acid–base reaction

acid–base reaction

Introduction

      a type of chemical process typified by the exchange of one or more hydrogen ions, H+, between species that may be neutral (molecules, such as water, H2O; or acetic acid, CH3CO2H) or electrically charged (ions, such as ammonium, NH4+; hydroxide, OH; or carbonate, CO2/3). It also includes analogous behaviour of molecules and ions that are acidic (acid) but do not donate hydrogen ions (aluminum chloride, AlCl3, and the silver ion AG+).

      Acids are chemical compounds that show, in water solution, a sharp taste, a corrosive action on metals, and the ability to turn certain blue vegetable dyes red. Bases (base) are chemical compounds that, in solution, are soapy to the touch and turn red vegetable dyes blue. When mixed, acids and bases neutralize one another and produce salts (salt), substances with a salty taste and none of the characteristic properties of either acids or bases.

      The idea that some substances are acids, whereas others are bases, is almost as old as chemistry; and the terms acid, base, and salt occur very early in the writings of the medieval alchemists. Acids were probably the first of these to be recognized, apparently because of their sour taste. The English word acid, the French acide, the German Säure, and the Russian kislota are all derived from words meaning sour (Latin acidus, German sauer, Old Norse sūur, and Russian kisly). Other properties associated at an early date with acids were their solvent, or corrosive, action; their effect on vegetable dyes; and the effervescence resulting when they were applied to chalk (production of bubbles of carbon dioxide gas). Bases (or alkalies) were characterized mainly by their ability to neutralize acids and form salts, the latter being typified rather loosely as crystalline substances soluble in water and having a saline taste.

      In spite of their imprecise nature, these ideas served to correlate a considerable range of qualitative observations; and many of the commonest chemical materials that early chemists encountered could be classified as acids (hydrochloric, sulfuric, nitric, and carbonic acids), bases (soda, potash, lime, ammonia), or salts (common salt, sal ammoniac, saltpetre, alum, borax). The absence of any apparent physical basis for the phenomena concerned made it difficult to make quantitative progress in understanding acid–base behaviour, but the ability of a fixed quantity of acid to neutralize a fixed quantity of base was one of the earliest examples of chemical equivalence: the idea that a certain measure of one substance is in some chemical sense equal to a different amount of a second substance. In addition, it was found quite early that one acid could be displaced from a salt with another acid, and this made it possible to arrange acids in an approximate order of strength. It also soon became clear that many of these displacements could take place in either direction according to experimental conditions. This phenomenon suggested that acid–base reactions are reversible (reversibility)—that is, that the products of the reaction can interact to regenerate the starting material. It also introduced the concept of equilibrium (chemical equilibrium) to acid–base chemistry: this concept states that reversible chemical reactions reach a point of balance, or equilibrium, at which the starting materials and the products are each regenerated by one of the two reactions as rapidly as they are consumed by the other.

      Apart from their theoretical interest, acids and bases play a large part in industrial chemistry and in everyday life. sulfuric acid and sodium hydroxide are among the products manufactured in largest amounts by the chemical industry, and a large percentage of chemical processes involve acids or bases as reactants or as catalysts. Almost every biological chemical process is closely bound up with acid–base equilibria in the cell, or in the organism as a whole, and the acidity or alkalinity of the soil and water are of great importance for the plants or animals living in them. Both the ideas and the terminology of acid–base chemistry have permeated daily life, and the term salt is especially common.

Theoretical definitions of acids and bases

Hydrogen and hydroxide ions
      The first attempt at a theoretical interpretation of acid behaviour was made by Antoine-Laurent Lavoisier (Lavoisier, Antoine-Laurent) at the end of the 18th century. Lavoisier supposed that all acids must contain oxygen, and this idea was incorporated in the names used for this element in the various languages; the English oxygen, from the Greek oxys (sour) and genna (production); the German Sauerstoff, literally acid material; and the Russian kislorod, from kislota (acid). Following the discovery that hydrochloric acid contained no oxygen, Sir Humphry Davy (Davy, Sir Humphry, Baronet) in about 1815 first recognized that the key element in acids was hydrogen. Not all substances that contain hydrogen, however, are acids, and the first really satisfactory definition of an acid was given by Justus von Liebig (Liebig, Justus, Freiherr von) of Germany in 1838. According to Liebig, an acid is a compound containing hydrogen in a form in which it can be replaced by a metal. This definition held the field for about 50 years and is still considered essentially correct, though somewhat outmoded. At the time of Liebig's proposal, bases were still regarded solely as substances that neutralized acids with the production of salts, and nothing was known about the constitutional features of bases that enabled them to do this.

      The whole subject of acid–base chemistry acquired a new look and a quantitative aspect with the advent of the electrolytic dissociation theory propounded by Wilhelm Ostwald (Ostwald, Wilhelm) and Svante August Arrhenius (Arrhenius, Svante August) (both Nobel laureates) in the 1880s. The principal feature of this theory is that certain compounds, called electrolytes, dissociate in solution to give ions (ion). With the development of this theory it was realized that acids are merely those hydrogen compounds that give rise to hydrogen ions (hydrogen ion) (H+) in aqueous solution. It was also realized at that time that there is a correspondence between the degree of acidity of a solution (as shown by effects on vegetable dyes and other properties) and the concentration of hydrogen ions in the solution. Correspondingly, basic (or alkaline) properties could then be associated with the presence of hydroxide ions (OH) in aqueous solution, and the neutralization of acids by bases could be explained in terms of the reaction of these two ions to give the neutral molecule water (H+ + OH → H2O). This led naturally to the simple definition that acids and bases are substances that give rise, respectively, to hydrogen and hydroxide ions in aqueous solution. This definition was generally accepted for the next 30 or 40 years. In purely qualitative terms, it does not offer many advantages over Liebig's definition of acids, but it does provide a satisfactory definition for bases.

      Nevertheless, there is a great advantage in the definition of acids and bases in terms of hydrogen and hydroxide ions, and this advantage lies in its quantitative aspects. Because the concentrations of hydrogen and hydroxide ions in solution can be measured, notably by determining the electrical conductivity of the solution (its ability to carry an electrical current), a quantitative measure of the acidity or alkalinity of the solution is provided. Moreover, the equations developed to express the relationships between the various components of reversible reactions can be applied to acid and base dissociations to give definite values, called dissociation constants. These constants can be used to characterize the relative strengths (degrees of dissociation) of acids and bases and, for this reason, supersede earlier semiquantitative estimates of acid or base strength. As a result of this approach, a satisfactory quantitative description was given at an early date of a large mass of experimental observations, a description that remains essentially unaffected by later developments in definitions of acid–base reactions.

      The success of these quantitative developments, however, unfortunately helped to conceal some ambiguities and logical inconsistencies in the qualitative definitions of acids and bases in terms of the production of hydrogen and hydroxide ions, respectively. For example, it was not clear whether a substance like anhydrous hydrogen chloride, which would not conduct electricity, should be regarded as an acid or whether it should be considered an acid only after it had come in contact with water. A modified definition of bases also seemed to be required that would be applicable to nonaqueous solutions, in which the anion (negatively charged ion) produced is not the hydroxide ion, as is the case in water, but varies from solvent to solvent, the methoxide ion (CH3O) acting as the basic anion in methanol (CH3OH), for example, and the amide ion (NH2) playing the same role in liquid ammonia (NH3). Even with acids the solvent is involved, since there is much evidence to show that the so-called hydrogen ion in solution does not exist as H+ but always contains at least one molecule of solvent, as H3O+ in water, CH3OH2+ in methanol, and NH4+ in liquid ammonia. These considerations led to the development of definitions of acids and bases that depended on the solvent (see below Alternative definitions (acid–base reaction)). In spite of this change, however, the difficulty still remained that typical acid–base properties, such as neutralization, indicator (vegetable-dye) effects, and catalysis, often took place in solvents such as benzene or chloroform in which free ions could barely be detected at all (by conductivity measurements). Even for aqueous solutions a particular ambiguity arises in the definition of bases, some of which (for example, metallic hydroxides) contain a hydroxyl group, whereas others (such as amines) do not. The latter produce hydroxide ions in solution by reacting with water molecules.

The Brønsted–Lowry (Brønsted–Lowry theory) definition
 In order to resolve the various difficulties in the hydrogen–hydroxide ion definitions of acids and bases, a new, more generalized definition was proposed in 1923 almost simultaneously by J.M. Brønsted (Brønsted, Johannes Nicolaus) and T.M. Lowry. Although the pursuit of exact verbal definitions of qualitative concepts is usually not profitable in physical science, the Brønsted–Lowry definition of acids and bases has had far-reaching consequences in the understanding of a wide range of phenomena and in the stimulation of much experimental work. The definition is as follows: an acid is a species having a tendency to lose a proton, and a base is a species having a tendency to gain a proton. The term proton means the species H+ (the nucleus of the hydrogen atom) rather than the actual hydrogen ions that occur in various solutions; the definition is thus independent of the solvent. The use of the word species rather than substance or molecule implies that the terms acid and base are not restricted to uncharged molecules but apply also to positively or negatively charged ions. This extension is one of the important features of the Brønsted–Lowry definition. It can be summarized by the equation A ⇄ B + H+, in which A and B together are a conjugate acid–base pair. In such a pair A must obviously have one more positive charge (or one less negative charge) than B, but there is no other restriction on the sign or magnitude of the charges. Several examples of conjugate acid–base pairs are given in Table 1—>.

      A number of points about the Brønsted–Lowry definition should be emphasized:

      1. As mentioned above, this definition is independent of the solvent. The ions derived from the solvent (H3O+ and OH in water and NH4+ and NH2 in liquid ammonia) are not accorded any special status but appear as examples of acids or bases in terms of the general definition. On the other hand, of course, they will be particularly important species for reactions in the solvent to which they relate.

 2. In addition to the familiar molecular acids, two classes of ionic acids emerge from the new definition. The first comprises anions derived from acids containing more than one acidic hydrogen—e.g., the bisulfate ion (HSO4, as in Table 1—>) and primary and secondary phosphate ions (H2PO4 and HPO42−) derived from phosphoric acid (H3PO4). The second and more interesting class consists of positively charged ions (cations (cation)), such as the ammonium ion (NH4+), which can be derived by the addition of a proton to a molecular base, in this case ammonia (NH3). The hydronium ion (H3O+), which is the hydrogen ion in aqueous solution, also belongs to this class. The charge of these ionic acids, of course, always must be balanced by ions of opposite charges, but these oppositely charged ions usually are irrelevant to the acid–base properties of the system. For example, if sodium bisulfate (Na+HSO4) or ammonium chloride (NH4+Cl) is used as an acid, the sodium ion (Na+) and the chloride ion (Cl) contribute nothing to the acidic properties and could equally well be replaced by other ions, such as potassium (K+) and perchlorate (ClO4), respectively.

      3. Molecules such as ammonia and organic amines are bases by virtue of their tendency to accept a proton. With metallic hydroxides such as sodium hydroxide, on the other hand, the basic properties are due to the hydroxide ion itself, the sodium ion serving merely to preserve electrical neutrality. Moreover, not only the hydroxide ion but also the anions of other weak acids (for example, the acetate ion) must be classed as bases because of their tendency to reform the acid by accepting a proton. Formally, the anion of any acid might be regarded as a base, but for the anion of a very strong acid (the chloride ion, for example) the tendency to accept a proton is so weak that its basic properties are insignificant and it is inappropriate to describe it as a base. Similarly, all hydrogen compounds could formally be defined as acids, but in many of them (for example, most hydrocarbons, such as methane, CH4) the tendency to lose a proton is so small that the term acid would not normally be applied to them.

 4. Some species, including molecules as well as ions, possess both acidic and basic properties; such materials are said to be amphoteric (amphoterism). As shown in Table 1—>, both water and ammonia are amphoteric, a situation that can be represented by the schemes H3O+–H2O–OH and NH4+–NH3–NH2. Another example is the secondary phosphate ion, HPO42−, which can either lose or accept a proton, according to the following equations: HPO42− ⇄ PO43− + H+ and HPO42− + H+ ⇄ H2PO4. The amphoteric properties of water are particularly important in determining its properties as a solvent for acid–base reactions.

      5. The equation A ⇄ B + H+, used in the Brønsted–Lowry definition, does not represent a reaction that can be observed in practice, since the free proton, H+, can be observed only in gaseous systems at low pressures. In solution, the proton always is attached to some other species, commonly a solvent molecule. Thus in water the ion H3O+ consists of a proton bound to a water molecule. For this reason all observable acid–base reactions in solution are combined in pairs, with the result that they are of the form A1 + B2 ⇄ B1 + A2. The fact that the process A ⇄ B + H+ cannot be observed does not imply any serious inadequacy of the definition. A similar situation exists with the definitions of oxidizing and reducing agents, which are defined respectively as species having a tendency to gain or lose electrons, even though one of these reactions never occurs alone and free electrons are never detected in solution (any more than free protons are).

Alternative definitions
      Although the Brønsted–Lowry concept of acids and bases as donors and acceptors of protons is still the most generally accepted one, other definitions are often encountered. Certain of these are adapted for special situations only, but the most important of these other definitions is in some respects more general than the Brønsted–Lowry definition. This definition was first proposed by the American chemist Gilbert N. Lewis (Lewis, Gilbert N.) in 1923.

      According to Lewis, an acid is a species that can accept an electron pair from a base with the formation of a chemical bond composed of a shared electron pair ( covalent bond). This classification includes as bases the same species covered by the Brønsted–Lowry definition, since a molecule or ion that can accept a proton does so because it has one or more unshared pairs of electrons, and therefore it also can combine with electron acceptors other than the proton. On the other hand, the typical Lewis acids need not (and usually do not) contain protons, being species with outer electron shells that are capable of expansion, such as boron trifluoride (BF3), sulfur trioxide (SO3), and silver ion (Ag+). Lewis originally based his ideas on the experimental fact that these nonprotonic acids often exhibit the properties regarded as typical of acids, such as neutralization of bases, action on indicators, and catalysis. Such substances often are electron acceptors, but this is not always the case; carbon dioxide (CO2) and nitrogen pentoxide (N2O5), for example, contain completed octets of electrons and, according to usual valence theory, cannot accept any more. In addition, hydrogen-containing substances that have always been regarded as acids (acetic acid, for example) are not obviously electron acceptors, being rather adducts of the proton (a true Lewis acid) and a base such as the acetate ion. They can only be brought logically into the Lewis scheme by appealing to the fact that the reaction between a proton acid, which may be designated as XH, and a base, denoted by B, passes through an intermediate hydrogen-bonded state, X-H . . . B (in which the dotted line indicates a hydrogen bond, a relatively weak secondary attractive force).

      Numerous lengthy polemical exchanges have taken place regarding the relative merits of the Brønsted–Lowry and Lewis definitions. The difference is essentially one of nomenclature and has little scientific content. In the remainder of this article the term acid is used to denote a proton donor (following the Brønsted–Lowry terminology), whereas the term Lewis acid is employed exclusively to refer to electron-pair acceptors. This choice is based partly on the logical difficulties mentioned in the last paragraph and partly on the fact (see below Acid–base equilibria (acid–base reaction)) that the quantitative description of acid–base reactions is much simpler when it is confined to proton acids. It also represents the commonest usage of the terms.

      The definition of Lewis acids and bases in terms of the gain or loss of electrons should not be confused with the definition of oxidizing and reducing agents in similar terms. In oxidation–reduction reactions (oxidation–reduction reaction) one or more electrons are transferred completely from the reducing agent to the oxidizing agent, whereas in a Lewis acid–base reaction an electron pair on the base is used to form a covalent link with the acid.

      Certain other acid–base definitions have been based upon reactions occurring in specific solvent systems. For proton acids in amphoteric solvents these are equivalent to the Brønsted–Lowry definition. It is sometimes convenient to have general terms for the cation and anion derived from the solvent molecule by the addition and removal of a proton, respectively. The terms lyonium and lyate ions are occasionally used in this way. In water, the lyonium and lyate ions are H3O+ and OH; in ethanol, C2H5OH2+ and C2H5O; and in liquid ammonia, NH4+ and NH2. For a given solvent, an acid can then be defined as a substance that increases the lyonium ion concentration (and correspondingly decreases the lyate ion concentration), whereas a base increases the lyate ion concentration (and decreases the lyonium ion concentration). This kind of definition, to be sure, really does not add anything to the concept of acids and bases as proton donors and proton acceptors.

      The idea that an acid is a solute that gives rise to cations characteristic of the solvent and that a base is a solute that gives rise to anions characteristic of the solvent has sometimes been extended to solvents where no protons are involved at all—for example, liquid sulfur dioxide, SO2. In this example, the solvent is supposed to ionize according to the equation 2SO2 ⇄ SO2+ + SO32−. Thionyl chloride, regarded as SO2+ + 2Cl, then can be considered an acid, and potassium sulfite, 2K+ + SO32−, can be considered a base. The species SO2+ and SO32− can certainly be regarded as Lewis acids and bases, but it is doubtful that they exist to any appreciable extent in liquid sulfur dioxide, a situation that makes the discussion somewhat artificial. Although this view of acids and bases has been useful in stimulating work in unusual types of solvent (for example, in carbonyl chloride, selenium oxychloride, antimony trichloride, and hydrogen cyanide), it has not met with general acceptance.

Acid–base reactions

Proton-transfers
      As already mentioned (The Brønsted–Lowry definition (acid–base reaction)), the reaction expressed by the Brønsted–Lowry definition, A ⇄ B + H+, does not actually occur in any solution processes. This is because H+, the bare proton, has an enormous tendency to add to almost all chemical species and cannot exist in any detectable concentrations except in a high vacuum. Apart from any specific chemical interaction, the very small size of the proton (about 10−15 metre) means that it exerts an extremely powerful electric field, which will polarize and therefore attract any molecule or ion it comes into contact with. It has been estimated that the dissociation of 19 grams of the hydronium ion H3O+ to give 1 gram of protons and 18 grams of water would require the expenditure of about 1,200,000 joules (290,000 calories) of energy, and thus it is an extremely unlikely process indeed.

      Typical acid–base reactions may be thought of as the combination of two reaction schemes, A1 ⇄ B1 + H+ and H+ + B2 ⇄ A2, leading to the combined form A1 + B2 ⇄ B1 + A2. This represents a proton-transfer reaction from A1 to B2, producing B1 and A2. A large number of reactions in solution, often referred to under a variety of names, can be represented in this way. This is illustrated by the following examples, in each of which the species are written in the order A1, B2, B1, A2.

Dissociation of molecular acids in water
      In this instance, water acts as a base. The equation for the dissociation of acetic acid, for example, is CH3CO2H + H2O ⇄ CH3CO2 + H3O+.

Dissociation of bases in water
      In this case, the water molecule acts as an acid and adds a proton to the base. An example, using ammonia as the base, is H2O + NH3 ⇄ OH + NH4+. Older formulations would have written the left-hand side of the equation as ammonium hydroxide, NH4OH, but it is not now believed that this species exists, except as a weak, hydrogen-bonded complex.

Dissociation of acids and bases in nonaqueous solvents
      These situations are entirely analogous to the comparable reactions in water. For example, the dissociation of acetic acid in methanol may be written as CH3CO2H + CH3OH ⇄ CH3CO2 + CH3OH and the dissociation of ammonia in the same solvent as CH3OH + NH3 ⇄ CH3O + NH4+.

Self-dissociation of amphoteric solvents
      In this case, one solvent molecule acts as an acid and another as a base. Self-dissociation of water and liquid ammonia may be given as examples:

      For a strong acid and a strong base in water, the neutralization reaction is between hydrogen and hydroxide ions—i.e., H3O+ + OH ⇄ 2H2O. For a weak acid and a weak base, neutralization is more appropriately considered to involve direct proton transfer from the acid to the base. For example, the neutralization of acetic acid by ammonia may be written as CH3CO2H + NH3 → CH3CO2 + NH4+. This equation does not involve the solvent; it therefore also represents the process of neutralization in an inert solvent, such as benzene, or in the complete absence of a solvent. (If one of the reactants is present in large excess, the reaction is more appropriately described as the dissociation of acetic acid in liquid ammonia or of ammonia in glacial acetic acid.)

hydrolysis of salts
      Many salts give aqueous solutions with acidic or basic properties. This is termed hydrolysis, and the explanation of hydrolysis reactions in classical acid–base terms was somewhat involved. In terms of the Brønsted–Lowry concept, however, hydrolysis appears to be a natural consequence of the acidic properties of cations derived from weak bases and the basic properties of anions derived from weak acids. For example, hydrolysis of aqueous solutions of ammonium chloride and of sodium acetate is represented by the following equations:

      The sodium and chloride ions take no part in the reaction and could equally well be omitted from the equations.

      The acidity of the solution represented by the first equation is due to the presence of the hydronium ion (H3O+), and the basicity of the second comes from the hydroxide ion (OH). The reverse reactions simply represent, respectively, the neutralization of aqueous ammonia by a strong acid and of aqueous acetic acid by a strong base.

      A superficially different type of hydrolysis occurs in aqueous solutions of salts of some metals, especially those giving multiply charged cations. For example, aluminum, ferric, and chromic salts all give aqueous solutions that are acidic. This behaviour also can be interpreted in terms of proton-transfer reactions if it is remembered that the ions involved are strongly hydrated in solution. In a solution of an aluminum salt, for instance, a proton is transferred from one of the water molecules in the hydration shell to a molecule of solvent water. The resulting hydronium ion (H3O+) accounts for the acidity of the solution:

Reactions of Lewis acids
      In the reaction of a Lewis acid with a base the essential process is the formation of an adduct in which the two species are joined by a covalent bond; proton transfers are not normally involved. If both the Lewis acid and base are uncharged, the resulting bond is termed semipolar or coordinate, as in the reaction of boron trifluoride with ammonia:

      Frequently, however, either or both species bears a charge (most commonly a positive charge on the acid or a negative charge on the base), and the location of charges within the adduct often depends upon the theoretical interpretation of the valences involved. Examples are:

      In another common type of process, one acid or base in an adduct is replaced by another:

      In fact, reactions such as the simple adduct formations above often are formulated more correctly as replacements. For example, if the reaction of boron trifluoride with ammonia is carried out in ether as a solvent, it becomes a replacement reaction:

      Similarly, the reaction of silver ions with ammonia in aqueous solution is better written as a replacement reaction:

      Furthermore, if most covalent molecules are regarded as adducts of (often hypothetical) Lewis acids and bases, an enormous number of reactions can be formulated in the same way. To take a single example, the reaction of methyl chloride with hydroxide ion to give methanol and chloride ion (usually written as CH3Cl + OH → CH3OH + Cl) can be reformulated as replacement of a base in a Lewis acid–base adduct, as follows: (adduct of CH3+ and Cl) + OH → (adduct of CH3+ and OH) + Cl. Opinions differ as to the usefulness of this extremely generalized extension of the Lewis acid–base-adduct concept.

      The reactions of anhydrous oxides (usually solid or molten) to give salts may be regarded as examples of Lewis acid–base-adduct formation. For example, in the reaction of calcium oxide with silica to give calcium silicate, the calcium ions play no essential part in the process, which may be considered therefore to be adduct formation between silica as the acid and oxide ion as the base:

      A great deal of the chemistry of molten-oxide systems can be represented in this way, or in terms of the replacement of one acid by another in an adduct.

Acid–base catalysis
      Acids (including Lewis acids) and bases act as powerful catalysts for a great variety of chemical reactions, in the laboratory, in industry, and in processes occurring in nature. Historically, catalytic action was regarded as one of the essential characteristics of acids, and the parallel occurrence of catalytic action and electrical conductivity was one of the compelling pieces of evidence in establishing the theory of electrolytic dissociation as the basis of acid–base behaviour at the end of the 19th century.

      Acid–base catalysis was originally thought of in terms of a mysterious influence of the acid or base, but it is now generally believed to involve an actual acid–base reaction between the catalyst and the reacting substance, termed the substrate, with the catalyst being regenerated at a later stage of the reaction. Moreover, knowledge of reaction mechanisms is now sufficient to suggest detailed sequences of reactions for many acid- or base-catalysis reactions, most of these sequences being at least plausible and in many instances well established.

      In most acid–base reactions the addition or removal of a proton does not bring about any drastic change in the structure of the molecule or in its stability or reactivity. It is a characteristic of reactions catalyzed by acids or bases, however, that the addition or removal of a proton either makes the substrate unstable, so that it decomposes or rearranges, or that it causes the substrate to become reactive toward some other species present in the system. In cases of rearrangement, the regeneration of the catalyst often involves the removal or addition of a proton at a site other than that at which the initial addition or removal took place. It is not necessary that the substrate in an acid- or base-catalyzed reaction should itself have marked acid–base properties, since even a very small extent of initial acid–base reaction may be enough to bring about the subsequent change.

      Instances of acid–base catalysis are numerous indeed; a few examples are given here, as follows:

isomerization of olefins, acid-catalyzed
      Unsaturated compounds frequently rearrange reversibly under the influence of acids to give products in which the double bond occurs in a new location. The interconversion of 2-butene and 1-butene is shown here:

      Reversible dehydration of alcohols, acid-catalyzed. Under the influence of acids, alcohols generally undergo loss of water to give olefinic products. The dehydration of ethanol (ethyl alcohol) to ethylene occurs as follows:

Keto–enol tautomerism, acid- and base-catalyzed
      Acids and bases both bring about the establishment of an equilibrium (chemical equilibrium) between ketones (ketone) (or aldehydes (aldehyde)) and their enol forms, which contain a hydroxyl group directly attached to a doubly bonded carbon atom:

The interconversion between the two forms is called keto–enol tautomerism. The reaction cannot always be observed directly, since the enol form may not reach measurable concentrations, even at equilibrium, but the highly active enol may be detected by its reaction with various reagents, notably the halogens (bromine, for example). Keto–enol tautomerization of acetone can be brought about by acid or base catalysis, as follows:

Aldol condensation, base-catalyzed
      Self-condensation of aldehydes, the so-called aldol condensation, occurs readily, when catalyzed by bases, to give β-hydroxy aldehydes. The prototype of this reaction is the conversion of acetaldehyde to β-hydroxybutyraldehyde, or aldol. The first step of this reaction is the production of an enolate ion (as in formation of the keto–enol tautomeric mixture), but this anion then reacts with a second molecule of acetaldehyde to give the product as shown below:

      These examples illustrate the importance of acid–base catalysis in organic reactions. The equations have been written in terms of H3O+ and OH as the acid and base catalysts, respectively, and these are certainly the most important catalysts in aqueous solution. For many of these reactions (especially isomerization of olefins and dehydration of alcohols), there is ample evidence that other acids or bases also can act as catalysts. This behaviour is known as general acid–base catalysis. It appears particularly clearly in inert solvents such as benzene, in which catalysis by molecular acids and bases is frequently observed despite the absence of detectable quantities of ions derived from the solvent. Acidic groups, such as sulfonic acid (−SO3H) and carboxylic acid groups (−CO2H), attached to a solid molecular framework (as in some ion exchange resins) also act as heterogeneous catalysts for many chemical reactions.

      The above examples show that proton-transfer processes can play a specific part in reaction mechanisms and, in these and similar instances, it is doubtful whether any uncatalyzed or spontaneous reaction of the same type can take place. Apparent evidence to the contrary can usually be explained by catalysis by solvent molecules or by adventitious acidic or basic impurities.

      Lewis acids can exert a catalytic effect in two different ways. In the first of these they interact with hydrogen-containing compounds present in the system to assist the release of a proton to the substrate. For example, the polymerization of olefins by Lewis acids, such as boron trifluoride (BF3), aluminum chloride (AlCl3), and titanium tetrachloride (TiCl4), is believed to be caused by their interaction with proton acids (for example, traces of water) and the olefin to give a carbonium ion, which then reacts further with more olefin:

      In the second mode of action, the Lewis acid acts directly on the substrate, and by withdrawing electrons converts it into a reactive form. A typical example is the action of catalysts like aluminum chloride on alkyl halides to produce carbonium ions: RCl + AlCl3 → R+ + [AlCl4]. The carbonium ion can then react further with other substances, for example, aromatic hydrocarbons. The same type of catalysis probably occurs with many solid oxide catalysts (for example, aluminosilicates), although it is often difficult to decide whether the catalytic action of these materials is due to centres with a deficiency of electrons or to acidic hydroxyl groups.

Acid–base equilibria
      Certain general principles apply to any solvent with both acidic and basic properties—for example, water, alcohols, ammonia, amines, and acetic acid. Denoting the solvent molecule by SH, proton transfer can give rise to the ions SH2+ and S, sometimes called lyonium and lyate ions, respectively (see above). In the pure solvent these are the only ions present, and they must be present in equal concentrations to preserve electrical neutrality. The equilibrium (chemical equilibrium) involved, therefore, is as follows: 2SH ⇄ SH2+ + S. The equilibrium constant (Ks′) for this reaction (the mathematical quantity that expresses the relationships between the concentrations of the various species present at equilibrium) would normally be given by the equation Ks′ = [SH2+] [S]/[SH]2, in which the square brackets denote the concentrations of the species within the brackets. In a given solvent, however, the concentration of the solvent, [SH], is a large and constant quantity, and it is therefore usual to eliminate this term and express the self-dissociation of the solvent by the equation Ks = [SH2+][S]. In this equation, Ks is termed the ion product or the autoprotolysis constant of the solvent. The concentrations are usually expressed in moles per litre, a mole being the molecular weight of the compound in grams. Since a solvent that is a good proton donor is normally a poor proton acceptor, and vice versa, the degree of ionization is generally low and Ks is usually a small quantity. It is about 10−14 for water at ordinary temperatures, and one of the largest Ks values known is 1.7 × 10−4 for 100 percent sulfuric acid. The above equation applies not only to the pure solvent, but also (with the same value of Ks) to any dilute solutions of acids, bases, or salts in the solvent in question. In these solutions [SH2+] and [S] need not be equal, since the condition of electrical neutrality involves the concentration of other ions as well, and it is obvious from the equation that a high value of [SH2+] must imply a low value of [S] and vice versa.

      If an acid A is added to the solvent SH it will be at least partly converted into the conjugate base B according to the reaction A + SH ⇄ B + SH2+, which would be characterized formally by an equilibrium constant [B][SH2+]/[A] [SH]. Again, however, it is usual to omit the term for the constant concentration of the solvent, [SH], from this expression, and to define a constant Ka by the equation

      which is known as the dissociation constant of the acid A in the solvent SH. Any acid–base reaction A1 + B2 ⇄ B1 + A2 will proceed from left to right almost completely if A1 is a much stronger acid than A2. It is a natural extension of this idea to use the equilibrium constant as a measure of the strength of A1 relative to A2. The dissociation constant is thus (apart from the constant factor [SH], which has been omitted) a measure of the acid strength of A relative to that of the lyonium ion SH2+.

      In some instances reaction goes so completely from left to right that it is not possible to measure the equilibrium constant. A is said then to be a strong acid in the solvent SH; similarly, acids with readily measurable dissociation constants (in practice less than about 0.1) are known as weak acids.

      Similar considerations apply to solutions of bases. The reaction involved in this case is SH + B ⇄ S + A, and the equilibrium constant Kb defined by

      is known as the dissociation constant of the base B. Apart from the omitted constant factor [SH], Kb represents the basic strength of B relative to that of the lyate ion S. Bases are termed strong and weak in the same way that acids are.

      The values of Ka and Kb for a conjugate acid–base pair A–B in a given solvent are not independent, since consideration of the dissociation constants of the solvent, acid, and base show that KaKb = [SH2+][S] = Ks in which Ks is the ion product of the solvent. It is therefore unnecessary to specify both Ka and Kb, and it has become common practice to characterize an acid–base pair by Ka only, which may be termed the acidity constant of A–B in the solvent SH. If the value of Kb is required it is readily obtained from Ka and Ks. Since readily accessible values of Ka are always much less than unity, it is often convenient to introduce a quantity pKa, sometimes called the acidity exponent, and defined by the relation pKa = −log10Ka. Values of pKa are generally of a more convenient magnitude.

      The above expressions for the various equilibrium constants depend only on the concentrations of the species concerned, which are tacitly assumed to exist in solution independently of one another. This is not always the case, and in exact treatments of these equilibria two modifications are frequently necessary. In the first place, some or all of the reacting species are ions and, because of the electrical forces between them, the law expressing their concentrations at equilibrium is not always valid. Corrections may be applied by multiplying the concentrations by certain factors called activity coefficients (activity coefficient), the values of which can be calculated theoretically or derived from other measurements. Furthermore, ions of opposite charge may attract one another so strongly that they no longer exist independently but are partly present as ion pairs (ion pair), thus altering the forms of the equilibrium equations. For many purposes, however, the simple equations given here are adequate, especially with regard to reactions in aqueous solutions.

Aqueous solutions
      Since aqueous solutions are of particular importance in the laboratory and in the physiology of animals and plants, it is appropriate to consider them separately. The ion product of water, Kw = [H3O+] [OH], has the value 1.0 × 10−14 mole2litre−2 at 25° C, but it is strongly temperature-dependent, becoming 1.0 × 10−15 at 0° C and 7 × 10−13 at 100° C. In principle the value of Kw can be determined by measuring the electrical conductance of very pure water, in which [H3O+] = [OH] = 10−7 at 25° C, but in practice it is derived from other measurements—for example, measurements of the degree of hydrolysis of salts.

      For an uncharged acid, in this example acetic, the dissociation constant is given by the following expression:

      For acetic acid, Ka has the value 1.76 × 10−5 at 25° C. The dissociation constant may be expressed in terms of the degree of dissociation of the acid. This quantity, represented by the Greek letter alpha, α, is equal to the fraction of the acid that appears in dissociated form—in this case as the ions CH3CO2 and H3O+. If the initial concentration of acid is designated by c, then the concentrations of the ions are each equal to αc, or [H3O+] = [CH3CO2] = αc, and the concentration of undissociated acid is equal to c(1 − α), or [CH3CO2H] = c(1 − α).

      Substituting these expressions into the equation giving the value of the dissociation constant gives

. From this equation it can be inferred that the degree of dissociation (α) increases with decreasing concentration (c). For small degrees of dissociation (α < < 1), the equation becomes
; whereas, at sufficiently low concentrations (c < < 1), α tends to unity (α → 1).

      Discussions exactly analogous to this apply to a number of other acid–base equilibria—for example, (1) the dissociation of ammonia in water, (2) the hydrolysis of ammonium salts, and (3) the hydrolysis of an acetate.

      For reaction (1—>) α is the degree of dissociation of ammonia, and the dissociation constant is Kb, the basic dissociation constant. In reaction (2—>), the hydrolysis of an ammonium salt (for example, ammonium chloride), α would be termed the degree of hydrolysis and K the hydrolysis constant. In terms of the general definition of acids and bases, however, K could equally be called the acidity constant for the acid–base pair NH4+–NH3, and this is a more rational way of describing the process. Finally, reaction (3—>) represents the hydrolysis of an acetate (for example, sodium acetate); the resulting equilibrium constant is termed the hydrolysis constant and can be seen to equal Kw/Ka, where Kw is the ion product of water and Ka the acidity constant for the acid–base pair CH3CO2H–CH3CO2 (i.e., the dissociation constant of acetic acid). The investigation of equilibria such as this is, in fact, one of the methods for determining the value of Kw (see above).

      The equilibria considered so far arise when one component of an acid–base pair is dissolved in water—if necessary, along with an ion, such as Na+ or Cl, having negligible acid–base properties. The direct consequence of this is that the two new species produced (for example, those on the right-hand sides of the equations [1–3—>] above) have equal concentrations (αc), and hence the previously given equation

is applicable.

      A solution of a more generally useful type can be obtained by deliberately varying the proportions of acid and base present; such a solution is called a buffered (buffer) solution or, somewhat more colloquially, a buffer. A buffered solution containing various concentrations of acetic acid and acetate ion, for example, can be prepared by mixing solutions of acetic acid and sodium acetate, by partially neutralizing a solution of acetic acid with sodium hydroxide, or by adding less than one equivalent of a strong acid to a solution of sodium acetate. Similarly, a buffer based on the pair NH4+–NH3 can be prepared by mixing solutions of ammonia and an ammonium salt, by partially neutralizing a solution of ammonia with a strong acid, or by adding less than one equivalent of sodium hydroxide to a solution of an ammonium salt. The hydrogen ion concentration in a buffer solution is, of course, still given by the usual equation, which is conveniently written as

      Since hydrogen ion concentrations are usually less than unity and cover an extremely wide range, it is often convenient to use instead the negative logarithm of the actual concentration, a figure that varies usually only in the range 1–13. This figure is termed the pH, and its definition is expressed by the equation pH = − log10[H3O+]. For example, in pure water [H3O+] = 1 × 10−7, with the result that the pH = 7.0. The same term can be applied to alkaline solutions; thus, in 0.1 molar sodium hydroxide [OH] = 0.1, [H3O+] = Kw/[OH] = 1 × 1014/0.1 = 10−13, and pH = 13.0.

      Applying the pH concept to buffered solutions gives the following equation:

known as the buffer ratio, can be calculated from the way in which the solution is prepared. According to this equation, the pH of the buffered solution depends only on the pKa of the acid and on the buffer ratio. Most particularly it does not depend on the actual concentrations of A and B. Therefore, the pH of a buffered solution is little affected by dilution of the solution. It is also insensitive to the addition of acid or alkali, provided that the amounts added are much smaller than both [A] and [B]. This so-called buffering action will be impaired if either [A] or [B] becomes too small; hence, buffer ratios must not deviate too far from unity, and the effective buffering range of a given acid–base system is roughly from pH = pKa + 1 to pH = pKa − 1, corresponding to buffer ratios from 0.1 to 10.

   Figure 1—> shows the relation between pH and composition for a number of commonly used buffer systems. Effective buffer action is confined to the central, steep portion of each curve, where the pH is least sensitive to the composition. Figure 1—> shows that an acid bearing several acidic hydrogens, such as phosphoric acid, can be used to prepare buffer solutions in several different pH ranges. Buffer action plays an important part in controlling the pH of many biological fluids; for example, the pH of the blood is controlled at about 7.4 by the carbonic acid–bicarbonate system shown in Figure 1—>. Buffers are widely used to control the pH in chemical or biological experiments. For the latter, the system H2PO4–HPO42− is particularly useful, being effective in the physiological pH range, 6–8.

      The same principles can be applied for the quantitative treatment of systems containing larger numbers of acid–base pairs; for example, in an aqueous solution of ammonium acetate, the following acid–base pairs must be considered: NH4+–NH3, CH3CO2H–CH3CO2, H3O–H2O, and H2O–OH. The situation is much more complicated in many solutions that are important in industry or in nature, but it is always possible to make a complete prediction of the state of the system in terms of the acidity constants Ka of each acid–base pair (provided, of course, that reactions other than proton transfers do not interfere).

Nonaqueous solvents
      Although acid–base properties have been investigated most thoroughly in aqueous solutions, partly because of their practical importance, water is in many respects an abnormal solvent. In particular, it has a higher dielectric constant (a measure of the ability of the medium to reduce the force between two electric charges) than most other liquids, and it is able itself to act either as an acid or as a base. The behaviour of acids and bases in several other solvents will be described briefly here.

      The effect of the solvent on the dissociation of acids or bases depends largely upon the basic or acidic properties of the solvent, respectively. Since many acid–base reactions involve an increase or decrease in the number of ions, they are also influenced by the dielectric constant of the solvent, for a higher dielectric constant favours the formation of ions. Finally, the specific solvation (or close association with the solvent) of particular ions (excluding the solvation of the proton to give SH2+, which is already included in the basicity of the solvent) may be important. It is usually not easy to separate these three effects and, in particular, the effects of dielectric constant and solvation merge into one another. These points are illustrated with examples of several of the more important solvents. In this discussion the solvents are classified as amphoteric (amphoterism) (both acidic and basic), acidic (in which the acidic properties are much more prominent than the basic), basic (in which the reverse is true), and aprotic (in which both acidic and basic properties are almost entirely absent). Finally, concentrated aqueous acids are mentioned as an example—a particularly important one—of mixed solvents.

Amphoteric solvents
      The most important nonaqueous solvents of this class are the lower alcohols methanol and ethanol (ethyl alcohol). They resemble water in their acid–base properties but, because of their lower dielectric constants, facilitate processes producing ions to a much smaller extent. In particular, the ion products of these solvents are much smaller (Ks = 10−17 for CH3OH and 10−19 for C2H5OH, compared with 10−14 for water), and the dissociation constants of molecular acids and bases are uniformly lower than in water by four to five powers of 10. Nitric acid, for example, which is almost completely dissociated in water (Ka about 20), has Ka = 2.5 × 10−4 in methanol. On the other hand, the equilibrium constants of processes such as NH4+ + ROH ⇄ NH3 + ROH2+ and CH3CO2 + ROH ⇄ CH3CO2H + RO are similar in all three solvents, since they do not involve any change in the number of ions.

Acidic solvents
      The most important strongly acidic solvent is sulfuric acid, which is able to protonate a wide variety of compounds containing oxygen or nitrogen. Thus, water, alcohols, ethers, ketones, nitro compounds, and sulfones all act as bases in sulfuric acid. This solvent must also possess some basic properties, because its ionic product is high ([H3SO4+] [HSO4] = 1.7 × 10−4), but the basicity of the solvent is obscured normally by its very high acidity. For example, carboxylic acids behave as strong bases in sulfuric acid, reacting almost completely according to the equation RCO2H + H2SO4 → RCO2H2+ + HSO4. Many substances undergo reactions in sulfuric acid that are more complicated than simple proton transfers, often yielding species important because of their chemical reactivity. Thus, some alcohols produce carbonium ions in sulfuric acid; with triphenylcarbinol, for example, the reaction is (C6H5)3COH + 2H2SO4 → (C6H5)3 C+ + H3O+ + 2HSO4. Nitric acid gives the nitronium ion, NO2+, according to the equation HNO3 + 2H2SO4 → NO2+ + H3O+ + 2HSO4. This ion frequently is the active agent in the nitration of organic compounds. Hydrogen fluoride has solvent properties resembling those of sulfuric acid but is less acidic and has negligible basic properties. acetic acid is another acidic solvent that has been extensively studied. Because of its low dielectric constant, ions exist in it largely in the form of ion pairs, and more complex associates are frequently formed. For this reason a quantitative interpretation of acid–base equilibria in acetic acid is often difficult, but some general conclusions can be drawn. In particular, it can be seen that all substances more basic in water solution than aniline react completely with acetic acid according to the equation B + CH3CO2H → BH+ + CH3CO2. All such bases therefore give solutions with indistinguishable acid–base properties; this is often referred to as a levelling effect of the solvent. The converse is true for acids; for example, the strong mineral acids, nitric, hydrochloric, sulfuric, hydrobromic, and perchloric (HNO3, HCl, H2SO4, HBr, and HClO4) are “levelled” in aqueous solution by complete conversion to the hydronium ion, but in acetic acid they are differentiated as weak acids with strengths in the approximate ratio 1:9:30:160:400.

Basic solvents
      The only basic solvent that has been investigated in any detail is liquid ammonia, which has the very low ion product [NH4+] [NH2] = 10−33. As might be expected, this solvent has a marked levelling effect upon acids; thus, for example, acetic, benzoic, nitric, and hydrochloric acids all give solutions with identical acidic properties, owing to the ion NH4+, although, of course, in water they behave very differently.

Aprotic solvents
      Strictly aprotic solvents include the hydrocarbons and their halogen derivatives, which undergo no reaction with added acids or bases. Acid–base equilibrium in these solvents can be investigated only when a second acid–base system is added; the usual reaction A1 + B2 ⇄ B1 + A2 then takes place. Most such investigations have employed an indicator as one of the reacting systems, but the results are often difficult to interpret because of association of both ions and molecules in these media of low dielectric constant.

      The term aprotic has been extended recently to include solvents that are unable to lose a proton, although they may have weakly basic properties. Some of these aprotic solvents have high dielectric constants (for example, N, N-dimethylformamide, dimethyl sulfoxide, and nitrobenzene) and are good solvents for a variety of substances. They have a powerful differentiating effect on the properties of acids and bases. In particular, basic anions are poorly solvated in these solvents and thus behave as very strong bases; for example, it has been estimated that sodium methoxide dissolved in dimethyl sulfoxide gives a solution 109 times as basic as in methanol.

Concentrated aqueous acids
      Dilute solutions of strong acids—for example, hydrochloric, sulfuric, and perchloric (HCl, H2SO4, HClO4)—in water behave essentially as solutions of the ion H3O+, and their acidity increases in proportion to their concentration. At concentrations greater than about one molar (that is, one mole of acid per litre of solution), however, the acidity, as measured by action on indicators or by catalytic ability, increases much more rapidly than the concentration. For example, a 10 molar solution of any strong acid is about 1,000 times as acidic as a 1 molar solution. This behaviour is undoubtedly largely due to the depletion of water with increasing concentration of acid; the hydronium ion, H3O+, is known to have a strong tendency to further hydration, probably mainly to the ion H3O+(H2O)3 (that is, H9O4+), and a decrease in water content increases the proton-donating power of the solution. The acidity of these concentrated solutions is commonly measured by the acidity function, H0, a quantity measured by the effect of the solvent on a basic indicator I. It is defined by H0 + pKih+ − log10 [IH+]/[I] and becomes equal to the pH in dilute solution. The acidity function H0 frequently is found to be independent of the nature of the indicator and to give an approximate measure of the catalytic power of the acid solution. Mixtures of sulfuric acid and water ranging from 10 to 100 percent sulfuric acid have H0 values between −0.3 and −11.1, which corresponds to an acidity range of nearly 11 powers of 10.

Lewis acids
      Much less information is available about Lewis acid–base equilibria than about ordinary acid–base equilibria (chemical equilibrium), but it is clear that the situation is less simple for the former than for the latter. When a given Lewis acid reacts with a series of similarly constituted bases the equilibrium constants often vary in parallel with the conventional basic strengths. This is the case when a zinc halide, ZnX2, for example, reacts with a series of amines. In general, however, it is not possible to arrange Lewis acids and bases in a unique order that will predict the extent to which a given pair will react. Thus, although the hydroxide ion (OH) is always a much stronger base than ammonia (NH3) in reactions with proton acids, in reactions with the Lewis acid Ag+, the complex Ag(NH3)2+ is fairly stable, whereas AgOH is completely dissociated. Similarly, for some metal cations complex formation increases in the order fluoride < chloride < bromide < iodide, whereas for other metal cations the order is the reverse of this.

      This kind of behaviour has led to a classification of Lewis acids and bases into “hard” and “soft” categories; as a rule, hard acids react preferentially with hard bases and, similarly, soft acids react with soft bases. The terms hard and soft are chosen to suggest that the atomic structures associated with hard acids and bases are rigid and impenetrable, whereas those associated with soft acids and bases are more readily deformable. Hard acids include the proton; sodium, calcium, and aluminum ions; and carbonium ions. The soft acids include cuprous, silver, mercurous, and the halogen cations. Typical soft bases are iodide, thiocyanate, sulfide, and triphenylphosphine; whereas hard bases include hydroxide, fluoride, and many oxyanions. The dividing line between the hard and soft categories is not a sharp one, and its theoretical interpretation is obscure. Nevertheless, a surprising amount of factual information can be coordinated on the basis of preferential reactions of hard acids with hard bases and soft acids with soft bases.

The effect of molecular structure
      Regularities and trends in the properties of the elements are best understood in terms of the periodic table, an orderly pattern seen when the elements are arranged in order of increasing atomic number. Comparing the hydrides (hydride) of the various elements in the table reveals appreciable acidity only among those of the elements on the right-hand side of the table, especially the halogen elements—fluorine, chlorine, bromine, and iodine. This generalization is borne out when the elements across the first period of the table are examined in order; as the right-hand side of the table is approached, the elements encountered are carbon, nitrogen, oxygen, and fluorine. The hydrides of these elements show increasing acidity. Methane, CH4, a hydride of carbon, has no detectable acidic properties, and the pKa decreases sharply in the series ammonia (NH3), 35; water (H2O), 16; and hydrogen fluoride (HF), 4. In any given group of the periodic table, the acidity of the hydrides increases as the group is descended. For example, the two groups at the right-hand side of the table include, respectively and in descending order, the elements oxygen, sulfur, selenium, and tellurium; and fluorine, chlorine, bromine, and iodine. The pKa's of the hydrides of the first group are as follows: water (H2O), 16; hydrogen sulfide (H2S), 7; hydrogen selenide (H2Se), 4; and hydrogen telluride (H2Te), 3. Similarly, hydrogen fluoride (HF) is a weak acid, whereas hydrogen chloride (HCl), hydrogen bromide (HBr), and hydrogen iodide (HI) are all completely dissociated (are strong acids) in aqueous solution. These trends are due to variations in bond strength, electronegativity (attractive power of the atomic nucleus for electrons), and ionic solvation energy, of which the first is the most important. When a hydride is able to lose two or more protons, the loss of the second is always more difficult because of the increased negative charge on the base—e.g., H2S − HS (pK 7), HS − S2− (pK 15); similarly, NH4+ − NH3 (pK 9.5), NH3 − NH2 (pK 35).

      A simple rule applies to the strengths of the oxyacids (oxyacid), which can be given the general formula XOn(OH)m, in which X is any nonmetal. In these compounds, the pK decreases with increasing n but does not depend significantly upon m. When n = 0 (e.g., ClOH, Si(OH)4), pKa is between 8 and 11; when n = 1 (e.g., HNO2, H2SO3) gives pKa 2–4; whereas with n = 2 or 3 (e.g., H2SO4, HClO4) the acids are completely dissociated in water (pKa < 0). These regularities are probably attributable to the sharing of the negative charge of the anion between n + 1 equivalent oxygen atoms; the more extensive the charge spread, the lower is the energy of the anion and hence the stronger the acid.

      The most important groups of organic acids are the alcohols (alcohol) (including the phenols (phenol)) and the carboxylic acids (carboxylic acid). The simple alcohols are very weak acids (pK 16–19); the phenols are considerably stronger (pK ∼ 10); and the carboxylic acids stronger still (pK ∼ 5). The strength of the carboxylic acids is due to the sharing of the negative charge between two equivalent oxygen atoms in the ion RCO2. The most important organic bases are the amines (amine), RNH2, R2NH, or R3N. Most of these are stronger bases than ammonia; i.e., their cations are weaker acids than the ammonium ion.

      The effect of substituents on the acid–base properties of organic molecules has been very extensively studied and is one of the main methods of investigating the nature of the electron displacements produced by substitution in these molecules. The simplest classification is into electron-attracting substituents (halogens, carbonyl, nitro, and positively charged groups) and electron-repelling groups (alkyl groups, negatively charged groups). The electron-attracting groups make acids stronger and bases weaker, whereas electron-repelling groups have the opposite effects. There are, however, often more specific electronic effects, especially in aromatic and unsaturated compounds, for which special explanations are needed.

Experimental studies of acidity and basicity

Dissociation constants in aqueous solution
      The classical method for determining the dissociation constant of an acid or a base is to measure the electrical conductivity of solutions of varying concentrations. From these the degree of dissociation (α, see above) can be determined and Ka calculated from the equation

      This method is unsuitable for acids with pK less than 2 because α is then close to unity and the value 1 − α is therefore subject to error. It also is unsuitable for acids of pK > 7 because impurities in the solvent may affect the conductivity or displace the dissociation equilibrium.

      It is often preferable to use a more specific method for determining the concentration of one of the species in the scheme A + H2O ⇄ B + H3O+. For example, a hydrogen electrode (or more commonly a glass electrode, which responds in the same way) together with a reference electrode, commonly the calomel electrode, serves to measure the actual hydrogen ion concentration, or the pH, of the solution. If E is the electromotive force (in volts) observed by the electrode, the equation giving the pH is as follows:

      In this equation, the value of E0 depends on the nature of the reference electrode and is usually obtained by calibration with a solution of known pH. Measurements can be made on aqueous solutions of the acid, in which case [B] = [H3O+], but it is better to use a series of buffer solutions with known ratios [A]/[B], since these are less sensitive to the presence of impurities. Such a series is obtained by successive additions of alkali to a solution of the acid (or of a strong acid to a solution of the base) and the procedure is then often termed a pH titration.

      If A and B have different optical properties—for example, if they differ in colour or in the absorption of ultraviolet light—this property can be used to measure the ratio [A]/[B], commonly by using an instrument called a spectrophotometer. Since [H3O+] must also be known, the commonest procedure is to measure [A]/[B] in a solution made by adding a small quantity of A or B to a standard buffer solution. If A and B do not have convenient optical properties—as is commonly the case—an indicator, that is, an acid–base system that does show a difference in colour in changing from A to B, is used. If a small quantity of indicator Ai–Bi, with acidity constant Ki, is added to a buffer solution A–B, it is easily shown that the following relation holds:

      in which [Ai]/[Bi] is measured spectrophotometrically, and all the other quantities on the right-hand side of the equation are known.

      If accurate values of K are required, it is necessary in all the above methods to take into account the effect of interionic forces upon the equation and the quantities measured. This factor can induce a considerable degree of complexity into the problem.

Selected values of acidity constants
       pKa's of representative acids and basesTable 2 (pKa's of representative acids and bases) contains acidity constants for selected substances. These are listed as acids or bases according to the nature of the uncharged species, but in each case the value given is (pKa) for the acid form (pKa and pKb for a conjugate acid–base pair being related by the equation

      for aqueous solutions at 25° C). For instances in which several values of pKa are given, these relate to successive dissociations; e.g., for phosphoric acid they correspond to dissociations of H3PO4, H2PO4, and HPO42−. All values given refer to aqueous solutions at or near 25° C; parentheses indicate values that have been estimated indirectly or are uncertain for other reasons.

Ronald Percy Bell

Additional Reading
Paul Walden, Salts, Acids, and Bases (1929), chronicles the early development of concepts in this field. R.P. Bell, Acids and Bases, 2nd ed. (1969), is an elementary account, while his The Proton in Chemistry, 2nd ed. (1973), is at a somewhat more advanced level; as is Edward J. King, Acid–Base Equilibria (1965). Also of interest are Edward Caldin and Victor Gold (eds.), Proton-Transfer Reactions (1975); William B. Jensen, The Lewis Acid-Base Concepts (1980); and Adrien Albert and E.P. Serjeant, The Determination of Ionization Constants: A Laboratory Manual, 3rd ed. (1984). Special topics covered in this article are discussed in Ludwig F. Audrieth and Jacob Kleinberg, Non-Aqueous Solvents (1953); J.J. Lagowski (ed.), The Chemistry of Non-Aqueous Solvents, 5 vol. in 6 (1966–78); and Colin H. Rochester, Acidity Functions (1970). Tables of dissociation constants can be found in G. Kortüm, W. Vogel, and K. Andrussow, Dissociation Constants of Organic Acids in Aqueous Solution (1961); D.D. Perrin, Dissociation Constants of Organic Bases in Aqueous Solution (1965), and Ionisation Constants of Inorganic Acids and Bases in Aqueous Solution, 2nd ed. (1982); E.P. Serjeant and Boyd Dempsey (eds.), Ionisation Constants of Organic Acids in Aqueous Solution (1979); and Kosuke Izutsu, Acid-Base Dissociation Constants in Dipolar Aprotic Solvents (1990).Ronald Percy Bell Ed.

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР

Look at other dictionaries:

  • Acid–base reaction — An acid base reaction is a chemical reaction that occurs between an acid and a base. Several concepts exist which provide alternative definitions for the reaction mechanisms involved and their application in solving related problems. Despite… …   Wikipedia

  • Acid-alkali reaction — An acid alkali reaction is a special case of an acid base reaction, where the base used is also an alkali. When an acid reacts with an alkali it forms a metal, salt and water. Acid alkali reactions are also a type of neutralisation reaction.In… …   Wikipedia

  • Acid–base homeostasis — Blood gas, acid base, gas exchange terms PaO2 Arterial oxygen tension, or partial pressure PAO2 Alveolar oxygen tension, or partial pressure PACO2 Arterial carbon dioxide tension, or partial pressure PaCO2 Alveolar carbon dioxide tension, or… …   Wikipedia

  • Acid-base titration — [ frame|100px|frame|Titration setup. The burette would normally be held by a clamp, not shown here. The pink is most likely caused by use of the phenolphthalein indicator.] An acid base titration is a method in chemistry that allows quantitative… …   Wikipedia

  • Brønsted–Lowry acid-base theory — In chemistry, the Brønsted–Lowry acid base theory is an acid base theory. It was proposed independently by Johannes Nicolaus Brønsted and Thomas Martin Lowry in 1923. In this system, an acid is defined as any chemical species (molecule or ion)… …   Wikipedia

  • Brønsted–Lowry acid–base theory — In chemistry, the Brønsted–Lowry theory is an acid base theory, proposed independently by Johannes Nicolaus Brønsted and Thomas Martin Lowry in 1923.[1][2] In this system, Brønsted acids and Brønsted bases are defined, by which an acid is a… …   Wikipedia

  • acid–base catalysis — ▪ chemistry       acceleration of a chemical reaction by the addition of an acid or a base, the acid or base itself not being consumed in the reaction. The catalytic reaction may be acid specific (acid catalysis), as in the case of decomposition… …   Universalium

  • Acid dissociation constant — Acetic acid, a weak acid, donates a proton (hydrogen ion, high …   Wikipedia

  • Acid — This article is about acids in chemistry. For the drug, see Lysergic acid diethylamide. For other uses, see Acid (disambiguation). Acidity redirects here. For the novelette, see Acidity (Novelette). Acids and Bases …   Wikipedia

  • Base (chemistry) — For the term in genetics, see base (genetics) Acids and Bases Acid dissociation constant Acid base extraction Acid–base reaction Acid–base titration …   Wikipedia

Share the article and excerpts

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