700, named for their absorption maxima—perform quantum capture. When a photon is absorbed, it excites an electron in these pairs from the ground state to a higher-energy singlet state (P680* or P700*), creating a strong reducing potential. This excitation initiates the subsequent electron transport process. Carotenoids in the antenna can also contribute to excitation of these centers under certain conditions, ensuring robust energy delivery.

Electron Transport and Z-Scheme

The light-dependent reactions of photosynthesis involve a series of electron transfers driven by absorbed light energy, initiating with the excitation of pigments in the reaction centers of two photosystems. Photosystem II (PSII), characterized by its reaction center chlorophyll pair P680, absorbs light at approximately 680 nm and initiates electron transport by ejecting an electron from the oxidized P680⁺ to the primary acceptor pheophytin. This electron is then passed through plastoquinone (PQ), a lipid-soluble carrier embedded in the thylakoid membrane, which becomes reduced to plastoquinol (PQH₂) after accepting two electrons and two protons. PQH₂ diffuses to the cytochrome b₆f complex, where it undergoes oxidation, releasing electrons to plastocyanin (PC), a soluble copper-containing protein in the thylakoid lumen. The cytochrome b₆f complex, consisting of cytochromes b₆ and f along with the Rieske iron-sulfur protein, facilitates this transfer while contributing to proton translocation across the membrane. Plastocyanin shuttles the electron to photosystem I (PSI), whose reaction center chlorophyll pair P700 absorbs light at 700 nm, boosting the electron to a higher state and reducing the final acceptor (Fd), an iron-sulfur protein. Reduced then donates the electron to NADP⁺ via ferredoxin-NADP⁺ reductase, forming NADPH. This linear, non-cyclic electron flow from PSII to PSI generates both ATP (through proton gradient-driven ) and NADPH, essential for the subsequent carbon fixation reactions. The Z-scheme provides a graphical representation of this electron transport pathway, plotting the redox potentials of carriers against their sequential order to illustrate the energy profile. Electrons originate at the high redox potential of +0.82 V (for the H₂O/O₂ couple) and descend through PSII (P680 at +1.1 V when oxidized), plastoquinone (0 V), cytochrome b₆f (+0.3 V), plastocyanin (+0.37 V), and PSI (P700 at ~+0.43 V when oxidized), before reaching the low potential of -0.32 V (at pH 7) at NADP⁺/NADPH. The "Z" shape arises from two light-induced upward jumps in potential at the photosystems, compensating for the overall energy drop required to drive the endergonic reduction of NADP⁺. This scheme, first proposed to link the two photosystems via cytochrome components, highlights how serial excitations maintain favorable thermodynamics. In addition to non-cyclic flow, a cyclic electron transport pathway operates around , where electrons from reduced return to the b₆f complex via the b₆f pathway, bypassing NADP⁺ and enhancing ATP without net NADPH formation. This cyclic mode, involving and , helps balance the ATP/NADPH ratio needed for carbon under varying light conditions.

Water Splitting and

In oxygenic photosynthesis, , also known as photolysis, occurs at the lumenal side of (PSII) and serves as the initial electron source for the . This process oxidizes two molecules to produce one dioxygen molecule, four protons, and four electrons, requiring the absorption of four photons per O₂ released:
$2 \mathrm{H_2O} \rightarrow \mathrm{O_2} + 4 \mathrm{H^+} + 4 e^-
The protons contribute to the transmembrane proton gradient essential for ATP synthesis, while the electrons are transferred via Z to the oxidized reaction chlorophyll ⁺, integrating into the broader .
The (OEC), embedded within PSII, catalyzes this challenging four-electron oxidation of under mild conditions. The OEC consists of a Mn₄CaO₅ cubane-like , where four (Mn1–Mn4) are bridged by five oxygen atoms, with a central calcium ion stabilizing the and facilitating substrate binding. This is ligated by residues from the and CP43 proteins of PSII, positioning it near the to release O₂ and protons directly into the space. The Mn₄CaO₅ configuration enables sequential accumulation of oxidizing equivalents, avoiding high-energy intermediates that could damage the protein matrix. The catalytic mechanism follows the , a four-step sequence of metastable S-states (S₀ to S₄) proposed by Bessel Kok and colleagues, where each state represents a progressive oxidation of the cluster by light-induced charge separation in PSII. Starting from the dark-stable S₁ state, absorption of a advances the : S₁ → S₂ → S₃ → (S₄ → S₀) + O₂, with S₄ being a transient peroxide-like intermediate that spontaneously decays to release O₂ and return to S₀, accompanied by proton release at specific transitions. Spectroscopic studies confirm distinct electronic configurations for each S-state, with oxidation states evolving from mixed valences, such as (III, , , ) in S₁, to higher ones, e.g., Mn()₄ in S₃, culminating in O–O bond formation likely between a Mn-bound and a or oxyl . This stepwise progression ensures efficient water oxidation with minimal .

Light-Independent Reactions

Carbon Fixation in the Calvin Cycle

The Calvin-Benson-Bassham (CBB) cycle, also known as the , is the primary pathway for carbon fixation in photosynthetic organisms, occurring in the stroma of eukaryotes and the of prokaryotes. This light-independent process assimilates atmospheric CO₂ into organic compounds, utilizing ATP and NADPH generated by the to drive the synthesis of carbohydrates. The cycle enables the conversion of inorganic carbon into biomass, supporting global . The CBB cycle operates through three sequential phases: , reduction, and regeneration of the CO₂ acceptor molecule. In the phase, CO₂ is fixed onto ribulose-1,5-bisphosphate (RuBP), a five-carbon sugar phosphate, in a catalyzed by the ribulose-1,5-bisphosphate carboxylase/oxygenase (). This forms an unstable six-carbon intermediate that rapidly hydrolyzes into two molecules of 3-phosphoglycerate (3-PGA), a three-carbon compound. , the most abundant protein on , constitutes up to 50% of soluble leaf protein in higher plants and is estimated to account for approximately 0.1% of global , underscoring its central role in carbon assimilation. The exhibits a specificity factor (τ) that favors CO₂ over O₂, typically ranging from 80 to 100 in C₃ plants, which determines the efficiency of relative to competing oxygenation reactions. During the reduction phase, the 3-PGA molecules are first phosphorylated by ATP to form 1,3-bisphosphoglycerate, with the group from ATP transferred to the carboxyl group of 3-PGA. This intermediate is then reduced by NADPH, transferring a to produce glyceraldehyde-3-phosphate (G3P), a triose . For every three CO₂ molecules fixed, six 3-PGA are produced, requiring six ATP and six NADPH to yield six G3P molecules. One of these G3P molecules is released as the net product of the cycle, while the remaining five are used in the subsequent phase. The regeneration phase recycles the five G3P molecules through a series of enzymatic reactions involving aldolase, , and other enzymes to reform three RuBP molecules, consuming an additional three ATP. This phase ensures the cyclic nature of the pathway, allowing continuous CO₂ fixation. The overall of the CBB cycle, representing three turns to produce one net G3P, is given by: $3 \, \mathrm{CO_2} + 9 \, \mathrm{ATP} + 6 \, \mathrm{NADPH} \rightarrow \mathrm{G3P} + 9 \, \mathrm{ADP} + 8 \, \mathrm{P_i} + 6 \, \mathrm{NADP^+} This balanced equation highlights the energy investment required for carbon assimilation. The net G3P output serves as a metabolic precursor for the of glucose and other carbohydrates, such as , through subsequent gluconeogenic pathways in the or . Two G3P molecules can condense to form one glucose , linking carbon fixation directly to and growth in photosynthetic organisms.

Alternative Carbon Pathways (C3, C4, CAM)

In most plants, the pathway serves as the default mechanism for carbon fixation, where directly incorporates CO₂ into ribulose-1,5-bisphosphate in the chloroplasts, but this exposes the to oxygen, leading to that reduces efficiency, particularly under hot and dry conditions with low CO₂ and high O₂ levels. can consume up to 25-30% of fixed carbon in plants in such environments, limiting productivity in tropical or arid regions. The pathway evolved as an adaptation to minimize through spatial separation of initial CO₂ fixation and the , featuring Kranz anatomy with distinct mesophyll and bundle sheath cells. In mesophyll cells, phosphoenolpyruvate (PEP) carboxylase—a high-affinity for CO₂ (as )—fixes CO₂ into the four-carbon compound oxaloacetate, which is then reduced to malate or transaminated to aspartate. These C4 acids diffuse to bundle sheath cells, where they are decarboxylated to release CO₂ for RuBisCO, creating a localized high-CO₂ that suppresses oxygenation. This mechanism enhances by 50-100% in C4 plants compared to under optimal conditions, with examples like (Zea mays) demonstrating improved water and nitrogen use. Although C4 plants represent only about 3% of species, they account for roughly 23% of global terrestrial primary productivity, dominating in warm, high-light grasslands and savannas. Crassulacean acid metabolism (CAM) provides a temporal separation strategy for carbon fixation, allowing plants to conserve water in extremely arid environments by opening stomata at night. During the night, CO₂ enters through open stomata and is fixed by PEP carboxylase in mesophyll cells into oxaloacetate, which is converted to malate and stored in vacuoles, causing a characteristic acidification. By day, with stomata closed to minimize , malate is decarboxylated, releasing CO₂ for the while maintaining high internal CO₂ levels to reduce . This pathway enhances water-use efficiency by up to 10-fold compared to plants, as seen in succulents like cacti (e.g., species) and ( comosus), though it limits maximum growth rates due to slower overall carbon assimilation. Both and pathways have arisen through evolutionary convergence, independently evolving multiple times in response to similar selective pressures for CO₂ concentration. photosynthesis has originated over 60 times across at least 19 angiosperm families, often involving gene recruitment and anatomical modifications from C3 ancestors. Similarly, has evolved independently more than 30 times in diverse lineages, including both and monocots, sharing genetic underpinnings like adaptations with . These parallel innovations highlight the repeatability of photosynthetic evolution under environmental stress.

Regulation and Environmental Factors

Kinetics and Reaction Order

The kinetics of photosynthesis follow a strict temporal sequence, with the occurring on ultrafast timescales of picoseconds to milliseconds, preceding the slower that span seconds to minutes. Initial excitation energy transfer in light-harvesting complexes and charge separation in proceed in picoseconds to nanoseconds, while subsequent transport steps, including reduction and oxidation, extend into microseconds to milliseconds. This rapid phase generates ATP and NADPH, which are then utilized in the , where enzymatic turnovers, particularly by , limit the overall rate to seconds per catalytic cycle, with complete regeneration of ribulose-1,5-bisphosphate requiring multiple iterations over minutes under steady-state conditions. Rate-limiting factors in photosynthesis primarily involve the capacity of electron transport chains and the catalytic efficiency of key enzymes in carbon fixation. The electron transport capacity, governed by the and photosystem turnover rates, can constrain ATP and NADPH production at high light intensities, though it typically operates near saturation under ambient conditions. In the light-independent phase, RuBisCO's low turnover rate of approximately 3 s⁻¹ serves as a major bottleneck, dictating the pace of CO₂ fixation and subsequent reductions in the cycle. Feedback regulation ensures coordination between light and dark phases through pH gradients and metabolite signaling. The proton gradient (ΔpH) across the thylakoid membrane, established during electron transport, regulates and b6f activity to prevent over-reduction, maintaining balance on timescales. Metabolite levels, such as reduced generated via , activate enzymes like fructose-1,6-bisphosphatase through thiol-disulfide exchanges, enhancing carbon fixation rates in response to light availability. RuBisCO adhere to Michaelis-Menten parameters, with an apparent Kₘ for CO₂ of approximately 9–15 μM in chloroplasts, reflecting its affinity under stromal conditions and influencing fixation efficiency at ambient CO₂ levels. The enzyme's underscore its role as a regulatory node, where substrate availability modulates the overall photosynthetic flux.

Light, Temperature, and CO2 Influences

Photosynthetic rates are highly sensitive to , which drives the by providing energy for electron excitation in . At low intensities, the rate of photosynthesis increases linearly with due to limited absorption, but it reaches a point where additional does not proportionally enhance carbon fixation because downstream processes, such as the , become limiting. For most terrestrial , this typically occurs around 500–1000 μmol m⁻² s⁻¹, well below full sunlight intensities of approximately 2000 μmol m⁻² s⁻¹. Beyond , excessive can induce , where damage , reducing and overall . The effectiveness of light also depends on wavelength, as chlorophyll pigments absorb specific spectra within the (PAR) range of 400–700 nm. Red light (around 660 nm) and far-red light (700–750 nm) are most efficient for driving photosynthesis, exhibiting the highest quantum yields due to strong absorption by and balanced excitation of I and II. (400–500 nm), while less efficient per for carbon fixation, plays a critical role in stomatal opening, particularly in , thereby regulating CO₂ influx and water loss to optimize overall photosynthetic performance. Temperature influences photosynthesis by affecting , , and of gases like CO₂ and O₂. For C₃ , which dominate temperate ecosystems, the optimal temperature range for maximum photosynthetic rates is typically 20–30°C, where activity and electron transport are balanced without significant thermal stress. In contrast, C₄ , adapted to warmer environments, exhibit higher optima around 30–40°C, benefiting from enhanced activity and reduced at elevated temperatures. Extremes beyond these ranges limit rates; low temperatures slow enzymatic reactions and increase CO₂ but risk chilling injury, while high temperatures above 40°C can denature proteins like , leading to irreversible declines in activity. CO₂ concentration governs the step in the via Michaelis-Menten kinetics, where photosynthetic rate rises hyperbolically with CO₂ until saturating at levels far above ambient. The half-saturation constant (Kₘ) for Rubisco's affinity for CO₂ is approximately 250–900 μmol mol⁻¹ () depending on species and , meaning current atmospheric levels of about 426 (as of ) support substantial but unsaturated fixation. Rising CO₂, as observed in recent decades, enhances net photosynthesis by 10–20% through increased efficiency and partial suppression of competing oxygenation reactions, promoting greater accumulation in many ecosystems. However, this fertilization effect can be constrained by limitations, particularly , as accelerated growth dilutes tissue concentrations and demands higher uptake without proportional availability.

Photorespiration and Efficiency Losses

Photorespiration arises from the oxygenase activity of the enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (), which competes with its carboxylase function in the presence of atmospheric oxygen. In this reaction, catalyzes the addition of O₂ to ribulose-1,5-bisphosphate (RuBP), yielding one molecule of 3-phosphoglycerate (3-PGA) and one molecule of 2-phosphoglycolate (2-PG), the latter being a toxic byproduct that must be metabolized. The oxygenase activity is favored under conditions of high O₂/CO₂ ratios, such as those prevalent in current atmospheric conditions (approximately 21% O₂ and 0.04% CO₂), which exacerbate the inefficiency relative to the carboxylase reaction. The 2-PG is dephosphorylated to glycolate in the , initiating the photorespiratory salvage pathway that spans multiple organelles to recover usable carbon. Glycolate is transported to the , where it is oxidized to glyoxylate and then aminated to ; is subsequently shuttled to the for conversion to serine, releasing CO₂ and in the process. This cycle, known as the photorespiratory pathway, recycles about 75% of the carbon from 2-PG back into 3-PGA for the , but incurs significant energetic costs, including the consumption of ATP and NADPH without net carbon gain. A key inefficiency is the net release of fixed carbon as CO₂ during the glycine decarboxylase reaction in the . The simplified for the salvage of two glycolate molecules is: $2 \text{ glycolate} + \text{O}_2 \rightarrow \text{serine} + \text{CO}_2 + \text{NH}_3 This results in a loss of 0.5 CO₂ per RuBisCO oxygenation event, as serine (C₃) is produced from two glycolate (2×C₂) units minus the released CO₂. In C₃ plants, leads to the release of 25–30% of recently fixed carbon as CO₂ under ambient conditions, substantially reducing and contributing to yield limitations in crops like and . This loss is particularly pronounced at higher temperatures, which increase the specificity of for O₂ over CO₂, as noted in studies of environmental influences on . Mitigation of photorespiration is achieved in C₄ and CAM plants through CO₂-concentrating mechanisms that elevate local CO₂ levels around RuBisCO, suppressing the oxygenase activity and reducing carbon losses to near negligible levels. Efforts to engineer similar improvements in C₃ plants include synthetic photorespiratory bypasses that redirect glycolate metabolism to avoid CO₂ release, such as pathways converting glycolate to Calvin cycle intermediates without decarboxylation, offering potential targets for enhancing photosynthetic efficiency. Recent advances as of 2025 include successful field trials of bypasses in rice, improving productivity by up to 33% and nitrogen uptake, as well as system-level analyses identifying carbon-fixing alternative pathways as promising for major crops.

Variations and Adaptations

Carbon Concentrating Mechanisms

Carbon concentrating mechanisms (CCMs) are biochemical and biophysical strategies evolved by photosynthetic organisms to increase the concentration of CO₂ at the of the , thereby enhancing the efficiency of carbon fixation in environments where ambient CO₂ levels are limiting. These mechanisms counteract the low solubility and diffusion rate of CO₂ in water and air, particularly under conditions of high where RuBisCO's oxygenase activity competes with . By elevating the CO₂/O₂ ratio around RuBisCO, CCMs minimize wasteful , which can otherwise reduce photosynthetic productivity by up to 30% in C₃ under current atmospheric conditions. In , CCMs primarily involve the formation of , proteinaceous microcompartments that encapsulate and within a selectively permeable shell. within the converts actively transported (HCO₃⁻) ions into CO₂, creating a localized high-CO₂ microenvironment that can exceed ambient levels by 10- to 100-fold. employ multiple HCO₃⁻ transporters, such as those encoded by the cmp, sbt, and bicA gene families, to actively uptake inorganic carbon from the extracellular medium, fueling this process even at low external CO₂ concentrations. This system not only boosts carboxylation rates but also suppresses by maintaining an optimal CO₂/O₂ balance inside the . In eukaryotic algae, particularly green algae like , CCMs center on , dense, phase-separated organelles where is sequestered and surrounded by a matrix that facilitates CO₂ diffusion while limiting O₂ entry. These function as CO₂ pumps through the action of plasma membrane and envelope transporters, including HCO₃⁻ influx systems mediated by CCM genes such as those in the LCIA and CCP1 families, which actively accumulate inorganic carbon intracellularly. Carbonic anhydrases then dehydrate HCO₃⁻ to CO₂ near the , achieving CO₂ concentrations up to 40 times higher than in the and enabling efficient fixation under low-CO₂ aquatic conditions. This pyrenoid-based CCM is inducible and dynamically regulated, enhancing photosynthetic rates by reducing photorespiratory losses. Among terrestrial plants, C₄ species employ a spatial CCM that minimizes CO₂ leakiness in bundle sheath cells, where RuBisCO is localized. In these plants, initial CO₂ fixation in mesophyll cells produces C₄ acids that are transported to the bundle sheath, where they are decarboxylated to release CO₂ in a compartment with reduced permeability to gases, often reinforced by suberized cell walls. This anatomical arrangement, exemplified in crops like and , concentrates CO₂ around by 10- to 60-fold, drastically lowering the oxygenase reaction and . In some aquatic plants, an analogous strategy involves the localization of glycine decarboxylase to specific cell layers, such as bundle sheath-like compartments, to recycle photorespiratory CO₂ and concentrate it for refixation, thereby adapting to submerged environments with limited CO₂ . Overall, CCMs across these organisms increase local CO₂ availability by 10- to 100-fold, which can reduce by 80-90% compared to non-concentrating systems, significantly improving net carbon assimilation and in CO₂-limited habitats.

Photosynthetic Efficiency Across Organisms

Photosynthetic efficiency refers to the fraction of incident converted into stored in through photosynthesis. In C3 plants, the theoretical maximum under ideal conditions (30°C and 380 ppm CO₂) is 4.6%, limited by factors such as the energy required for carbon fixation and losses from photorespiration. However, in field conditions, the average realized efficiency drops to approximately 1-2% due to environmental stresses, suboptimal light distribution within canopies, and maintenance costs. A key metric of efficiency at the photochemical level is the , which measures the number of evolved per absorbed. In oxygenic photosynthesis, the quantum yield is approximately 0.1, corresponding to 8-10 required per O₂ produced, as four are minimally needed for each (PSII and ) to drive and NADP⁺ reduction. This yield is reduced from its theoretical maximum by losses including (re-emission of energy as light), (heat dissipation), and antenna spillover (inefficient energy transfer between ). These mechanisms protect the photosynthetic apparatus from excess light but limit overall energy capture. Comparisons across organisms highlight variations in efficiency tied to biochemical pathways. in , which uses only one and alternative electron donors like , achieves lower overall efficiencies of about 1-2%, as it generates less reducing power per photon without . In contrast, C4 plants enhance efficiency by concentrating CO₂ at , minimizing ; their theoretical maximum is 6%, and laboratory measurements have reached 6-7% under controlled conditions. Recent advances in have pushed efficiencies beyond natural limits in . Engineered strains, modified to optimize light-harvesting antennas and carbon concentrating mechanisms, have achieved photosynthetic efficiencies exceeding 10% in settings, representing a significant improvement over wild-type (typically 8-10%) and enabling higher productivity for biofuels. These developments, building on post-2020 genetic tools, focus on reducing energy losses in the light reactions to enhance outputs like ATP and NADPH.

Evolutionary History

Origins in Ancient Prokaryotes

Photosynthesis first emerged in ancient prokaryotes as anoxygenic forms, utilizing electron donors other than water, such as hydrogen sulfide or iron, to fix carbon in the absence of oxygen. Geological evidence from carbon isotopic ratios in Archean rocks, showing depletions in ¹³C consistent with biological carbon fixation, suggests that anoxygenic photosynthesis arose around 3.5 billion years ago. These isotopic signatures, preserved in metasedimentary rocks from sites like the Pilbara Craton in Australia, indicate the presence of microbial communities capable of light-driven metabolism during the Paleoarchean era. The core machinery of these early photosynthetic systems involved reaction centers classified as Type I and Type II, which evolved through gene duplication events in bacterial lineages. Type I reaction centers, found in green sulfur bacteria and heliobacteria, reduce ferredoxin using low-potential electron acceptors, while Type II centers, present in purple bacteria, drive cyclic electron flow to generate ATP via quinone reduction. Phylogenetic analyses reveal that the heterodimeric structure of Type II centers (subunits L and M) resulted from independent gene duplications of ancestral proteins, with similar duplications giving rise to the core polypeptides in Type I systems. These innovations allowed prokaryotes to harness light energy in anaerobic environments, marking a pivotal step in microbial evolution. The transition to oxygenic photosynthesis, which uses as an and produces oxygen as a , occurred between 3.4 and 2.4 billion years ago in , involving the coupling of two distinct (PSI and PSII) derived from ancestral Type I and Type II centers. This evolutionary innovation likely arose through genomic rearrangements and the integration of oxygen-evolving complexes, enabling water oxidation and boosting energy yields. Fossil evidence from in the approximately 2.55 billion-year-old Nauga Formation in , featuring laminated microbial mats with isotopic and mineralogical signatures of oxygen production, supports the early emergence of this process. Additionally, banded iron formations (BIFs) from the , such as those in the Hamersley Basin dated to around 2.5 billion years ago, record the precipitation of iron oxides linked to rising oxygen levels from cyanobacterial activity, preceding the .

Endosymbiosis and Eukaryotic Development

The serial endosymbiosis theory posits that chloroplasts originated from a primary endosymbiotic event in which a heterotrophic eukaryotic engulfed a photosynthetic cyanobacterium approximately 1.5 billion years ago, leading to the establishment of the lineage, which includes glaucophytes, , , and land plants. Over time, the engulfed cyanobacterium lost autonomy, with its genes progressively transferred to the host , transforming it into a semi-autonomous dependent on nuclear-encoded proteins for function. This event marked a pivotal step in eukaryotic development, enabling oxygenic photosynthesis in diverse multicellular forms and reshaping global biogeochemical cycles. Compelling evidence for this cyanobacterial origin includes the striking similarity between chloroplast genomes and those of free-living , such as shared gene content, circular DNA structure, and 70S ribosomes, as revealed by phylogenetic analyses of conserved genes like rbcL and psaB. Further support comes from the presence of cyanobacterial-derived proteins in eukaryotic genomes, confirming the endosymbiotic ancestry. The spread of photosynthesis to other eukaryotic lineages occurred through secondary and tertiary endosymbioses, where eukaryotic were engulfed by non-photosynthetic hosts; for instance, a single secondary endosymbiosis involving a red alga gave rise to complex plastids in chromalveolates (including stramenopiles, , and haptophytes), characterized by additional surrounding membranes and nucleomorph remnants in some groups like cryptophytes. Tertiary events, such as in certain dinoflagellates, involved further engulfment of secondary , leading to diverse plastid morphologies. A hallmark of endosymbiotic integration was massive gene transfer from the cyanobacterial to the host , with modern chloroplast genomes retaining only about 5-10% of the original ~3,000 , while approximately 90% now reside in the and are targeted back to the via transit peptides. This process, known as endosymbiotic gene transfer (EGT), occurred over time and continues at low rates, as evidenced by nuclear plastid DNA (NUPT) insertions detected in genomes. Recent genomic studies from 2023-2025 have illuminated the complexity of plastid acquisitions in protists, confirming multiple independent secondary endosymbioses—such as distinct events in cryptophytes and ochrophytes (stramenopiles)—and repeated tertiary acquisitions in lineages like Kareniaceae, challenging earlier single-origin models for chromalveolate plastids. These findings, derived from comparative phylogenomics and metagenomic surveys, underscore the dynamic, reticulate of photosynthetic organelles across eukaryotic diversity.

Diversification in Modern Lineages

Photosynthesis has diversified across modern prokaryotic and eukaryotic lineages, reflecting adaptations to varied environments and ecological niches. In prokaryotes, represent the primary group performing oxygenic photosynthesis, utilizing water as an and producing oxygen as a byproduct through two (PSI and PSII). These organisms, including genera like Synechococcus and Prochlorococcus, are ubiquitous in aquatic and terrestrial habitats and form symbiotic associations, such as in lichens and reefs. In contrast, predominates in other prokaryotes, such as green sulfur bacteria (Chlorobiaceae) and (Chromatiaceae), which use alternative s like or organic compounds and lack the oxygen-evolving capability, restricting them largely to anaerobic environments like stratified lakes and sediments. Among eukaryotes, photosynthesis arose via endosymbiosis and has radiated into several major lineages. The lineage, encompassing chlorophytes () and embryophytes (land plants), relies on chlorophyll a and b in chloroplasts derived from a primary algal , enabling adaptations from freshwater plankton to terrestrial vascular systems with specialized structures like leaves and stomata. The lineage includes rhodophytes () with phycobiliproteins for light harvesting in deeper waters, and secondary acquisitions in chromalveolates such as diatoms and dinoflagellates, which feature silica frustules or unique pigment compositions for marine productivity. Euglenids, within the excavate supergroup, acquired algal-derived plastids secondarily and exhibit mixotrophic lifestyles, combining photosynthesis with in freshwater and habitats. Photosynthetic capability has been lost in numerous eukaryotic lineages, particularly among parasitic and protists, where plastids persist as non-photosynthetic organelles retaining genes for essential functions like . For instance, holoparasitic such as (dodder) and exhibit massive plastome reductions, eliminating most photosynthesis-related genes while conserving a core set for roles, reflecting evolutionary trade-offs for host dependency. This gene retention underscores the plastid's indispensable role beyond energy production. Overall, photosynthetic eukaryotes comprise only about 10% of known eukaryotic diversity, dominated by algae and plants, yet they drive much of global primary production. Cyanobacteria, in particular, account for up to 25% of oceanic net primary productivity, underpinning marine food webs and carbon cycling through abundant picocyanobacterial forms like Prochlorococcus. This diversification highlights photosynthesis's pivotal role in sustaining biodiversity and biogeochemical processes across modern ecosystems.

Discovery and Research History

Initial Observations and Experiments

In 1771, English chemist conducted pioneering experiments demonstrating that plants could restore air rendered unfit for or . He enclosed a burning in a glass vessel until the flame extinguished, indicating the air was "vitiated," and then introduced a sprig of mint. After several days of exposure to , the air regained its purity, as evidenced by the rekindling of a or the survival of a placed inside. Priestley interpreted this as plants absorbing impurities from the air, though he initially attributed the effect to nocturnal processes rather than light. Building on Priestley's findings, physician in 1779 established the essential role of light and the specificity of plant parts in this air-purifying process. Through over 500 experiments, Ingenhousz showed that only the green portions of plants—such as leaves—released the restorative gas (later identified as oxygen) when exposed to , while non-green parts like roots or flowers had no such effect, and the process ceased in darkness or shade. He concluded that drove the purification, distinguishing it from mere plant , which occurred at night and actually impaired air quality. Swiss pastor and botanist Jean Senebier advanced these observations in 1782 by elucidating the involvement of , then termed "fixed air." Senebier demonstrated that plants absorb fixed air during the day and release purified air in proportions inversely related to the fixed air consumed, with the rate of restoration increasing with . His experiments confirmed Ingenhousz's light dependency while quantifying how greater illumination enhanced the process, laying groundwork for understanding in photosynthesis. These early discoveries were shaped by the prevailing phlogiston theory, which posited that combustion released a substance called phlogiston, leaving air "phlogisticated" and impure. Priestley and contemporaries viewed fixed air as phlogisticated common air, with plants purportedly purifying it by absorbing excess phlogiston, a misconception that persisted until the oxygen theory supplanted phlogiston in the late 18th century.

Key Conceptual Advances

In the 1860s, Julius von Sachs advanced the understanding of photosynthesis by developing the starch test, which demonstrated that is the first visible product of the process in green leaves. Using iodine staining, Sachs showed that leaves exposed to light in the presence of accumulate granules in chloroplasts, confirming carbohydrate synthesis as a direct outcome of photosynthetic activity. He further established the separation of light-dependent and light-independent phases through experiments where plants deprived of light consumed stored , while illuminated leaves produced it only in green tissues containing . These findings, detailed in Sachs' works from 1862 and 1864, provided that photosynthesis requires light and is localized to chloroplasts. Theodor Wilhelm Engelmann's 1882 experiment introduced the concept of the action spectrum, revealing how specific light wavelengths drive photosynthesis. By projecting a light spectrum onto a filamentous alga (Cladophora) and observing the accumulation of oxygen-seeking aerobic , Engelmann demonstrated that bacteria clustered primarily in the red and blue regions, corresponding to chlorophyll absorption maxima. This visual method quantified across wavelengths, showing that oxygen evolution peaks where light is most effectively utilized by pigments, thus linking spectral absorption to biological . Published in Botanische Zeitung, the study marked a pivotal shift toward spectroscopic analysis in . Frederick Frost Blackman's 1905 formulation of the law of limiting factors clarified how environmental variables constrain photosynthetic rates. Observing that increasing initially boosted carbon assimilation in but plateaued due to temperature or CO₂ shortages, Blackman proposed that the overall rate is governed by the factor closest to its minimum threshold. This principle highlighted the multi-step nature of photosynthesis, distinguishing rapid light reactions from slower enzymatic carbon fixation, and emphasized that no single factor acts in isolation. His quantitative approach, based on rate measurements in leaves, influenced subsequent models of resource optimization in . Otto 's measurements in the early 1920s provided the first estimates of in terms of light energy utilization. Using suspensions and manometric techniques to track per absorbed , Warburg and Negelein initially reported a quantum requirement of approximately 4–5 per O₂ evolved, suggesting high under optimal conditions. However, these findings sparked debate, as subsequent refinements and independent studies, including those by , established a more consistent initial estimate of around 12 per O₂, reflecting real-world losses in the process. Warburg's work, published in Biochemische Zeitschrift in , pioneered precise quantification and underscored the photochemical basis of photosynthesis.

Recent Developments and Refinements

In the mid-20th century, significant strides in understanding photosynthesis were made through isotopic tracing techniques. and his team at the , utilized radioactive (¹⁴C) to elucidate the Calvin-Benson cycle, identifying the biochemical pathway by which CO₂ is fixed into organic compounds in plants. This work, spanning the 1940s and 1950s, pinpointed ribulose-1,5-bisphosphate (RuBP) as the key acceptor molecule and mapped the cycle's enzymatic steps, earning the 1961 for his contributions to the chemical understanding of biological processes. Parallel advancements focused on the light-dependent reactions. British biochemist Robin Hill demonstrated in the 1930s and 1940s that isolated chloroplasts could produce oxygen and reduce an like in the presence of light, establishing the "Hill reaction" as evidence for photosynthetic independent of carbon fixation. By the 1950s, this laid the groundwork for the Z-scheme model, proposed by Robert Hill and Fay Bendall in 1960, which describes the sequential energy transfers between I and II, integrating potentials to explain and NADP⁺ reduction. The 1960s brought discoveries of photosynthetic variations that enhanced efficiency in certain environments. Marshall Davidson Hatch and Charles Roger Slack identified the C4 pathway in tropical grasses like , where CO₂ is initially fixed into four-carbon compounds in mesophyll cells before being concentrated around in bundle sheath cells, minimizing and boosting carbon fixation rates under high temperatures and low CO₂. This mechanism, detailed in their 1966-1967 publications, explained higher productivity in plants and inspired agricultural applications for crops like . Since the 2010s, efforts to enhance have included under projects like RIPE (Realizing Increased Photosynthetic Efficiency) at the University of . For instance, in 2022, researchers accelerated recovery from photoprotection in soybeans using genetic modifications, improving photosynthesis and by up to 30% in field trials. Quantum biology studies, including those by Graham Fleming's group at UC from 2007 onward, have confirmed long-lived quantum coherence in light-harvesting complexes, enabling near-100% efficiency in energy transfer and informing bio-inspired solar technologies. As of 2025, these advances continue to integrate into climate models, with IPCC assessments projecting CO₂ fertilization could increase global gross primary productivity by 10-30% by 2100 under moderate emissions scenarios, balanced against limitations like water stress.

References

  1. [1]
    The Photosynthetic Process - Life Sciences
    Photosynthesis is the physico-chemical process by which plants, algae and photosynthetic bacteria use light energy to drive the synthesis of organic compounds.
  2. [2]
    Photosynthesis - PMC - PubMed Central - NIH
    Oct 26, 2016 · By facilitating conversion of solar energy into chemical energy, photosynthesis acts as the primary energy input into the global food chain.
  3. [3]
    What is Photosynthesis | Smithsonian Science Education Center
    Apr 12, 2017 · The whole process of photosynthesis is a transfer of energy from the Sun to a plant. In each sugar molecule created, there is a little bit ...
  4. [4]
    Intro to photosynthesis (article) - Khan Academy
    Jul 17, 2016 · Photosynthesis is the process in which light energy is converted to chemical energy in the form of sugars. In a process driven by light energy, ...Missing: authoritative | Show results with:authoritative
  5. [5]
    Photosynthesis - National Geographic Education
    Feb 26, 2025 · Photosynthesis is the process by which plants use sunlight, water, and carbon dioxide to create oxygen and energy in the form of sugar.
  6. [6]
    The origin of atmospheric oxygen on Earth: The innovation ... - PNAS
    The evolution of O2-producing cyanobacteria that use water as terminal reductant transformed Earth's atmosphere to one suitable for the evolution of aerobic ...
  7. [7]
    The Great Oxidation Event: How Cyanobacteria Changed Life
    Feb 18, 2022 · The great oxidation event, which released oxygen into Earth's atmosphere, was catalyzed by cyanobacteria and ultimately led to the evolution ...
  8. [8]
    The strange history of atmospheric oxygen - PMC - NIH
    Mar 28, 2022 · The first atmospheric oxygen came from the metabolism of microorganisms, the cyanobacteria, that used photosynthesis, but for which oxygen was an unwanted by‐ ...
  9. [9]
    Global primary productivity - [carbon] - Biosphere - BNID 100241
    About 1.0×10^14kg (100 billion tons) of carbon are fixed annually into organic compounds by photosynthetic organisms (often called the net primary productivity) ...
  10. [10]
    [PDF] Carbon budgets - Intergovernmental Panel on Climate Change
    No, natural carbon sinks have taken up a near constant fraction of our carbon dioxide (CO2) emissions over the last six decades. However, this fraction is ...
  11. [11]
    Membrane Dynamics in Phototrophic Bacteria - Annual Reviews
    Sep 8, 2020 · Photosynthetic membranes are typically densely packed with proteins, and this is crucial for their function in efficient trapping of light ...
  12. [12]
    The initial steps of biogenesis of cyanobacterial photosystems occur ...
    It is well established that the site for photosynthetic electron transport in cyanobacteria is the thylakoid membrane. Hence, both PSI and PSII exhibit their ...
  13. [13]
    Structure and evolution of photosystem I in the early-branching ...
    The available membrane space for the photosynthetic apparatus in the thylakoid-free Gloeobacteriales is restricted to the plasma membrane (16, 26, 40). To ...<|separator|>
  14. [14]
    Architecture and mechanism of the light-harvesting apparatus of ...
    May 26, 1998 · In most purple bacteria, the photosynthetic membranes contain two types of light-harvesting complexes, light-harvesting complex I (LH-I) and ...
  15. [15]
    Anoxygenic photosynthesis and the delayed oxygenation of Earth's ...
    Jul 9, 2019 · Anoxygenic photosynthesizers (photoferrotrophs) translates into diminished global photosynthetic O 2 release when the ocean interior is Fe(II)-rich.
  16. [16]
    Chloroplasts and Photosynthesis - Molecular Biology of the Cell
    Chloroplasts and photosynthetic bacteria obtain high-energy electrons by means of photosystems that capture the electrons that are excited when sunlight is ...
  17. [17]
    Fine structure of granal thylakoid membrane organization using cryo ...
    The space between the thylakoid membrane and the chloroplast envelope is called the stroma. The details of the thylakoid membrane folding and the ...Missing: double | Show results with:double
  18. [18]
    The multiple roles of light-harvesting chlorophyll a/b-protein ...
    Higher plant chloroplasts possess highly conserved pigment–protein complexes for the outer light-harvesting antenna of PSI (LHCI) and PSII (LHCII). In PSII, ...Missing: review | Show results with:review
  19. [19]
    (PDF) Chlorophylls and carotenoids: Pigments of photosynthetic ...
    Aug 7, 2025 · This chapter presents detailed information on chlorophylls and carotenoids to give practical directions toward their quantitative isolation and determination.<|separator|>
  20. [20]
    8.3.2: Absorption of Light - Biology LibreTexts
    Dec 16, 2021 · Chlorophylls and carotenoids are the two major classes of photosynthetic pigments found in plants and algae; each class has multiple types ...Missing: primary | Show results with:primary
  21. [21]
    [PDF] Chapter 9 The Photosynthetic Pigments
    These can be divided into three major classes (see Tables. 9.1-9.3): (1) chlorophylls and (2) carotenoids--both water insoluble, and. (3) phycobilins-water- ...Missing: sources | Show results with:sources
  22. [22]
    Hyperspectral and Chlorophyll Fluorescence Analyses of ...
    The chlorophyll curve shows peaks in the blue (approximately ≈ 430–450 nm) and red (approximately ≈ 660–680 nm) regions, which are typical absorption ...
  23. [23]
    Spectrum of Light as a Determinant of Plant Functioning - NIH
    Mar 17, 2020 · ... photosynthetic action spectrum coincides with the absorption spectrum of chlorophyll. This finding implied the involvement of chlorophyll in ...
  24. [24]
    Light Absorption and Energy Transfer in the Antenna Complexes of ...
    In this review, we describe the underlying photophysical principles by which this energy is absorbed, as well as the mechanisms of electronic excitation energy ...
  25. [25]
    Photosynthetic Light-Harvesting (Antenna) Complexes—Structures ...
    Jun 3, 2021 · LHCII is the main antenna complex comprising about half of the thylakoid protein and binding roughly 50% of the Chls. LHCII forms homo- and ...
  26. [26]
    O2 and Other High-Energy Molecules in Photosynthesis: Why Plants ...
    Photon energy absorbed by P700 (a chlorophyll dimer with an absorption maximum near 700 nm) loosens an electron, forming excited P700*, which is easily ionized ...
  27. [27]
    Short-Range Effects in the Special Pair of Photosystem II Reaction ...
    At the heart of photosynthetic reaction centers (RCs) are pairs of chlorophyll a (Chla), P700 in photosystem I (PSI) and P680 in photosystem II (PSII) of ...<|control11|><|separator|>
  28. [28]
    Function of the Two Cytochrome Components in Chloroplasts - Nature
    HILL, R., BENDALL, F. Function of the Two Cytochrome Components in Chloroplasts: A Working Hypothesis. Nature 186, 136–137 (1960). https://doi.org/10.1038 ...
  29. [29]
    A New Leaf Copper Protein 'Plastocyanin', a Natural Hill Oxidant
    The discovery and function of plastocyanin: A personal account. Sakae Katoh. Photosynthesis Research (1995) ; Zur Rolle von Plastocyanin und Cytochrom f im ...<|control11|><|separator|>
  30. [30]
    Ferredoxin and Photosynthetic Phosphorylation - Nature
    It has now been shown that ferredoxin can catalyse, by two distinct photochemical reactions, the production of ATP in cell-free photosynthetic systems.Missing: discovery | Show results with:discovery
  31. [31]
    Structural Changes of the Oxygen-evolving Complex in Photosystem ...
    Kok cycle. The classical Kok cycle with the intermediate S states in water oxidation is shown. Proposed oxidation states for the manganese atoms are indicated.
  32. [32]
    Closing Kok's cycle of nature's water oxidation catalysis
    Jul 16, 2024 · The Mn4CaO5(6) cluster in photosystem II catalyzes water splitting through the Si state cycle (i = 0–4). Molecular O2 is formed and the ...
  33. [33]
    On the origin of oxygenic photosynthesis and Cyanobacteria
    Oct 9, 2019 · Oxygenic photosynthesis is one of the most important metabolisms to have evolved on Earth as it has enabled complex life forms to emerge.
  34. [34]
    Engineering the Calvin–Benson–Bassham cycle and hydrogen ...
    Dec 11, 2020 · In the CBB cycle, RuBisCO catalyzes the carboxylation reaction converting ribulose-1,5-diphosphate and CO2 to generate 2 molecules of 3- ...
  35. [35]
    Review The most abundant protein in the world - ScienceDirect.com
    The most abundant protein in nature is probably the chloroplast enzyme ribulose bisphosphate carboxylase/oxygenase (Fraction I protein).
  36. [36]
    Modeling the Calvin-Benson cycle - PMC - PubMed Central
    Nov 3, 2011 · The Calvin-Benson cycle is a central part of the carbon metabolism in oxygenic photosynthesis, involving 11 different enzymes that catalyze 13 ...
  37. [37]
    The Path from C3 to C4 Photosynthesis - PMC - NIH
    C4 PHOTOSYNTHESIS. The C4 photosynthetic carbon cycle is an elaborated addition to the C3 photosynthetic pathway. It evolved as an adaptation to high light ...Missing: paper | Show results with:paper
  38. [38]
    Nature's green revolution: the remarkable evolutionary rise of C4 ...
    Plants with the C 4 photosynthetic pathway dominate today's tropical savannahs and grasslands, and account for some 30% of global terrestrial carbon fixation.
  39. [39]
    The evolution of C4 photosynthesis - Sage - 2004 - New Phytologist
    Dec 23, 2003 · PEP carboxylase is localized in cytoplasmic pockets at the periphery of each cell, while Rubisco and the decarboxylating enzymes are located in ...
  40. [40]
    C4 Cycles: Past, Present, and Future Research on C4 Photosynthesis
    The cycle is restarted when PEP is transported from the M chloroplast to the M cell cytoplasm to combine once again with CO2. The key features of the NADP-ME ...
  41. [41]
    New research changes understanding of C4 plant evolution
    Nov 15, 2010 · C4 plants compose only 3 percent of flowering plant species, yet account for about 25 percent global terrestrial productivity. About 60% of C4 ...<|separator|>
  42. [42]
    On the Evolutionary Origin of CAM Photosynthesis - PubMed Central
    CAM plants fix CO2 via phosphoenolpyruvate carboxylase (PEPC) during the night when it is cooler and less water is lost. The resulting organic acids, ...
  43. [43]
    CAM photosynthesis: the acid test - New Phytologist Foundation
    Oct 12, 2021 · The overwhelming body of evidence is consistent with net dark CO2 fixation in CAM plants being mediated via the canonical pathway of malic acid ...
  44. [44]
    Crassulacean Acid Metabolism - an overview | ScienceDirect Topics
    Crassulacean acid metabolism (CAM) is defined as a photosynthetic adaptation that allows plants to maximize CO2 uptake and recapture respiratory CO2 under ...Physiology And Metabolism · 2.6 Crassulacean Acid... · 2.6. 2 Regulation Of Cam<|separator|>
  45. [45]
    Evolutionary trajectories, accessibility and other metaphors: the case ...
    Apr 16, 2019 · This review highlights what we have learned about the recurring assembly of C 4 and CAM, focusing on the increasingly predictable stepwise evolutionary ...III. Background on CAM and... · IV. The C4 evolutionary... · The CAM evolutionary...
  46. [46]
    Shared origins of a key enzyme during the evolution of C4 and CAM ...
    Mar 17, 2014 · We show that independent origins of both CAM and C 4 photosynthesis in Caryophyllales co-opted the same genes for PEPC through similar adaptive changes.
  47. [47]
    Molecular dynamics simulations in photosynthesis - PMC - NIH
    In this membrane, the initial steps of photosynthesis, collectively known as the light reactions, take place. ... nanoseconds up to the ~ microseconds timescale.
  48. [48]
    Turnover time of metabolites of the Calvin–Benson cycle
    Turnover time of metabolites of the Calvin–Benson cycle ; 0.1 to 1.0 Sec · Unspecified · Stitt M, Lunn J, Usadel B. Arabidopsis and primary photosynthetic ...Missing: minutes | Show results with:minutes
  49. [49]
    Rubisco deactivation and chloroplast electron transport rates co-limit ...
    May 17, 2023 · Rubisco deactivation and chloroplast electron transport rates co-limit photosynthesis above optimal leaf temperature in terrestrial plants.
  50. [50]
    Photosynthesis: The Calvin Cycle - Heineke - Wiley Online Library
    Mar 15, 2009 · The Calvin cycle is a reductive process in the stroma of chloroplasts responsible for the synthesis of carbohydrates from carbon dioxide.Missing: stoichiometry CO2
  51. [51]
    Relationship between photosynthetic electron transport and pH ...
    We calculated the light-dependent pH gradient (Delta pH) across the thylakoid membrane in intact leaves. These Delta pH values were correlated with the ...
  52. [52]
    Chloroplast thioredoxin systems dynamically regulate ... - NIH
    Apr 15, 2019 · Here, we review the current knowledge of regulation of photosynthesis by chloroplast TRXs and assess the prospect of improving plant photosynthetic efficiency.Missing: feedback | Show results with:feedback
  53. [53]
    Bisphosphate Carboxylase/Oxygenase (Rubisco) Activase Using
    To correct for the influence of CO2 on Rubisco activity, the turnover number was adjusted to reflect the lower CO2 using a Km(CO2) of 9 um. This allowed ...
  54. [54]
    [PDF] Photosynthesis: Physiological and Ecological Considerations
    Light compensation points of sun plants range from 10 to. 20 µmol m–2 s–1, whereas corresponding values for shade plants are 1 to 5 µmol m–2 s–1. Why are light ...
  55. [55]
    PHOTOINHIBITION OF PHOTOSYNTHESIS IN NATURE!
    Although photoinhibition decreases capacity for light-limited pho tosynthesis at the leaf level, significant decreases in canopy photosynthetic. C-uptake on ...Missing: threshold | Show results with:threshold
  56. [56]
    Photosynthetic Physiology of Blue, Green, and Red Light - Frontiers
    Mar 4, 2021 · At low light, red light has higher quantum yield. At high light, green light can be more efficient due to uniform absorption, while at low ...Introduction · Materials and Methods · Results · Discussion
  57. [57]
    ISWS - Alternative Crop Suitability Maps - Climate
    Mar 15, 2021 · The C3 plants generally originate from temperate regions and photosynthesis is optimized at temperatures from 59-86°F (15-30°C). Certain grasses ...
  58. [58]
    The temperature response of C3 and C4 photosynthesis - SAGE
    Jun 15, 2007 · In general, photosynthesis can function without harm between 0 and 30 °C in cold-adapted plants that are active in winter and early spring, or ...MECHANISMS... · ACCLIMATION OF C3... · TEMPERATURE RESPONSE...
  59. [59]
    [PDF] A guide to photosynthetic gas exchange measurements - RIPE project
    Dec 31, 2023 · These include the incident light intensity (Iinc; often also referred to as Q) at which A becomes light‐ saturated (Isat), the light ...
  60. [60]
    Mauna Loa carbon dioxide forecast for 2025 - Met Office
    Jan 17, 2025 · As a result, we forecast the 2025 annual average CO2 concentration at Mauna Loa to be 426.6 ± 0.6 ppm (Figure 1). This will continue the ongoing ...
  61. [61]
    A constraint on historic growth in global photosynthesis due to rising ...
    Nov 27, 2023 · Our analysis suggests that CO 2 fertilization increased global annual terrestrial photosynthesis by 13.5 ± 3.5% or 15.9 ± 2.9 PgC (mean ± sd) between 1981 and ...
  62. [62]
    Atmospheric CO2 Concentration and Other Limiting Factors in ... - NIH
    By increasing the level of CO2, photosynthesis increases, but only if there is no other limiting factor. For example, nitrogen is often in short supply to ...
  63. [63]
    Photorespiration Revisited - PMC - PubMed Central - NIH
    Photorespiration results in the loss of up to 25% of the carbon that is fixed during photosynthetic carbon assimilation (Ludwig and Canvin, 1971).
  64. [64]
    Design and in vitro realization of carbon-conserving photorespiration
    Nov 20, 2018 · Photorespiration recycles ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) oxygenation product, 2-phosphoglycolate, back into the ...
  65. [65]
    Photorespiration connects C3 and C4 photosynthesis
    Feb 22, 2016 · It consumes ATP and NADPH and leads to a net loss of CO2 for the plant. This reduces the efficiency of carbon fixation in plants by up to 30 ...Abstract · Introduction · Summary
  66. [66]
  67. [67]
    concentrating Mechanisms in Plants to Improve Yields
    Pyrenoid-based CCMs actively elevate CO2 concentration, potentially improving CO2 assimilation and crop yields in plants.
  68. [68]
    Functions, Compositions, and Evolution of the Two Types of ...
    Carboxysomes are specialized protein microcompartments composed of a polyhedral protein shell within which cyanobacteria concentrate CO2 around their primary ...
  69. [69]
    concentrating-mechanism (CCM): functional components, Ci ...
    In cyanobacteria, Rubisco is encapsulated in unique micro-compartments known as carboxysomes. Cyanobacteria can possess up to five distinct transport systems ...
  70. [70]
    [PDF] Cyanobacterial CO2-concentrating mechanism components
    To overcome RuBisCO's limitations, a number of terrestrial plant species have evolved CO2- concentrating mechanisms (CCMs), including C4 photo- synthesis ...
  71. [71]
    Review From algae to plants: understanding pyrenoid-based CO 2
    Pyrenoid-based CO2-concentrating mechanisms (pCCMs) enhance CO2 fixation by concentrating CO2 near Rubisco, improving photosynthetic CO2 uptake.
  72. [72]
    Modelling the pyrenoid-based CO2-concentrating mechanism ...
    May 19, 2022 · In algae, such a CO2-concentrating mechanism occurs within a phase-separated organelle called the pyrenoid. Pyrenoid-based CO2-concentrating ...
  73. [73]
    Bundle sheath suberisation is required for C4 photosynthesis in a ...
    Feb 26, 2021 · The pathway is characterised by a biochemical CO2 concentrating mechanism that operates across mesophyll and bundle sheath (BS) cells and relies ...
  74. [74]
    C4 Photosynthesis (The CO2-Concentrating Mechanism and ... - NIH
    A high [CO2] is maintained in the bundle sheath compartment in maize until Ci decreases below approximately 100 [mu]bar. The results from these gas exchange ...
  75. [75]
    Systems analysis of the CO2 concentrating mechanism in ... - eLife
    Apr 29, 2014 · Mangan and Brenner found that containing the enzymes within a carboxysome increased the concentration of carbon dioxide inside the cell by an ...
  76. [76]
    What is the maximum efficiency with which photosynthesis ... - PubMed
    The maximum conversion efficiency of solar energy to biomass is 4.6% for C3 photosynthesis at 30 degrees C and today's 380 ppm atmospheric [CO2], but 6% for C4 ...
  77. [77]
    Photosynthetic efficiency - Wikipedia
    ... theoretical maximum efficiency of solar energy conversion is approximately 11%. In actuality, however, plants do not absorb all incoming sunlight (due to ...Typical efficiencies · C3 vs. C4 and CAM plants · Research
  78. [78]
    C3 and C4 Photosynthesis – Implications for Crop Production
    Feb 28, 2024 · Although they comprise less than 3% of known plant species, their greater efficiency allows them to carry out 20-30% of global terrestrial ...
  79. [79]
    THE QUANTUM YIELD OF PHOTOSYNTHESIS1 - Annual Reviews
    The quantum yield of photosynthesis is derived from measurements of light intensity and rate of photosynthesis. Appropriate measurements of light intensity ...Missing: O2 | Show results with:O2
  80. [80]
    On the evolution of the concept of two light reactions and two ... - NIH
    A personal commentary on the 1960 paper of Robert (Robin) Hill and Fay Bendall. It was on 9 April 1960 that the Z-scheme of Hill and Bendall (1960) was ...
  81. [81]
    [PDF] The controversy over the minimum quantum requirement for oxygen ...
    O2 molecule at the lowest light intensity used; a linear extrapolation to zero light intensity would give a value of. 8–10 quanta per O2 molecule! Although ...
  82. [82]
    An overview of anoxygenic phototrophic bacteria and their ...
    Anoxygenic phototrophic bacteria (APB) are a phylogenetically diverse group of organisms that can harness solar energy for their growth and metabolism.
  83. [83]
    Systematic Comparison of C3 and C4 Plants Based on Metabolic ...
    We demonstrated that in contrast to C3, C4 plants have less dense topology, higher robustness, better modularity, and higher CO 2 and radiation use efficiency.
  84. [84]
    Emerging technologies for advancing microalgal photosynthesis ...
    Microalgae can convert a higher percentage of the sun's energy to biomass energy, estimated to be around 8 ~ 10%, and theoretically produce up to 77 g/m2/d of ...
  85. [85]
    Advances in Genetic Engineering in Improving Photosynthesis and ...
    Jan 18, 2023 · This review describes the molecular events that happen during photosynthesis and microalgal productivity through genetic engineering and discusses future ...
  86. [86]
    Synthetic biology and metabolic engineering paving the way for ...
    Aug 20, 2025 · The genetic modification of GM algae leads to improved photosynthetic efficiency and enhanced lipid accumulation and cell rupture through ...
  87. [87]
    Early Evolution of Photosynthesis - PMC - NIH
    Oct 6, 2010 · There is suggestive evidence that photosynthetic organisms were present approximately 3.2 to 3.5 billion years ago, in the form of ...
  88. [88]
    The paleobiological record of photosynthesis
    Jul 7, 2010 · Taken as a whole, the evidence available indicates that O2-producing photosynthetic microorganisms originated earlier than 2,450 Ma ago; that ...The Great Oxidation Event · Cellular Fossils · The Carbon Isotopic...
  89. [89]
    Gene duplication and the evolution of photosynthetic reaction center ...
    We investigate the evolutionary relationships between photosynthetic reaction center proteins (D1, D2, L and M) and demonstrate that the pattern of ...Missing: type | Show results with:type
  90. [90]
    A fresh look at the evolution and diversification of photochemical ...
    In other words, the heterodimeric character of all known Type II reaction centers evolved twice after two separate gene duplication events: one that produced ...
  91. [91]
    Evolutionary relationships between “Q-type” photosynthetic reaction ...
    Aug 23, 1990 · Strong evidence is found for independent gene duplications having produced the L and M subunits of the photosynthetic purple bacterial reaction ...
  92. [92]
    [PDF] Evolution of Oxygenic Photosynthesis - CalTech GPS
    May 11, 2016 · Abstract. The origin of oxygenic photosynthesis was the most important metabolic innovation in Earth history. It allowed life to generate ...<|control11|><|separator|>
  93. [93]
    When did oxygenic photosynthesis evolve? - PMC - NIH
    U–Th–Pb isotopes in metasomatized and metamorphosed turbidites perhaps represent the earliest evidence of oxygenic photosynthesis. ... 3.5-billion-year-old ...
  94. [94]
    Evidence for benthic oxygen production in Neoarchean lacustrine ...
    May 9, 2022 · Geochemical, petrographic, and sedimentary data from 2.7 Ga Hartbeesfontein stromatolites support the evolution of oxygenic photosynthesis ...
  95. [95]
    Genomic Insights into Plastid Evolution - Oxford Academic
    May 13, 2020 · In this review, we provide an overview of recent advances in our understanding of the origin and spread of plastids from the perspective of comparative ...
  96. [96]
    The endosymbiotic origin, diversification and fate of plastids - Journals
    Mar 12, 2010 · Red algae appear to have been taken up only once, giving rise to a diverse group called chromalveolates. Additional layers of complexity come ...
  97. [97]
    The plastid ancestor originated among one of the major ... - Nature
    Sep 15, 2014 · A plastid origin occurring during the divergence of one of the major cyanobacterial lineages that include N 2 -fixing filamentous cyanobacteria.
  98. [98]
    Evolutionary analysis of Arabidopsis, cyanobacterial, and ... - PNAS
    Chloroplasts were once free-living cyanobacteria that became endosymbionts, but the genomes of contemporary plastids encode only ≈5–10% as many genes as ...
  99. [99]
    Genomics and chloroplast evolution: what did cyanobacteria do for ...
    Plant chloroplast originated, through endosymbiosis, from a cyanobacterium, but the genomic legacy of cyanobacterial ancestry extends far beyond the chloroplast ...
  100. [100]
    New Model and Dating for the Evolution of Complex Plastids of Red ...
    The Chromalveolata hypothesis proposes a single red alga endosymbiosis that involved the ancestor of all the Chromalveolata lineages: cryptophytes, haptophytes, ...
  101. [101]
    Suppression of plastid-to-nucleus gene transfer by DNA double ...
    May 16, 2025 · Plant nuclear genomes contain thousands of genes of mitochondrial and plastid origin as the result of endosymbiotic gene transfer (EGT).
  102. [102]
    repeated endosymbiotic acquisitions in kareniacean dinoflagellates
    Mar 18, 2024 · Here, we explore kareniacean plastid evolution as a model for the endosymbiotic establishment of plastids and the spread of photosynthesis ...
  103. [103]
    Anoxygenic photosynthesis with emphasis on green sulfur bacteria ...
    Compared to previous studies with tubes, desulfurization efficiency has decreased, but energy efficiency has increased (Hurse and Keller, 2004; Cui et al., 2021) ...
  104. [104]
    Review The Puzzle of Plastid Evolution - ScienceDirect.com
    Jan 27, 2009 · The possibility of a truly ancient red algal secondary endosymbiosis in a putative 'chromalveolate'–rhizarian ancestor has been discussed 88 ...
  105. [105]
    A molecular timescale for eukaryote evolution with implications for ...
    Mar 25, 2021 · There are three main lineages with primary plastids: red algae, green algae (including land plants) and glaucophytes—altogether forming a ...The Phylogeny Of Eukaryotes · The Eukaryote Tree Of Life... · Phylogenomic Dataset...
  106. [106]
    Red macroalgae in the genomic era - Borg - 2023 - New Phytologist
    Aug 30, 2023 · The resulting primary plastids are evident in three main eukaryotic lineages – the red algae, the green algae (and their land plant ...Iv. Red Algal Cell Biology · Vi. Reproductive Biology · Viii. Red Algal Genomics
  107. [107]
    Phylogenetic Relationships and Morphological Character Evolution ...
    Nov 7, 2014 · Photosynthetic euglenids acquired chloroplasts by secondary endosymbiosis, which resulted in changes to their mode of nutrition and affected the evolution of ...
  108. [108]
    Extensive plastome reduction and loss of photosynthesis genes in ...
    Oct 2, 2019 · The reduction in plastid genome size and gene content in parasitic plants predominantly results from loss of photosynthetic genes.
  109. [109]
    The loss of photosynthesis pathway and genomic locations of the ...
    May 8, 2020 · Our results suggest coordinated loss of photosynthesis related functions in the plastid and nuclear genomes of a holoparasitic plant.
  110. [110]
    Genomic reconfiguration in parasitic plants involves considerable ...
    Losing photosynthesis and relying entirely on heterotrophic nutrient supply from another plant is a derived stage of an obligate parasitic lifestyle. More ...
  111. [111]
    Evolution of Photosynthetic Eukaryotes; Current Opinion, Perplexity ...
    There are about ten lineages of photosynthetic eukaryotes, including Chloroplastida, Rhodophyta, and Cryptophyta. Mechanistically, eukaryotic photosynthesis ...
  112. [112]
    Present and future global distributions of the marine Cyanobacteria ...
    The Cyanobacteria Prochlorococcus and Synechococcus account for a substantial fraction of marine primary production. Here, we present quantitative niche ...
  113. [113]
    Fraction of global primary biomass production - BioNumbers
    Some have estimated the cyanobacterial contribution as 10% of oceanic NPP (ref 68) or 25% (ref 67). As carbon fixation is estimated to be ~45% of total NPP, ...
  114. [114]
    Highlights in photosynthesis research - Nobel Prize
    Jan Ingenhousz, The Netherlands, demonstrates that the plant in Priestley's experiment is dependent on light and its green parts.
  115. [115]
    a commentary on Priestley (1772) 'Observations on different kinds of ...
    Priestley's account of fixed air in 'Observations' [1, pp. 148–170] expanded ... But Priestley soon loosened phlogiston from its tradition ...
  116. [116]
    [PDF] Dr Jan IngenHousz, or why don't we know who discovered ...
    Nov 17, 2007 · August 1771 Joseph Priestley, chemist from Birmingham, England, performed his famous experiment with the mouse and the mint plant. This ...Missing: primary | Show results with:primary
  117. [117]
    Photosynthesis: basics, history and modelling - PMC - NIH
    In this review, we emphasize that mathematical modelling is a highly valuable tool in understanding and making predictions regarding photosynthesis.<|control11|><|separator|>
  118. [118]
    [PDF] Photosynthesis: The Power Plant and the Chemical Factory of Life
    Senebier's "fixed air" became carbon dioxide, CO₂, Priestley's "vital air" became oxygen, O2, and "plant nourriture" became organic matter, that is, chemical ...
  119. [119]
    Julius Sachs (1832–1897) and the experimental physiology of plants
    Jul 6, 2015 · The German biologist Julius Sachs was the first to introduce controlled, accurate, quantitative experimentation into the botanical sciences.
  120. [120]
    [PDF] Photosynthesis: basics, history and modelling - Life Sciences
    The role of CO2 in photosynthesis was shown by Jean Senebier (1782), whereas the synthesis of starch was shown by Julius von Sachs (1862,. 1864). However ...Missing: test | Show results with:test
  121. [121]
    Highlighting Theodor W. Engelmann's “Farbe und Assimilation ...
    Oct 22, 2021 · In 1883, Theodor Wilhelm Engelmann, a German scientist, wrote his essay “color and assimilation” (Ger.: “Farbe und Assimilation”) describing ...
  122. [122]
    Optima and Limiting Factors - Semantic Scholar
    Optima and Limiting Factors · F. F. Blackman · Published 1 April 1905 · Biology · Annals of Botany.Missing: original | Show results with:original
  123. [123]
    The maximum quantum yield controversy. Otto Warburg and ... - NIH
    This is standard textbook information, with QY taken as about 0·1 O2 per photon or, put the other way (the quantum requirement = 1/QY), about 10 photons ...Missing: 1910s | Show results with:1910s