Fact-checked by Grok 2 weeks ago

Retinal

Retinal, also known as retinaldehyde, is the form of () and serves as the essential for vertebrate vision. With the molecular formula C₂₀H₂₈O, it features a β-ionone ring connected to a polyene chain ending in an aldehyde group, enabling light absorption in the . In photoreceptor cells of the , the 11-cis- of retinal covalently binds to proteins via a protonated linkage at residue 296, forming visual pigments such as in rods and iodopsins in cones. Upon absorption, 11-cis-retinal undergoes rapid to all-trans-retinal, triggering a conformational change in the opsin that activates the G-protein and initiates the phototransduction signaling cascade, ultimately leading to hyperpolarization of the photoreceptor and neural transmission of visual information. To sustain continuous vision, retinal participates in the , a series of enzymatic reactions primarily occurring in the (RPE) and photoreceptors. All-trans-retinal released from activated visual pigments is reduced to all-trans-, transported to the RPE, and esterified by lecithin:retinol acyltransferase (LRAT) before isomerohydrolase converts it to 11-cis-retinol, which is then oxidized by short-chain /reductase enzymes like retinol 5 (RDH5) to regenerate 11-cis-retinal. This cycle ensures a steady supply of the , with binding proteins such as cellular retinaldehyde-binding protein (CRALBP) and interphotoreceptor retinoid-binding protein (IRBP) facilitating transport and preventing toxicity from free retinal, which can accumulate and cause retinal degeneration if unregulated. Disruptions in retinal metabolism, such as mutations in RPE65 or LRAT, lead to inherited retinal dystrophies like , underscoring its critical role in ocular health. Beyond vision, retinal serves as a precursor to , which influences via retinoid X receptors (RXRs) and retinoic acid receptors (RARs), contributing to embryonic development and epithelial maintenance, though its primary biological significance lies in phototransduction.

Chemical Structure and Properties

Molecular Composition and Isomers

Retinal, also known as retinaldehyde, has the molecular formula C<sub>20</sub>H<sub>28</sub>O and is a of characterized by a monocyclic diterpenoid structure. It features a β-ionone —a with a conjugated , a dimethyl substitution at position 1, and a at position 5—connected via a to a linear polyene chain. This chain consists of four conjugated carbon-carbon and terminates in an (-CHO) at carbon 15, enabling its role as a due to the extended conjugation. The carbon skeleton of retinal spans 20 atoms, with the β-ionone ring encompassing carbons 1–6, followed by the polyene side chain (carbons 7–15) where the double bonds occur at positions 7–8, 9–10, 11–12, and 13–14. The aldehyde group at C15 imparts polarity and reactivity, while the polyene chain's conjugation delocalizes electrons, facilitating light absorption and . In text representation, the structure can be described as a β-ionone ring linked to -CH=CH-C(CH<sub>3</sub>)=CH-CH=CH-C(CH<sub>3</sub>)=CH-CHO, with specific at the double bonds determining the . Retinal primarily exists as geometric isomers differing in the configuration of the polyene chain's double bonds, with all-trans-retinal and 11-cis-retinal being the most biologically relevant. In all-trans-retinal, all four double bonds exhibit E (trans) stereochemistry, resulting in a fully extended linear conformation. In contrast, 11-cis-retinal has Z (cis) configuration at the 11–12 double bond, introducing a bend in the chain, while the other double bonds remain E. This cis configuration strains the molecule, enabling rapid photoisomerization upon light absorption, where rotation occurs specifically around the 11–12 bond to yield the all-trans form./Photoreceptors/Chemistry_of_Vision/Cis-Trans_Isomerization_of_Retinal) The all-trans isomer is thermodynamically more stable than the 11-cis isomer by approximately 4 kcal/, owing to reduced steric hindrance in its extended form. Consequently, 11-cis-retinal displays higher reactivity, being more prone to both and photochemical conversion to the all-trans configuration compared to other mono-cis isomers. This instability contributes to its selective accumulation in biological systems through enzymatic control.

Physical and Spectroscopic Properties

Retinal is a lipophilic molecule characterized by poor solubility in water, approximately 0.1 μM at room temperature and pH 7.3, reflecting its nonpolar polyene structure that favors partitioning into lipid environments over aqueous media. In contrast, it exhibits good solubility in organic solvents such as ethanol (>10 mg/mL), chloroform, and diethyl ether, facilitating its extraction and handling in laboratory settings. All-trans-retinal typically presents as yellow to orange crystals or a crystalline powder, with a melting point of 62–65 °C; its boiling point is estimated at 367 °C under standard pressure, though it often decomposes before reaching this temperature due to thermal instability. The UV-Vis absorption spectrum of retinal features a prominent λ_max at approximately 380 nm for the all-trans isomer in solvents like , arising from π–π* transitions in its extended of four double bonds and the carbonyl. This absorption shifts modestly with , such as to around 370 nm for the 11-cis form. reveals weak emission from retinal, with a broad spectrum peaking near 500 nm upon excitation at 380 nm, attributed to its low (<0.01) but useful for detecting trace amounts and probing excited-state dynamics in non-aqueous environments. yields distinctive vibrational signatures for structural elucidation, including a strong all-trans indicator band at 960 cm⁻¹ from C–H out-of-plane wagging and C=C stretching modes around 1550–1580 cm⁻¹, enabling differentiation of configurational isomers without . The functionality at C-15 imparts high reactivity to retinal, rendering it susceptible to aerial oxidation, which converts it to via enzymatic or non-enzymatic pathways, and to forming protonated Schiff bases with primary amines such as ε-amino groups of residues. These reactions underpin retinal's role in processes but also contribute to its instability, necessitating storage under inert atmospheres to prevent degradation.

Biosynthesis and Metabolism

Dietary Sources and Conversion

Retinal, a key form of , is primarily obtained through dietary sources that provide either preformed or its provitamin precursors. Preformed , including retinal and , is found in animal-derived foods such as liver, fish, eggs, dairy products, and fish oils, where concentrations are highest in organ meats like beef liver. In contrast, provitamin A , particularly , serve as the main plant-based sources and are abundant in and green vegetables like carrots, , sweet potatoes, and leafy greens, as well as in fruits such as apricots and . These must undergo enzymatic conversion in the body to yield retinal, making plant sources an indirect but significant contributor to status, especially in vegetarian diets. The conversion of provitamin A carotenoids to retinal occurs mainly in the through the action of the β-carotene 15,15'-monooxygenase 1 (BCMO1), which catalyzes the oxidative central cleavage of at its 15,15' . This symmetric cleavage reaction stoichiometrically produces two molecules of all-trans-retinal from one molecule of , with the 's activity dependent on molecular oxygen and iron as cofactors. However, the efficiency of this bioconversion varies widely among individuals due to genetic, , and health factors, with a typical weight-based ratio of approximately 12:1 (μg to 1 μg equivalents), though it can range from 3.6:1 to 28:1 in humans. Retinal produced via this pathway is rapidly reduced to by retinal reductases to facilitate further transport. Following cleavage, retinal-derived retinol is esterified in enterocytes and incorporated into chylomicrons for lymphatic and delivery to the liver, the primary storage site for . From the liver, is mobilized into the bloodstream bound to (RBP), often complexed with for stability, enabling distribution to peripheral tissues. This process supports an , where a portion of retinyl esters is recycled through to enhance overall and maintain steady-state levels.

Interconversions with Vitamin A Derivatives

Retinal, the aldehyde form of vitamin A, undergoes reversible oxidation from retinol, the primary circulating form, through the action of retinol dehydrogenases (RDHs), which are predominantly members of the short-chain dehydrogenase/reductase (SDR) superfamily. This bidirectional reaction is catalyzed by enzymes such as RDH10, which exhibits high catalytic efficiency with a Km value of approximately 0.035 μM for all-trans-retinol, making it a key player in maintaining retinal levels. The reverse reaction, reducing retinal back to retinol, is facilitated by retinal reductases like RDH11 and RDH12, which prefer NADP(H) as a cofactor and show Km values around 0.12 μM for retinaldehyde. These enzymes are widely distributed, with RDH10 being ubiquitously expressed, particularly during embryogenesis, while RDH11 displays broad tissue presence including liver and lung. The irreversible oxidation of to , the transcriptionally active form of , is mediated by retinal dehydrogenases (RALDHs), also known as ALDH1A enzymes. Key isoforms include RALDH1 (ALDH1A1), primarily expressed in adult liver; RALDH2 (ALDH1A2), crucial for embryonic development with expression in mesodermal tissues; and RALDH3 (ALDH1A3), found in adult tissues like skin and lung. Kinetic parameters vary, with RALDH2 showing a of 0.66 μM for retinaldehyde and high efficiency (Vmax around 200 nmol/min/mg), while RALDH3 has a of 3.9 μM but superior overall activity (Vmax 306 nmol/min/mg). This step commits retinal to signaling pathways, including brief roles in gene regulation via nuclear receptors. For long-term storage, retinol is esterified to retinyl esters primarily in hepatic stellate cells through the enzyme lecithin:retinol acyltransferase (LRAT), which transfers an acyl group from phosphatidylcholine to retinol. These esters accumulate in lipid droplets within stellate cells, accounting for over 70% of the body's vitamin A reserves, and are mobilized via hydrolysis when needed. LRAT expression is highest in liver, ensuring regulated storage independent of immediate metabolic demands. Regulatory enzymes in these interconversions include alcohol dehydrogenases (ADHs), such as ADH1 and ADH4, which provide auxiliary oxidation of to retinal under high-vitamin A conditions, with Km values typically in the 10-100 μM range for . The SDR dominates physiological regulation, encompassing multiple RDH isoforms with tissue-specific kinetics; for instance, RoDH4 (SDR9C8) has a Km of about 1 μM for and is liver-enriched. These enzymes collectively fine-tune retinal availability, preventing toxicity from excess accumulation.

Role in Animal Vision

Binding to Opsins and Pigment Formation

Opsins are a family of G-protein-coupled receptors characterized by seven transmembrane α-helices, serving as the protein moiety in visual pigments found in the photoreceptor cells of the vertebrate retina. In rod cells, the primary opsin is , while cone cells express three distinct types of opsins responsible for color discrimination. The 11-cis-retinal binds covalently to a conserved residue—specifically Lys296 in bovine —via a protonated linkage, forming the functional light-sensitive pigment. This binding results in the formation of visual pigments with specific absorption properties. exhibits an absorption maximum at approximately 500 , enabling high to dim in rod-mediated . In contrast, cone opsins, historically referred to as iodopsins, display absorption maxima tuned to different spectral regions: short-wavelength-sensitive (SWS) opsins around 420 for , middle-wavelength-sensitive (MWS) opsins around 530 for , and long-wavelength-sensitive (LWS) opsins around 560 for , facilitating trichromatic in humans and many . Isomer specificity is crucial for stable pigment formation, as only the 11-cis of retinal forms a tight, functional with opsins; the all-trans , produced upon , dissociates readily and does not regenerate the under physiological conditions. This selectivity ensures that the maintains a reservoir of the correct for prompt response to stimuli. opsins act as inverse agonists when bound to 11-cis-retinal, stabilizing the inactive state until occurs. The absorption maxima of these pigments are fine-tuned by the protonation state of the and interactions with the surrounding protein . The protonated inherently absorbs around 440 nm, but electrostatic interactions with charged residues, such as the counterion glutamate (Glu113 in ), and hydrophobic packing within the pocket induce bathochromic shifts, extending absorption into the . In cone opsins, variations in residues near the —particularly in the retinal binding pocket—further modulate these shifts to achieve discrimination for , with key tuning sites identified in transmembrane helices.

Visual Cycle and Phototransduction

The visual cycle in vertebrate photoreceptors begins with the absorption of a by 11-cis-retinal bound to in , triggering an ultrafast to all-trans-retinal within 200 femtoseconds, which initiates a series of conformational changes leading to the primary photoproduct bathorhodopsin. This , occurring in less than 1 , stores the photon's energy and propagates through intermediates such as lumirhodopsin and metarhodopsin I, culminating in the active metarhodopsin II (also known as R*) within milliseconds. Metarhodopsin II serves as the signaling state, binding and activating the by catalyzing the exchange of GDP for GTP on its α-subunit. The activated transducin-α-GTP complex then stimulates the effector enzyme cGMP phosphodiesterase (PDE6), which hydrolyzes cyclic guanosine monophosphate (cGMP) in the rod outer segment, rapidly decreasing its concentration from approximately 5 μM in the dark to below 1 μM upon illumination. This drop in cGMP closes cGMP-gated cation channels on the plasma membrane, reducing the inward flux of Na⁺ and Ca²⁺, which hyperpolarizes the rod photoreceptor from a dark potential of about -40 mV to -70 mV, thereby modulating release to cells. The process amplifies the signal, with a single activating hundreds of molecules and thousands of PDE6 catalytic events, ensuring high sensitivity for low-light detection. All-trans-retinal is subsequently released from , marking the decay of metarhodopsin II and the termination of the phototransduction cascade through GTP hydrolysis and PDE6 inhibition. Regeneration of 11-cis- occurs via the to restore sensitivity. In rods, all-trans- is reduced to all-trans- by retinol dehydrogenases (RDH8 and RDH12) in the photoreceptor, then transported to the adjacent (RPE) bound to interphotoreceptor retinoid-binding protein (IRBP). In the RPE, all-trans- is esterified by lecithin: acyltransferase (LRAT) and isomerized to 11-cis-retinol by the RPE65, a membrane-associated isomerohydrolase that uses all-trans-retinyl esters as substrate and requires a iron cofactor for activity. The 11-cis- is oxidized to 11-cis-, which returns to the photoreceptor for rebinding to , completing the cycle in minutes under typical conditions. While rods primarily rely on this RPE-dependent cycle, cones exhibit a faster regeneration pathway involving an intra-retinal mediated by Müller glial cells, enabling dark adaptation approximately 10 times quicker than in to support rapid and adaptation to varying light intensities. This cone-specific cycle utilizes enzymes like retinosome-associated proteins in the , bypassing the slower RPE transport, and is essential for maintaining function under bright, dynamic lighting.

Functions in Microorganisms

Microbial Rhodopsins and Mechanisms

Microbial rhodopsins, classified as Type I rhodopsins, are a diverse family of light-activated retinal-binding proteins found primarily in , , and some eukaryotes, distinct from Type II rhodopsins in that function as G-protein-coupled receptors for visual signaling. Unlike Type II rhodopsins, which utilize 11-cis-retinal and undergo slower signaling cascades, Type I rhodopsins employ all-trans-retinal as the and exhibit rapid photocycles enabling direct transport or sensory responses without dissociation of the retinal. A prototypical example is bacteriorhodopsin (BR), discovered in the archaeon Halobacterium salinarum, where it forms a light-driven proton pump in the purple membrane. Upon absorption of green light, the all-trans-retinal in BR isomerizes to 13-cis, triggering a series of conformational changes in the seven-transmembrane helix structure that translocates a proton from the cytoplasm to the extracellular space, establishing a proton gradient used for ATP synthesis via ATP synthase. This photocycle, completing in milliseconds, includes intermediates such as the red-shifted K state and the deprotonated M state, restoring the original configuration. Halorhodopsin (HR), also from H. salinarum, exemplifies anion transport among microbial rhodopsins, functioning as a pump to maintain intracellular balance under high-salinity conditions. Its mechanism mirrors BR's but directs inward flux, with the retinal facilitating anion binding and release through analogous photocycle states. , such as ChR2 from the green alga , represent light-gated cation channels, allowing rapid influx of Na⁺, K⁺, Ca²⁺, and H⁺ upon illumination. The 13-cis opens a conduction pathway, enabling for applications like , with channel opening kinetics in the microsecond range. Sensory rhodopsins (SRs), including SRI and SRII in H. salinarum and fungal SRs, mediate phototaxis by modulating flagellar activity rather than ion pumping. Light-induced isomerization in SRs alters interactions with transducer proteins, propagating signals to alter motility toward or away from light sources, as seen in SRII's avoidance response in archaea. In microbes, retinal is synthesized de novo from β-carotene, derived via the mevalonate pathway from isopentenyl pyrophosphate, with lycopene β-cyclase (encoded by crtY in H. salinarum) converting lycopene to β-carotene as a key step before oxidative cleavage to retinal. Some bacteria, like those in marine environments, may acquire retinal externally from environmental sources, while archaea like Halobacterium rely predominantly on internal biosynthesis triggered by apoprotein expression. The diversity of microbial rhodopsins extends to over 7,000 identified sequences across prokaryotes and eukaryotes, encompassing proton pumps like proteorhodopsin in ocean , sodium pumps in marine microbes, and anion channels, reflecting adaptations to varied and ionic environments.

Evolutionary and Biotechnological Roles

Retinal-based proteins, known as microbial rhodopsins, trace their evolutionary origins to ancient , where they first enabled light-driven proton pumping for energy generation in extreme environments. Subsequent events disseminated these genes to and eukaryotes, fostering widespread adoption across domains of life and contributing to the diversification of phototrophic mechanisms. This genetic mobility underscores the versatility of retinal as a , allowing its integration into diverse protein scaffolds for sensing and energy transduction. The primordial role of retinal predates the emergence of animal vision by billions of years, serving primarily in microbial phototrophy to harvest and facilitate atmospheric oxygenation on . In like , utilized retinal for efficient proton translocation, a process that likely influenced the evolution of oxygenic by competing for resources in niches. This ancient functionality highlights retinal's foundational impact on life's adaptation to before its co-option into complex visual systems in metazoans. In , retinal-binding have revolutionized , enabling precise optical control of neural activity. Pioneering work demonstrated that expressing channelrhodopsin-2 in mammalian neurons allows millisecond-precision activation via , as shown in hippocampal slices and mouse brain circuits where light pulses reliably evoked action potentials and modulated behavior. This approach has facilitated mapping of neural pathways, with applications extending to therapeutic modulation of circuits in models of neurological disorders. Bacteriorhodopsin, another retinal protein, finds industrial use in and owing to its robust photocycle, which achieves near-unity in proton pumping. In , oriented films of exhibit efficiencies up to 20%, enabling reversible data recording through light-induced conformational changes. For solar devices, biohybrid photovoltaic cells incorporating generate photocurrents with power conversion efficiencies around 0.5-2%, leveraging the protein's stability and directional charge separation for sustainable energy harvesting. Synthetic biology efforts have engineered retinal variants and mutants to enhance selectivity, expanding their utility in cellular tools. For instance, kalium from Klebsormidium nitens exhibit over 100-fold K⁺ preference over Na⁺, enabling light-gated flux for precise membrane hyperpolarization in optogenetic applications. These proteins, informed by cryo-EM structures, optimize retinal's interaction with the protein pocket to tune conductance properties without compromising photocycle kinetics.

Physiological and Clinical Aspects

Health Implications of Deficiency and Excess

Deficiency in , which limits the production of retinal, primarily manifests as night blindness due to impaired regeneration of , the light-sensitive pigment in photoreceptor cells. Prolonged deficiency progresses to , a spectrum of ocular disorders including conjunctival dryness, corneal ulceration, and potentially irreversible blindness. Additionally, it compromises immune function, elevating the risk of severe infections such as , , and respiratory illnesses, particularly in children. Globally, remains a significant issue, affecting more than half of countries, predominantly in developing regions where an estimated 250,000–500,000 vitamin A-deficient children become blind every year, half of whom die within 12 months of losing their sight, and it increases mortality from common infections. However, global prevalence has been decreasing, with incident cases dropping from approximately 127 million in 1990 to 23 million in 2019, though it remains a concern in regions like and (as of 2024 data). levels below 20 μg/dL serve as a key for deficiency, reflecting depleted liver stores and systemic inadequacy. Excess intake of preformed , leading to , causes acute symptoms such as , , and vertigo, while chronic exposure results in liver toxicity, including and . Elevated levels from overload also pose teratogenic risks, inducing congenital malformations in the , heart, and limbs when consumed during . The tolerable upper intake level for adults is 3,000 μg/day to prevent these adverse effects. Vitamin A bioavailability interacts with minerals like and iron; hinders retinol-binding protein synthesis, impairing retinal transport, while disrupts iron mobilization and hemoglobin formation, exacerbating . Supplementation with these minerals can enhance status indicators in deficient populations.

Therapeutic Applications and Research

Retinal, as a key component of the , plays a central role in therapies aimed at restoring or supporting vision in degenerative eye diseases. The Age-Related Eye Disease Study (AREDS) and its follow-up AREDS2 demonstrated that high-dose supplements containing vitamins C and E, , , , and can reduce the risk of progression from intermediate to advanced () by approximately 25% in high-risk patients. The original AREDS formulation included beta-carotene (a provitamin A precursor that converts to retinal), but AREDS2 replaced it with and due to risk in smokers; these formulations, now standard for management, provide antioxidant support for photoreceptor health. Derivatives of retinal, such as 13-cis-retinoic acid (), have established therapeutic roles beyond . is FDA-approved for severe recalcitrant nodular , where it reduces sebum production, prevents follicular hyperkeratinization, and exhibits effects by modulating via receptors. Clinical trials show remission in up to 80% of patients after a 4-5 month course at 0.5-1 mg/kg/day, though it requires strict monitoring for teratogenicity and . In genetic disorders, targets defects in the retinal ; for (LCA) caused by mutations, which impair 11-cis-retinal regeneration, subretinal delivery of adeno-associated viral (AAV) vectors encoding functional has restored enzyme activity and improved in phase I/II trials. Patients showed sustained pupillary light responses and mobility improvements up to 3 years post-treatment, marking the first approved retinal (Luxturna). Retinal prosthetics offer bionic alternatives for end-stage retinal diseases like , where photoreceptor loss disrupts phototransduction. The Argus II Retinal Prosthesis System, an epiretinal implant, bypasses damaged photoreceptors by converting camera-captured light into electrical pulses delivered to surviving retinal ganglion cells via a 60- array, enabling patients to perceive light patterns, motion, and large objects. Implanted in over 350 patients worldwide since FDA approval in 2013, it improves functional vision in daily tasks, though resolution remains low (20/1260 equivalent). Ongoing refinements focus on higher electrode counts and wireless power to enhance spatial resolution. Emerging research extends retinal's derivatives to oncology and neurodegeneration. All-trans-retinoic acid (ATRA), synthesized from retinal, induces in (APL) by targeting PML-RARα fusion proteins, achieving complete remission rates over 90% when combined with . Preclinical studies explore its role in solid tumors like via similar differentiation pathways. In neurodegeneration, 2020s investigations link /retinal homeostasis to (AD); mouse models deficient in retinal dehydrogenase show amyloid-beta accumulation and synaptic loss, while dietary supplementation modulates to reduce and cognitive decline. A 2024 study in Frontiers in Nutrition reported that -enriched diets altered intestinal transcriptomes, lowering AD biomarkers in transgenic models, suggesting preventive potential. Retinal imaging also serves as a non-invasive AD biomarker, with thinning of inner retinal layers correlating to disease progression in cohort studies.

Historical Development

Early Isolation and Identification

In the early 1900s, investigations into dietary deficiencies revealed the existence of a fat-soluble essential for growth and health. In 1913, Elmer V. McCollum and Marguerite Davis isolated this factor from butter fat and egg yolk, demonstrating through rat feeding experiments that it prevented conditions like and supported normal development, marking the discovery of . The link between this and emerged from prior studies on retinal pigments. As early as 1877, Wilhelm Kühne isolated visual purple—later identified as —from frog retinas, observing its purple color in the dark and its bleaching to a yellow intermediate upon light exposure, suggesting a photochemical role in sight; this work was revisited in the 1920s and 1930s as researchers like Alfred Kühn explored its regeneration in insects and amphibians. Advancing chemical characterization, Paul Karrer determined the structure of () in 1931 and synthesized it the following year, confirming its polyene chain via degradation and spectroscopic methods. In 1933, extracted from mammalian retinas and verified its identity through , showing maxima at 325 nm matching synthetic standards, while collaborating with Karrer to analyze samples from , sheep, and pigs. These efforts relied on bioassays measuring growth restoration in vitamin A-deficient rats and color reactions like the antimony trichloride test for potency, alongside rudimentary UV-visible to detect carotenoid-like bands. The specific isolation of retinal, the aldehyde form critical to visual pigments, occurred in the mid-1940s. In 1944, Richard A. Morton and T. W. Goodwin oxidized A₁ to yield crystalline retinene₁, characterizing its UV-vis absorption maximum at approximately 380 nm and its color reaction absorption at 664 nm in the test, confirming its role as a . By 1946, S. Ball, Goodwin, and Morton established retinene₁ as the of A₁ through reduction back to and comparative , solidifying its structural identity without altering the polyene backbone. Later biochemical confirmations validated these findings through enzymatic assays.

Key Discoveries and Milestones

In the 1950s and 1960s, Ruth Hubbard and made pivotal advances in elucidating the , demonstrating that retinal undergoes a series of enzymatic conversions involving oxidation and reduction to regenerate the visual pigment after light exposure. Their work established that all-trans-retinal, produced upon absorption, is reduced to all-trans-retinol and transported to the for re-isomerization back to 11-cis-retinal, which then binds to to reform . Hubbard and Wald specifically identified 11-cis-retinal as the key isomer bound to in the dark-adapted state, confirming its role in the cycle through spectroscopic and biochemical analyses. This understanding of the cycle earned Wald the in or in 1967, shared with Ragnar Granit and Haldan Keffer Hartline, for foundational insights into . The 1970s and 1980s saw the discovery of retinal's roles beyond vertebrate vision, with Walther Stoeckenius and Dieter Oesterhelt identifying in 1971 as a light-driven in halobium, marking the first microbial . This protein, containing all-trans-retinal as its , undergoes to 13-cis-retinal upon light absorption, generating a proton gradient for ATP synthesis and expanding retinal's functional repertoire to microbial energy transduction. Subsequent studies in the 1980s revealed diverse microbial rhodopsins, such as , which facilitated ion flux and highlighted evolutionary conservation of retinal-based photobiology. During the 1990s and 2000s, advanced retinal research significantly, with Krzysztof Palczewski and colleagues determining the first of bovine at 2.8 in 2000, revealing how 11-cis-retinal binds within the seven-transmembrane bundle of the G-protein-coupled receptor. This structure elucidated the molecular basis of retinal's linkage to and its role in stabilizing the inactive state, providing a template for understanding phototransduction signaling. In 2005, , , and colleagues pioneered by expressing channelrhodopsin-2—a microbial retinal-binding protein—in mammalian neurons, enabling precise optical control of neural activity with millisecond precision and transforming tools. In the 2010s and 2020s, clinical and biophysical milestones emerged, including the 2017 FDA approval of (Luxturna), the first for inherited retinal dystrophy caused by RPE65 mutations, which restores the enzyme essential for 11-cis-retinal production in the and improves vision in affected patients. Concurrently, studies using and time-resolved have uncovered coherent vibrational and electronic effects during retinal's ultrafast , occurring in femtoseconds and involving conical intersections that enhance in both and . These insights, from works like Nogly et al. in 2018, reveal quantum mechanical underpinnings of retinal's , informing models of in biological systems.

References

  1. [1]
    Retinal | C20H28O | CID 638015 - PubChem - NIH
    All-trans-retinal is a retinal in which all four exocyclic double bonds have E- (trans-) geometry. It has a role as a gap junctional intercellular ...Missing: function | Show results with:function
  2. [2]
    Structural biology of 11-cis-retinaldehyde production in the classical ...
    The vitamin A derivative 11-cis-retinaldehyde plays a pivotal role in vertebrate vision by serving as the chromophore of rod and cone visual pigments.
  3. [3]
    Chemistry of the Retinoid (Visual) Cycle - PMC - NIH
    Many different compounds can be generated from this monocyclic diterpenoid, which contains a β-ionone ring and polyene chain with a C15 aldehyde group.
  4. [4]
    Chemistry of the Retinoid (Visual) Cycle | Chemical Reviews
    ... retinal. Many different compounds can be generated from this monocyclic diterpenoid, which contains a β-ionone ring and polyene chain with a C15 aldehyde group.
  5. [5]
    Shedding new light on the generation of the visual chromophore
    Aug 5, 2020 · Since 11-cis-retinol is thermodynamically less stable than all-trans-retinol by ∼4 kcal/mol, it has been argued that ester cleavage, which has a ...Abstract · Sign Up For Pnas Alerts · Different Opsins And Their...
  6. [6]
    The stability of 11-cis-retinal and reactivity toward nucleophiles
    The prosthetic group of visual pigments, 11-cis--retinal, is the least resistant of the mono-cis isomers to isomerization and is converted to all-trans-retinal.Missing: stereochemistry photoisomerization
  7. [7]
    116-31-4 CAS MSDS (Retinal) Melting Point Boiling Point Density ...
    ChemicalBook Provide 116-31-4(Retinal)Melting Point Boiling Point ... Boiling point: 366.92°C (rough ... all-trans-Retinal,99% all trans-Retinal,VitaminA ...<|separator|>
  8. [8]
    All-Trans Retinal Mediates Light-Induced Oxidation In Single ... - NIH
    Because its absorption spectrum has a λmax ~ 380 nm and extends well into the visible range, all-trans retinal can mediate oxidation by visible light.
  9. [9]
    Tuning the Electronic Absorption of Protein-Embedded all-trans ...
    ... λmax = 508 nm, a 68 nm red-shift as compared to the PSB of retinal in ethanol (λmax = 440 nm). In contrast to the rhodopsins and our earlier engineered ...
  10. [10]
    Vitamin A aldehyde-taurine adduct and the visual cycle | PNAS
    Sep 21, 2020 · Inefficient removal of the NRPE isomers is a hazard since this Schiff base conjugate can react with a second molecule of all-trans-retinal ...
  11. [11]
    Vitamin A and Carotenoids - Health Professional Fact Sheet
    Mar 10, 2025 · Preformed vitamin A is found in foods from animal sources, including dairy products, eggs, fish, and organ meats [1,2]. Provitamin A carotenoids ...
  12. [12]
    Provitamin A metabolism and functions in mammalian biology
    Provitamin A carotenoids such as β-carotene are the major source for retinoids (vitamin A and its derivatives) in the human diet.
  13. [13]
    The Human Enzyme That Converts Dietary Provitamin A ...
    β-Carotene 15–15′-oxygenase (BCO1) catalyzes the oxidative cleavage of dietary provitamin A carotenoids to retinal (vitamin A aldehyde).
  14. [14]
    β-Carotene Is an Important Vitamin A Source for Humans1–3
    The oxidative cleavage of β-carotene, the major carotenoid of human diets, is achieved by BCMO1, which cleaves β-carotene into 2 molecules of all-trans-retinal ...
  15. [15]
    Bioconversion of dietary provitamin A carotenoids to vitamin A in ...
    These data show that the conversion efficiency of dietary β-carotene to retinol is in the range of 3.6–28:1 by weight.
  16. [16]
    Absorption and retinol equivalence of β-carotene in humans is ...
    These results show that while less cleavage of β-carotene occurred due to vitamin A supplementation, higher absorption resulted in larger molar vitamin A ...
  17. [17]
    Vitamin A: Overlapping Delivery Pathways to Tissues from the ...
    Provitamin A carotenoids can be absorbed intact from the diet and are found in the circulation both postprandially in chylomicrons and under fasting conditions ...Missing: enterohepatic | Show results with:enterohepatic
  18. [18]
    Vitamin A Transporters in Visual Function: A Mini Review on ... - NIH
    The transport of hepatic retinol within the serum is facilitated through its binding to retinol-binding protein 4 (RBP4). RBP4, the transport protein ...Missing: enterohepatic | Show results with:enterohepatic
  19. [19]
    Vitamin A Metabolism: An Update - MDPI
    This review will focus on recent advances for understanding retinoid metabolism that have taken place in the last ten to fifteen years. Keywords: chylomicron; ...Missing: enterohepatic | Show results with:enterohepatic
  20. [20]
    Enzymology of retinoic acid biosynthesis and degradation
    This review summarizes current knowledge about the roles of various biosynthetic and catabolic enzymes in the regulation of retinoic acid homeostasisMissing: interconversions | Show results with:interconversions
  21. [21]
    Kinetic Analysis of Human Enzyme RDH10 Defines the ... - NIH
    RDH10 has a relatively high apparent Km value for NAD+ (∼100 μm) but the lowest apparent Km value for all-trans-retinol (∼0.035 μm) among all NAD+-dependent ...Missing: Km RALDH isoforms
  22. [22]
    Early activation of hepatic stellate cells induces rapid initiation of ...
    Jul 26, 2024 · Lecithin:retinol acyltransferase (LRAT) is the main enzyme producing retinyl esters (REs) in quiescent hepatic stellate cells (HSCs).
  23. [23]
    Retinoid Absorption and Storage Is Impaired in Mice Lacking ...
    Approximately 66–75% of dietary retinoid is taken up and stored as retinyl ester in the liver (9, 10), primarily in the nonparenchymal hepatic stellate cells ( ...
  24. [24]
    Ultrafast structural changes direct the first molecular events of vision
    Mar 22, 2023 · The structure of rhodopsin consists of seven transmembrane (TM) α-helices with an 11-cis retinal chromophore covalently bound through a ...<|control11|><|separator|>
  25. [25]
    The counterion–retinylidene Schiff base interaction of an ... - Nature
    May 13, 2019 · The opsin has a fold consisting of seven transmembrane α-helices, where retinal is covalently attached to a Lys residue (Lys296, numbering ...
  26. [26]
    Retinal orientation and interactions in rhodopsin reveal a two-stage ...
    Sep 2, 2016 · Its light-sensitive retinal chromophore is covalently bound via a protonated Schiff's base (PSB) linkage to Lys2967.43 (superscripts denote ...
  27. [27]
    Absorption spectrum of rhodopsin: 500 nm absorption band - PubMed
    Absorption spectrum of rhodopsin: 500 nm absorption band.Missing: maximum | Show results with:maximum
  28. [28]
    Characterization of mutant rhodopsins responsible for autosomal ...
    May 5, 1993 · However, each of the Thr-58 and Arg-135 mutants bound 11-cis-retinal to form a pigment with a visible absorbance maximum (lambda max) of 500 nm.
  29. [29]
    Spectral Tuning of Pigments Underlying Red-Green Color Vision
    Variations in the absorption spectra of cone photopigments over the spectral range of about 530 to 562 nanometers are a principal cause of individual ...
  30. [30]
    A rhodopsin exhibiting binding ability to agonist all-trans-retinal
    Vertebrate rhodopsins are able to bind the inverse agonist 11-cis-retinal but are unable to bind the agonist all-trans-retinal, indicating that vertebrate ...
  31. [31]
    An all-trans-retinal-binding opsin peropsin as a potential dark-active ...
    Feb 23, 2018 · The absorbance around 390 nm decreased by approximately 40% with a slight blue shift of the peak after illumination for 15 min, in agreement ...Missing: wavelength | Show results with:wavelength
  32. [32]
    Molecular mechanisms and evolutionary robustness of a color ...
    Jan 24, 2024 · The general color tuning mechanism of retinal proteins, i.e., how the protein environment shifts the absorption maximum of the bound chromophore ...
  33. [33]
    Mechanisms of spectral tuning in blue cone visual pigments. Visible ...
    Sep 18, 1998 · Spectral tuning by visual pigments involves the modulation of the physical properties of the chromophore (11-cis-retinal) by amino acid side chains.
  34. [34]
    The First Step in Vision: Femtosecond Isomerization of Rhodopsin
    The first step in vision, the 11-cis→11-trans torsional isomerization of the rhodopsin chromophore, is essentially complete in only 200 femtoseconds.<|control11|><|separator|>
  35. [35]
    Phototransduction in Rods and Cones by Yingbin Fu - Webvision
    Jul 30, 2018 · Activation of rod phototransduction cascade that results in the closure of cGMP-gated channels on the plasma membrane (from dark to light state) ...
  36. [36]
    Phototransduction - Basic Neurochemistry - NCBI Bookshelf
    In the cGMP-cascade mechanism, hydrolysis of G protein-bound GTP to GDP inactivates G protein. Thermal decay of R*, presumably meta-rhodopsin II, to opsin ...
  37. [37]
    RPE65 is the isomerohydrolase in the retinoid visual cycle - PMC - NIH
    Although it is known that RPE65 is critical for regeneration of 11-cis retinol in the visual cycle, the function of RPE65 is elusive. Here we show that ...
  38. [38]
    The Cone-specific Visual Cycle - PMC - PubMed Central
    Thus, the retina visual cycle could explain the substantially faster dark adaptation of cones compared to that of rods, which rely only on the RPE visual cycle ...
  39. [39]
    The Retina-Based Visual Cycle - Annual Reviews
    Sep 18, 2024 · The continuous function of vertebrate photoreceptors requires regeneration of their visual pigment following its destruction upon activation ...
  40. [40]
  41. [41]
  42. [42]
  43. [43]
  44. [44]
    Microbial rhodopsins: functional versatility and genetic mobility
    Here, we discuss the evolution of the type 1 microbial rhodopsins and document five cases of lateral gene transfer (LGT) between domains.
  45. [45]
    The Evolutionary Kaleidoscope of Rhodopsins | mSystems
    Sep 19, 2022 · ... microbial rhodopsins horizontally transferred between archaea and bacteria. ... Horizontal gene transfer in eukaryotic evolution. Nat Rev Genet 9 ...
  46. [46]
    Early evolution of purple retinal pigments on Earth and implications ...
    Oct 11, 2018 · We propose that retinal-based phototrophy arose early in the evolution of life on Earth, profoundly impacting the development of photosynthesis.Missing: vision | Show results with:vision
  47. [47]
    Millisecond-timescale, genetically targeted optical control of neural ...
    Aug 14, 2005 · We demonstrate reliable, millisecond-timescale control of neuronal spiking, as well as control of excitatory and inhibitory synaptic transmission.
  48. [48]
    Diffraction efficiency of bacteriorhodopsin films for holography ...
    May 1, 1992 · Diffraction efficiency of bacteriorhodopsin films for holography containing bacteriorhodopsin wildtype BRWT and its variants BRD85E and BRD96N.Missing: solar | Show results with:solar
  49. [49]
    Recent advances in bacteriorhodopsin-based energy harvesters ...
    In this review, recent advances in bR-based hybrid electrodes are summarized for their applications toward solar cells, water-splitting devices, and fuel cells.
  50. [50]
    Physiology, Night Vision - StatPearls - NCBI Bookshelf
    Rods have a single photopigment, rhodopsin, which utilizes the protein scotopsin and the Vitamin A–derived retinol cofactor.[1] This cascade is essential for ...Introduction · Cellular Level · Mechanism · Pathophysiology
  51. [51]
    Vitamin A Deficiency - Nutritional Disorders - Merck Manuals
    Deficiency impairs immunity and hematopoiesis and causes rashes and typical ocular effects (eg, xerophthalmia, night blindness). Diagnosis is based on typical ...
  52. [52]
    Vitamin A deficiency - World Health Organization (WHO)
    A plasma or serum retinol concentration <0.70 μmol/L indicates subclinical vitamin A deficiency in children and adults, and a concentration of <0.35 µmol/L ...
  53. [53]
    Analysis for policy to overcome barriers to reducing the prevalence ...
    Jun 26, 2021 · Nearly 30% of children under the age of five are estimated to suffer from vitamin A deficiency (VAD) worldwide, and 190 million preschool ...<|control11|><|separator|>
  54. [54]
    Vitamin A Deficiency - StatPearls - NCBI Bookshelf
    Vitamin A deficiency can lead to ophthalmological, dermatological, and immune impairment. This activity addresses the complications of vitamin A deficiency and ...
  55. [55]
    Vitamin A Toxicity - StatPearls - NCBI Bookshelf
    Sep 2, 2023 · The toxicity symptoms include nausea, vomiting, headache, dizziness, irritability, blurred vision, and muscular incoordination. Acute toxicity ...
  56. [56]
    Vitamin A and Pregnancy: A Narrative Review - MDPI
    The mechanism of action by which vitamin A exerts teratogenicity is attributed to the influence of high concentrations of certain retinoic acid metabolites ( ...
  57. [57]
    Vitamin A - The Nutrition Source
    Preformed vitamin A comes from animal products, fortified foods, and vitamin supplements. Carotenoids are found naturally in plant foods. There are other types ...
  58. [58]
    Zinc | Linus Pauling Institute | Oregon State University
    Zinc and vitamin A interact in several ways. Zinc is a component of retinol-binding protein, a protein necessary for transporting vitamin A in the blood. Zinc ...Function · Deficiency · Disease Prevention · Disease Treatment
  59. [59]
    [PDF] Influence of Provitamin A Carotenoids on Iron, Zinc - HarvestPlus
    For example, when vitamin A is deficient in the diet, iron metabolism is negatively affected and iron is not incorporated effectively into hemoglobin. (Hodges ...
  60. [60]
    Iron and zinc supplementation improves indicators of vitamin A ...
    Supplementation with zinc, iron, or both improved indicators of vitamin A status. The results of this study agree with previous observations of a metabolic ...
  61. [61]
    Age-Related Eye Disease Studies (AREDS/AREDS2)
    Dec 27, 2024 · The AREDS studies were designed to learn more about the natural history and risk factors of age-related macular degeneration (AMD) and cataract.AREDS 2 Supplements · AREDS Frequently Asked... · Clinical Trials
  62. [62]
    AREDS/AREDS2 Clinical Trials - National Eye Institute - NIH
    Dec 3, 2024 · The NEI conducted the Age-Related Eye Disease Study (AREDS) and the follow-on AREDS2 to study cataract and age-related macular degeneration (AMD)<|control11|><|separator|>
  63. [63]
    The use of isotretinoin in acne - PMC - NIH
    Systemic Isotretinoin for the Treatment of Acne. Oral isotretinoin (13-cis-retinoic acid) was first approved as treatment for severe acne by the US Food and ...
  64. [64]
    Oral isotretinoin therapy for acne vulgaris - UpToDate
    Apr 30, 2025 · Oral isotretinoin (13-cis retinoic acid) is a retinoid most often used for the treatment of acne vulgaris, particularly severe or ...
  65. [65]
    Effect of Gene Therapy on Visual Function in Leber's Congenital ...
    Subretinal delivery of recombinant adeno-associated virus vector containing the RPE65 cDNA results in improved retinal function and improved vision, as ...
  66. [66]
    Human RPE65 Gene Therapy for Leber Congenital Amaurosis
    Human gene therapy with rAAV2-vector was performed for the RPE65 form of childhood blindness called Leber congenital amaurosis.
  67. [67]
    Argus II retinal prosthesis system: a review of patient selection ...
    Jun 13, 2018 · The Argus II retinal prosthesis system is one such artificial vision device approved for patients with RP. This review focuses on the factors important for ...Patient Selection · Surgical Procedure · Special Surgical...Missing: phototransduction | Show results with:phototransduction
  68. [68]
    Argus II: The 'Bionic Eye' An Incredible Breakthrough for People with ...
    The Argus II is a three-part device that allows some perception of light and motion in patients who have lost their vision due to retinitis pigmentosa.Missing: phototransduction | Show results with:phototransduction
  69. [69]
    Retinal Prostheses: Engineering and Clinical Perspectives for Vision ...
    Schematic representation of the main components of the ARGUS II retinal prosthesis system, including the camera, video processing unit, and electrode array.Retinal Prostheses... · 4.2. 1. Epiretinal... · 4.3. 1. Epiretinal...
  70. [70]
    Retinoic acid and cancer treatment - PMC - NIH
    Nov 28, 2014 · It is currently understood that retinoic acid plays important roles in cell development and differentiation as well as cancer treatment.
  71. [71]
    Current Role and Future Directions of the Retinoic Acid Pathway in ...
    Retinoids and their naturally metabolized and synthetic products (e.g., all-trans retinoic acid, 13-cis retinoic acid, bexarotene) induce differentiation in ...
  72. [72]
    Therapeutic insights elaborating the potential of retinoids ... - Frontiers
    Increased ingestion of pro-vitamin A carotenoids is linked to a decreased risk of various brain disorders, including Alzheimer's disease, according to ...Missing: 2020s | Show results with:2020s
  73. [73]
    Dietary vitamin A shows promise in Alzheimer's disease intervention ...
    Mar 28, 2024 · Study explored the impact of dietary vitamin A on gut microbiota and intestinal transcriptome, revealing its potential in Alzheimer's ...Missing: 2020s | Show results with:2020s
  74. [74]
    Retinal manifestations and their diagnostic significance in ...
    Aug 10, 2025 · A meta-analysis conducted by Mejia-Vergara et al. (2020) on using OCT in MCI and AD found that OCT can detect inner retinal degeneration. ...Missing: 2020s | Show results with:2020s
  75. [75]
    The color purple: milestones in photochemistry - The FASEB Journal
    Dec 1, 2008 · and Wald, G. (1952) Cis-trans isomers of vitamin A and retinene in the rhodopsin system. J. Gen. Physiol. 36, 269–315. 10.1085/jgp.36.2.269.<|control11|><|separator|>
  76. [76]
    [PDF] Carotenoids, flavins and vitamin A and B2 - Nobel Prize
    In a lecture which I gave in 1932 I remarked: "The chemistry of the vi- tamins has made great progress in the last few years; it has overtaken the. Page 16. 448.
  77. [77]
    Vitamin A in the Retina - Nature
    Vitamin A in the Retina. Download PDF. Letter; Published: 26 August 1933. Vitamin A in the Retina. GEORGE WALD. Nature volume 132, pages 316–317 (1933)Cite this ...<|control11|><|separator|>
  78. [78]
    George Wald | Biographical Memoirs: Volume 78
    In Zürich, Wald collected retinas from cattle, sheep, and pigs; extracted them with organic solvents; and with Karrer confirmed the presence of vitamin A in all ...
  79. [79]
    The Nobel Prize in Physiology or Medicine 1967 - NobelPrize.org
    The Nobel Prize in Physiology or Medicine 1967 was awarded jointly to Ragnar Granit, Haldan Keffer Hartline and George Wald for their discoveries.Missing: Hubbard cycle
  80. [80]
    Rhodopsin-like Protein from the Purple Membrane of Halobacterium ...
    Sep 29, 1971 · Rhodopsin-like Protein from the Purple Membrane of Halobacterium halobium. DIETER OESTERHELT &; WALTHER STOECKENIUS. Nature New Biology volume ...Missing: discovery | Show results with:discovery
  81. [81]
    The Work of Walther Stoeckenius - ScienceDirect
    In 1971, Walther Stoeckenius discovered that Halobacterium halobium contains a purple pigment that is chemically similar to rhodopsin and works as a ...
  82. [82]
    the history of halobacterial and microbial rhodopsin research | FEMS ...
    Oesterhelt & Stoeckenius, 1971). Because of this cofactor, Oesterhelt and Stoeckenius named the protein 'bacteriorhodopsin' in their 1971 paper (Oesterhelt ...The first microbial rhodopsin... · The genetics of microbial... · Proteorhodopsin...
  83. [83]
    Crystal structure of rhodopsin: A G protein-coupled receptor - PubMed
    GPCRs share many structural features, including a bundle of seven transmembrane alpha helices connected by six loops of varying lengths.
  84. [84]
    LUXTURNA - FDA
    Jun 9, 2022 · June 8, 2022 Approval Letter - LUXTURNA · December 19, 2017 Approval Letter - LUXTURNA · December 18, 2017 Summary Basis for Regulatory Action - ...
  85. [85]
    Retinal isomerization in bacteriorhodopsin captured by a ... - Science
    In this work, we used XFEL radiation to study the structural dynamics of retinal isomerization in the light-driven proton-pump bacteriorhodopsin (bR).