Fact-checked by Grok 2 weeks ago

Acid

In chemistry, an acid is defined as a molecular entity or capable of donating a hydron (proton, H⁺) or forming a with an . This encompasses the Brønsted-Lowry concept of proton donation and the concept of electron-pair acceptance, providing a unified framework for understanding acid behavior across various solvents and reaction conditions. Acids are fundamental to numerous chemical processes, exhibiting properties such as a less than 7 in aqueous solutions, the ability to turn paper red, and a sour taste in dilute forms./03:_Acid-Base_Chemistry/3.02:_Brnsted_and_Lewis_Acids_and_Bases) The concept of acids evolved through key theoretical advancements in the late 19th and early 20th centuries. In 1884, proposed the first modern definition, describing acids as substances that increase the concentration of hydrogen ions (H⁺) when dissolved in , laying the groundwork for understanding in aqueous solutions./03:_Acid-Base_Chemistry/3.02:_Brnsted_and_Lewis_Acids_and_Bases) This Arrhenius model was expanded in 1923 by Johannes Brønsted and Thomas Lowry, who redefined acids as proton (H⁺) donors in any acid-base reaction, independent of the solvent and applicable to a broader range of chemical systems./03:_Acid-Base_Chemistry/3.02:_Brnsted_and_Lewis_Acids_and_Bases) Concurrently, introduced a more general perspective in 1923, classifying acids as electron-pair acceptors, which extended the theory to non-protonic reactions and coordination chemistry./03:_Acid-Base_Chemistry/3.02:_Brnsted_and_Lewis_Acids_and_Bases) These definitions—Arrhenius, Brønsted-Lowry, and Lewis—remain central to contemporary acid-base chemistry, with the IUPAC Gold Book integrating them into its current nomenclature. Acids are classified by their strength, source, and structure, influencing their reactivity and applications. Strong acids, such as (HCl) and (H₂SO₄), fully dissociate in to yield H⁺ ions, resulting in high conductivity and corrosive properties./05:_Molecules_and_Compounds/5.09:_Naming_Acids) In contrast, weak acids like acetic acid (CH₃COOH) partially dissociate, establishing with their conjugate bases and exhibiting milder effects./05:_Molecules_and_Compounds/5.09:_Naming_Acids) Structurally, acids include binary acids (e.g., , HCl), which consist of bonded to a ; oxyacids (e.g., HNO₃, H₂SO₄), containing oxygen; and organic acids (e.g., , ), typically featuring a carboxyl group (-COOH) and prevalent in biological systems. Common examples also encompass (H₂CO₃) from dissolved CO₂ and (H₃PO₄) used in food additives. Beyond fundamental reactions like neutralization with bases to form salts and , acids play pivotal roles in industry and daily life./07:_Acids_and_Bases/7.08:_Acids_and_Bases_in_Industry_and_in_Daily_Life) , the most industrially produced chemical worldwide, is essential for manufacturing s (e.g., phosphate-based), refining, metal extraction, and production, with global output of approximately 261 million metric tons annually as of 2024. is vital for to remove rust and in adjustment for , while supports explosives and ./07:_Acids_and_Bases/7.08:_Acids_and_Bases_in_Industry_and_in_Daily_Life) In and , organic acids like citric and act as preservatives, flavor enhancers, and metabolic intermediates, underscoring acids' ubiquity in sustaining and enabling diverse technological advancements.

Definitions

Arrhenius Acids

The Arrhenius theory of acids, developed by Swedish chemist in his 1884 doctoral dissertation, provided the first definition by linking acidic properties to the electrolytic of substances in water. This groundbreaking work explained how acids behave through the production of charged particles, earning Arrhenius the in 1903 for his contributions to understanding electrolytes. According to the Arrhenius definition, an acid is a substance that increases the concentration of hydrogen ions (H⁺, often represented as the hydronium ion in notation) when dissolved in water. The general for an Arrhenius acid can be expressed as: \text{HA(aq)} \rightarrow \text{H}^+(\text{aq}) + \text{A}^-(\text{aq}) This process occurs fully for strong acids and partially for weak acids, leading to observable properties like sour taste, reaction with metals, and neutralization with bases. Representative examples include hydrochloric acid (HCl), a strong acid that completely dissociates in water to produce H⁺ and Cl⁻ ions, and acetic acid (CH₃COOH), a weak acid that partially dissociates to yield H⁺ and CH₃COO⁻ ions. These dissociations directly contribute to the increased H⁺ concentration characteristic of acidic solutions. However, the Arrhenius definition is restricted to aqueous solutions and fails to account for acidic behavior in non-aqueous solvents or for substances that exhibit acidity without producing ions, such as certain metal cations./07%3A_Acids_and_Bases/7.02%3A_Acids_and_Bases) Later theories, like Brønsted-Lowry, expanded on this by focusing on proton transfer in various media./07%3A_Acids_and_Bases/7.02%3A_Acids_and_Bases)

Brønsted–Lowry Acids

The Brønsted–Lowry theory defines an acid as a substance that donates a proton (H⁺ ) to another substance, termed a , which accepts the proton. This proton-transfer mechanism forms the core of acid-base reactions under this framework, expanding applicability beyond aqueous solutions to any medium where proton donation occurs. The theory was independently proposed in 1923 by Danish chemist and British chemist , providing a broader perspective than earlier models by emphasizing relative rather than specific production./Acids_and_Bases/Acid/Bronsted_Concept_of_Acids_and_Bases) In a Brønsted–Lowry acid-base reaction, the acid (HA) donates a proton to the base (B), yielding the conjugate base (A⁻) and conjugate acid (HB⁺). This process is reversible and represented by the general equilibrium: \ce{HA + B ⇌ A^- + HB^+} The conjugate acid-base pair consists of species differing by one proton, such as HA and A⁻, where the strength of the acid inversely relates to the strength of its conjugate base. This theory generalizes proton dissociation in water as a specific case of broader proton transfer./Acids_and_Bases/Acid/Bronsted_Concept_of_Acids_and_Bases) Certain substances exhibit , acting as both Brønsted–Lowry acids and depending on the reaction conditions, due to their ability to either donate or accept protons. The bicarbonate ion (HCO₃⁻) is a classic example of an amphoteric species. As a , it accepts a proton from : \ce{HCO3^- + H2O ⇌ H2CO3 + OH^-} As an acid, it donates a proton to : \ce{HCO3^- + H2O ⇌ CO3^{2-} + H3O^+} These reactions highlight bicarbonate's role in buffering systems, such as in biological fluids. Representative examples illustrate proton donation in non-aqueous or varied contexts. The ammonium ion (NH₄⁺) functions as a Brønsted–Lowry acid by transferring a proton to the hydroxide ion: \ce{NH4^+ + OH^- ⇌ NH3 + H2O} Here, NH₄⁺ is the acid, OH⁻ is the base, NH₃ is the conjugate base, and H₂O is the conjugate acid. Similarly, the hydrogen sulfate ion (HSO₄⁻) demonstrates amphoterism: it acts as an acid by donating a proton to water to form sulfate and hydronium ions (\ce{HSO4^- + H2O ⇌ SO4^{2-} + H3O^+}), or as a base by accepting a proton to form sulfuric acid (\ce{HSO4^- + H2O ⇌ H2SO4 + OH^-}), though the latter is less common.-plus-oh-(aq)-greater-nh3(aq)-plus-h2o(l)-is-nh4plus-a-bronsted-lowry-acid-a-bronsted-lowry-base-or-neither/) The Brønsted–Lowry framework ties directly to through the (K_a), which quantifies the position of proton donation for weak acids in : K_a = \frac{[A^-][\ce{H^+}]}{[HA]} A larger K_a indicates a stronger tendency to donate protons, reflecting greater within this theory. This expression underpins quantitative analysis of conjugate pair behaviors.

Lewis Acids

In 1923, proposed a general theory of acid-base reactions that defines a acid as any species capable of accepting an electron pair from a base to form a coordinate covalent bond, broadening the scope beyond proton transfer mechanisms. This definition emphasizes the role of electron deficiency in the acid, allowing it to complete its valence shell through donation from a base. The general reaction can be represented as: \text{A (acceptor)} + :\text{B (donor)} \rightarrow \text{A}–\text{B} where A is the Lewis acid and :B denotes the lone pair on the base. A classic example is the reaction between boron trifluoride (BF₃) and ammonia (NH₃), where the electron-deficient boron atom in BF₃ accepts the lone pair from nitrogen in NH₃ to form the adduct F₃B–NH₃. Another prominent application occurs in organic synthesis, such as Friedel-Crafts alkylation reactions, where aluminum chloride (AlCl₃) acts as a Lewis acid by coordinating with the halogen of an alkyl halide to generate a carbocation electrophile. Lewis acids play crucial roles in catalysis, particularly in biological systems where metal ions like Zn²⁺ function as electron-pair acceptors to activate substrates. For instance, in the enzyme , Zn²⁺ coordinates with water to facilitate its , enhancing the of . This definition extends to non-protonic species, including metal cations such as Fe³⁺, which accept electron pairs from ligands due to their high , and carbocations like (CH₃)₃C⁺, which seek stabilization through electron donation. Protonic acids represent a subset of Lewis acids, as the H⁺ itself acts as an electron-pair acceptor.

Properties

Dissociation and Equilibrium

In aqueous solutions, acids dissociate by ionizing to produce hydrogen ions (H⁺) and their conjugate bases, as originally conceptualized in the Arrhenius definition of acids. For a general acid HA, this process is represented as HA ⇌ H⁺ + A⁻, where the extent of determines whether the acid is strong or weak. Strong acids, such as (HCl), undergo complete dissociation in , meaning nearly 100% of the molecules ionize to form H⁺ and Cl⁻ ions, with no significant established. In contrast, weak acids partially ionize, resulting in an mixture of undissociated HA, H⁺, and A⁻. The for weak acid is quantified by the , K_a, defined as K_a = \frac{[H^+][A^-]}{[HA]}, where the concentrations are those at and activities are approximated by concentrations in dilute solutions. This constant reflects the position of the ; a smaller K_a indicates less and a weaker tendency to produce H⁺. For example, acetic acid (CH₃COOH) has K_a \approx 1.8 \times 10^{-5} at 25°C, meaning only a small ionizes in typical solutions. Pure also exhibits a related autoionization : \ce{H2O ⇌ H^+ + OH^-}, governed by the ion product constant K_w = [H^+][OH^-] = 1.0 \times 10^{-14} at 25°C, which establishes a baseline concentration of H⁺ and OH⁻ ions even in neutral conditions. The concentration from weak can be approximated for initial calculations when the concentration C is much greater than the dissociated amount, yielding [H^+] \approx \sqrt{K_a \cdot C}; this simplification assumes [H⁺] = [A⁻] and negligible change in [HA] from the initial value, valid for moderately dilute solutions where is less than 5%. External factors influence this per Le Châtelier's principle: dilution decreases concentrations of all species, shifting the toward greater to restore , thereby increasing the percent . changes alter K_a itself, as is typically endothermic; higher temperatures favor the forward reaction, increasing K_a and [H⁺].

Acid Strength

Acid strength quantifies the extent to which an donates a proton (H⁺) in , primarily measured by the K_a, defined for the equilibrium \ce{HA ⇌ H+ + A-} as K_a = \frac{[\ce{H+}][\ce{A-}]}{[\ce{HA}]}. The value, given by \mathrm{p}K_a = -\log_{10} K_a, provides a convenient scale where a lower corresponds to a stronger due to greater proton donation tendency. In aqueous solutions, acids are classified as strong if they fully dissociate ( < 0), such as hydrochloric acid (HCl, ≈ -7), which exists entirely as \ce{H+} and \ce{Cl-}. Weak acids, with > 0, partially dissociate; for example, (HF, = 3.17) ionizes only to a limited extent due to the strong H–F bond and poor stabilization of the \ce{F-} conjugate base. Several factors influence by affecting the stability of the conjugate or the ease of proton release. Bond strength plays a key role: weaker H–A bonds favor stronger acids, as seen in the hydrogen halides where (strong H–F bond) is much weaker than (weaker H–I bond, ≈ -9). Inductive effects from electron-withdrawing groups, such as on a carbon chain, stabilize the negative charge on the conjugate by withdrawing electron density, increasing acidity (e.g., is stronger than acetic acid). stabilization is particularly effective, delocalizing the conjugate charge over multiple atoms, as in carboxylic acids where the ion's charge spreads across two oxygen atoms, making them more acidic than alcohols. For polyprotic acids, which can donate multiple protons, successive values increase because each subsequent conjugate is less willing to lose a proton; for (H₂SO₄), pKa₁ ≈ -3 (strong first to \ce{HSO4-}) while pKa₂ ≈ 2 (weaker second to \ce{SO4^2-}). In non-aqueous solvents, which are less basic than , acid strengths can differ markedly due to reduced leveling effects; for instance, in acetic acid, the order reverses from aqueous behavior, with HCl weaker than HBr (and strongest) as the solvent's lower proton-accepting ability allows differentiation based on inherent bond polarities and conjugate . Superacids, developed in the mid-20th century, exceed the strength of concentrated (H₀ ≈ -12, where H₀ is the extending for highly acidic media); the "magic acid" system of (HSO₃F) with (SbF₅) achieves H₀ < -20, enabling protonation of weak bases like hydrocarbons.

Nomenclature

The nomenclature of acids has evolved from early trivial names based on sensory properties or origins, such as "vinegar" for acetic acid, to systematic conventions established in the late 18th and 19th centuries by chemists like and , who emphasized compositional elements, with the International Union of Pure and Applied Chemistry (IUPAC) formalizing rules in the 20th century to promote precision and universality. This shift addressed ambiguities in pre-modern naming, where acids were often described by their sources or effects rather than structure, leading to the adoption of substitutive and additive methods that reflect molecular composition. For inorganic acids, binary acids—those composed of hydrogen and a single nonmetal—are named using the prefix "hydro-" followed by the stem of the nonmetal and the suffix "-ic acid," as in for HCl. Oxyacids, which include oxygen, follow traditional naming based on the corresponding oxyanion: the suffix "-ic acid" denotes the anion with more oxygen or higher oxidation state (e.g., for H₂SO₄, derived from sulfate), while "-ous acid" indicates fewer oxygen atoms or lower oxidation state (e.g., for H₂SO₃, from sulfite); additional prefixes like "per-" (highest oxygen, as in , HClO₄) and "hypo-" (lowest, as in , HClO) refine these distinctions. IUPAC also endorses additive nomenclature for clarity, listing ligands alphabetically around the central atom (e.g., tetraoxidosulfuric acid for H₂SO₄), though traditional names remain widely retained. Organic acids employ substitutive nomenclature, prioritizing the principal functional group as the suffix. Carboxylic acids, featuring the -COOH group, are named by identifying the longest carbon chain including the carboxyl carbon and appending "-oic acid," with the chain numbered from the carboxyl group; for instance, CH₃COOH is (preferred IUPAC name, or PIN), though the retained common name is acceptable in general use. Sulfonic acids, with the -SO₃H group, similarly use the suffix "-sulfonic acid" attached to the parent hydrocarbon chain or ring, such as for CH₃SO₃H or for C₆H₅SO₃H. Polyprotic acids, capable of donating multiple protons, extend these rules to their anions through "hydrogen" prefixes indicating remaining ionizable hydrogens, as seen in dihydrogen phosphate for H₂PO₄⁻ (from phosphoric acid, H₃PO₄) or hydrogen phosphate for HPO₄²⁻; this convention treats partial deprotonation systematically while aligning with oxyanion naming patterns. Overall, IUPAC distinguishes preferred systematic names (e.g., ethanoic acid) from retained trivial ones (e.g., acetic acid) to balance innovation with established terminology, ensuring nomenclature supports both educational and practical applications without implying acid strength differences solely through naming conventions.

Chemical Behavior

Monoprotic and Polyprotic Acids

Monoprotic acids are those capable of donating a single proton (H⁺) per molecule in aqueous solution, resulting in a single acid dissociation equilibrium characterized by one acid dissociation constant, K<sub>a</sub>. Representative examples include hydrochloric acid (HCl), a strong monoprotic acid that fully dissociates, and acetic acid (CH₃COOH), a weak monoprotic acid with K<sub>a</sub> ≈ 1.8 × 10<sup>−5</sup>. In contrast, polyprotic acids can donate more than one proton per through successive steps. Diprotic acids, such as (H₂SO₄), release two protons, while triprotic acids like (H₃PO₄) release three. For a generic diprotic acid denoted as H₂A, the stepwise equilibria are: \ce{H2A ⇌ H+ + HA-} \quad K_{a1} = \frac{[\ce{H+}][\ce{HA-}]}{[\ce{H2A}]} \ce{HA- ⇌ H+ + A^{2-}} \quad K_{a2} = \frac{[\ce{H+}][\ce{A^{2-}}]}{[\ce{HA-}]} Triprotic acids follow analogous stepwise processes for each proton. The acid dissociation constants for successive steps decrease markedly (K<sub>a1</sub> ≫ K<sub>a2</sub> ≫ K<sub>a3</sub>), so pK<sub>a1</sub> < pK<sub>a2</sub> < pK<sub>a3</sub>; this occurs because each subsequent proton is removed from an increasingly negatively charged anion, which experiences greater electrostatic repulsion and holds the proton more tightly. For , for instance, K<sub>a1</sub> = 1.0 × 10<sup>3</sup> while K<sub>a2</sub> = 1.2 × 10<sup>−2</sup>, and for , K<sub>a1</sub> = 7.1 × 10<sup>−3</sup>, K<sub>a2</sub> = 6.3 × 10<sup>−8</sup>, and K<sub>a3</sub> = 4.2 × 10<sup>−13</sup>. The relative concentrations of the various from a polyprotic in depend on the , with predominance shifting across the pK<sub>a</sub> values. For , the dihydrogen phosphate species (H₂PO₄⁻) predominates in solutions with pH between approximately 2 and 7, the range spanning its first and second pK<sub>a</sub> values (2.1 and 7.2). A key biological example is (H₂CO₃), a diprotic acid formed from CO₂ and H₂O in , where the ion (HCO₃⁻) is the dominant species at physiological pH (around 7.4), contributing to the that stabilizes pH between 7.35 and 7.45.

Neutralization Reactions

Neutralization reactions occur when an reacts with a to form a and , involving the combination of ions (H⁺) from the acid and hydroxide ions (OH⁻) from the base to produce water. The general for such a reaction is represented as HA + BOH → BA + H₂O, where HA is the acid, BOH is the base, BA is the salt, and H₂O is . These reactions are typically exothermic, particularly for pairs of strong acids and strong bases, where the of neutralization is approximately -57 kJ/ at 25°C, reflecting the formation of from fully dissociated ions. The stoichiometry of neutralization reactions depends on the number of ionizable protons in the acid and hydroxide groups in the . For monoprotic acids, such as (HCl), the reaction follows a 1:1 molar ratio with a monohydroxy like (NaOH): HCl + NaOH → NaCl + H₂O. Polyprotic acids require multiple equivalents of ; for example, (H₂SO₄), a diprotic acid, reacts with two moles of NaOH: H₂SO₄ + 2NaOH → Na₂SO₄ + 2H₂O. The salts formed in neutralization reactions derive their properties from the strengths of the parent acid and base, specifically their conjugate pairs. Salts from strong acids and strong bases, such as NaCl from HCl and NaOH, are neutral with a pH of 7 in aqueous solution. In contrast, salts from strong acids and weak bases, like (NH₄Cl) from HCl and NH₃, are acidic (pH < 7) due to the of the conjugate acid of the . Representative examples illustrate these principles. The reaction of HCl with NaOH produces NaCl and water, a classic strong acid-strong neutralization. Historically, neutralization has been applied in soap production through , where fatty acids from animal fats or vegetable oils react with (NaOH) to form salts and .

Weak Acid–Weak Base Equilibria

The reaction between a weak acid (HA) and a (B) proceeds incompletely, establishing an described by the equation: HA + B \rightleftharpoons A^- + HB^+ The K for this reaction is given by K = \frac{K_a K_b}{K_w}, where K_a is the of HA, K_b is the base dissociation constant of B, and K_w is the ion product of . This relationship arises because the forward reaction involves proton transfer from HA to B, with the position of favoring the side containing the weaker and the weaker (i.e., the side where the pK_a of the is higher and the pK_b of the is lower). Such equilibria form the basis of buffer solutions, which are mixtures of a weak acid and its conjugate base (or a and its conjugate acid) that resist changes in upon addition of small amounts of strong acid or base. For instance, a can be prepared by mixing acetic acid (CH₃COOH) with its conjugate base (CH₃COO⁻), maintaining a stable through the reversible proton exchange. The of an acidic buffer is calculated using the Henderson-Hasselbalch equation: \text{pH} = \text{p}K_a + \log_{10} \frac{[A^-]}{[HA]} This equation, derived from the K_a expression, allows prediction of buffer pH based on the ratio of conjugate to acid concentrations, assuming activity coefficients near unity. A practical example is the mixture of acetic acid and (NH₃), where the equilibrium produces and ions (NH₄⁺), creating a system effective around neutral pH. In biological contexts, the —comprising (H₂CO₃) and (HCO₃⁻)—maintains blood pH near 7.4 by buffering metabolic acids and CO₂-derived protons. Buffers are most effective within approximately pK_a ± 1 unit, where the concentrations of the acid and conjugate are within a 10:1 , maximizing resistance to pH shifts. Beyond this range, buffering capacity diminishes significantly.

Measurement

Titration

is an analytical technique used to determine the concentration of an by gradually adding a of known concentration and monitoring the of the . The typically involves placing a known volume of the acid in an and titrating it with the from a , recording the after each incremental addition using a until the is reached. This method relies on the neutralization reaction between the and , allowing for precise quantification of the acid's molarity. The resulting titration curve plots against the volume of added, providing a visual representation of the reaction progress. For a strong titrated with a strong , the curve is sigmoidal, characterized by a low initial , a gradual increase, and a sharp rise near the due to excess . In contrast, the curve for a weak titrated with a strong features gentler slopes and plateaus, reflecting the buffering capacity of the system; the strength influences the curve's shape, with weaker acids producing less pronounced changes in . regions appear midway to the , where the approximates the pK_a of the , as the contains roughly equal concentrations of the and its conjugate , resisting changes. The occurs when the moles of acid equal the moles of base added for monoprotic acids, marking complete neutralization. For strong acid-strong base titrations, this point is at 7, as the resulting is . In weak acid-strong base titrations, the exceeds 7, typically around 8-9, because the conjugate base of the weak acid hydrolyzes to produce excess OH^-. For polyprotic acids such as H_2SO_4, the displays two distinct corresponding to each proton donation, with inflection breaks at a low around 2–3 (first, forming HSO_4^-, determined by the pK_a of HSO_4^- ≈ 2) and 9 (second, forming SO_4^{2-}). To calculate the volume of required to reach the for a monoprotic , use the : V_{\text{eq}} = \frac{C_{\text{acid}} \times V_{\text{acid}}}{C_{\text{base}}} where C_{\text{acid}} and C_{\text{base}} are the concentrations, and V_{\text{acid}} is the initial volume of ; this assumes a 1:1 and follows from the equality of moles at equivalence. For polyprotic s, the is adjusted by the number of equivalents, but remains based on stoichiometric balance.

pH and Indicators

The concept of pH was introduced in 1909 by Danish biochemist Søren Peder Lauritz Sørensen while working at the Carlsberg Laboratory, providing a practical scale to quantify the acidity of solutions based on hydrogen ion activity. Sørensen's innovation addressed the need for a logarithmic measure during biochemical research on enzyme activity in brewing. pH is formally defined by the International Union of Pure and Applied Chemistry (IUPAC) as the negative base-10 logarithm of the activity of ions in :
\mathrm{pH} = -\log_{10} a(\mathrm{H^+})
where a(\mathrm{H^+}) represents the effective concentration accounting for non-ideal behavior. In dilute aqueous s at 25°C, this approximates to \mathrm{pH} = -\log_{10} [\mathrm{H^+}], with the scale conventionally spanning 0 (highly acidic, [H⁺] = 1 M) to 14 (highly basic, [H⁺] = 10⁻¹⁴ M), and pH 7 indicating neutrality due to water's K_w = 10^{-14}. Values below 0 or above 14 occur in concentrated strong acids or bases, but the 0–14 range applies to most aqueous systems under standard conditions.
Acid-base indicators are typically weak organic acids or bases that undergo a , resulting in a visible color shift near their pKₐ value, allowing qualitative assessment. The color transition occurs over a narrow interval (usually 1–2 units) where the indicator's protonated and deprotonated forms coexist in comparable amounts. A common example is , a weak acid with pKₐ ≈ 9.3, which remains colorless below 8.2 in its protonated form and turns pink above 10.0 in its deprotonated form due to extended conjugation in the basic state. For precise quantitative measurement, glass pH electrodes are widely used, consisting of a thin, ion-selective glass membrane that develops a potential proportional to the external [H⁺] relative to an internal reference solution. This potential adheres to the Nernst equation for the hydrogen ion half-cell:
E = E_0 - 0.059 \log_{10} a(\mathrm{H^+})
at 25°C, corresponding to a change of 59 mV per pH unit, with the electrode potential increasing by 59 mV as the pH decreases by one unit.
Despite their utility, pH measurements face limitations in non-aqueous solvents, where the absence of water alters ion activity and hydration, rendering standard scales and glass electrodes unreliable without solvent-specific calibrations or alternative probes. Universal indicators, blends of multiple dyes such as methyl red, bromothymol blue, and thymol blue, overcome some precision needs by displaying a continuous color gradient across pH 1–14 (red for acidic, green for neutral, purple for basic), facilitating broad-range visual approximations without equipment.

Types of Acids

Mineral Acids

Mineral acids, also known as inorganic acids, are water-soluble acids derived from inorganic minerals and lack carbon in their molecular structure. They encompass , such as (\ce{HCl}), and oxyacids, including (\ce{H2SO4}), (\ce{HNO3}), and (\ce{H3PO4}). These acids are typically strong, meaning they fully dissociate in water to release ions, contributing to their high reactivity and corrosive nature. Sulfuric acid is one of the most industrially significant mineral acids, produced via the , which involves the of (\ce{SO2}) to (\ce{SO3}), followed by absorption in . Global of reached approximately 200 million metric tons per year in the 2020s, underscoring its role as a cornerstone of chemical . is synthesized through the , where is oxidized over a platinum-rhodium catalyst to (\ce{NO}), then further oxidized and absorbed in to form the acid. Annual global output for was around 58 million metric tons in 2024. Hydrochloric acid is commonly produced by reacting () with , generating gas that is then dissolved in . Its global volume stood at about 15 million metric tons in 2024. These acids exhibit high strength and corrosiveness, capable of rapidly degrading metals, tissues, and materials upon contact due to their ability to donate protons and, in some cases, act as oxidizing agents. For instance, sulfuric and nitric acids are among the strongest mineral acids, with values indicating near-complete , while is similarly potent but non-oxidizing. , however, is relatively weaker, with multiple dissociation steps yielding a lower acidity (pKa1 ≈ 2.14), making it less corrosive and suitable for applications like production where milder reactivity is preferred. A notable mixture involving mineral acids is , composed of concentrated and in a 3:1 ratio, which generates nascent and to dissolve noble metals like that resist individual acids.

Organic Acids

Organic acids are carbon-containing compounds that exhibit acidic properties, primarily through the presence of functional groups capable of donating protons. The most prevalent class is carboxylic acids, characterized by the general formula \ce{RCOOH}, where R is an organic substituent, such as in (\ce{HCOOH}), the simplest member. Sulfonic acids, with the formula \ce{RSO3H}, represent another key class, exemplified by (\ce{CH3SO3H}); these are notably stronger acids due to the electron-withdrawing sulfonyl group. Other classes include enols, which feature a hydroxyl group attached to a carbon-carbon (\ce{C=C-OH}), though they are less stable and often exist in tautomeric equilibrium with keto forms. In IUPAC , carboxylic acids are named by replacing the -e ending of the parent with -oic acid, such as ethanoic acid for \ce{CH3COOH}. Compared to acids, most acids are weaker, with carboxylic acids typically having pKa values around 4-5, indicating partial dissociation in ; for instance, acetic acid (\ce{CH3COOH}), found in , has a pKa of 4.76./Carboxylic_Acids/Nomenclature_of_Carboxylic_Acids) in is generally high for short-chain variants due to hydrogen bonding, but decreases with longer hydrophobic chains; benzoic acid (\ce{C6H5COOH}), used as a preservative, has a pKa of 4.20 and limited in pure but improved in conditions. In contrast, sulfonic acids like exhibit pKa values around -1.9, approaching the strength of acids. Many organic acids play vital roles in biological systems, often derived as intermediates in the Krebs cycle (also known as the ), a central in aerobic organisms that generates energy through the oxidation of . Key examples include , which initiates the cycle, and succinic and malic acids, which facilitate and . Halogenated organic acids, such as (\ce{CF3COOH}), deviate from the typical weakness of carboxylic acids; its pKa of 0.23 results from the high of the three atoms, which stabilize the conjugate base by inductive withdrawal of .

Specialized Acids

Superacids represent a class of exceptionally strong acids that exceed the acidity of pure , defined by a H_0 value less than -12. These media enable the of notoriously unreactive hydrocarbons, such as alkanes, which was first demonstrated in the 1960s by George A. Olah using - mixtures. A prominent example is , a 1:1 molar mixture of (FSO₃H) and (SbF₅), which achieves an H_0 of approximately -19.2 and facilitates the formation of alkanium ions like the ethyl cation from . Olah's pioneering work on these systems, including the direct observation of protonated alkanes via , earned him the 1994 for contributions to chemistry. Vinylogous acids feature extended conjugation through groups, allowing the acidic proton to be delocalized over a longer chain, which modulates their reactivity compared to simple analogs. Ascorbic acid () exemplifies this, functioning as a vinylogous where the hydroxyl group's acidity is enhanced by involving the distant carbonyl, resulting in a pKa of about 4.1 for the enolic proton. This enables ascorbic acid's role as an , with the conjugated system facilitating electron transfer. Nucleic acids, such as DNA and RNA, incorporate phosphoric acid derivatives in their sugar-phosphate backbones, forming phosphodiester linkages that confer polyanionic character. The phosphate groups in these biopolymers exhibit two relevant pKa values: approximately 1 for the primary dissociation (yielding the monoanion) and around 6 for the secondary dissociation (to the dianion), as observed in nucleotide monophosphates like AMP. These pKa values ensure that the backbone remains negatively charged at physiological pH, stabilizing the helical structures through electrostatic repulsion and interactions with counterions. Certain acids operate exclusively under the Lewis definition, accepting electron pairs without proton donation, and boron-based compounds like , B(OH)₃, illustrate this behavior. acts as a weak Lewis acid by coordinating to Lewis bases such as or via its electron-deficient atom, forming tetrahedral adducts, though it shows minimal Brønsted acidity with a exceeding 9. This property underpins its applications in complexation chemistry. In the 2020s, have found emerging roles in , particularly for degrading persistent fluorinated pollutants. For instance, a novel developed in 2023 enables the conversion of non-biodegradable perfluorocarbons, akin to Teflon, into harmless ions under mild conditions, addressing environmental contamination from fluorochemicals. Similarly, halogen-substituted silicon-based Lewis , reported in 2025, offer potential for sustainable by promoting selective C-H activations without hazardous solvents.

Applications and Roles

Industrial and Catalytic Uses

Sulfuric acid (H₂SO₄) is the most widely produced industrial chemical, with approximately 55% of global output used in the manufacture of phosphate fertilizers such as and . Another significant application is in lead-acid batteries, where it serves as the to facilitate electrochemical reactions for in vehicles and backup power systems. Hydrochloric acid (HCl) plays a key role in the production of (PVC), acting as a precursor in the synthesis of monomer through processes like the balanced salt process. In catalysis, acids enable essential reactions in organic synthesis and petrochemical processing. The Fischer esterification, developed by Emil Fischer and Arthur Speier, involves the acid-catalyzed reaction of carboxylic acids with alcohols to form esters, a foundational method for producing biodiesel and pharmaceutical intermediates since its introduction in 1895. Zeolites, as solid acid catalysts, are extensively used in fluid catalytic cracking (FCC) units in petroleum refineries to break down heavy hydrocarbons into gasoline and diesel fractions, improving yield and selectivity through their microporous structure. Notable examples of acid applications include (HNO₃) in the nitration of organic compounds to produce explosives such as and , where it acts as both an oxidizing and nitrating agent. (H₃PO₄) is industrially employed in the beverage sector, particularly for production, where it provides acidity and stabilizes the formulation during large-scale manufacturing. In the , there has been a notable shift toward sustainable acid production, with bio-based carboxylic acids like succinic and gaining traction through microbial processes to replace petroleum-derived counterparts, driven by advancements in for reduced carbon footprints.

Food and Physiological Roles

Acids play essential roles in food and human , contributing to , preservation, and digestive processes. In , , a weak , is naturally abundant in fruits such as lemons and limes, where it imparts a characteristic tart flavor and contributes to the low of fruit juices, typically ranging from 2 to 3. This acidity not only enhances taste but also aids in the of certain minerals. Similarly, acetic acid in , at concentrations around 5%, is widely used for vegetables, where it lowers the pH to create an environment that preserves flavor and texture while extending shelf life. , produced through bacterial of sugars in foods like , , and , adds a tangy profile and supports content, promoting gut health. In , the low created by these acids inhibits bacterial growth and spoilage. For instance, maintaining a of 4.6 or lower in acidic foods prevents the of harmful bacterial spores, such as those from , ensuring safety without high-heat processing. Ascorbic acid, known as , serves as an in fruits and , preventing oxidation that leads to browning and loss; it is particularly effective in preserving the freshness of juices and cut produce by scavenging free radicals. Physiologically, hydrochloric acid (HCl) in the maintains a highly acidic environment, with concentrations around 0.1 M and a of 1 to 2, which is crucial for . This acidity activates pepsinogen into active , the primary enzyme for breaking down proteins into peptides, and kills ingested pathogens by denaturing their proteins and disrupting cellular functions when the drops below 3. The of sour taste arises from hydrogen ions (H⁺) stimulating specific proton-selective channels, such as OTOP1, in cells on the , signaling acidity to the brain. Deficiencies in stomach acid, known as hypochlorhydria, impair these functions and are linked to of nutrients, including proteins, vitamins (such as B12), and minerals like iron and calcium, potentially leading to deficiencies and digestive disorders.

Biological and Environmental Significance

In biological systems, acids play essential roles in and function. such as , which contains an acidic , contribute to the overall charge and folding of proteins, often acting as proton donors in active sites to facilitate . In uricotelic non-mammalian vertebrates such as and reptiles, serves as the primary nitrogenous waste product, allowing birds to excrete ammonia-derived waste in a semi-solid form that conserves and minimizes during flight and arid adaptations. Environmentally, acid rain—precipitation with a pH below 5.6 resulting from atmospheric reactions involving sulfur dioxide (SO₂) and nitrogen oxides (NOₓ)—has significantly impacted ecosystems since gaining widespread scientific awareness in the 1970s. These acids leach essential nutrients like calcium and magnesium from forest soils while mobilizing toxic aluminum, leading to reduced tree growth, defoliation, and in affected regions such as the and . Ocean acidification represents another critical environmental threat, where increased atmospheric CO₂ dissolves in to form (H₂CO₃), causing surface to drop by approximately 0.1 units since the . This shift, with research intensifying in the post-2000s era, impairs shell formation in organisms like and corals by reducing carbonate ion availability, disrupting food webs and coastal economies. The evolutionary origins of may trace back to acidic conditions in a , where prebiotic chemistry in warm, acidic pools facilitated the synthesis of organic molecules like , as hypothesized in early experiments simulating Earth's early atmosphere. Remediation efforts for acidified ecosystems often involve applying () to neutralize acidity in soils and waters, restoring and nutrient balance in forests and lakes affected by . Nucleic acids, such as and , exemplify specialized biological acids that store and transmit genetic information essential for all forms.

Safety and Health Impacts

Household and Laboratory Handling

Acids commonly found in households include (H₂SO₄) used in fluid and some drain cleaners, and (HCl) present in toilet bowl cleaners and pool maintenance products. In household settings, acids should be stored in original containers or compatible or bottles, kept in a cool, dry, well-ventilated area away from bases, oxidizers, and foodstuffs, with secondary containment like trays to catch spills. Before disposal, dilute acids with water and neutralize if possible, following local regulations for . When handling household acids, wear chemical-resistant gloves, safety goggles, and protective clothing to prevent and , as acids can cause corrosive burns due to their proton-donating properties. For spills, immediately neutralize with a mild base like baking soda () to form a and water, then absorb and dispose properly while ventilating the area. In laboratories, acids are stored in dedicated corrosive-resistant cabinets that are ventilated and equipped with spill containment trays, using or containers compatible with specific acids to prevent reactions. Laboratory handling requires including or gloves, chemical splash goggles, face shields, lab coats, and closed-toe shoes; respiratory protection may be needed for volatile acids. Procedures for volatile acids like (HNO₃) mandate use within a to contain fumes and vapors, ensuring the sash is lowered and airflow is verified before starting work. When diluting concentrated acids, always add the acid to slowly while stirring, never the reverse, to avoid exothermic splattering and potential explosions. For in both settings, flush skin or eyes exposed to acids with copious lukewarm for at least 15-20 minutes, removing contaminated clothing, and seek immediate medical attention. In cases of acid ingestion, do not induce ; instead, dilute by giving small amounts of or if the person is conscious and able to swallow, then contact poison control or emergency services immediately.

Acidity in Human Physiology

In human physiology, acidity plays a critical role in maintaining the body's balance, particularly in the , where the normal is 7.35 to 7.45. This narrow range is primarily regulated by the , which involves the equilibrium between (H₂CO₃) and ions (HCO₃⁻), helping to neutralize excess acids or bases produced during . Deviations below pH 7.35 result in , a condition that can impair function and cellular processes; for instance, , often seen in uncontrolled , arises from the accumulation of that lower pH. The lungs and kidneys are essential organs for pH regulation, controlling the levels of (CO₂) and (HCO₃⁻). The adjusts pH rapidly by altering rates to expel CO₂, a volatile acid formed from dissociation, while the kidneys provide longer-term control by reabsorbing or excreting HCO₃⁻ and excreting hydrogen ions over hours to days. Imbalances in these mechanisms contribute to acid-base disorders; for example, in like , compensatory reduces CO₂ to mitigate the pH drop. Acids also have direct medical applications in . Aspirin, or acetylsalicylic acid, serves as an by inhibiting enzymes, reducing synthesis that sensitizes receptors, thereby alleviating and . In cases of , a condition characterized by insufficient (HCl) secretion in the , HCl therapy can be administered to restore gastric acidity, aiding and . Diseases related to acidity imbalances highlight its physiological impact. Gastroesophageal reflux disease (GERD) occurs when excess stomach acid refluxes into the esophagus, causing irritation, inflammation, and symptoms like heartburn due to the corrosive nature of gastric HCl. Gout, conversely, stems from hyperuricemia leading to the deposition of uric acid crystals in joints, triggering acute inflammatory arthritis. In the 2020s, proton pump inhibitors (PPIs), which suppress acid production, have seen widespread use for managing conditions like GERD.

Environmental Effects

Acid rain primarily forms when emissions of sulfur dioxide (SO₂) and nitrogen oxides (NOₓ, including NO₂) from sources such as power plants and vehicles react with , oxygen, and other chemicals in the atmosphere to produce sulfuric and nitric acids, which then fall as precipitation. These acids, derived from sources, lower the of to levels as low as 4.2-5.0, far below the typical 5.6 of unpolluted . In soils, this acidity leaches essential minerals and nutrients while mobilizing toxic aluminum, which binds to roots and impairs growth; in water bodies, it increases aluminum concentrations that disrupt and eggs, leading to die-offs in lakes and streams where drops below 5.0. Aquatic ecosystems suffer cascading effects, with sensitive species like and populations declining sharply in acidified waters. Regulatory efforts have significantly curbed these impacts. The U.S. Clean Air Act, initially enacted in with key amendments in 1990 targeting , has reduced SO₂ emissions from power plants by over 90% since 1990 through cap-and-trade programs and technologies like , which capture up to 95% of SO₂ before release. Internationally, the 1999 Gothenburg Protocol under the UNECE Convention on Long-Range Transboundary Air Pollution mandates emission reductions for SO₂ (up to 63% by 2010 from 1990 levels in participating countries), NOₓ (41%), and other pollutants to combat acidification across and . These measures, including widespread adoption of in coal-fired plants, have led to measurable recovery in affected ecosystems. A notable case is Germany's , where in the damaged nearly 50% of trees through needle loss, , and bark erosion, sparking widespread "Waldsterben" () alarm and prompting stricter emission controls. Post-regulation, SO₂ emissions in fell by about 90% from levels, enabling recovery; by the 2000s, tree health improved significantly, with reduced crown damage and stabilization observed in monitoring plots. Beyond atmospheric acids, rising atmospheric CO₂ drives by forming upon dissolution in , with models projecting a global surface decline of 0.3 to 0.4 units by 2100 under high-emission scenarios, threatening and calcification. As of 2025, concerns have arisen over a potential return of in the U.S. due to rollbacks in emission regulations, though overall reductions remain significant.