Small Molecule
HO
14280-30-9
The hydroxide ion ($OH^{-}$), cataloged in the DrugBank database under the accession number DB14522, represents a unique case study in pharmacological classification. It is not a therapeutic agent in its own right but rather a fundamental diatomic anion indispensable to the fields of chemistry, biology, and medicine. This report provides an exhaustive analysis of the hydroxide ion, clarifying its identity and deconstructing its multifaceted roles. Its pharmacological relevance is not derived from its use as a standalone drug, but from its function as a reactive chemical moiety within a range of established pharmaceutical compounds, most notably the metal hydroxides employed as antacids. Furthermore, it serves as a critical excipient in drug formulation and manufacturing, primarily for pH adjustment and, in the case of aluminum hydroxide, as a vaccine adjuvant.
Physiologically, the hydroxide ion is a cornerstone of acid-base homeostasis; its concentration, in a delicate inverse relationship with the hydrogen ion, dictates the pH of all bodily fluids, a parameter that must be maintained within a narrow range for enzymatic reactions and cellular processes to function. Its toxicology is entirely context-dependent, varying from the extreme corrosivity of concentrated sodium hydroxide solutions to the relatively benign, though metabolically significant, side effects of poorly soluble antacids like magnesium hydroxide. This report will systematically explore the physicochemical properties of the hydroxide ion, its integral role in human physiology, its applications within hydroxide-containing therapeutic agents, its function as a pharmaceutical excipient, and its toxicological profile. By synthesizing information from chemical databases, physiological literature, and clinical pharmacology, this document aims to provide a definitive and nuanced understanding of this simple yet profoundly important chemical entity.
A comprehensive understanding of the hydroxide ion's role in physiology and pharmacology begins with a detailed characterization of its fundamental chemical and physical properties. This section consolidates data from various chemical and drug information resources to establish a definitive profile.
The hydroxide ion is recognized across scientific disciplines and regulatory databases by a consistent set of identifiers. Its primary designation in the context of pharmacology is its DrugBank Accession Number, DB14522.[1] In the broader chemical literature, it is universally identified by its CAS Registry Number, 14280-30-9.[2] For regulatory and substance identification purposes, it is assigned the Unique Ingredient Identifier (UNII) 9159UV381P.[1]
The entity is known by several synonyms that reflect its chemical nature and context. The most common generic name is "Hydroxide ion".[1] Other frequently used terms include "Hydroxide," "Hydroxyl anion," and "Hydroxy ion".[1] The systematic IUPAC name is simply "hydroxide," though more formal names such as "oxidanide" and "hydridooxygenate(1−)" exist; "oxidanide" is noted as a systematic name that is not commonly recommended in practice.[2] The chemical formula is consistently represented as $OH^{-}$ or $HO^{-}$, with the negative sign indicating its anionic nature.[2] In computational chemistry formats, its structure is denoted by the SMILES string [O-][H].[1]
The hydroxide ion is one of the simplest diatomic anions, composed of a single oxygen atom covalently bonded to a single hydrogen atom.[3] The overall negative electric charge ($z = -1$) arises from the gain of an extra valence electron compared to the neutral hydroxyl radical ($HO\cdot$).[8] This additional electron primarily localizes on the highly electronegative oxygen atom, giving it a full octet of valence electrons.[8]
The Lewis structure for the hydroxide ion depicts the oxygen and hydrogen atoms sharing one pair of electrons to form a single covalent bond. The oxygen atom also possesses three lone pairs of electrons, satisfying the octet rule.[11] The hydrogen atom, with its shared pair, achieves a stable, full outer shell.[11] This electronic configuration explains its formal charge of -1 and its valency of 1, which dictates its ability to form ionic compounds with cations. For instance, it combines with a monovalent cation like sodium ($Na^{+}$) to form $NaOH$ or with a divalent cation like calcium ($Ca^{2+}$) to form $Ca(OH)_{2}$.[8]
As a diatomic species, its molecular geometry is necessarily linear.[12] The molecular weight of the hydroxide ion is reported with minor variations across sources, reflecting differences in isotopic abundance calculations. The average molecular weight is consistently cited as being approximately 17.007 to 17.008 g/mol, while its monoisotopic mass is approximately 17.003 g/mol.[1]
The hydroxide ion is a chemically versatile species with several key functions. It is a strong Brønsted-Lowry base, meaning it readily accepts a proton ($H^{+}$) to form its conjugate acid, water ($H_{2}O$).[3] It also functions as a Lewis base by donating one of its lone pairs of electrons to a Lewis acid.[2] This property allows it to act as a ligand, forming coordination complexes with metal ions, such as the tetrahydroxoaluminate ion, $[Al(OH)_{4}]^{-}$.[2] Furthermore, its electron-rich nature makes it a potent nucleophile and a catalyst in many organic and inorganic reactions.[3] Its conjugate base is the oxide anion ($O^{2-}$).[9]
In aqueous solution, the hydroxide ion is colorless and highly soluble.[8] Certain physical properties listed in chemical databases, such as a boiling point of 100°C and a vapor pressure of 24.5 mmHg at 25°C, are characteristic of its aqueous solution (i.e., water) rather than the ion itself, a distinction that is important for accurate characterization.[14] Computational predictions from DrugBank provide further insight into its properties, with a predicted logP of -0.65, confirming its high hydrophilicity, and a pKa of 15.7 for its conjugate acid (water), underscoring its nature as a strong base.[1]
In an aqueous environment, the hydroxide ion does not exist as a simple, isolated species. It is strongly solvated through the formation of hydrogen bonds with surrounding water molecules.[3] The oxygen atom of the hydroxide ion is a very strong hydrogen bond acceptor, leading to the formation of stable hydrated clusters.[16] The first water molecule binds strongly to form the $H_{3}O_{2}^{-}$ complex. The most stable hydrated form in the gas phase and likely in solution is believed to be a tetrahedral structure, $H_{7}O_{4}^{-}$, where the hydroxide ion's oxygen atom accepts hydrogen bonds from three water molecules, while its own hydrogen atom donates a weak hydrogen bond to a fourth water molecule.[16] This complex hydration shell is dynamic and fundamental to its behavior, influencing its transport mechanism (related to the Grotthuss mechanism for protons) and its reactivity in biological systems.[18]
Table 1: Physicochemical Properties of the Hydroxide Ion (DB14522)
| Property Category | Parameter | Value / Description | Source(s) |
|---|---|---|---|
| Identifiers | Generic Name | Hydroxide ion | 1 |
| DrugBank ID | DB14522 | 1 | |
| CAS Number | 14280-30-9 | 2 | |
| UNII | 9159UV381P | 1 | |
| IUPAC Name | hydroxide | 1 | |
| Synonyms | Hydroxide, Oxidanide, Hydroxyl anion | 1 | |
| Structural Data | Chemical Formula | $OH^{-}$ (or $HO^{-}$) | 2 |
| SMILES | [O-][H] | 1 | |
| InChI | InChI=1S/H2O/h1H2/p-1 | 1 | |
| InChI Key | XLYOFNOQVPJJNP-UHFFFAOYSA-M | 1 | |
| Molecular Data | Average Mol. Weight | 17.0073 g/mol | 1 |
| Monoisotopic Mass | 17.0027 g/mol | 1 | |
| Valency / Charge | -1 | 8 | |
| Physicochemical Properties | pKa (Strongest Acidic) | 15.7 (Predicted, for conjugate acid H₂O) | 1 |
| Conjugate Acid | Water ($H_{2}O$) | 3 | |
| Conjugate Base | Oxide anion ($O^{2-}$) | 3 | |
| Solubility | Highly soluble in water | 8 | |
| Chemical Roles | Primary Functions | Base, Ligand, Nucleophile, Catalyst | 3 |
The tight regulation of hydroxide ion concentration is a fundamental requirement for life. Its influence extends across all levels of biological organization, from the structure of macromolecules to the function of entire organ systems. This section explores the central role of the hydroxide ion in maintaining physiological homeostasis.
The concept of acid-base balance is fundamentally a matter of regulating the concentrations of hydrogen ions ($H^{+}$) and hydroxide ions ($OH^{-}$) in the body's aqueous compartments.[19] These two ions exist in a dynamic equilibrium governed by the autoionization of water: $H_{2}O \rightleftharpoons H^{+} + OH^{-}$.[21] The equilibrium constant for this reaction, known as the ion product of water ($K_{w}$), is approximately $10^{-14}$ at physiological temperatures. This constant dictates a strict inverse relationship: as the concentration of $H^{+}$ rises, the concentration of $OH^{-}$ must fall, and vice versa.[3] The pH scale is a logarithmic measure of the $H^{+}$ concentration, and thus it is also an indirect measure of the $OH^{-}$ concentration.[9]
The human body maintains the pH of arterial blood within an exceptionally narrow range of 7.35 to 7.45.[23] Even minor deviations from this range can have catastrophic consequences. A decrease in pH (acidemia) or an increase in pH (alkalemia) alters the charge states of amino acid residues in proteins, leading to denaturation and loss of function for critical enzymes, ion channels, and structural proteins.[23]
To prevent such fluctuations, the body employs a sophisticated, multi-layered defense system of chemical buffers. These systems are designed to absorb excess acid ($H^{+}$) or base ($OH^{-}$).[25] When an excess of a base, such as hydroxide, enters the system, it is immediately neutralized by the weak acids of these buffer systems.
The existence of these complex, interconnected physiological systems—spanning from the molecular level (hemoglobin) to the organ level (kidneys and lungs)—underscores a central principle of biology. All of these intricate regulatory networks are ultimately dedicated to managing the concentration of two of the simplest chemical species: the proton and its counterpart, the hydroxide ion. This reveals that much of physiological complexity has evolved to safeguard the thousands of pH-sensitive biochemical reactions essential for life from the disruptive potential of an unregulated $H^{+}$ or $OH^{-}$ ion. The maintenance of this simple chemical equilibrium is, therefore, a primary organizing principle of physiological function.
The activity of virtually every enzyme in the body is critically dependent on pH.[19] The concentration of hydroxide ions directly influences the ionization state of acidic and basic amino acid side chains within an enzyme's active site and throughout its structure.[28] These charge states are essential for maintaining the precise three-dimensional conformation of the enzyme and for the specific ionic and hydrogen bonds required for substrate binding and catalysis.[28]
Each enzyme exhibits a characteristic optimal pH at which its catalytic activity is maximal. Deviations from this optimum, caused by an increase or decrease in $OH^{-}$ concentration, lead to a rapid decline in reaction rate. Extreme pH levels can cause irreversible denaturation of the enzyme's protein structure, resulting in a complete loss of function.[28]
Beyond this general influence, the hydroxide ion can also play a direct role as a nucleophile in certain enzymatic mechanisms.
The role of the hydroxide ion in cellular respiration is inextricably linked to the generation of the proton-motive force (PMF) that drives ATP synthesis. During aerobic respiration, the electron transport chain, located in the inner mitochondrial membrane, harnesses the energy from electrons carried by NADH and FADH₂ to actively pump protons ($H^{+}$) from the mitochondrial matrix into the intermembrane space.[32]
This process creates a powerful electrochemical gradient across the membrane. This gradient has two components: a chemical gradient due to the difference in proton concentration (a pH gradient) and an electrical gradient due to the separation of positive charge.[33] While this is often referred to simply as a "proton gradient," it is equally a "hydroxide gradient" in the opposite direction. The accumulation of $H^{+}$ in the intermembrane space makes it more acidic (lower pH, lower $OH^{-}$ concentration) relative to the matrix, which becomes more alkaline (higher pH, higher $OH^{-}$ concentration).
The potential energy stored in this gradient is the PMF. This force drives protons to flow back down their electrochemical gradient into the matrix through a specific protein channel: ATP synthase.[32] The flow of protons through this molecular turbine provides the energy to phosphorylate ADP, forming ATP in a process known as oxidative phosphorylation.[33] The transport of this newly synthesized ATP out of the matrix and ADP into the matrix is also coupled to the gradient, often via an ATP-ADP translocase that functions as an antiport with the hydroxide ion (or a symport with the proton), further highlighting the direct involvement of the $H^{+}/OH^{-}$ balance in the cell's energy economy.[35]
The classification of the hydroxide ion in the DrugBank database under accession number DB14522 as an "investigational" small molecule presents a significant point of ambiguity that requires careful deconstruction.[1] An ion as fundamental as hydroxide is not, and cannot be, an investigational drug in the conventional sense. A critical analysis of the data associated with this entry reveals that its purpose is not to denote a therapeutic agent but to serve as a chemical or structural reference tag, linking various distinct and complex pharmaceutical products that share a hydroxide moiety.
The designation of the hydroxide ion as "investigational" within DrugBank is a classification artifact. Large-scale databases often employ ontological structures that categorize entries based on chemical components. In this framework, DB14522 likely functions as a root entry for any drug that is chemically a hydroxide salt or contains a hydroxide group integral to its structure or formulation. This allows for systematic chemical indexing but can lead to misinterpretation if viewed solely from a pharmacological perspective. The clinical trials associated with this entry are not testing the efficacy of hydroxide ions themselves, but rather the complex drugs to which the tag has been applied.
A systematic examination of the clinical trials linked to DB14522 confirms that the hydroxide ion is never the primary therapeutic agent under investigation.
The consistent pattern across these examples demonstrates the function of the DB14522 entry. It acts as a chemical linker, not a pharmacological descriptor. This distinction is crucial. A non-expert user, observing that an "investigational drug" named "Hydroxide ion" is being tested for "Cognitive Dysfunctions," could form the wildly inaccurate conclusion that simple alkaline solutions are being studied as a treatment for dementia. This highlights a significant challenge in the field of bioinformatics: the potential for "semantic traps" where the logical structure of a database, sound from a chemical indexing standpoint, creates misleading pharmacological inferences. An essential function of expert analysis is to bridge this gap, translating the structural data into accurate functional knowledge and preventing such misinterpretations.
While the hydroxide ion itself is not a drug, it is the key functional component of an important class of medicines: the inorganic hydroxide antacids. In these compounds, the hydroxide moiety is central to the mechanism of action, providing symptomatic relief for acid-related gastrointestinal disorders. Additionally, specific hydroxide compounds have found specialized and vital roles in other fields of medicine, such as dentistry.
Antacids containing metal hydroxides are over-the-counter medications used to treat conditions such as heartburn, acid indigestion, and gastroesophageal reflux disease (GERD).[40] Their primary mechanism of action is the direct chemical neutralization of excess hydrochloric acid (HCl) secreted in the stomach.[42] The hydroxide ions ($OH^{-}$) from the antacid compound react with the hydrogen ions ($H^{+}$) from the gastric acid in a classic neutralization reaction to form water ($H^{+} + OH^{-} \rightarrow H_{2}O$).[44] This reaction raises the gastric pH. Increasing the pH from 1.5 to 3.5 can reduce the concentration of gastric acid by 100-fold.[42] This provides several therapeutic benefits: it alleviates the pain and burning sensation caused by acid irritating the esophageal and gastric mucosa, it inhibits the activity of the proteolytic enzyme pepsin (which is most active at a very low pH), and it may increase the tone of the lower esophageal sphincter, further reducing reflux.[41]
Table 2: Comparative Pharmacology of Common Hydroxide-Based Antacids
| Feature | Aluminum Hydroxide (Al(OH)3) | Magnesium Hydroxide (Mg(OH)2) |
|---|---|---|
| Mechanism of Action | Neutralizes HCl to form $AlCl_{3}$ and $H_{2}O$; binds phosphate | Neutralizes HCl to form $MgCl_{2}$ and $H_{2}O$; osmotic laxative at higher doses |
| Relative Neutralizing Power | Modest | High |
| Onset of Action | Slow | Rapid |
| Duration of Action | Prolonged | Short |
| Primary GI Side Effect | Constipation | Diarrhea |
| Key Systemic Concern | Aluminum toxicity and hypophosphatemia (especially in renal failure) | Hypermagnesemia (especially in renal failure) |
| Common Formulation Strategy | Often combined with $Mg(OH)_{2}$ to balance GI effects | Often combined with $Al(OH)_{3}$ to counteract diarrhea |
| Source(s) | 43 | 43 |
Beyond antacid therapy, certain hydroxide compounds are indispensable in other medical specialties.
In addition to their role in active pharmaceutical ingredients (APIs), hydroxide compounds are widely used as excipients—substances included in a drug formulation for purposes other than direct therapeutic action. These roles are critical for ensuring the stability, efficacy, and safety of the final medicinal product.
One of the most common applications of hydroxide compounds in pharmaceutical manufacturing is as alkalinizing agents.[63] Strong bases like sodium hydroxide ($NaOH$), potassium hydroxide ($KOH$), and to a lesser extent, ammonium hydroxide ($NH_{4}OH$), are used to precisely adjust and maintain the pH of liquid formulations.[64]
The control of pH is paramount for several reasons. For many APIs, stability is highly pH-dependent; an incorrect pH can lead to rapid degradation of the drug, reducing its shelf life and efficacy. Solubility is also frequently influenced by pH. Adjusting the pH can keep a weakly acidic or basic drug in its ionized, more soluble form, which is essential for parenteral (injectable) solutions and some oral liquid formulations.[64] For intravenous solutions, adjusting the pH to a near-physiological level is crucial to prevent pain, irritation, and tissue damage at the injection site. Sodium hydroxide is a key excipient used in the manufacturing processes of many widely used medicines, including aspirin and various anticoagulants, to ensure these stability and solubility requirements are met.[66]
Aluminum hydroxide serves a unique and vital role as a pharmaceutical excipient in the field of vaccinology. It is one of the most widely used adjuvants in human vaccines.[69] An adjuvant is an ingredient that enhances the body's immune response to the antigen (the active component of the vaccine), leading to a more robust and durable immunity.
The mechanism by which aluminum hydroxide acts as an adjuvant is multifaceted. It is believed to create a "depot" or "repository effect" at the injection site. The antigen adsorbs onto the surface of the insoluble aluminum hydroxide particles, which leads to a slow release of the antigen over time. This prolonged exposure allows for a more sustained interaction with the immune system.[49] More recent research has shown that aluminum hydroxide also actively stimulates the innate immune system. It promotes the recruitment of immune cells to the injection site and enhances the uptake of the antigen by antigen-presenting cells, such as macrophages. This activation is thought to involve the NLRP3-inflammasome pathway, a key component of the innate immune response, which ultimately leads to a stronger T-cell and antibody response to the vaccine antigen.[49]
The versatility of hydroxide compounds allows for their use in other specialized excipient functions. For example, calcium hydroxide can be used as a wet binder in the manufacturing of tablets, helping to impart the necessary mechanical strength to the granules during the compression process.[70] Various hydroxides are also found in topical preparations and oral suspensions where they may contribute to pH control or the overall stability of the formulation.[69]
Table 3: Summary of Hydroxide Compounds as Pharmaceutical Excipients
| Hydroxide Compound | Primary Excipient Function | Mechanism / Purpose | Example Applications | Source(s) |
|---|---|---|---|---|
| Sodium Hydroxide ($NaOH$) | Alkalinizing Agent / pH Adjuster | Strong base used to increase the pH of a formulation. | Adjusting pH of injectable solutions, stabilizing APIs in liquid formulations (e.g., aspirin, anticoagulants). | 64 |
| Potassium Hydroxide ($KOH$) | Alkalinizing Agent / pH Adjuster | Strong base, similar function to NaOH. | pH adjustment in various liquid preparations. | 63 |
| Aluminum Hydroxide ($Al(OH)_{3}$) | Vaccine Adjuvant | Creates a "depot effect" for slow antigen release; stimulates the innate immune system (e.g., NLRP3-inflammasome). | Adjuvant in numerous human vaccines (e.g., Hepatitis A, Hepatitis B, DTaP). | 49 |
| Calcium Hydroxide ($Ca(OH)_{2}$) | Wet Binder / pH Modifier | Provides alkaline stability; can be used to aid granulation in tablet manufacturing. | Antacid formulations, topical preparations, tablet manufacturing. | 70 |
| Ammonium Hydroxide ($NH_{4}OH$) | Alkalinizing Agent / pH Adjuster | Weak base used for pH regulation in various chemical and pharmaceutical processes. | pH control in textile processing, chemical synthesis, and some formulations. | 63 |
To fully contextualize the role of hydroxide compounds in pharmaceuticals, it is useful to understand their industrial-scale production and their widespread use across various sectors. Sodium hydroxide, in particular, is a commodity chemical produced in massive quantities globally.
The predominant method for the industrial production of sodium hydroxide is the chlor-alkali process.[3] This is an electrochemical process that involves the electrolysis of a concentrated sodium chloride solution, commonly known as brine.[73] An electric current is passed through the brine, leading to a series of reactions that co-produce three valuable chemicals: sodium hydroxide ($NaOH$), chlorine gas ($Cl_{2}$), and hydrogen gas ($H_{2}$).[72] The overall reaction can be summarized as: $2NaCl(aq) + 2H_{2}O(l) \rightarrow 2NaOH(aq) + Cl_{2}(g) + H_{2}(g)$.[73]
There are three main types of electrolytic cells historically used for this process:
The utility of hydroxide compounds extends far beyond medicine. Their chemical properties make them essential raw materials and processing agents in a vast array of industries.
The toxicological profile of the hydroxide ion is a clear illustration of the principle that "the dose makes the poison." The hazard associated with hydroxide-containing compounds is not determined by the mere presence of the $OH^{-}$ moiety but is critically dependent on the compound's physicochemical properties—primarily its solubility and concentration—which dictate the bioavailable concentration of free hydroxide ions at a biological interface.
Concentrated solutions of strong, highly soluble bases such as sodium hydroxide ($NaOH$) and potassium hydroxide ($KOH$) are extremely corrosive to all living tissues.[77] Upon contact, they cause rapid tissue damage through several mechanisms. The high concentration of hydroxide ions leads to saponification of fats in cell membranes, disrupting their integrity. It also causes hydrolysis and denaturation of proteins, leading to liquefaction necrosis, a type of cell death that results in deep, penetrating tissue injury.[77] The dissolution of solid hydroxides is also highly exothermic, generating significant heat that can cause severe thermal burns in addition to the chemical burns.[79]
This extreme local toxicity stands in stark contrast to the safety profile of therapeutic hydroxides like magnesium hydroxide. The critical difference lies in solubility. Sodium hydroxide is highly soluble, instantly releasing a high concentration of corrosive $OH^{-}$ ions. Magnesium hydroxide, conversely, has very low solubility in water.[53] It acts as a solid reservoir, only dissolving and releasing hydroxide ions as it reacts with and is neutralized by stomach acid. This slow, controlled release prevents the buildup of a high local concentration of free $OH^{-}$, thus avoiding corrosive damage and allowing it to function safely as an antacid. This demonstrates that the toxicity is a function of kinetics and concentration, not merely chemical identity.
Given the severe hazards of concentrated hydroxides, strict safety protocols are essential in industrial and laboratory settings.
When used therapeutically, the primary toxicological concerns are not local corrosivity but systemic metabolic disturbances that can arise from chronic or excessive use, particularly in susceptible populations.
Table 4: Summary of Toxicological Profile and Exposure Limits
| Hydroxide Compound | Primary Hazard Type | Key Toxic Effects / Adverse Effects | Occupational Exposure Limit (Ceiling) |
|---|---|---|---|
| Sodium Hydroxide ($NaOH$) | Corrosive | Severe chemical and thermal burns to skin, eyes, and respiratory tract upon contact. Ingestion causes catastrophic GI tract damage. Not systemically toxic, but local effects can be fatal. | 2 mg/m³ |
| Potassium Hydroxide ($KOH$) | Corrosive | Similar to NaOH; highly corrosive to all tissues. | 2 mg/m³ (ACGIH TLV) |
| Aluminum Hydroxide ($Al(OH)_{3}$) | Metabolic | GI: Constipation. Systemic: Hypophosphatemia with chronic use. In renal failure, risk of aluminum toxicity (encephalopathy, osteomalacia). | Not established for therapeutic use; considered nuisance dust. |
| Magnesium Hydroxide ($Mg(OH)_{2}$) | Metabolic | GI: Diarrhea (dose-dependent). Systemic: In renal failure, risk of hypermagnesemia (CNS depression, cardiac arrhythmias, respiratory depression). | Not established for therapeutic use; considered nuisance dust. |
| Calcium Hydroxide ($Ca(OH)_{2}$) | Irritant / Corrosive | Corrosive as a solid or slurry. In dental use, high pH is therapeutic but can damage surrounding tissues if not properly contained. | 5 mg/m³ (OSHA PEL) |
| Source(s) | 49 | 82 |
The hydroxide ion, identified in DrugBank as DB14522, is a quintessential example of a chemical entity whose significance in pharmacology is defined not by its direct action as a drug, but by its fundamental and versatile roles across chemistry and biology. The analysis of its DrugBank entry reveals that its classification as an "investigational" substance is an artifact of chemical indexing, serving to link disparate pharmaceutical products through a common structural moiety rather than to denote a novel therapeutic agent.
In its physiological context, the hydroxide ion is a central player in the maintenance of life itself. Its concentration, inextricably linked to that of the hydrogen ion, defines the pH of the body's internal environment. The preservation of this delicate acid-base balance is the ultimate objective of complex, multi-organ regulatory systems, as even minor deviations can disrupt the function of countless enzymes and proteins.
The therapeutic utility of the hydroxide ion is realized through its incorporation into inorganic compounds, primarily the metal hydroxide antacids. In this capacity, its basicity is harnessed to neutralize gastric acid, providing relief from common gastrointestinal ailments. The distinct pharmacological profiles of aluminum hydroxide and magnesium hydroxide—with their opposing effects on gut motility and unique systemic risks—highlight the critical importance of the associated cation in determining the overall clinical behavior of the compound. Beyond this, specialized hydroxides like calcium hydroxide have become indispensable tools in modern endodontics, where its extreme alkalinity is leveraged for its potent antimicrobial and tissue-stimulating properties.
Furthermore, the hydroxide ion is a workhorse in pharmaceutical manufacturing, where compounds like sodium hydroxide are essential excipients for pH adjustment, ensuring the stability and solubility of active ingredients. The unique application of aluminum hydroxide as a vaccine adjuvant underscores its capacity to interact with the immune system in a beneficial way, a function far removed from simple acid neutralization.
Finally, the toxicology of the hydroxide ion is a lesson in context. The same chemical entity that is a component of a safe over-the-counter medication can, in the form of a concentrated, highly soluble salt like sodium hydroxide, be an extremely dangerous and corrosive agent. This duality emphasizes that risk is a function of concentration, solubility, and kinetics, not just chemical identity.
In conclusion, the Hydroxide ion (DB14522) is not a drug, but a fundamental building block. Its journey from a simple diatomic anion to a key determinant of physiological homeostasis, a reactive center in widely used medications, and a functional component in advanced pharmaceutical formulations illustrates its profound and multifaceted importance. A deep and nuanced understanding of its properties is therefore essential for progress in fields ranging from clinical medicine and pharmacology to toxicology and industrial chemistry.
Published at: October 29, 2025
This report is continuously updated as new research emerges.
Empowering clinical research with data-driven insights and AI-powered tools.
© 2025 MedPath, Inc. All rights reserved.